Sie sind auf Seite 1von 12

Engineering Fracture Mechanics 108 (2013) 251–262

Contents lists available at SciVerse ScienceDirect

Engineering Fracture Mechanics


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

A comparative study of the fatigue behavior of two heat-treated


nodular cast irons
R. Konečná a,⇑, G. Nicoletto b, L. Bubenko a, S. Fintová a
a
University of Žilina, Faculty of Mechanical Engineering, Department of Materials Engineering, Univerzitná 1, 01026 Žilina, Slovak Republic
b
University of Parma, Department of Industrial Engineering, Parco Area delle Scienze, 181/A, 43100 Parma, Italy

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

Keywords: The mechanical strength of nodular cast iron (NCI) can be improved by heat treatment. Iso-
Heat treated ductile iron thermal Ductile Iron (IDI) competes with Austempered Ductile Iron (ADI) for applications
Austempered ductile iron subject to dynamic loading. The preliminary metallographic analysis of ADI and IDI showed
Fatigue
similar graphite nodule characteristics but different metallic matrix structure. An extensive
Fatigue crack growth
mechanical characterization of the two materials determined and compared (i) the high
Microstructure
Heterogeneity cycle fatigue strength, (ii) the fatigue crack growth rates and threshold DKth, (iii) the typical
crack propagation mechanisms, and (iv) the material heterogeneity in terms of local tensile
tests.
Ó 2013 Elsevier Ltd. All rights reserved.

1. Introduction

Nodular cast irons (NCIs) are widely used to cast structural parts when easy castability, high mechanical and fatigue prop-
erties and wear resistance are simultaneously required [1]. NCIs combine the castability of gray cast iron with the high
toughness of cast steels being considerably less expensive than cast steels to produce. The mechanical properties of NCI de-
pend on many factors, such as charge composition, chemical composition, inoculation method, size of graphite nodules, pro-
portion of ferrite in the matrix and many others [1–5]. Typical matrix structures of as-cast NCI parts are characterized by
fully ferrite or pearlite or different ferrite–pearlite content [2–5]. A post-cast heat treatment, such as austempering, can sig-
nificantly improve mechanical and fatigue properties. Austempered Ductile Iron (ADI) represents the top grade of the NCI
family range [6–8]. The conventional process consists of austenization in the temperature range of 871–982 °C for sufficient
time to get a fully austenitic matrix, and then quenching to an austempering temperature range of 240–400 °C to avoid for-
mation of pearlite. The part is maintained at this temperature for 2–4 h depending on the section size [7–9]. In general, sec-
tion sizes greater than 19 mm require addition of alloying elements such as Cu, Ni and Mo [7,8]. Mechanical properties of ADI
depend on austenization and austempering temperatures and times, as-cast structure, chemical composition and section
size, with the austempering temperature as the most important [8]. High austempering temperatures result in high ductility,
high fatigue and impact strengths and relatively low yield and tensile strengths. At low austempering temperatures, ADI dis-
plays high yield and tensile strengths, high wear resistance and lower ductility and impact strength [10–12]. Chemical com-
position of austempered ductile iron is related to hardenability and austemperability of NCI.
The excellent properties of ADI are attributed to its unique microstructure consisting of high carbon residual austenite
and acicular ferrite, i.e. bainitic matrix often called ausferrite [13]. However, for its structural complexity it is difficult to
identify the individual microstructural features that influence the fatigue properties and the fatigue crack initiation mech-
anisms. Austempering to high temperatures produces a coarser ferrite but a lower volume fraction of ferrite. This is accom-

⇑ Corresponding author. Tel.: +421 944097048.


E-mail addresses: radomila.konecna@fstroj.uniza.sk (R. Konečná), gianni.nicoletto@unipr.it (G. Nicoletto).

0013-7944/$ - see front matter Ó 2013 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.engfracmech.2013.04.017
252 R. Konečná et al. / Engineering Fracture Mechanics 108 (2013) 251–262

