Sunteți pe pagina 1din 10

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/225735164

On the fatigue crack growth–microstructure relationship in ultrafine-grained


interstitial-free steel

Article  in  Journal of Materials Science · January 2010


DOI: 10.1007/s10853-010-4511-7

CITATIONS READS

28 472

5 authors, including:

Demircan Canadinc I. Karaman


Koc University Texas A&M University
97 PUBLICATIONS   1,310 CITATIONS    371 PUBLICATIONS   8,012 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Shape memory and strain glass alloys View project

Hydrogen Embrittlement - Understanding and research framework View project

All content following this page was uploaded by Demircan Canadinc on 16 January 2014.

The user has requested enhancement of the downloaded file.


J Mater Sci (2010) 45:4813–4821
DOI 10.1007/s10853-010-4511-7

ULTRAFINE GRAINED MATERIALS

On the fatigue crack growth–microstructure relationship


in ultrafine-grained interstitial-free steel
T. Niendorf • F. Rubitschek • H. J. Maier •

D. Canadinc • I. Karaman

Received: 3 February 2010 / Accepted: 9 April 2010 / Published online: 27 April 2010
Ó Springer Science+Business Media, LLC 2010

Abstract The crack growth behavior in an ultrafine- Motivation and significance


grained (UFG) interstitial-free (IF) steel processed by equal
channel angular pressing (ECAP) was investigated utilizing Recently, ultrafine-grained (UFG) and nano-crystalline
miniaturized compact-tension specimens with different (NC) materials have gained remarkable attention due to
microstructural characteristics. The current results dem- their extraordinary properties [1–5]. By reducing the grain
onstrate that both the ECAP processing route and the size into the submicron regime, the strength of both pure
direction of crack growth with respect to the extrusion metals and alloys can be increased considerably, while the
direction dictate the crack growth behavior in UFG IF ductility of these materials is preserved [2, 6]. The tech-
steel. Specifically, the highest crack growth rates and the niques for achieving NC materials and UFG materials are
lowest threshold values were observed for the lowest grain fundamentally different: ‘‘bottom-up’’ approaches are
size. Moreover, an unusual deviation from the expected employed for achieving NC materials with grain sizes less
direction of crack expansion was observed, where the than 100 nm, while UFG materials with grain sizes ranging
deviation depended on the processing route and direction from 100 nm up to 1 lm are produced by ‘‘top–down’’
of crack growth. This deviation is attributed to the presence techniques [2, 7, 8]. The latter techniques do less suffer
of elongated structures in the microstructure, which were from specific process-induced imperfections, such as
mainly detectable in the UFG IF steel following a small impurities, and the processing of relatively large quantities
number of pressings, and to a smaller extent in the opti- of material is feasible [1, 8]. High pressure torsion (HPT)
mized microstructures. Specifically, these elongated struc- [9], accumulative roll bonding [10], and equal channel
tures formed parallel to the material’s plastic flow during angular pressing (ECAP) [1, 2, 7, 11] are the most com-
ECAP processing and moved the crack away from the monly utilized techniques, which all make use of severe
expected direction of growth due to the high stress con- plastic deformation for achieving the high degree of grain
centration zones they created along with the process- refinement [1, 2]. The homogenously deformed volume
induced damages. fraction of material is the highest upon ECAP processing,
and a good repeatability is achieved regarding the final
grain size and the resulting texture with this process,
T. Niendorf (&)  F. Rubitschek  H. J. Maier making it mostly the method of choice for obtaining UFG
Lehrstuhl für Werkstoffkunde (Materials Science), University microstructures [1, 2, 11].
of Paderborn, 33095 Paderborn, Germany Many of the studies to date focused on the microstruc-
e-mail: niendorf@mail.uni-paderborn.de
tural evolution during ECAP processing and the resulting
D. Canadinc monotonic properties [1, 2, 11], and results are now avail-
Advanced Materials Group, Department of Mechanical able for various materials featuring different microstruc-
Engineering, Koc University, Sariyer, 34450 Istanbul, Turkey tural characteristics, including high-purity face-centered
cubic (fcc) metals [2, 12], hexagonal closed packed (hcp)
I. Karaman
Department of Mechanical Engineering, Texas A&M University, biomaterials [13, 14], and body-centered cubic (bcc) steels
College Station, TX 77843, USA [15, 16]. However, establishing a deep knowledge of the

