Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

Engineering Fracture Mechanics 127 (2014) 181–193

Contents lists available at ScienceDirect

Engineering Fracture Mechanics


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

Notch ductility of steels for automotive components


M. Faccoli ⇑, G. Cornacchia, M. Gelfi, A. Panvini, R. Roberti
Department of Mechanical and Industrial Engineering, University of Brescia, via Branze 38, 25123 Brescia, Italy

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

Article history: The aim of this work is to experimentally evaluate the notch ductility of a TWinning
Received 3 October 2013 Induced Plasticity (TWIP) steel, a Dual Phase (DP) steel and two Quenching and Partitioning
Received in revised form 16 May 2014 (Q&P) steels for automotive components. The notch ductility was investigated by studying
Accepted 7 June 2014
the resistance to fracture initiation in tensile specimens with notches of various root radii.
Available online 26 June 2014
Fracture toughness is discussed in terms of the critical J-integral Jc. It was found that Jc
progressively increases with the notch tip radius q for all the studied steels, demonstrating
Keywords:
a high sensitivity of ductility upon notch severity. The relationship between Jc and q
Notch ductility
TWinning Induced Plasticity (TWIP) steel
allowed to calculate the maximum strain at fracture initiation emax, that is proposed as a
Quenching and Partitioning (Q&P) steels deformation limit parameter useful to rank the forming capacity of the steels. It can be con-
Dual Phase (DP) steel sidered as the maximum strain that can be sustained ahead of the notch tip up to the onset
Double Edge Notched Tension (DENT) of fracture advancement. Afterwards, a metallographic analysis by means of light optical
specimen microscope and scanning electron microscope was carried on etched specimens. The frac-
Thin sheets ture mechanisms were also investigated by means of scanning electron microscope.
Ó 2014 Elsevier Ltd. All rights reserved.

1. Introduction

In order to fulfil ever-growing requirements associated with high oil price and environmental pollution issues, the
automobile industry has been since long encouraging considerable efforts to reduce the weight of vehicles and in recent
times is assessing the opportunity to introduce new Advanced High Strength Steels (AHSS), such as Al-added TWinning
Induced Plasticity (TWIP) steels and Quenching and Partitioning (Q&P) steels, with improved properties of strength and
ductility.
Currently, TWIP steels become one of the most attractive and potential materials for manufacturing vehicle components.
TWIP steels are high-Mn austenitic steels, with (15–25)% of Mn. In these steels Mn is the most important alloy element: it
stabilizes the austenitic phase at room temperature, preventing the formation of the ferrite and the manganese carbides
foreseeable by the Fe–C–Mn equilibrium ternary phase diagram. Another important function of Mn is to reduce the austenite
stacking fault energy, favoring the TWIP effect during deformation [1]. TWIP is characterized by the formation of deforma-
tion twins, that lead to strain hardening phenomena as they act as obstacles for dislocation glide. In these steels a high C
percentage is necessary to increase yield and tensile strength. Recently, TWIP steels in which the Mn content is reduced,
while Al is added, have also been developed. In these steels the Al addition, typically less than 3%, results in a slight reduction
in density which is beneficial to weight reduction of car body. Al also stabilizes the austenitic phase and plays an important
role promoting twinning inside grains [2–5] leading to excellent formability.

⇑ Corresponding author. Tel.: +39 030 371 5572; fax: +39 030 3702448.
E-mail address: michela.faccoli@ing.unibs.it (M. Faccoli).

http://dx.doi.org/10.1016/j.engfracmech.2014.06.007
0013-7944/Ó 2014 Elsevier Ltd. All rights reserved.
182 M. Faccoli et al. / Engineering Fracture Mechanics 127 (2014) 181–193

Nomenclature

a length of edge notches


2b ligament between the notches
B specimen thickness
E Young’s modulus
J J-integral
Jc critical elasto-plastic J-integral parameter at crack initiation
KI stress intensity factor
L specimen length
N strain hardening modulus of the steel
P applied load
PN load normalized by the crack length function
2W specimen width
Da crack advancement
Dlpl plastic load line displacement
DPN change in the normalized load due to crack growth
eY strain up to yield load
emax critical notch root strain
efracture strain to fracture
emax load strain up to maximum load
C mathematical gamma function
q notch tip radius
qeff critical notch radius
rY yield stress
up plastic displacement