Nomenclature

A elongation to rupture
ADI Austempered Ductile Iron
AFM atomic force microscopy
B thickness of CT specimen geometry
CT compact tension
E Young’s modulus of elasticity
EC eutectic cell
EF effective ferrite
IDI Isothermal Ductile Iron
FCG fatigue crack growth
HBW Brinell hardness
Ka stress intensity factor amplitude
Kath threshold stress intensity factor amplitude
Kmax maximum stress intensity factor
Kmin minimum stress intensity factor
LOM light optical microscope
Nf number of cycles to failure
NCI nodular cast iron
R fatigue load ratio
Rp0.2 conventional yield stress
Rm tensile strength
Sa stress amplitude
SEM scanning electronic microscope
Sw fatigue limit
W width of CT specimen geometry
DK stress intensity factor range
DKth threshold stress intensity factor ran

panied by lower yield strength. Austempering carried out at lower temperatures produce finer and greater volume fraction of
ferrite and higher yield strength [14–17]. If austempering time is very short, the amount of the austenite transformation is
less than 100% [18].
The fatigue crack initiation behavior in ADI was investigated in [15]. It showed that crack initiation at the microstructural
level mainly occurred at pores, either surface or sub-surface, rather than at graphite nodules. The number of primary initi-
ation sites had also a profound impact on the mechanism of failure and ultimately the lifetime of the specimen. Long crack
fatigue propagation, on the other hand, is the result of a complex mechanism consisting partly of the initiation and growth
backwards of small cracks started at surface irregularities of the graphite nodules. Initiation of these cracks is apparently
activated by the stress concentration produced when the tip of the main crack is at a sufficiently short distance from the
nodule [18]. This microcrack initiation apparently depends on one or more of microstructure, load ratio and stress range
[19]. These small cracks eventually coalesce with the main crack front that continues to grow in the normal way until a
new nodule is reached. Since several nodules can be involved in the growth process at different portions of the crack front,
the average growth rate is affected by the size, shape and distribution of graphite nodules [18]. Decohesion of the interface
between the graphite nodules and matrix is likely to be caused by mechanical property mismatch occurring in the stress field
ahead of the advancing crack, leading to the subsequent initiation of microcracks [19]. In the near-threshold fatigue crack
growth regime, the probability of occurrence of those conditions necessary for the initiation of cracks starting from nearby
nodules is reduced [18]. Furthermore, microstructural features such as the boundaries between prior austenite grains in the
ausferrite (bainite) microstructure were observed to significantly retard (by blocking planar slip along the austenite {1 1 1}
plane) the propagation of a short fatigue crack, owing to the additional requirement to tilt and twist the crack plane for crys-
tallographic propagation into the next grain [20,21].
The cost increase of alloying elements used in the low-strength ADI grades has recently motivated the development of
other heat treated NCI such as so-called Isothermal Ductile Iron (IDI), which is an intermediate grade between the low-
strength ADI and pearlitic NCI [22]. Since IDI is a rather new material compared to ADI, the mechanisms of fatigue crack ini-
tiation and propagation are not yet fully described and understood. Compared with the pearlitic DI grades, IDI combines sim-
ilar strength with higher toughness properties as a result of the isothermal heat treatment, performed after casting of a
special preconditioning of the metal bath. The heat treatment basically consists of heating the DI casting above the critical
temperature followed by cooling at a rate that promotes the pearlite formation. In addition, no alloying elements are added
for the IDI production, thus benefiting in terms of cost, as well as in the technical performance, because the absence of alloy-
ing elements (Mo in particular) implies less segregation and, as a consequence, lower thickness sensitivity. The resulting
R. Konečná et al. / Engineering Fracture Mechanics 108 (2013) 251–262 253

microstructure of IDI predominantly consists of ferrite and pearlite, with different distribution compared to the as-cast DI,
and graphite nodules. The IDI matrix is denominated ‘‘Perferritic’’ [22].
In general, ADI and IDI castings compete for applications characterized by dynamic service loading. Determining the
properties used in fatigue design and understanding the fatigue damage mechanisms is thus necessary.
The aim of this paper is to present the perferritic IDI material and the characterization of its fatigue behavior in compar-
ison with an ADI of equivalent strength. An extensive metallographic analysis of ADI and IDI specimens is preliminarily per-
formed to investigate the base material structure. The mechanical characterization of the materials under study consisted in
(i) cyclic plane bending fatigue tests to determine the role of microstructure and defects on the fatigue crack initiation phase,
(ii) fatigue crack growth tests to investigate crack propagation behavior and crack tip interaction with material structure, and
(iii) tensile tests on multiple specimens extracted from FCG specimens to assess the extent of material homogeneity within a
casting block. After testing, fatigue crack paths were investigate to highlight the preferred crack advance mechanism through
the microstructure. The scanning electron microscope SEM and Atomic Force Microscope AFM were also used to investigate
the fatigue fracture surfaces.