123
4814 J Mater Sci (2010) 45:4813–4821

materials’ monotonic and cyclic deformation response is Experimental procedures


crucial for any of the envisaged applications, such as
medical implants and light-weight automotive components. The chemical composition of the IF steel investigated in
Consequently, the investigation of fatigue behavior of UFG this study is tabulated in Table 1 [16]. The initial material
materials became of interest to many researchers, and in was received in form of a hot rolled plate with a thickness
particular, the microstructural evolution in the high-cycle of 25 mm. For ECAP processing, billets of 25 9 25 9
fatigue (HCF) and the low-cycle fatigue (LCF) regimes has 175 mm3 were cut by a low-speed band-saw, such that the
been recently studied in detail [14, 16–19]. longitudinal direction of each billet coincided with the
So far, the contributions of crack initiation and propa- rolling direction of the original plate. The billets were
gation were considered in a collective manner in the ECAP processed at room temperature at a processing speed
analyses of fatigue behavior [16, 18], and there is a limited of 2.5 mm/s. The ECAP die had a 25 9 25 mm2 channel
amount of data on the crack growth behavior of UFG with a sharp 90° channel cross-section angle, and benefited
materials in correlation with the microstructure [17]. To the from the ‘‘sliding walls concept’’ [21]. The IF steel samples
best of the authors’ knowledge, no distinction has been were extracted from billets processed along the ECAP
made between the crack initiation and crack propagation routes 2A, 8BC, 8E, and 16E.2 Coarse-grained (CG) IF
behavior for bcc UFG materials, with the exception of one steel was also considered in this study for comparison
study carried out on bcc UFG low-carbon steels [19], purposes.
where the material was investigated upon being heat trea- Miniature compact tension (CT) specimens were
ted. However, it has been previously shown that heat extracted from the ECAP-processed billets similar to those
treatment has a significant influence on the fatigue prop- in ASTM standards [23], however, with slightly different
erties of UFG interstitial-free (IF) steel [20], and thus, the dimensions, as presented in Fig. 1. CG IF steel samples of
generalization of the results obtained in [19] to all bcc UFG the same dimensions were obtained by heat treatment of
steels is questionable. 2A material according to [20]. The samples were cut by
This study was undertaken with this motivation, in order wire electro-discharge machining (EDM) in such a way
to focus on the crack propagation behavior in UFG alloys that crack growth only took part in the uniformly processed
with a bcc structure, and the crack growth behavior of a sections of the ECAP billets [22]. The miniature specimens
UFG bcc IF steel was investigated in as-ECAP-processed employed do not meet the minimum thickness criterion for
condition. For this purpose, IF steel samples processed the determination of the critical stress intensity factor (KIc),
through different processing routes, and thus, with different and thus, the crack growth curves presented in this study do
microstructural characteristics were tested. In particular, not go beyond the end of the Paris regime, so that linear
crack growth rates were determined as a function of stress elastic fracture mechanics was still applicable. In order to
intensity factor range (DK), and the constants of the allow for the investigation of the role of the microstructural
Paris law were calculated. Mechanical experiments were morphology induced by ECAP processing, the specimens
accompanied by thorough microstructural investigations. were cut parallel to all three side surfaces of the billets, as
The focus was placed on the crack growth–microstructure depicted in the inset of Fig. 2. Following EDM, the sur-
relationship with an emphasis on the influence of the ECAP faces of the CT specimens were mechanically ground down
shear plane.1 The current results indicate that the ECAP to a grit size of 5 lm, such that the detrimental effects of
processing route is a major parameter dictating the crack EDM on the sample surfaces were eliminated. In order to
growth behavior. Specifically, a higher susceptibility to monitor the fatigue-induced changes in the crack wake,
crack growth was observed for UFG IF steel samples that some samples were electropolished prior to cyclic loading
underwent a higher number of ECAP passes. Furthermore, to ensure the surface quality needed for electron back-
in all UFG microstructures, the cracks grew along elon- scatter diffraction (EBSD) analysis. Electropolishing was
gated grain structures, which were primarily found on the conducted utilizing a 5% perchloric acid solution under an
flow plane parallel to the last ECAP shear plane. applied potential of 30 V at -40°C.
The mechanical tests were carried out on a servo-
hydraulic load frame at room temperature. For measuring
the crack length, the potential drop method [24] was uti-
lized, where the current was limited to 5 A in order to
avoid temperature increase in the samples [25]. The
1
potential drops were measured using a voltmeter with
Here ‘‘inclination plane’’ which is also denoted as ‘‘grain elongation
nanovolt resolution. Commercial LabView software was
plane’’ [11] would be a more appropriate term; however, ‘‘shear
plane’’ is used instead in this text as it is the commonly utilized
2
terminology in the literature (Figs. 2 and 4). For more information on ECAP processing routes, see [22].