Al-added TWIP steels are an evolution of the Fe–18Mn–0.6C steel, that is a baseline TWIP steel shown to have the desir-
able twinning mechanism. However Fe–18Mn–0.6C steel had several problems as a low yield strength in the annealed con-
dition (around 250 MPa), the occurrence of cementite precipitation during cooling after hot rolling and/or annealing and the
susceptibility to hydrogen delayed fracture [6]. Alloying Fe–18Mn–0.6C with Al was introduced to overcome these deficien-
cies. Jin and Lee [6] found that Al addition of 1.5–2% suppresses cementite precipitation because of the decrease in both
activity and diffusivity of C in the austenite. According to [7], the stacking fault energy increases linearly with Al additions.
For example, the stacking fault energy of Fe–18Mn–0.6C increases from 13 mJ/m2 to 30 mJ/m2 with the addition of 1.5% Al
[8]. As a consequence this steel does not exhibit the TRansformation Induced Plasticity (TRIP) phenomenon, associated with
the martensitic transformation under cold working operations, because it is observed only in very low stacking fault steels
(below 20 mJ/m2) [9]. This is an important effect, as formation of martensite tends to reduce ductility and is generally not
desirable in a TWIP steel. Other effects of the increased stacking fault energy are the increase in yield strength and the
decrease of the strain hardening of the steel [6]. Furthermore, recent investigations [10,11] show that Al improves the resis-
tance to hydrogen embrittlement of TWIP steels. The reasons are still unclear, some authors suggest that Al inhibits the
dynamic and static strain aging by increasing both the activation energy for C diffusion and the stacking fault energy. The
inhibition of dynamic strain aging decreases the dislocation density [11] and lowers the frequency of dislocation pinning
by C and results in a decrease in the material strength [12]. Static and dynamic strain aging inhibition is considered to reduce
the hydrogen embrittlement susceptibility of these steels. A recent study suggests that Fe–18Mn–1.5Al–0.6C is relatively
immune to hydrogen embrittlement [13].
Even though TWIP steels exhibit good formability when evaluating the deep drawing capacity, the edge stretchability is
weak and is not better than other AHSS grades [14]. This suggests that notch sensitivity, considering roughness and other
types of surface defects, is a concern that needs to be addressed in order to minimize this sensitivity by adjusting the man-
ufacturing process. Up today, very little data is available in literature on this topic. This latter issue has been recently studied
by Chung et al. [15] who confirmed that TWIP steels are highly sensitive to the surface conditions.
Quenching and Partitioning is a heat treatment, recently developed by Speer [16], of great interest from both the research
community and the industry, because it allows for a better combination of steel strength and ductility. The process starts
with a partial or full austenitisation, followed by quenching austenite below the martensite-start temperature (Ms), but
above the martensite-finish temperature (Mf), with the aim to form a controlled amount of martensite. The steel then under-
goes an isothermal treatment (partitioning) at the same or higher temperature, in order to produce the diffusion of carbon
from the supersatured martensite to the remaining austenite. The material is finally quenched to room temperature: the less
stable austenite becomes martensite, whereas the austenite that is sufficiently C enriched retains at room temperature. The
obtained microstructure is formed by ferrite (in the case of partial austenitisation), martensite and retained austenite. The
M. Faccoli et al. / Engineering Fracture Mechanics 127 (2014) 181–193 183

martensite provides the strength, whereas the retained austenite and the ferrite improve the ductility. Carbon-enriched
retained austenite is considered beneficial, because in these steels TRansformation Induced Plasticity (TRIP) during deforma-
tion can positively contribute to work hardening, formability and fracture toughness [17]. The process is referred to as
Quenching and Partitioning (Q&P) to distinguish it from conventional Quenching and Tempering (Q&T) of martensite, where
carbide precipitation and decomposition of retained austenite (to ferrite plus cementite) are typical. In Q&P steels the C con-
centration is about 0.2% because low-carbon steels are less susceptible to carbide precipitation than high-carbon steels and
have better weldability. Carbide precipitation has to be minimized during the partitioning step because carbides act as sink
of carbon that is no longer available to enrich the austenite [18]. This phenomenon can be avoided also by alloying with Si, Al
or P [19,20]. Up to date, Si is the alloying element that has shown the highest success in this aspect. Property advancements
continue to be made through research on this emerging technology.
Automobile industry is particularly focused on optimizing the formability of AHSS, therefore investigations of materials
formability, formability limit parameters and early detection of potential failure are important issues in designing forming
processes to avoid failure of metal parts during forming. The present study consists of a notch ductility characterization of a
TWIP steel, two Q&P steels and a traditional automotive DP1000 steel aiming at determining a limiting parameter for failure
initiation. This property is investigated by studying the resistance to fracture initiation in notched tensile specimens with
notches of various root radii, in order to evaluate its effect on the cracking initiation in thin sheets. The results from notched
tensile specimens will allow to graduate the materials behavior with respect to formability, by means of the maximum strain
that can be sustained at the notch root before fracture initiation. The fracture mechanisms were also investigated in all stud-
ied steels by means of scanning electron microscope (SEM).

2. Materials and methods

The TWIP steel, the DP1000 steel and the Q&P steels investigated in the present paper have the chemical compositions
shown in Table 1. All steels were provided as sheets of different thickness.
The tensile curves and the mechanical properties of the studied steels are shown in Fig. 1 and in Table 2, respectively.
The studied TWIP steel is an Al-added TWIP steel, having 1.5% Al, not yet widely used. The DP1000 steel is a traditional
automotive steel, obtained from a heat treatment which consist of heating at an intercritical temperature and then quench-
ing. Both the studied Q&P steels are Si-added Q&P steel, furthermore the Q&P2 steel contains Mo. The presence of Mo in the
chemical composition has been shown to result in increased strength (Fig. 1), as also found by Santofimia et al. [18]. They are
experimental materials, not industrialized up to now, subjected to the thermal cycles detailed in Table 3, using a Gleeble
3800 thermomechanical simulator.
The metallographic samples were machined out of the sheets, were polished and etched with a specific etchant to reveal
the microstructure in the rolling direction. The TWIP steel was etched with hydrochloric acid (0.5%), dodecyl-benzene sul-
phonic acid (3%) and a saturated solution of picric acid at a temperature of 60 °C. The Q&P steels were etched with 0.5% Nital
and 2% Picral. The DP1000 steel was etched with 0.5% Nital.
The steels microstructures were examined using a Leica DMI 5000 M light optical microscope, equipped with Leica Appli-
cation Suite (LAS) image analyzer. SEM observations and semi-quantitative chemical analyses, using a LEO EVO-40XVP scan-
ning electron microscope with a Link Analytical eXL microprobe, were as well carried out. In the present research work paper
the notch ductility is investigated by studying the resistance to fracture initiation in Double Edge Notched Tension (DENT)
specimens with notches having different root radius. The DENT geometry is particularly well suited to thin sheets, due to the
symmetry and the absence of buckling. Geometry and size of DENT specimens are shown in Fig. 2.
Specimen length L and width 2W are equal to 130 mm and 25 mm respectively. Thickness B is equal to 1.45 mm for TWIP
steel specimens, 1 mm for Q&P steels specimens and 1.85 mm for DP1000 steel specimens, according to the different sheets
made available by the manufacturing steel work. Symmetric edge notches of length a equal to 4 mm are separated by a lig-
ament 2b equal to 17 mm. The edge notches have a finite tip radius q. Various notch tip radii were machined (in the range
0.05–0.5 mm) with the aim to investigate the effect of q on fracture initiation in thin sheets.
Moreover, in order to evaluate the effect of material anisotropy due to the rolling process, the TWIP steel specimens and
the DP1000 ones were machined in both the longitudinal (T) and transverse (L) direction of the sheets, with their notches
perpendicular and parallel to the rolling direction respectively. Due to the small size of the Gleeble sheets, the Q&P steels
specimens could be machined only in the longitudinal direction.