2. Materials and experimental details

The Italian company Zanardi Fonderie, an established producer of NCI and ADI castings, supplied both IDI and ADI 800 in
the form of cast and heat treated NCI blocks. The chemical composition of the materials is presented in Table 1. The main
difference is in the content of alloying elements that makes IDI cost competitive with respect to ADI.
After casting, the different NCI blocks were subjected to a heat treatment according to Zanardi Fonderie internal stan-
dards. ADI treatment consisted in heating up to 900 °C for 90 min then quenching in oil then austempering up to 330 °C
for 90 min in a salt bath. The IDI treatment consisted in austenitizing un unalloyed NCI above critical temperature at
785 °C for 120 min then cooling promoting perlite formation up to 378 °C for 90 min.
Specimens for a variety of mechanical tests (i.e. tensile, fatigue and FCG) and for the metallographic inspection were ma-
chined from heat treated ADI and IDI blocks.

2.1. Microstructure and defect characterization

2.1.1. Microstructures
Preliminarily, an extensive metallographic analysis of polished ADI and IDI specimens was performed on light microscope
NEOPHOT 32.
The microstructure of austempered ductile iron ADI 800, see Fig. 1a, consists of coarse needle-shaped acicular ferrite
(bainitic ferrite) in carbon-enriched retained austenite and nodular graphite particles. Detail of microstructure shows clus-
ters of ferrite needles with the same orientation, Fig. 1b. The IDI microstructure is a mixture of proeutectoid ferrite and per-
lite, see Fig. 2. Eutectic cells of comparable size in both materials were observed at low magnification, Fig. 3. Distribution of
ferrite and pearlite is nonuniform with high content of ferrite in the middle of eutectic cells (ECs), Fig. 4a, and high content of
pearlite on the boundary of eutectic cells, Fig. 4b.
As far as the graphite nodule characterization, inspection of metallographic sections showed that graphite nodule distri-
bution was reasonably random in ADI and IDI, Fig. 2. Additional parameters are proposed for quality control in ductile iron
like size, eccentricity and solidity of the nodules [4]. Here the average size of graphite particles according ASTM A247-10 [23]
was measured using image analysis software NIS Element 5.0 and found to range from 30 to 60 lm corresponding to the
level 6 etalon [23]. Graphite nodule counting was performed using image analysis software too. The nodule count of ADI
was about 75 nodules/mm2. The nodule count of IDI was higher than ADI, i.e. about 90 nodules/mm2. However, both data
were rather low for a NCI where the suggested nodule count is around 150 nodules/mm2 [24]. The graphite fraction in
the volume, i.e. around 12.5%, was similar for ADI and IDI specimens. Graphite globularity (i.e. shape factor) should not
be smaller than 0.75 (where a perfect sphere = 1) because the mechanical properties of NCIs decrease when graphite nodules
show an irregular shape. Nodularity (percentage of nodules with globularity higher than 80%) was determined using image
analysis software.
Small casting defects, such as micro-shrinkages, and inclusions were found predominantly at the boundary of eutectic
cells in both materials, see Fig. 5. This is consistent with the fact that the boundary of a eutectic cell solidifies last. Defect
sizes were generally small but occasionally some of them exceeded 100 lm in size. Examples of such rather large defects
are shown in Fig. 6. They are apparently larger than the graphite nodules and of a sharp, elongated shape. Therefore they
could act as preferred fatigue crack initiation places.

Table 1
Chemical composition of ADI 800 and IDI (in weight%).

Material C Si Mn S P Ni Cu Mg Mo
ADI 800 3.62 2.63 0.66 0.008 0.023 – 0.34 0.042 –
IDI 3.69 2.53 0.17 – – 0.05 0.08 – 0.012
254 R. Konečná et al. / Engineering Fracture Mechanics 108 (2013) 251–262

Fig. 1. (a) Typical microstructure of ADI 800. (b) Detail of ferrite clusters in ADI matrix (etched 5% Nital).

Fig. 2. Typical microstructure of perferritic IDI, etched 5% Nital.

(a) ADI (b) IDI

Fig. 3. Eutectic cells, etched 5% Nital.

2.2. Mechanical testing

2.2.1. Reversed bending fatigue testing


High cycle fatigue tests were performed here on flat ADI and IDI specimens (i.e. geometry in Fig. 7) subjected to planar
four-point bending cycles at 20 Hz frequency. The fatigue testing machine was a Ambrosi RB H01-3D44X equipped with a
load-cell DACELL JMMA-K50. Tests were programmed through a PC-based Virtual Instrument interface and data were mon-
itored and acquired using a NI LabVIEW DAQ card. The stress amplitude Sa applied to the specimen was calculated from the
applied cyclic force monitored by the load cell. The fatigue load ratio R = Smin/Smax was fully reversed (R = 1). When the
specimen broke the number of cycles to failure Nf was recorded. A fatigue test was interrupted if the specimen did not break
within 2  106 cycles.
R. Konečná et al. / Engineering Fracture Mechanics 108 (2013) 251–262 255

(a) middle part of eutectic cell (b) boundary of eutectic cell


Fig. 4. Detail of IDI matrix, etched 5% Nital.