123
J Mater Sci (2010) 45:4813–4821 4815

employed to determine the actual crack length from the electron microscope operating at 20 kV equipped with an
measured potential drop and to calculate the actual stress EBSD detector, a confocal laser scanning microscope
intensity factor. Experiments with a constant frequency of (CSM) equipped with a 408 nm violet laser, and a X-Ray
20 Hz were conducted under DK control with a positive diffractometer (XRD) with a Cu–Ka source operated at
stress ratio of R = 0.1, where the starting load level cor- 40 kV were utilized for microstructural investigation.
responded to DK = 18 MPaHm. Finally, a scanning

Table 1 Chemical composition of the IF steel investigated in this Experimental results


study [16]
The crack growth behavior of the IF steel was strongly
Element Fe C S N O Ti Al
dependent on the microstructure, which itself was dictated
Weight Balance 0.0023 0.0077 0.0018 0.002 0.065 0.05 by the ECAP processing route. Figure 2 depicts crack
percent growth curves of CG and UFG IF steel samples. The crack
growth curves shown in Fig. 2 were all obtained on sam-
ples extracted parallel to the transverse plane (TP), as
shown in the inset of Fig. 2. Obviously, crack growth rates
increased and fatigue threshold values (DKth) decreased in
the following order: CG, 2A, 8E, 16E and finally 8BC. The
threshold value for the CG IF steel was 9.5 MPaHm,
whereas DKth for the UFG IF steel from route-8BC was
5.0 MPaHm. Differences in crack growth behavior were
pronounced in the near-threshold region, while in the Paris
regime the behavior of IF steel samples with different
microstructures was rather similar. Thus, the same Paris
law constants were assigned within this region for all UFG
IF steel samples: C = 7.5 9 10-11 m/cycle, m = 2.3. The
inset shown in the right lower corner of Fig. 2 schemati-
cally depicts the variation in crack growth direction for
Fig. 1 Geometry of the miniature CT specimens used in this study different IF steel samples. Crack growth perpendicular to
[25] the loading direction was observed for CG IF steel and

Fig. 2 Crack growth rates as a


function of DK for UFG steel
samples extracted parallel to
TP. The insets depict the sample
orientation and the direction of
crack growth for different TP IF
steel microstructures

123
4816 J Mater Sci (2010) 45:4813–4821

UFG IF steel samples processed along routes 2A and 8E, microstructures at various stages of fatigue. Obviously,
and deviations from this normal crack growth direction crack growth direction differs already in the very early
were noted in the case of route-16E- and 8BC-processed stages directly after crack initiation from the notch. The
UFG IF steel samples. For the latter UFG condition a crack growth is very similar in route-2A and 8E UFG IF
maximum deviation of about 30° from the normal crack steels, i.e., perpendicular to the loading direction. No
growth direction was observed, which has to be considered pronounced differences are visible, but for the given
in the evaluation of the measured crack growth rates, as loading conditions (corresponding to DK = 18 MPaHm),
will be discussed later in this paper. the crack path in the route-2A UFG IF steel turns out
The SEM micrographs shown in Fig. 3 depict the slightly more tortuous. In comparison to these samples, the
cracks grown in UFG IF steel samples with different crack growth behavior in the UFG IF steel samples pro-
cessed along both ECAP routes 16E and 8BC differs sig-
nificantly. The crack growth direction is unusual, such that
the crack deviates about 10° and 30° from the expected
crack growth direction in route-16E and route-8BC mate-
rial, respectively. The crack path in both cases is as tor-
tuous as in the case of the route-8E UFG IF steel.
In order to examine the influence of the crack growth
direction and sample orientation with respect to the
extrusion direction and the last ECAP shear plane, a set of
crack growth experiments were conducted on route-2A and
8BC UFG IF steel samples. The curves displaying the crack
growth behavior on three different planes are presented in
Fig. 4. As already shown in Fig. 2, the route-8BC material
is characterized by larger crack growth rates and lower
threshold values as compared to the route-2A UFG IF steel.
Comparison of the results obtained from the differently
Fig. 3 Micrographs depicting the crack paths in UFG IF steel
samples under loading in the vertical direction. Route-2A and 8E oriented samples of the route-2A UFG IF steel yields a
materials exhibit normal crack growth, while route-8BC and 16E significant variation. Threshold values are in the range of
materials are characterized by an unusual crack path 6.8–8.6 MPaHm, and the sample orientation seems to have