Table 1
Chemical compositions in weight percent of the studied steels.

Steel C Mn Si P S Cr Ni Mo Cu V Al Ti B Nb
TWIP 0.65 18.3 0.05 0.02 0.01 0.02 0.41 a) 0.05 0.04 1.5 b) b) b)
DP1000 0.18 2.3 0.18 0.007 <0.001 0.51 0.015 a) 0.01 0.002 0.034 0.002 0.0007 0.025
Q&P1 0.206 1.682 1.69 0.0109 0.013 0.079 0.014 0.002 0.012 0.006 0.043 0.003 0.0013 0.01
Q&P2 0.195 1.833 1.543 0.0186 0.016 0.048 0.015 0.206 0.013 0.006 0.003 0.003 0.0022 0.007

a) Not detected, b) not measured.


184 M. Faccoli et al. / Engineering Fracture Mechanics 127 (2014) 181–193

Fig. 1. Tensile curves of the TWIP steel and the DP1000 steel from [21] and of the Q&P steels.

Table 2
Yield stress (YS), ultimate tensile stress (UTS) and elongation to fracture (El.) of the TWIP steel and the DP1000 steel from
[21] and of the Q&P steels.

Steel YS (MPa) UTS (MPa) El.%


TWIP 465 960 47
DP1000 705 1013 13
Q&P1 636 1071 20
Q&P2 685 1228 13

Table 3
Conditions of the thermal treatments carried out on the Q&P steels.

Steel Austenitizing Quenching Partitioning


T (°C) t (s) T (°C) t (s) T (°C) t (s)
Q&P1 900 180 240 10 350 10
Q&P2 850 180 220 20 350 60

Fig. 2. Double Edge Notched Tension (DENT) geometry of specimens.

Tensile fracture tests were carried out under displacement control, by means of a servo-hydraulic testing system at a con-
trolled crosshead speed equal to 0.5 mm/min. The test specimen displacement was measured across the minimum ligament
section by means of an extensometer with an initial gauge length of 25 mm.
Since from preliminary tests fracture was observed to advance under monotonically increasing load, to point out the load
– load line displacement values corresponding to fracture initiation it was initially decided to adopt of the simplified key-
curve method originally pointed out by Ernst et al. [22] and subsequently developed by Reese and Schwalbe [23] on the basis
of the normalization method proposed by Landes et al. [24–27]. Given the small thickness of the specimens, a nearly plane
stress condition could be foreseen as well as a notably curved crack front during fracture advancement, with crack initiation
occurring in the mid-thickness of the specimen as already well documented by El-Soudani [28] for monotonically loaded
ductile crack advance. Therefore it was also decided to keep under control the crack advancement on one side of the spec-
imen, not with the purpose of detecting crack initiation, but just to limit lateral crack growth to nearly 0.10–0.15 mm in the
case of an excessive extension of the loading curve beyond the maximum load. For this reason each test was monitored by
means of a camera equipped with a zoom, acquiring one shot every two seconds, i.e. every roughly 16.7 lm increase in load
point displacement.
In all tests ductile fracture initiation and propagation were observed, the fracture toughness is therefore assessed in terms
of the critical elasto-plastic J-integral parameter at crack initiation, Jc. The J values were derived for the DENT geometry from
the experimental load/displacement record using the analytical relation proposed by Rice et al. [29]:
M. Faccoli et al. / Engineering Fracture Mechanics 127 (2014) 181–193 185

 Z 
K 2I 1
J Rice ¼ þ  2  Pdup  Pup ð1Þ
E 2b  B

where KI is the stress intensity factor, E is the Young’s modulus, P is the applied load and up is the plastic displacement. KI for
a DENT plate was calculated using the formula reported by Anderson [30].
While adopting J integral as the characterizing parameter for fracture initiation it was considered that in the case of large
scale yielding the J integral approach could be no more applicable [30]. Nevertheless the analysis carried out by Anderson
[30] indicated that notwithstanding ductile fracture toughness is size dependent in panel loaded in tension, the size depen-
dence is not nearly as severe as it is for cleavage. In addition, some experimental studies reported that in the plane stress case
the initiation toughness is independent of specimen geometry [31,32]. Moreover Sun et al. [33] showed that for a stationary
crack the stress–strain fields are independent of specimen geometry and in good agreement with the HRR solution [34,35].
More recently Yuan and Yang [36] found that the region of J-dominance is much larger in plane stress than in plane strain,
and the crack-tip fields can be better described by the J-integral in plane stress than in plane strain, in agreement with the
fact that for plane stress the in-plane geometry effect on the crack-tip fields is negligible. Finally Yan and Mai [37] studied
the effect of constraint on stable crack growth in plane stress and concluded that for both stationary and growing cracks in
plane stress, the distributions of opening stress, equivalent stress and equivalent strain are not sensitive to specimen geom-
etry and the triaxiality in the crack-tip region is independent of specimen geometry. Taken into account the fact that the
relationship between J integral at fracture initiation, notch root radius and the maximum strain at the notch root established
by Rice [35] was at the origin of the design of the experimental research, it was definitely concluded that J integral is the
most appropriate parameter for the study of the influence of notch root radius on fracture initiation.
For Q&P1 steel specimens, as well as for Q&P2 steel specimens having the largest notch root radius, during quasi static
loading crack initiation was immediately followed by an abrupt decrease in load caused by a ductile shear instability
propagation. Therefore Jc was directly calculated in correspondence of the maximum load registered during the test. On
the contrary, in almost all other tests on the experimented steels crack initiation was followed, at least initially, by quasi
static ductile propagation. Only in a limited number of cases ductile propagation turned into a shear instability propagation,
especially for the larger notch root radii.
Finally, the fractographic appearance of some broken specimens was investigated by means of scanning electron micro-
scope, in order to identify the fracture mechanisms in the studied steels.