(a) ADI (b) IDI


Fig. 5. Casting defects at the boundary of eutectic cells. Etched 5% Nital.

(a) ADI (b) IDI


Fig. 6. Casting defects (in black) and graphite nodules (gray) against the matrix (white), no etched.

Fig. 7. Geometry of fatigue test specimen.


256 R. Konečná et al. / Engineering Fracture Mechanics 108 (2013) 251–262

2.2.2. Fatigue crack growth testing


Repeated fatigue crack growth experiments were conducted here on three CT specimens (ASTM standard geometry with
thickness B = 10 mm, width W = 50 mm) machined from ADI and IDI blocks, respectively. A chevron starter notch was ma-
chined in the specimens to favor fatigue pre cracking. Fatigue crack propagation experiments were performed according to
the ASTM standard E 647-08 [25], in the electromagnetic resonant testing machine Roell Amsler HFP 5100 at initial cyclic
frequency of 100 Hz.
A sinusoidal waveform with the constant load ratio R = 0.1 was applied. To provide sufficient visibility of fatigue crack
propagation, lateral specimen surfaces were polished by 1-lm-grain-size diamond paste. CCD cameras were used to monitor
the propagating cracks, while the current crack length (a) was measured by digital micrometers and recorded simulta-
neously with number of cycles (N). The fatigue crack propagation curves da/dN vs. Ka where Ka = (Kmax Kmin)/2 were thus
determined [25].
The threshold stress intensity factor amplitude Kath of the different materials was determined here using the load shed-
ding technique [25]. This procedure involves a stepwise reduction of the stress intensity amplitude due to stepwise reduc-
tion of the applied load after the crack has grown through the corresponding plastic zone length at the previous Ka level. The
drop of the stress amplitude in the last steps before the determination of the threshold value was max. 5% of the preceding
value. The crack growth threshold values Kath were then identified as the values of Ka at which the crack growth rate was of
the order of 10 10 m/cycle. At the R-ratio used here, the threshold stress intensity factor amplitude Ka is equal one half of the
threshold stress intensity factor range DKth. Therefore, the results presented in the following sections are readily converted
for comparison to published data.

2.2.3. Fatigue crack paths and surfaces


The light microscope NEOPHOT 32 was used to investigate the fatigue crack paths to highlight the preferred crack ad-
vance mechanism through the etched microstructure. The scanning electron microscope SEM was used to investigate the
fatigue crack surfaces and the mechanisms of fatigue crack growth. The atomic force microscope (AFM) was also used to
map the micro-roughness of the fatigue fracture surface of IDI CT specimens.

2.2.4. Tensile testing


As it will be show in the results section of the paper, the FCG properties showed considerable specimen-to-specimen
dependence at near threshold crack growth rates. Therefore, tensile tests using specimens extracted from the broken CT
halves were performed to investigate possible material variability and other influencing factors. Multiple tensile specimens
(constant cross section 6  3 mm2 and length L0 = 22 mm) were extracted from CT halves. Tensile tests according ASTM E8
standard were performed using a 25 kN MTS 810 testing machine operated at displacement velocity of 0.01 mm/s while
strain was monitored with an extensometer MTS model 632.31F-24.

3. Results and discussion

This section is organized as follows: the high cycle fatigue tests results are presented first in order to identify and discuss
the crack initiation mechanisms, then the FCG test results are presented to analyze the material resistance to long crack
growth.
The section continues with an extensive investigation of the fatigue crack advance mechanisms using LOM and SEM. At
the end of the section, the degree of material homogeneity is quantified using the local tensile tests.