Fig. 4 Crack growth rates as a


function of DK for UFG steel
samples processed along ECAP
routes 2A and 8BC. The insets
depict the different sample
orientations with respect to the
last ECAP shear plane, and the
direction of crack growth for
each of the IF steel variants

123
J Mater Sci (2010) 45:4813–4821 4817

only a slight influence on this value. But as shown in the The EBSD grain maps shown in Fig. 6 depict the
inset in the lower right corner of Fig. 4, the growth microstructures of route-2A and 8BC UFG IF steels in the
directions of the cracks on the TP and flow plane (FP) crack wake on the FP. For the route-2A material it is
differ significantly: on the TP the crack growth occurs obvious that crack growth follows elongated structures,
perpendicular to the loading axis, while an offset of about which are indeed agglomerations of subgrains only divi-
30° is present on the FP. This, of course, leads to an ded by low-angle grain boundaries (LAGBs). Crack
underestimation of the crack growth rate on the FP (see advance on these boundaries seems to be very easy, and
Discussion for details). Specifically, crack growth takes the crack growth direction changed only when it approa-
place almost parallel to the last ECAP shear plane on the ched a region characterized by more equiaxed grains sep-
FP. For the route-8BC UFG IF steel on all planes (TP, FP arated mainly by high-angle grain boundaries (HAGBs)
and longitudinal plane (LP)) a similar direction of crack (upper images in Fig. 6), where a kink evolved. Since the
growth is observed, such that even the crack growth rates grain refinement in the route-8BC IF steel is much more
as a function of DK are similar. On FP and LP, crack effective, all aspects of crack growth were studied on a
growth direction can be linked to ECAP shear planes, and smaller scale (lower images of Fig. 6). Accordingly, the
it can be correlated to the last ECAP shear plane on FP and large elongated structures present in the route-2A material
to the penultimate shear plane on LP. were not present anymore, but still the microstructure con-
In order to investigate the crack growth behavior on a sisted of elongated structures, and the crack growth took
smaller scale, SEM and CSM studies were carried out place parallel to these structures.
(Fig. 5). The SEM images show the typical crack growth Since crack growth is a bulk phenomenon, investigation
behavior for the route-8BC UFG IF steel, and a step-like of the crack growth–microstructure relationship only on the
structure is visible on all surfaces. In general the crack is samples’ surfaces is not sufficient. Therefore, the crack
characterized by the deviation of about 30° from the surfaces of all samples were investigated, as well. Figure 7
expected crack growth direction as described above, displays the crack surface of a route-8BC CT sample (TP).
however; regular small increments are present where the For better monitoring of the damage on this surface, the
crack propagates along the originally expected direction, angle of view was set at 70°. The crack surface shown in
which is perpendicular to the loading direction. In addition, Fig. 7 is very distorted, and a secondary crack parallel to
kinking bifurcation of the crack is often detectable. As the crack surface and delamination of a large area are
shown in the three-dimensional (3D) surface plot obtained visible. The height of the delaminated area is about 25 lm,
from CSM, the surface of the sample is characterized by a very well exceeding the average grain size of the route-8BC
significantly distorted topography when the crack is UFG IF steel. Such delaminations were present in all UFG
kinked. IF steel variants regardless of the loading conditions.

Fig. 5 SEM and CSM micrographs depicting the crack paths in


route-8BC UFG IF steel in differently oriented samples. The SEM
micrographs show a step-like crack path at different length scales on Fig. 6 SEM micrographs (the images are foreshortened as samples
both flow and transverse planes. The CSM results indicate increased are mounted at a 70° angle for EBSD analyses) and the corresponding
surface topography in a region where the crack is kinked, the EBSD grain maps. Both EBSD analyses were carried out on the flow
maximum topography shown in the CSM map is 1 lm plane. The crack follows the elongated grain structures in both cases