3. Results and discussion

The results of the metallographic examinations carried out using the light optical microscope and the scanning electron
microscope are summarized in Figs. 3–6.
The microstructure of the TWIP steel is shown in Fig. 3: the metallic matrix is constituted by small polygonal austenitic
grains, with evenly distributed titanium nitrides some microns maximum in size (Fig. 3a) and stringers of manganese
sulphides (Fig. 3b with SEM–EDS analysis) locally coupled to oxide inclusions. Some deformation banding along the rolling
direction is also evidenced by the chemical etching.
The microstructure of the DP1000 steel is shown in Fig. 4: it consists of a very fine matrix given by a soft ferrite phase
containing islands of martensite as the secondary strengthening phase.
The very fine microstructures of both the Q&P steels (Figs. 5 and 6) are composed of ferrite (not transformed to austenite
during the partial austenitisation), stress relieved martensite and retained austenite as the result of the Quenching and
Partitioning treatments.

Fig. 3. Microstructure of the TWIP steel with a titanium nitride (a) and a SEM–EDS analysis of a manganese sulphide stringer embedding some oxide
inclusions (b).
186 M. Faccoli et al. / Engineering Fracture Mechanics 127 (2014) 181–193

Fig. 4. Light optical microscope (left image) and SEM (right image) appearance of the DP1000 steel microstructure.

Fig. 5. Light optical microscope (left image) and SEM (right image) appearance of the Q&P1 steel microstructure.

Fig. 6. Light optical microscope (left image) and SEM (right image) appearance of the Q&P2 steel microstructure.

Table 4
Average non metallic inclusion spacing in the studied steels – ASTM E45-97 (2002) Standard.

Steel Average inclusion spacing (lm)


TWIP 162
DP1000 154
Q&P1 109
Q&P2 114

Complying with ASTM E45-97 (2002) Standard, the average spacing among larger non metallic inclusions was measured
in all steel. The results are summarized in Table 4.
M. Faccoli et al. / Engineering Fracture Mechanics 127 (2014) 181–193 187

Fig. 7. Normalized load - load point displacement curves of the studied steels, with marked notch tip radii of the specimens.

The load - load point displacement curves recorded for each of the tested specimen are shown in Fig. 7. In the test diagram
the load and the load point displacement were normalized with respect to the initial ligament area and the initial extensom-
eter gauge length respectively. The so obtained normalized test curves superimpose each other, at least in the initial part of
the loading history.
As already anticipated for all the Q&P1 steel specimens, as well as for the Q&P2 steel specimens having the largest notch
root radius, during quasi static loading crack initiation was immediately followed by an abrupt decrease in load. In any case
the abrupt decrease in load did not correspond to a brittle fracture, but was the consequence of an unstable shear propaga-
tion by a ductile shear mechanism which arrested before breaking apart the whole ligament. A similar behavior was also
observed in the case of the DP1000 specimens with the largest notch root radius. For all these specimens, therefore, Jc
was directly calculated in correspondence of the maximum load registered during the test, i.e. in correspondence of the
abrupt decrease in load and correspondingly at the onset of the unstable shear propagation. An abrupt decrease of load
was also recorded in some other cases, however it occurred well after the achievement of a quasi-static maximum load
and also after an initial crack advancement, as frequently observed also in the case of sheet metal forming with advanced
high strength steels [38].
For these specimens and for all other tests where crack initiation was followed by quasi static ductile propagation, as pre-
viously reported, it had initially been decided to point out crack initiation by the simplified key-curve method proposed by
Reese and Schwalbe [23], based upon the early key curve analysis developed by Joyce et al. [22] and upon the principle of
load separation [39]. According to the latter principle, the load P may be expressed as a function of crack length a and plastic
load line displacement Dlpl, by two separate multiplicative functions, i.e.:
P ¼ Gða=WÞ  HðDlpl =WÞ ð2Þ
188 M. Faccoli et al. / Engineering Fracture Mechanics 127 (2014) 181–193