3.1. High cycle fatigue response

In the traditional stress-life approach to fatigue design the number of cycles to initiation of a fatigue crack is assumed to
be approximately equal to the total number of cycle to complete failure Nf. This is appreciably true when the specimen sur-
face is smooth and the number of cycles is high, (i.e. Nf > 5 104 cycles). Fig. 8 shows the S–N data of the present cyclic bending
fatigue tests of ADI and IDI. S–N curves are drawn through the data points as best fit lines. Data points with arrows indicate
samples that did not fail within 2  106 cycles. The estimated fatigue limit under fully reversed loading are Sw = 230 MPa for
ADI and Sw = 170 MPa for IDI. The scatter of each data set is relatively low. Fatigue behavior of ADI is consistently superior to
that of IDI in accordance to the respective ultimate strengths. In a later section this initial observation will be corroborated by
extensive local tensile strength tests.
Comparison of the present data with published information is possible for ADI, which is a well studied material. Fatigue
data about IDI are scarce and due mainly to Zanardi Fonderie internal activity. Rotating bending fatigue results of ADI at dif-
ferent austempering temperature and times was given in [26–28]. It was observed that the fatigue strength increased as the
temperature increased from 230 to 330 °C. However, the fatigue strength decreased with increasing temperature. For similar
treatment conditions as in this case [26] reports fatigue strength of 222 MPa which is good agreement with the present tests.
R. Konečná et al. / Engineering Fracture Mechanics 108 (2013) 251–262 257

Fig. 8. High cycle fatigue tests results for ADI and IDI.

3.2. Fatigue crack growth behavior

3.2.1. FCG rates and threshold SIF amplitude Ka


Differently to the traditional stress-life approach to fatigue design, the fracture mechanics approach assumes that crack-
like defects exist in materials and structures. Use of fatigue crack propagation laws based on stress intensity factor K is then
used to predict the service life of the component. Experimental characterization of fatigue crack growth behavior of a mate-
rial leads to the determination of the relationship between fatigue crack propagation rate da/dN and the stress intensity fac-
tor amplitude Ka. For intermediate crack growth rates, typically in the range of 10 6–10 3 mm/cycle and often called Paris
region, the log (da/dN) vs. log Ka dependence can be approximated by a straight line. At the low Ka values, the crack prop-
agation curve tends to the threshold value of stress intensity factor amplitude Kath. Below this threshold, the fatigue crack
will not propagate.
The log–log plot of Fig. 9 shows the fatigue crack growth data of ADI 800 and of IDI. Table 2 summarizes the threshold
values Kath and the slope of the linear portion of the crack growth rate data of the repeated tests in ADI and IDI specimens.
The data of the three ADI specimens are observed to fall on a single crack growth curve with a narrow scatter band. The aver-
age threshold value of Kath = 3.8 MPa m1/2 (i.e. DKth = 8.9 MPa m1/2).
On the contrary, tests of the three IDI specimens resulted in three distinct fatigue crack growth curves as shown in Fig. 9.
Large differences occur especially in the threshold DK region with Kath ranging from ca. 6 MPa m1/2 (IDI1 specimen),
8 MPa m1/2 (IDI2) and 10 MPa m1/2 (IDI3) (i.e. namely the range covers DKth = 13.3–22 MPa m1/2). Again, the expected behav-
ior that threshold crack propagation condition decreases with increasing tensile strength is confirmed with threshold FCG of
ADI considerably lower that the threshold FCG of IDI.
The FCG behavior of ADI is consistent and in line with expected values from the literature when the ultimate strength of
the present material and the typical limited scatter in this kind of test is considered [28–29]. For the loading condition
R = 0.1, the microstructure influence is negligible. In [3] an increase of the stress ratio when testing ferritic–pearlitic (50–

Fig. 9. Fatigue crack propagation in ADI and IDI.


258 R. Konečná et al. / Engineering Fracture Mechanics 108 (2013) 251–262

Table 2
FCG rate slope and threshold Kath from repeated tests.

Kath (MPa m1/2) Slope


ADI 1 – –
ADI 2 3.82 3.01
ADI 3 3.82 3.03
IDI 1 5.93 3.59
IDI 2 7.68 3.73
IDI 3 9.81 5.54

50%) and the austempered ductile iron was considered to cause an increase in the microstructure influence associated to of
brittle cracking mechanisms superposed to the predominant ductile mechanisms especially at high DK.
The FCG behavior of IDI showing large specimen-to-specimen variation is more complex to explain. Potential contributing
factors could be related to (i) local material heterogeneities that heavily affect crack tip deformation, (ii) the experimental
load shedding methodology that modifies the crack closure pattern, etc. Possible contributions of microstructure and of
crack–microstructure interaction to the substantially different behavior were therefore investigated and are discussed in
the next section.