123
4818 J Mater Sci (2010) 45:4813–4821

similar trend was observed for the routes optimized for


high strength and superior fatigue performance (16E and
8E), while the crack growth rates were smaller for the non-
optimized route (2A) and the CG material. It is well known
that total fatigue life can be subdivided into three phases:
crack initiation, crack propagation and the final failure of
the material. As crack growth rates for the optimized
ECAP routes are higher, the superior fatigue lives of the
optimized UFG IF steel variants can only be attributed to a
Fig. 7 SEM micrograph displaying the crack surface of a fatigued significantly longer phase of fatigue crack initiation. Sim-
route-8BC TP sample. As depicted in the schematic, the micrograph
was recorded at an angle of view of about 70°. A secondary crack is ilar results have been shown for other UFG materials, such
visible parallel to the surface and at a distance of 25 lm, which is as copper and aluminum alloys [17, 26–28]. In all cases,
much larger than the average grain size the optimized routes exhibited superior fatigue perfor-
mance yet larger crack growth rates, regardless of the
Discussion specimen geometry.
Two main reasons have been forwarded to explain the
Sample geometry observed phenomenon: the crack paths in UFG materials
are less tortuous, and roughness-induced crack closure is
In this study, the stage three of the crack growth behavior not as pronounced as in the CG material. Accordingly, a
was not investigated due to the geometry of the miniatur- more tortuous crack path in any material leads to an
ized CT samples, which do not meet the minimum thick- effectively longer crack path, and eventually to more
ness criterion formulated in the former version of the actual energy consumption during fatigue crack growth [17, 29,
ASTM standard. The ligament beyond the phase of stable 30]. In the current material, all crack paths seem to be
crack growth is quite small leading to final failure that is nearly straight (Fig. 3), which has also been observed in
mainly characterized by plastic deformation, so failure UFG IF steel samples of different geometries (Niendorf
does not occur in a fracture mechanics manner [25]. (2010) Unpublished data). The presence of crack closure
However, in the range of DK considered in the experi- mechanisms decreases the range of DK with a fully opened
ments, the boundary conditions of fracture mechanics crack, which in fact is a prerequisite for crack advance [19,
testing are fulfilled, i.e., the plastic zone is small with 30–32]. Thus, any mechanism leading to crack closure can
respect to the crack length, the ligament, and the overall slow down crack growth [31, 32].
sample dimensions. The highest crack growth rates and the
lowest threshold value were exhibited by the route-8BC Crack propagation direction
UFG IF steel, where the crack growth was not perpendic-
ular to the loading direction (Figs. 2 and 3). Consequently, In order to support the conclusions made in the following
the potential drop method used for the measurement of paragraphs it is important to characterize the condition of
crack length leads to an underestimation of the crack the IF steel prior to ECAP processing. The material
length, since it primarily detects the resistance of the investigated was hot rolled and cooled down slowly, per-
smallest cross-section of the CT specimen, i.e., the liga- mitting a full recrystallization. As it is known from bcc
ment in front of the crack. Eventually, this underestimation materials this may lead to the evolution of an inherent
of the actual crack length leads to lower crack growth rates. texture, which could result in failure along specific planes
However, both values can be corrected by multiplication after subsequent deformation processing. Therefore, tex-
with a constant factor, which is defined by the angle of ture of the CG IF steel prior to ECAP processing has been
deflection. Since this correction factor is close to one, it investigated (Fig. 8). The inverse pole figures shown in
was not applied to the results shown in Figs. 2 and 4 in this Fig. 8 clearly demonstrate that only weak texture is present
study. and texture is not changed significantly upon ECAP pro-
cessing in case of the IF steel. Thus, initial texture cannot
Fatigue performance and crack propagation be the reason for the unusual crack propagation directions
found.
It is interesting to note that crack growth resistance of The current results (Figs. 2 and 3) demonstrate that the
route-8BC UFG IF steel, which has been shown to exhibit crack growth direction on TP in the route-16E and 8BC
superior fatigue performance both in the LCF and HCF UFG IF steel can deviate from the expected one in this
regimes [16, 20], (Niendorf (2010) Unpublished data) is case. The ECAP shear plane cannot play a significant role
indeed inferior, based on the current results (Fig. 2). A since the ECAP shear plane is not visible on TP. In order to