where G(a/W) and H(Dlpl/W) can be defined as geometry function and deformation functions. Landes et al. [27] demonstrated
that G(a/W) can be determined from the J calibration for the specimen geometry. As a consequence the load normalized by
the geometry function, PN, is a function of only plastic displacement:
PN ¼ P=Gða=WÞ ¼ HðDlpl =WÞ ð3Þ
Landes and co-workers attempted to correlate the normalized load with the plastic deformation via the H(Dlpl/W) function. If
the deformation function is known, the relationship between load, load line displacement and crack length is known at every
point.
Reese and Schwalbe [23] proposed an alternative approach, also based on the principle of load separation. The procedure
needs the load versus load-line displacement curve and the final crack length. It correlates the change in the normalized load
due to crack growth DPN to the crack advancement Da. In this way the linear relationship that exists between DPN and Da is
used, allowing the determination of Da values at any load level in the plastic regime. The geometry function G(a/W) can be
taken from the literature [27] and DPN can be expressed by:
DPN ¼ PN ðaf Þ  PN ða0 Þ ð4Þ
where a0 is the initial crack length and af is the final crack length. For the present test results a simplified approach was used.
It was first considered that PN(a0) versus the normalized load line displacement can be regarded as the normalized load curve
of a specimen with a stationary non propagating crack. On the other hand PN(ai) versus the normalized load line displacement
corresponds to the load curve recorded during the test. When ai = af, the difference between the two curves corresponds to
Eq. (4), while for ai = a0 the difference between the two curves vanishes being PN(a0) = PN(ai). Therefore the normalized loading
curve of a crack propagating specimen deviates from the loading curve of a non propagating crack specimen exactly in cor-
respondence of the onset of crack propagation. In other words, provided that the difference between two specimens is given
only by the notch root radius, and that owing to the notch severity fracture initiation is anticipated in the specimen with the
smaller notch root radius, their recorded loading curves must coincide in the initial part, up to the onset of fracture.
With reference to Fig. 7, it was argued that for each set of tests the normalized load of the largest notch root radius spec-
imen can be considered, up to crack initiation, as the stationary non propagating crack PN(a0) curve. For specimens with smal-
ler notch root radii, the deviation of the normalized load curve from such PN(a0) curves is the consequence of crack initiation.
The Jc values for specimens having a notch root radius smaller than the largest one have hence been calculated at the point
where their normalized curves deviate from the normalized curve of the largest notch root radius specimen, i.e. of a non
propagating crack specimen.
The Jc values, obtained from the tests, for the investigated steels are shown in Fig. 8 as a function of the DENT specimens
notch tip radius q.
Fig. 8 shows that Jc progressively increases with the notch tip radius q for all the studied steels, demonstrating a notable
sensitivity of ductility upon notch severity. No effect of the rolling direction on the notch ductility was observed both in the
TWIP steel and in the DP1000 steel. In particular a linear relationship can be observed between crack initiation Jc in the stud-
ied blunt notch specimens and the notch tip radius q, with a straight line passing through the origin, as also theoretically

Fig. 8. Critical J-integral Jc as a function of the notch root radius q.


M. Faccoli et al. / Engineering Fracture Mechanics 127 (2014) 181–193 189

demonstrated by Rice [40] in his original paper where J was proposed as a crack or notch tip field parameter. Actually, the
physical meaning of such a relationship cannot be extended down to the origin as this would mean that for a precracked
specimen, whose notch root radius can be considered null (q = 0), a fracture toughness equal to zero should be measured.
Indeed, as already found by many authors [41–53], the fracture toughness is independent of q when q is smaller than a crit-
ical notch radius qeff, up to which a blunt notch has the same effect of a crack, and is linearly related to q for q > qeff. A linear
square fit, including the origin, was then used to account for the dependence of Jc on q.
The slope of the straight line through the origin can be related to the maximum strain acting at the notch tip, emax, by the
following equation proposed by Rice [40]:

 1=ð1þNÞ
ðN þ 1=2ÞðN þ 3=2ÞCðN þ 1=2Þ J
emax  eY ð5Þ
Cð1=2ÞCðN þ 1Þ rY eY q

where N is the strain hardening exponent of the steel, C is the mathematical gamma function, rY is the yield strength of the
steel and eY is the strain up to the yield. This equation was derived assuming a power law flow rule, r = rY(e/eY)N for the plas-
tic behavior of the material, which can be considered to account for the overall mechanisms influencing the hardening
behavior of the steel matrix. Eq. (5) can be interpreted as stating that if a critical notch root strain, emax, is the appropriate
parameter for crack initiation at the root of a notch, then a plot of the applied J at the onset of crack growth vs. notch root
radius will be a straight line passing through the origin. In plane strain tests on different microstructure steels failing either
by ductile or brittle mechanisms, both qeff and emax were found to be directly related to the flow properties and microstruc-
tural characteristics of the tested steel [47,48]. In particular qeff was found to be equal to the mean interspacing distance
between major non metallic inclusions for steels failing by ductile microvoid nucleation and coalescence [46,47] or to the
mean grain size in the case of brittle intergranular fracture [42,43]. Akourri et al. [51] found that qeff is approximately equal
to the size of the process fracture zone.
In the present research work the value of qeff can be inferred from the experimental data only for the TWIP steel, for which
specimens having a notch root radius smaller than qeff were tested. In this case qeff appears to be of the same order of mag-
nitude as the mean interparticle spacing between major inclusions (approximately 160 lm, as reported in Table 4). No rela-
tionship between qeff and inclusion spacing can be established in the case of DP1000 and Q&P steels, for which qeff cannot be
determined from the experimental data. Tests on DENT samples having notch tip radius smaller than those evaluated in the
present work are needed for DP1000 and Q&P steels in order to find their qeff value.
In the case of plane strain ductile fracture of C–Mn steels the correspondence of qeff and the mean interparticle spacing
can be interpreted on the basis of a ductile fracture model that considers the development of microvoids ahead of the crack
front and their coalescence for the advancement of the crack front [49,52]. A preliminary investigation of the fracture sur-
faces was then carried out in order to verify the fracture mechanism responsible for the crack advancement. From a macro-
scopic point of view (Fig. 9a, c and e) a more or less developed flat triangular region is observed in front of the notch tip,
which then twists into the final slant fracture, i.e. a mixed Mode I & III failure at 45°, as is traditionally described by most
fracture mechanics textbooks for metal sheets fracture [54,55]. The extension of the triangular flat fracture is not uniquely
dependent on the specimen thickness or on the fracture resistance of the steel. It is most likely influenced by the complex
stress–strain field ahead of the notch root and hence is also influenced by the notch root radius, the yield strength and the
strain hardening behavior of the steel. Fig. 9b, d, and f show at a higher magnification the fracture surface in the central por-
tion of the specimen just ahead of the notch root, that is in correspondence of the initial part of the triangular flat region.
From a microscopic point of view, the fracture initiation and propagation occur by a ductile microvoid nucleation and coa-
lescence mechanism. Fracture initiation ahead of the notch root evidently occurs under Mode I in the flat triangular region.
The latter, however, occupies only a limited portion of the specimen thickness and starts to develop in correspondence of a
very limited through thickness contraction. In fact crack initiation from the notch root occurs well before deformation and/or
crack could be observed on the pictures taken with the side camera during the test. This does not mean that fracture initiates
under overall plane strain condition, but that at least in the central portion of the specimen, just for microvoid nucleation
and growth a certain level of stress triaxiality is anyway needed. The shape of larger microvoids nucleated at isolated, not
aligned or elongated inclusions is in fact spherical and not ellipsoidal, as a result of an enlargement under an almost local
triaxial stress field [56]. During further loading the through thickness stress rzz ceases to grow at the center of the specimens
because of the limited thickness and the state of stress turns to a substantially biaxial state with crack propagation occurring
by tear on a 45° plane.
The TWIP steel’s fracture surface (Fig. 10) presents many large microvoids mainly nucleated at groups of small titanium
nitrides (also observed in Fig. 3a using the light optical microscope), which are often aligned along the rolling direction. It is
supposed that coalescence of large microvoids, however, does not occur by continuous increase of their size but through the
formation of notably smaller microvoids. No indications have been found in the literature to confirm this hypothesis, as none
research work has been found on fracture modes in TWIP steels notched samples. Up today few works have also been carried
out on the study of fracture mechanisms in unnotched samples [57,58] and no clear conclusion can be drawn. It seems that
twinning, dynamic strain ageing and strain rate sensitivity play an important role regarding the occurrence of a slant frac-
ture. A recent publication by Lorthios et al. [59] shows that fracture consists of a large strings of voids elongated along the
rolling direction. However, their volume fraction is low and it cannot account for the final fracture. It is thus supposed that
190 M. Faccoli et al. / Engineering Fracture Mechanics 127 (2014) 181–193