3.2.2. Fatigue crack paths


The fatigue crack paths in the ADI and IDI CT specimens were investigated with special attention placed at two crack
lengths: (i) that corresponding to near-threshold condition Kath (i.e. da/dN = 10 9 m/cycles) and (ii) intermediate crack
growth rates (i.e. at Ka = 10 MPa m1/2).

(a) ADI 2

(b) IDI 2

(c) detail of crack growth path in ADI 2 (d) detail of crack growth path in IDI 2
9
Fig. 10. Fatigue fracture profiles at da/dN = 10 m/cycles, direction of crack propagation is from the right side, etched 3% Nital, Nomarski.
R. Konečná et al. / Engineering Fracture Mechanics 108 (2013) 251–262 259

Fig. 10 shows the fracture profiles typical for all ADI and of IDI specimens for near-threshold growth rate. The profile is
macroscopically planar in both materials; only when the crack paths locally interacts with graphite nodules deep dimples
with or without graphite nodules can be seen, Fig. 10a and b. Crack tilting due to its interaction with the microstructure re-
duces the driving force and the resulting fracture roughness activates closure mechanisms that also reduce the driving force.
Therefore, crack path roughness cannot explain the difference in fatigue crack growth behavior between ADI and IDI.
At higher magnification, Fig. 10c shows transcrystalline crack propagation in ADI. Because of the typical austempered
structure characterized by the presence of bainitic ferrite clusters, the crack propagated between bainitic ferrite needles ori-
ented parallel to the fracture profile or through clusters of bainitic ferrite needles, having specific orientation. An extensive
discussion of the role of graphite nodules on crack propagation in ADI is reported in [3]. Here the crack paths, adjacent to the
graphite nodules, followed the phase interface between graphite nodules and matrix, Fig. 10c. Locally, when graphite nod-
ules are irregular, the crack propagates locally through graphite. In the case of IDI specimens the fatigue crack growth cor-
responding to the rate of 10 9 m/cycles was characterized by transcrystalline propagation through the ferrite and pearlite
grains. Therefore, the difference in microstructure (i.e. bainite vs. perlite/ferrite) rather than the crack roughness may explain
the higher threshold of Ka of IDI compared to ADI.
When the crack paths at the crack length associated to Ka = 10 MPa m1/2 is examined the following differences are found.
The fatigue fracture profile of ADI 2, Fig. 11a, shows higher roughness compared to the condition da/dN = 10 9 m/cycles,
Fig. 10a because crack interaction with bainitic needles affects the crack growth direction. When the crack reaches a bound-
ary of a needle cluster, it reorients itself in accordance to the direction of the next cluster. On the other hand, secondary
cracks on the fracture profile were observed in ADI specimens.
The fatigue profiles of individual IDI specimens are different and are influenced by the presence of eutectic cells (ECs),
Fig. 11b, c, d. Therefore the fatigue crack propagates through a heterogeneous matrix with local ferrite/perlite distribution
difference. The mechanism of crack propagation depends on the position of the crack whether the pearlite-rich boundary
were crack growth is easier or the ferrite-rich center where crack growth rate is affected by ferrite plasticity. In addition,
in all the cases numerous secondary cracks were found in IDI specimens.
Since these observations of microstructure-fatigue crack path interaction cannot satisfactorily explain the specimen-to-
specimen difference of DKth in IDI shown in Fig. 9, two other types of tests were performed: (i) fracture surface roughness
measurements by Atomic force microscopy (AFM); (ii) local tensile tests.
Here the micro-roughness of the fatigue fracture surface of the two CT specimens of IDI was measured to identify a link
between surface roughness and specimen-to-specimen difference in Kath. Typical experimental output consisting in the 3D
fracture roughness maps of ADI and IDI are shown in Fig. 12. The fracture surface area investigated was about 2  2 mm2.
The average roughness Ra was numerically computed from the maps taken at locations on the crack surfaces associated
respectively to Kath and Ka = 10 MPa m1/2. For IDI1, average roughness Ra = 8.1 lm and 15.4 lm, respectively, while for

(a) specimen ADI 2 (b) specimen IDI 1

(c) specimen IDI 2 (d) specimen IDI 3


Fig. 11. Fatigue fracture profile, Ka = 10 MPa m1/2, direction of crack propagation is from the right side, etched 3% Nital.
260 R. Konečná et al. / Engineering Fracture Mechanics 108 (2013) 251–262

IDI3 Ra = 9.4 lm and 12.2 lm respectively. The micro-roughness is coherent with Kath values and with the crack growth rates
at Ka = 10 MPa m1/2.