123
J Mater Sci (2010) 45:4813–4821 4819

sample shows bifurcations, which lead to deflections of the


crack growth direction, and the surface of the TP sample
shows pronounced step-like structures, again leading to
deviation from the original direction of crack growth
(Fig. 5). This step-like structure of crack advance clearly
indicates that the deviation in the crack growth direction in
UFG steel is not caused by a deteriorated shear strength of
the material, but by microstructural features, such as
elongated grains.
Fig. 8 X-ray diffraction results displaying the texture of IF steel in
The micrograph presented in the lower part of Fig. 5
the as-received (CG) condition shows the surface topography of a route-8BC FP sample.
The plastic zone in the crack wake is clearly visible, the
relate crack growth to the last ECAP shear plane, speci- thickness of the sample is slightly decreased in this area,
mens were cut parallel to FP (investigation of the role of leading to a lower topography level in the CMS plot, and
the last shear plane) and LP (investigation of the role of the there is a kinked area characterized by a higher level
penultimate shear plane). topography. The presence of extrusions after fatigue
It should be noted that the crack growth curves of the loading, which stem from microstructural imperfections,
route-8BC UFG IF steel are very similar in both the Paris such as large areas dominated by LAGBs and elongated
regime and the near-threshold region (Fig. 4). Even though structures, has already been shown for UFG samples in the
crack growth in the samples cut along three different ori- LCF regime [33, 34]. In the current route-2A material, the
entations not perpendicular to the loading direction was crack obviously follows the elongated grain structures,
monitored (Figs. 4 and 5), the deviation from the expected which are an agglomeration of subgrains only divided by
crack growth direction is quite similar in all cases, with LAGBs (Fig. 6). This finding holds true for the route-8BC
slightly larger inclinations for the FP and LP samples. UFG IF steel, as well. However, it is not seen as clearly due
Therefore, the crack growth rates in FP and LP specimens to the optimized processing and the corresponding lower
are slightly higher upon correction. degree of elongation. As can be deduced primarily from the
In addition, crack growth always occurs in a direction EBSD map shown for the route-2A material, the elongated
parallel to the last ECAP shear plane, leading to the con- structures are ordered in a parallel fashion, dictating the
clusion that the ECAP shear plane has a deteriorating direction of crack growth. Kinking is only observed in
influence on the crack growth behavior, even for an opti- areas featuring grain structures with more equiaxed shapes
mized ECAP route. Specifically, in terms of crack propa- and different misorientations.
gation direction, the role of sample orientation is more It has been previously shown that elongated grain
pronounced for the route-2A UFG IF steel (Fig. 4). The structures play a dominant role in the crack initiation
higher crack growth rates and the lower fatigue threshold processes [34]. Specifically, the locally increased stresses
values were observed for the FP samples in general, but the induced by the elongated grains in the matrix dictate the
variation was much more pronounced in this material crack initiation sites. Similarly, in the case of crack
(Fig. 4). For the TP samples of 2A material, crack growth advance, this stress rising effect along elongated structures
was perpendicular to the loading direction, but in FP drives the crack growth process (Fig. 6). These structures
samples, crack growth was similar to the behavior in the are preferentially ordered along the direction of material
route-8BC samples. So in the case of non-optimized ECAP flow during ECAP [2, 26], and therefore, the influence of
processing the elongated microstructure, which has been the last ECAP shear planes is pronounced (visible on FP
described in former studies [16, 33, 34], features a and LP, c.f. Figs. 4, 5, and 6). Currently, the role of defects,
mechanical behavior dependent on sample orientation with i.e., micro pores [36], introduced during ECAP processing
respect to the loading direction [35]. on the crack growth behavior and the reason for the unu-
sual crack growth behavior on TP in route-8BC and 16E
materials are still unexplored, and work is underway to
Crack growth–microstructure relationship clarify these issues.
Finally, a careful evaluation of the current results pro-
The crack growth behavior depicted in high magnification vides important guidelines regarding the role of ECAP
is presented in Fig. 5. The SEM images in the upper part of processing on the microstructure–fatigue crack propaga-
Fig. 5, and Figs. 6 and 7 show a step-like crack growth tion, overall fatigue performance relationship. A strong
behavior in route-8BC UFG IF steel samples in two dif- relationship between sample orientation and crack deflec-
ferent orientations (FP and TP). The surface of the FP tion has already been observed in fracture toughness