Fig. 9. Fracture surfaces of: (a and b) TWIP steel’s sample with q = 0.05 mm; (c and d) DP1000 steel’s sample with q = 0.18 mm; (e and f) Q&P1 steel’s
sample with q = 0.27 mm.

Fig. 10. Fracture surface of a TWIP steel’s sample and a detail of titanium nitrides in the squared zone.
M. Faccoli et al. / Engineering Fracture Mechanics 127 (2014) 181–193 191

Fig. 11. Fracture surface of a DP1000 steel’s sample and a detail of the fibrous appearance of the fracture in the squared zone.

Fig. 12. Fracture surface of a Q&P1 steel’s sample and a detail of a secondary crack in the squared zone.

the latter occurs by a sudden intense nucleation and rapid coalescence of secondary microvoids, resulting in microdimples at
the fracture surface.
The DP1000 steel manifested a fibrous appearance of the fracture in the orthogonal direction to the rolling (Fig. 11) with
elongated microvoids nucleated in correspondence of alignments of small non metallic inclusions.
The Q&P1 steel’s fracture surface is shown in Fig. 12: the larger microvoids develop in correspondence of the numerous
micro-cracks oriented along the rolling direction and perpendicular to the main fracture plane of the sample. These second-
ary cracks are probably caused by a decohesion phenomenon whose causes are yet to be investigated.
From a macroscopic point of view, with the exception of the small central portion of the flat triangular region, fracture
develops along the oblique planes of maximum shear stress under a state of stress approaching a plane stress condition
for TWIP and Q&P steel specimens. In the case of DP1000 specimens, the thickness of 1.85 mm is large enough to cause a
more apparent mixed fracture, flat and oblique, with the former corresponding a more pronounced development of a mod-
erately triaxial state of stress somewhat closer to plane strain at the middle volume of the specimen through section.
In any case, notwithstanding in the present fracture tests nucleation and coalescence of microvoids at the onset of crack
advancement do not occur under a fully developed plane strain triaxial state of stress as in the case of the previously inves-
tigated ductile steels [48], the correspondence between inter-particle spacing and qeff that has been proved for the TWIP
steel and that must be further investigated for the DP100 and Q&P steels, most probably can be interpreted on the basis
of the previously proposed ductile fracture mechanism [49]. However, the fracture surfaces of the studied steels deserve
to be investigated with further and more exhaustive details in order to ascertain the effective role of non metallic inclusions
in the development of an advancing crack front ahead of a fatigue precrack or of a finite notch root radius smaller than qeff.
With reference to Fig. 8, when q is greater than the critical notch radius qeff, then fracture initiation seems to be controlled by
the achievement of a critical strain at the root of the notch, emax. This critical notch root strain can be viewed as the maximum
strain that can be sustained ahead of the notch tip up to the onset of fracture advancement.
emax, whose value can be directly calculated from the slope of the J-q plot for q > qeff by means of Eq. (5), depends on both
the yield and flow properties of the steel and on its fracture resistance and can be ultimately considered as an appropriate
parameter to characterize the notch ductility of a steel.
emax values were calculated for the experimental data in Fig. 8. The obtained values are reported in Table 5 along with
strain to fracture, efracture, and strain up to maximum load, emax load, measured from tensile tests.
192 M. Faccoli et al. / Engineering Fracture Mechanics 127 (2014) 181–193

Table 5
Calculated emax, with efracture and emax load measured from tensile tests.