3.2.3. Material heterogeneity and tensile response


The variability in the threshold behavior especially of IDI shown in Fig. 9 motivated further investigation of the homoge-
neity of the macroscopic strength of the material by local tensile specimen extraction. Microstructural variables (such as vol-
ume fraction of austenite, carbon content in austenite, ferritic cell size as well as the total carbon content in the matrix) affect

Fig. 12. Typical micro-roughness maps of the fracture surface obtained by AFM.
R. Konečná et al. / Engineering Fracture Mechanics 108 (2013) 251–262 261

Fig. 13. Tensile tests of ADI and IDI specimens.

Table 3
Average mechanical properties for ADI and IDI from tensile tests of Fig. 13.

Rm (MPa) Rp0.2 (MPa) E (GPa) A (%) HBW


ADI 1 801 708 152.4 2.5 270
ADI 2 810 714 141.2 1.9 270
ADI 3 820 730 144.3 2.2 268
IDI 1 640 500 164.0 6.5 223
IDI 2 730 472 149.3 8.1 210
IDI 3 680 425 140.3 7.5 205

fatigue crack growth behavior of ADI in the near threshold region according to [30]. In addition to the general structural char-
acterization reported in a previous section, specimen-to-specimen observations of graphite nodules and ferrite content were
preliminarily performed especially in IDI specimens. Differences in graphite nodule globularity was found in the case of IDI
specimens: in specimens IDI1 and IDI3 only 30% of particles had globularity higher than 80% and in the specimen IDI2, the
graphite particles globularity higher than 80% was observed for 50% of graphite particles. In the case of the specimen ADI1
only 30% of particles had globularity higher than 80%, for the specimen ADI2 40% and for the specimen ADI3 50% of graphite
particles with globularity higher than 80% was observed. The volume fraction of ferrite (effective ferrite, EF), which may also
be correlated to strength and ductility; was evaluated for each specimen using image analysis software (i.e. EF = 29% in IDI 1,
27% in IDI 2 and 34% in IDI 3). Measurement of porosity located on the boundary of EC in pearlitic part of microstructure of
IDI specimens was performed on no etched metallographic surfaces using image analysis software. Such porosity may influ-
ence local condition of crack propagation Results showed differences in total porosity among specimens (i.e. porosity was
0.305% in IDI 1, 0.131% in for IDI 2 and 0.155% in for IDI 3) but even more important, local porosity and especially the size
local of micro-shrinkages was different.
All the stress–strain curves from localized tensile tests of specimens extracted from the three CT specimens of ADI and IDI
are presented in Fig. 13. The average data are summarized in Table 3. They show, on one hand, that ADI has a higher ultimate
strength and lower ductility than IDI. The yield stress of ADI, as well, is significantly higher than that of IDI. Quantitatively,
the tensile strength data of ADI are more scattered (i.e. 660–900 MPa) than that of IDI (i.e. 620–800 MPa). Elongation of ADI
is in the range 2–4% and elongation of IDI 6–9%.
The measured specimen-to-specimen differences in terms of mechanical properties are quite significant in both materi-
als. So, apparently, a unequivocal correlation between the variability in Kath values and variability of static properties is not
found. Nonetheless, the demonstrated heterogeneity within a single CT specimen may explain that the near threshold con-
dition for a fatigue crack may be affected differently in the three CT specimens because of local differences in strength and
ductility. Furthermore the lower yield strength and higher ductility of IDI compared to ADI explains the higher Kath of IDI
compared to ADI as well on cyclic plastic zone size and crack opening displacements. Analogous differences in strength
are found within ADI but apparently the influence in Kath is not remarkable.

4. Conclusions

The fatigue behavior of the perferritic IDI material in comparison with that of ADI of equivalent strength was investigated.
The extensive metallographic analysis of ADI and IDI showed similar graphite nodule characteristics but different metallic
matrix structure. The mechanical characterization of the materials revealed:
262 R. Konečná et al. / Engineering Fracture Mechanics 108 (2013) 251–262

(i) High cycle fatigue strength of ADI is higher than IDI in accordance with a recorded difference in tensile strength.
(ii) Fatigue crack growth and threshold DKth of ADI is significantly lower than IDI. The difference is higher than expected
in the basis of strength considerations.
(iii) The DKth values of IDI as determined using three CT specimens considerably vary while the DKth values of ADI are
constant.
(iv) Crack propagation mechanisms identified studying the crack paths are different in ADI and IDI microstructures.
(v) Fatigue fracture roughness is low at Kath in both materials because the microstructure is very fine. The roughness
increase with applied Ka and is higher in IDI than ADI because of the softer phase of the former.
(vi) Local tensile tests performed on specimens extracted from the CT halves showed a considerable material heterogene-
ity both in IDI and ADI.