123
4820 J Mater Sci (2010) 45:4813–4821

measurements conducted on UFG HPT Armco iron, where crack closure is not as pronounced as in the CG
elongated structures were also observed [37]. The sche- counterparts.
matic presented in Fig. 9 summarizes the current findings
obtained from samples on the FP, and accordingly, the
Conclusions
crack growth behavior is very sensitive to ECAP process-
ing. Specifically, in its CG state, the IF steel samples
The role of ECAP on the crack growth behavior in UFG IF
exhibit the best resistance against crack propagation, even
steel was investigated. Miniature compact tension speci-
though they exhibit poor crack initiation and overall fatigue
mens extracted from IF steel processed along several
lives [16]. A similar trend is valid for the UFG variants
ECAP routes with different order of material flow during
obtained by non-optimized ECAP processing, where the
ECAP processing were used to evaluate both crack growth
overall fatigue performance is better than that of the CG
rates as a function of loading amplitude and direction of
material [16, 33]. In the case of UFG IF steel obtained by
crack growth. Mechanical testing was accompanied by
optimized ECAP processing, however, the crack growth
thorough microstructural investigations using scanning
rates are the highest (Fig. 9), despite the superior fatigue
electron microscopy, confocal laser scanning microscopy,
performance and longest crack initiation periods [16, 33].
and electron backscatter diffraction. The following con-
The current microstructural findings indicate that the
clusions can be drawn from the work presented herein:
microstructure dictates this behavior, such that the grain
shape and grain misorientation become of utmost impor- 1. Crack growth rates increase and fatigue threshold
tance. The elongated grain structures drive the crack away values decrease concomitant with decreasing grain size
from the expected direction, mainly due to added stress in UFG IF steel. The fatigue threshold value decreases
concentration they impose upon the matrix (Fig. 9). These by a factor of two comparing coarse-grained and route-
elongated grain structures feature a large volume fraction 8BC UFG materials.
of LAGBs, and thus, the crack progresses along the 2. The ECAP processing route significantly influences
direction the LAGBs are cut through, i.e., along the major the crack advance. Specifically, a significant deviation
axes of the grains. As for the UFG material obtained upon from the expected direction of crack growth normal to
optimized ECAP processing, the regularly shaped grains the loading direction is notable, where the deviation
are divided mainly by HAGBs; yet, the already large stress from the normal direction is a function of the ECAP
concentration aids the crack growth. Furthermore, as the processing route.
crack approaches these zones a deviation of the crack from 3. Sample orientation is another factor dictating the crack
the initial propagation direction is observed (Fig. 9). This, advance. Direction of crack growth on flow and
however, is absent in the CG material, and the crack paths longitudinal planes depends on the ECAP shear plane
in UFG materials are less tortuous, and roughness-induced in all UFG IF steel variants.

Fig. 9 Schematic summarizing the microstructure–crack propaga- represent the relative crack initiation, crack propagation, and overall
tion–overall fatigue performance relationship in UFG IF steel. fatigue performances of different microstructures. The weak micro-
Samples on the FP are considered. The dashed line indicates the structural alignment in the route-8BC material (Fig. 6) has been
expected crack propagation direction. The bars schematically omitted in the schematic