Steel emax efracture emax load


TWIP 0.66 0.47 0.46
DP1000 0.34 0.13 0.07
Q&P1 0.20 0.21 0.16
Q&P2 0.19 0.12 0.08

According to the tensile curves of Fig. 1 and the data in Table 5, the TWIP steel shows the lowest tensile strength, the
highest ductility and it is also in absolute the least sensible to the presence of a notch. The strain hardening rate of the TWIP
steel is lower than that of the DP1000 and the Q&P steels, but its local hardening capacity is higher, resulting in an increased
uniform elongation. The excellent capacity of local hardening of TWIP steels is mainly related to their single phase structure
compared with DP, as also observed by Ding et al. [60], and Q&P steels. However, the strengthening and deformation mech-
anisms controlling the strain hardening behavior of these alloys are still unclear. The different ranking obtained from the test
on notched specimens with respect to the tensile behavior is most probably to be related to the different state of stress. This
aspect, however, deserve to be further investigated. On the basis of the proposed method to measure a steel notch ductility
starting from fracture tests on notched specimens, the Q&P steels show the worst behavior, notwithstanding the ductility
parameters from the tensile test are better than those of the DP1000 steel.

4. Conclusions

In this work the notch ductility of a TWIP steel, a DP steel and two Q&P steels for automotive components was investi-
gated by studying the resistance to fracture initiation in notched tensile specimens with notches of various root radii q. The
results from notched tensile specimens allow to graduate the materials behavior with respect to formability, in terms of the
maximum strain that can be sustained ahead of the notch tip up to fracture initiation, emax. This parameter can be inferred
from the relationship between critical J integral and notch root radius q proposed by Rice.
The fracture mechanisms were also investigated in all studied steels by means of scanning electron microscope. The
obtained results can be summarized as follows.
General conclusions:

 In all tests fracture initiation and propagation occur by a ductile microvoid nucleation and coalescence mechanism, the
fracture toughness was therefore assessed in terms of the critical elasto-plastic J-integral parameter at crack initiation, Jc.
 Jc progressively increases with the notch tip radius q for all the studied steels, demonstrating a notable sensitivity of duc-
tility upon notch severity.
Further conclusions have been found for the TWIP steel.
 When q is smaller than a critical notch radius qeff, the fracture toughness is independent of q and the blunt notch has the
same effect of a crack.
 When q is greater than a critical notch radius qeff, fracture initiation seems to be controlled by the achievement of a crit-
ical strain at the root of the notch, emax, which can be viewed as the maximum strain that can be sustained ahead of the
notch tip up to the onset of fracture advancement. emax was calculated from the slope of the J-q plot. It depends on both
the yield and flow properties of the steel and on its fracture resistance and can be ultimately considered as an appropriate
parameter to characterize the notch ductility of a steel. According to this parameter the TWIP steel is the least sensible to
the presence of a notch, whereas the Q&P steels show the worst behavior.
 qeff appears to be of the same order of magnitude as the mean interparticle spacing between major inclusions (approx-
imately 160 lm).

The fracture surfaces of the studied steels deserve further investigations in order to ascertain the effective role of non
metallic inclusions in crack initiation and propagation.

Acknowledgements

This work was financed by the Italian Government (Project number 2008PSH7YZ_004). The authors are grateful to Prof.
De Sanctis, University of Pisa, for providing the tensile curves and mechanical properties of the Q&P steels and to Valentina
Ferrari for her collaboration in carrying out the experiments.

References

[1] Allain S, Chateau JP, Bouaziz O. Mater Sci Eng A 2004;387–389:143–7.


[2] Huang BX, Wang XD, Rong YH, Wang L, Jin L. Mater Sci Eng A 2006;438:306–40.
[3] Jang JM, Kim SJ, Kang NH, Cho KM, Suh DW. Met Mater Int 2009;15:909–16.
M. Faccoli et al. / Engineering Fracture Mechanics 127 (2014) 181–193 193