Acknowledgement

Part of this research was supported by project VEGA Grant No. 1/0196/12.

References

[1] Davis JR. Cast irons: ASM specialty handbook. 2nd ed. Materials Park (OH): ASM International; 1996.
[2] Konečná R, Hadzimová B, Nicoletto G, Matejka M. Role of ferrite content on the impact strength of nodular cast irons. Mater Engng 2003;X(2):31–38.
[3] Cavallini M, Di Bartolomeo O, Iacoviello F. Fatigue crack propagation damaging micromechanisms in ductile cast irons. Engng Fract Mech
2008;75:694–704.
[4] De Santis A, Di Bartolomeo O, Iacoviello D, Iacoviello F. Quantitative shape evaluation of graphite particles in ductile iron. J Mater Process Technol
2004;196:292–302.
[5] Konečná R, Bujnová P, Nicoletto G. Failure mechanism in ferritic–pearlitic nodular cast iron. Engng Mech 2004;11(5):319–23.
[6] Labrecque C, Gagne M. Ductile iron: fifty years of continuous development. Can Metall Quart 1998;37(5):343–78.
[7] Hayrynen KL. The production of austempered ductile iron (ADI). In: World Conference on ADI Louisville, KY; 2002.
[8] Harding RA. The production, properties and automotive applications of austempered ductile iron. Kovove Materialy 2007;45(1):1–16.
[9] Dorazil E. High strength austempered ductile iron. 2nd ed. Praha: Academia; 1991.
[10] Kovacs BV. Mod Cast 1990;80(3):38–41.
[11] Sohi MH, Ahmadabadi MN, Vahdat AB. J Mater Process Technol 2004;153–154:203–8.
[12] Trudel A, Gagné M. Can Metall Q 1997;36(5):289–98.
[13] Putatunda SK. Development of austempered ductile cast iron (ADI) with simultaneous high yield strength and fracture toughness by novel two-step
austempering process. Mater Sci Engng 2001;315(1–2):70–80.
[14] Gundlach RB, Janowak JF. Met Prog 1985;12:231–6.
[15] Janowak JF, Gundlach RB, Eldis GT, Rohrting K. AFS Int Cast Metal J 1982;6:28–42.
[16] Moore DJ, Noun TB, Rundman KB. AFS Trans 1987;87:165–74.
[17] Bartosiewicz L, Krause AR, Kovacs BV, Putatunda SK. AFS Trans 1992;92:135–42.
[18] Greno GL, Otegui JL, Boeri RE. Mechanisms of fatigue crack growth in austempered cast iron. Int J Fatigue 1999;21(1):35–43.
[19] Strokes B, Gao N, Reed PAS. Mater Sci Engng A 2007;445–446:374–85.
[20] Marrow TJ, Buffiere JY, Withers PJ, Johnson G, Engelberg D. Int J Fatigue 2004;26:717–25.
[21] Marrow TJ, Cetinel H. Fatigue Fract Engng Mater Struct 2000;23:425–34.
[22] Massagia S. The development of ADI and IDI in Italy. Procedia Engng 2010;2(1):1459–76.
[23] ASTM A247–10 Standard test Method for Evaluating the Microstructure, 2010.
[24] Karsay SI, Ductile iron: I. Production. Trenčín. Fompex; 1995 (in Slovak).
[25] ASTM Standard E 647-08: Standard Test Method of Measurement of Fatigue Crack Growth Rates, 2008.
[26] Tayanc M, Aztekin K, Bayram A. Mater Des 2007;28(3):797–803.
[27] Lin CK, Lai PK, Shih TS. Influence of microstructure on the fatigue properties of austempered ductile irons-I. High-cycle fatigue. Int J Fatigue
1996;18:297–307.
[28] Bartosiewicz L, Krause AR, Alberts FA, Singh I, Patunda SK. Influence of microstructure on high-cycle fatigue behavior of austempered cast iron. Mater
Charact 1993;30:221–34.
[29] Greno GL, Pardo EL, Boeri RE. Fatigue of austempered ductile iron. Am Found Soc Trans 1998(1):31–7.
[30] Yang J, Putatunda SK. Near threshold fatigue crack growth behavior of austempered ductile cast iron (ADI) processed by a novel two-step
austempering process. Mater Sci Engng A 2005;393:254–68.

Das könnte Ihnen auch gefallen