123
J Mater Sci (2010) 45:4813–4821 4821

4. Overall, the microstructure induced upon ECAP pro- 15. Shin DH, Park KT (2005) Mater Sci Eng A410–411:299. doi:
cessing dictates the crack growth behavior. In partic- 10.1016/j.msea.2005.08.025
16. Niendorf T, Canadinc D, Maier HJ, Karaman I, Sutter SG (2006)
ular, in the case of crack propagation, elongated Int J Mater Res 97:1328
structures oriented parallel to the direction of material 17. Vinogradov A (2007) J Mater Sci 42:1797. doi:10.1007/s10853-
flow during ECAP processing form local stress fields 006-0973-z
that direct the crack as it grows. 18. Mughrabi H, Höppel HW (in press) Int J Fat. doi:10.1016/
j.ijfatigue.2009.10.007
19. Kim HK, Choi MI, Chung CS, Shin DH (2003) Mater Sci Eng
Acknowledgements The help of B. Gorny with CSM is acknowl- A340:243
edged. The German part of this study was supported by Deutsche 20. Niendorf T, Canadinc D, Maier HJ, Karaman I (2008) Int J
Forschungsgemeinschaft, within the Research Unit Program ‘‘Mecha- Fatigue 30:426. doi:10.1016/j.ijfatigue.2007.04.015
nische Eigenschaften und Grenzflächen ultrafeinkörniger Werkstoffe.’’ 21. Segal VM, Goforth RE, Hartwig KT (1995) Texas A&M Uni-
The U.S. part of this study was supported by the National Science versity, U.S. Patent No. 5,400,633
Foundation, contract no. CMMI-0900187, Materials and Surface 22. Barber RE, Dudo T, Yasskin PB, Hartwig KT (2004) Scripta
Engineering Program, Directorate of Engineering, Arlington, VA. Mater 51:373. doi:10.1016/j.scriptamat.2004.05.022
23. ASTM Standard E 647-08, American society for testing and
materials
24. Sander M, Richard HA (2006) Int J Fat 28:583. doi:10.1016/
References j.ijfatigue.2005.05.012
25. Niendorf T, Rubitschek F, Maier HJ, Niendorf J, Richard HA,
1. Segal VM (1995) Mater Sci Eng A197:157 Frehn A (2010) Mater Sci Eng A527:2412. doi:10.1016/j.msea.
2. Valiev RZ, Langdon TG (2006) Progr Mater Sci 51:881. doi: 2009.12.012
10.1016/j.pmatsci.2006.02.003 26. May J (2008) Dissertation, University of Erlangen-Nuremberg
3. Valiev RZ, Korznikov AV, Mulyukov RR (1993) Mater Sci Eng 27. May J, Dinkel M, Amberger D, Höppel HW, Göken M (2007)
A A168:141 Metall Mater Trans 38A:1941. doi:10.1007/s11661-007-9110-0
4. Meyers MA, Mishra A, Benson DJ (2006) Progr Mater Sci 28. Hübner P, Kiessling R, Biermann H, Hinkel T, Jungnickel W,
51:427. doi:10.1016/j.pmatsci.2005.08.003 Kawalla R, Höppel HW, May J (2007) Metall Mater Trans
5. Gleiter H (1989) Progr Mater Sci 33:233 38A:1926. doi:10.1007/s11661-007-9297-0
6. Valiev RZ, Alexandrov IV, Zhu YT, Lowe TC (2002) J Mater 29. Meyer LW, Sommer K, Halle T, Hockauf M (2008) J Mater Sci
Res 17:5 43:7426. doi:10.1007/s10853-008-2725-8
7. Valiev RZ, Estrin Y, Horita Z, Langdon TG, Zehetbauer MJ, Zhu 30. Vinogradov A, Nagasaki S, Patlan V, Kitagawa K, Kawazoe M
YT (2006) JOM 58:33. doi:10.1007/s11837-006-0213-7 (1999) Nano Struc Mater 11:925. doi:10.1016/S0965-9773(99)
8. Koch CC, Youssef KM, Scattergood RO, Murty KL (2005) Adv 00392-X
Eng Mater 7:787. doi:10.1002/adem.200500094 31. Elber W (1970) Eng Fract Mech 2:37
9. Zhilyaev AP, Nurislamova GV, Kim BK, Baró MD, Szpunar JA, 32. Suresh S (1998) Fatigue of materials. University Press, Cam-
Langdon TG (2003) Acta Mater 51:753. doi:10.1016/S1359- bridge, UK
6454(02)00466-4 33. Niendorf T, Canadinc D, Maier HJ, Karaman I (2007) Metall
10. Höppel HW, May J, Göken M (2004) Adv Eng Mater 6:781. doi: Mater Trans 38A:1946. doi:10.1007/s11661-007-9154-1
10.1002/adem.200306582 34. Niendorf T, Dadda J, Canadinc D, Maier HJ, Karaman I (2009)
11. Zhu YT, Lowe TC (2000) Mater Sci Eng A291:46. doi: Mater Sci Eng A517:225. doi:10.1016/j.msea.2009.04.053
10.1016/S0921-5093(00)00978-3 35. Niendorf T, Marten T, Maier HJ, Karaman I (2008) Mater Sci For
12. Kawasaki M, Horita Z, Langdon TG (2009) Mater Sci Eng 584–586:864
A524:143. doi:10.1016/j.msea.2009.06.032 36. Lapovok R, Tomus D, Mang J, Estrin Y, Lowe TC (2009) Acta
13. Semenova LP, Valiev RZ, Yakushina EB, Salimgareeva GH, Mater 57:2909. doi:10.1016/j.actamat.2009.02.042
Lowe TC (2008) J Mater Sci 43:7354. doi:10.1007/s10853- 37. Hohenwarter A, Pippan R (2010) Mater Sci Eng A527:2649. doi:
008-2984-4 10.1016/j.msea.2009.12.033
14. Niendorf T, Canadinc D, Maier HJ, Karaman I (2009) Scripta
Mater 60:344. doi:10.1016/j.scriptamat.2008.10.033

123

View publication stats

S-ar putea să vă placă și