[4] Ahn K, Yoo D, Seo MH, Park SH, Chung K. Met Mater Int 2009;15:637–47.
[5] Koyama M, Sawaguchi T, Ogawa K, Kikuchi T, Murakami M. Mater Sci Eng A 2008;497:353–7.
[6] Jin JE, Lee YK. Acta Mater 2012;60(4):1680–8.
[7] Dumay A, Château JP, Allain S, Migot S, Bouaziz O. Mater Sci Eng A 2008;483–484:184–7.
[8] De Cooman BC, Chen L, Kim HS, Estrin Y, Kim SK, Voswinckel H, State of the science of high manganese TWIP steels for automotive applications 2009,
In: Haldar A, Suwas S, Bhattacharjee D, editors, Microstructure and Texture in Steels, Springer. [Chapter 10]
[9] Gutierrez-Urrutia I, Raabe D. Acta Mater 2011;59:6449–62.
[10] Kwon O, Proc. of 1st Int. Conf. on High Mn steels, KIM, Seoul, 2011, CD-ROM.
[11] Chun YS, Park KT, Lee CS. Scr Mater 2012;66:960.
[12] Shun T, Wan CM, Bryne JG. Acta Mater 1992;40:3407.
[13] So KH, Kim JS, Chun YS, Park KT, Lee YK, Lee CS. ISIJ Int 2009;49(12):1952–9.
[14] Bouaziz O, Allain S, Scott CP, Cugy P, Barbier D. Curr Opin Solid State Mater Sci 2011;15(4):141–68.
[15] Chung K, Ahn K, Yoo D, Chung K, Seo M, Park S. Int J Plasticity 2011;27:52–81.
[16] Speer JG, Streicher AM, Matlock DK, Rizzo F, Krauss G. Quenching and partitioning: a fundamentally new process to create high strength trip sheet
microstructures. Materials Science and Technology 2003 Meeting, eds. Damm E.B. & Merwin M.J., 9–12 November 2003, p. 505.
[17] Wang L, Speer JG. Metallogr Microstruct Anal 2013;2:268–81.
[18] Santofimia MJ, Zhao L, Sietsma J. Metall Trans A 2011.
[19] Santofimia MJ, Zhao L, Sietsma J. Metall Trans A 2009;40A:46–57.
[20] Santofimia MJ, Speer JG, Clarke AJ, Zhao L, Sietsma J. Acta Mater 2009;57:4548–57.
[21] Matteis P, Scavino G, D’Aiuto F, Firrao D. Steel Res Int 2012;83:950–6.
[22] Joyce JA, Ernst HA and Paris PC, Direct evaluation of J-Resistance Curves from load displacement records, In: Fracture Mechanics: Twelfth Conference,
ASTM STP 700, American Society for Testing and Materials, Philadelphia, 1980, pp. 222-236.
[23] Reese ED, Schwalbe KH. Fatigue Fract. Engng Mater. Struct. 1993;16:271–80.
[24] Landes JD, Herrera R. Int J Fracture 1988;36:R9–R14.
[25] Herrera R, Landes JD. J Test Eval 1988;16(5):427–49.
[26] Sharobeam M, Landes JD. Int J Fracture 1991;47:81–104.
[27] Zhou Z, Lee K, Herrera R, Landes JD. Normalization: an experimental Method for developing J-R curves, elastic-plastic fracture test methods: the user’s
experience. In: Joyce JA, editor. ASTM STP 1114, vol. 2. Philadelphia: American Society for testing and Materials; 1991. p. 42–56.
[28] El-Soudani SM. Quantitative fractography and fracture mechanics characterization. JOM 1990:20–7.
[29] Rice JR, Paris PC, Merkle JG. Some further results of J-integral analysis and estimates. Progress if flaw growth and fracture toughness testing. ASTM STP
1973;536:231–45.
[30] Anderson TL. Int J Fracture 1989;41:79–104.
[31] Mai YW, Cotterell B. Eng Fract Mech 1995;21:123–8.
[32] Deng H, Wang T. Advanced in Fracture Research ICF-7 1989:323–31.
[33] Sun J, Deng Z, Tu M. Eng Fract Mech 1990;37:675–80.
[34] Hutchinson JW. J Mech Phys Solids 1968;16:13–31.
[35] Rice JR, Rosengren GF. J Mech Phys Solids 1968;16:1–12.
[36] Yuan FG, Yang S. Int J Fracture 1997;85:131–55.
[37] Yan C, Mai Y-W. Int J Fracture 1998;91:117–30.
[38] Li Y, Luo M, Gerlach J, Wierzbicki T. J Mater Process Technol 2010;210:1858–69.
[39] Ernst H, Paris PC, Landes JD, Estimations on J-integral and tearing modulus from a single specimen test record. In: Richard Roberts editor, Thirteenth
Conference, Fracture Mechanics, ASTM STP 743, American Society for Testing and Materials, Philadelphia, 1981. p. 476–02.
[40] Rice JR. J Appl Mech, Trans ASME E 1968;35:379–86.
[41] Tetelman AS, Wilshaw TR, Rau CA. Int J Fract Mech 1968;4:147.
[42] Ritchie RO, Geniets LCE, Knott JF. In: The microstructure and design of alloys, proc 3rd int conf strength metals and alloys, vol. 1, London: Institute of
Metals/Iron and Steel Institute; 1973. p. 124–28.
[43] Ritchie RO, Knott JF, Rice JR. J Mech Phys Solids 1973;21:395.
[44] Ritchie RO, Francis B, Server WL. Metall Trans A 1976;7A:831–8.
[45] Ritchie RO, Server WL, Wullaert RA. Metall Trans A 1979;10A:1557–70.
[46] Firrao D, Begley JA, De Benedetti B, Roberti R, Silva G. Scripta Metall Mater 1980;14:519.
[47] Firrao D, Begley JA, Roberti R, Silva G, De Benedetti B. Metall Trans 1982;13A:1003.
[48] Firrao D, Roberti R. Metall Sci Technol 1983;1:5.
[49] Firrao D, Roberti R, On the mechanism of ductile fracture nucleation ahead of sharp cracks. In: Wells J, Landes J editors, proceedings of the symposium
‘‘Fracture: Interactions of microstructures, Mechanisms and Mechanics’’, 113th AIME Annual Meeting, Los Angeles, California, TMS-AIME, Warrendale,
PA, 1984. p. 165–67.
[50] Cao WD, Liu XP. Metall Trans A 1984;20:1689.
[51] Akourri O, Louah M, Kifani A, Gilgert G, Pluvinage G. Eng Fract Mech 2000;65:491–505.
[52] Roberti R, Silva G, Firrao D, De Benedetti B. Int J Fatigue 1981;3:133–41.
[53] Firrao D, Roberti R, Silva G. In: Maurer KL, Matzer FE editors, Fracture and the role of microstructure, Emas, Warley, vol. ll, 1982, p. 727–35.
[54] Knott JF. Fundamentals of fracture mechanics. Butterworth & Co Publishers; 1973.
[55] Anderson TL. Fracture mechanics: fundamentals and applications. 3rd ed. CRC Press; 2004.
[56] Rice JR, Tracey DM. J Mech Phys Solids 1969;17:201–17.
[57] Bayraktar E, Khalid FA, Levaillant C. J Mater Process Technol 2004;147:145–54.
[58] Abbassi M, Kheirandish S, Kharrazi Y, Hejazi J. J Mater Sci Eng A 2009;513–514:72–6.
[59] Lorthios J, Nguyen F, Gourgues AF, Morgeneyer TF, Cugy P. Scripta Mater 2009;57(13):3978–88.
[60] Ding H, Tang ZY, Li W, Wang M, Song D. J Iron Steel Res Int 2006;13(6):66–70.

You might also like