Download as pdf or txt
Download as pdf or txt
You are on page 1of 7

Int J Fract (2009) 160:143–149

DOI 10.1007/s10704-009-9414-8

ORIGINAL PAPER

The effect of aluminum alloying on ductile-to-brittle


transition in Hadfield steel single crystal
E. G. Astafurova · Yu. I. Chumlyakov ·
H. J. Maier

Received: 9 April 2009 / Accepted: 5 October 2009 / Published online: 22 October 2009
© Springer Science+Business Media B.V. 2009

Abstract The ductile-to-brittle transition (DBT) in aluminum partially suppresses twinning in steel (II).
Fe-13Mn-1.3C (Hadfield steel, I) and Fe-13Mn-2.7 Twinning sets in only after a certain amount of disloca-
Al-1.3C (Hadfield steel, II) (wt.%) single crystals ori- tion slip, but still influences the fracture mechanism of
ented along [011], [1̄44], and [1̄11] directions was steel (II).
investigated under tension in the temperature inter-
val of 77 to 673 K. The DBT temperature interval Keywords Ductile-to-brittle transition · Hadfield
was found to be independent of single crystal orienta- steel · Slip · Twinning
tion. The DBT temperatures were estimated (1) as the
mean value between the temperature corresponding to
the minimum crystal ductility and the one coinciding 1 Introduction
with the onset of the plateau of the ε(T)-dependence
(TDBT1 ); and (2) as the temperature where the vol- Pure BCC metals and alloys often exhibit a ductile-
ume fraction of brittle failure on the fracture surfaces to-brittle transition (DBT). Below the ductile-to-brittle
was 50% (TDBT2 ). The DBT temperatures estimated transition temperature (DBTT) the strength increases
this way, do not coincide for both steels. Mechani- significantly and little plastic deformation occurs prior
cal twinning has been reported as the primary reason to failure. The DBT in BCC metals often corresponds to
for the occurrence of the DBT in austenitic high-car- a change in fracture mode from a ductile fibrous one to
bon Hadfield steel and appears to account for the dif- cleavage. The fast decrease in flow stress with increase
ference in DBT temperatures as well. Alloying with in temperature is typical for these materials, and drop of
stresses for plastic flow suppresses cleavage and favors
E. G. Astafurova (B)
ductile void growth (Riedel 1993). FCC materials show
Laboratory of Physical Materials Science, Institute of
Strength Physics and Materials Science SB RAS, mainly a gradual change in ductility and flow stress as
Akademichesky prospect 2/4, 634021 Tomsk, Russia a function of test temperature (Riedel 1993; Christian
e-mail: astafe@ispms.tsc.ru and Mahajan 1995). Still, a DBT occurs in some FCC
Y. I. Chumlyakov
metals and alloys, for example, in iridium (Panfilov
Laboratory of Physics of Plasticity and Strength, Siberian and Yermakov 2004), precipitation hardened Cu-Al-Co
Physical Technical Institute, Novosobornaya sq. 1, 634050 alloys and austenitic stainless nitrogen-bearing steels
Tomsk, Russia (Tomota et al. 1998a,b; Müllner 1997). Austenitic Had-
H. J. Maier
field steel (10–14 wt.% Mn, 1–1.4 wt.% C, Fe balance)
Lehrstuhl f. Werkstoffkunde (Materials Science), University also demonstrates a DBT, and in the low-temperature
of Paderborn, Pohlweg 47-49, 33098 Paderborn, Germany region it shows brittle intragranular fracture.

123
144 E. G. Astafurova et al.

Table 1 Slip and twinning


Axis orientation Slip systems SF for slip Twin systems SF for twinning
systems with the highest
Schmid factors (SF) for          
1̄11 [011] 1̄1̄1 , 1̄10 1̄1̄1 , 0.27 1̄21 1̄1̄1 0.31
single crystals oriented
  for        
tension along 1̄11 , 1̄44 , 1̄01 1̄11̄ , [011] 1̄11̄ , 1̄12 1̄11̄
     
[011] directions 1̄10 (111), 1̄01 (111) 2̄11 (111)
       
1̄44 1̄01 (111), 1̄10 (111) 0.42 2̄11 (111) 0.50
     
[011] 1̄01 (111), 1̄10 (111), 0.41 2̄11 (111) 0.47
     
[110] 1̄11 , [101] 1̄11 [211] 1̄11

Hadfield steel is an austenitic manganese steel, Twinning, dislocation slip and strain aging all con-
which possesses a high strain hardening rate, excellent tribute to the fracture behavior in Hadfield steel.
wear resistance and toughness. This combination The relevance of the individual effects is, however,
makes this steel attractive for industrial applications not clearly understood. In austenitic stainless steels
in the fields of earthmoving, mining, quarrying, oil that underwent twinning, cleavage-like fracture was
well drilling, steelmaking and railroading. Both the observed to arise due to the intersection of mechanical
high manganese and carbon contents in Hadfield steel twins (Müllner et al. 1994). The process of twin–twin
reduce the martensite start temperature, and thus the or twin–slip interaction is accompanied by emission of
austenite is stable down to the temperature of liq- dislocations inside and outside of the twin acting as the
uid nitrogen (77 K). The unusually high strain hard- obstacle or along the twin boundary. When the thick-
ening of Hadfield steel was attributed to the forma- ness of a twin is lower than the critical value needed to
tion of twin boundaries, which are strong barriers to nucleate a dislocation inside or outside of the twin, then
dislocation motion (Adler et al. 1986; Raghavan et shear transfer over the twin boundary becomes difficult
al. 1969). Another reason that explains the unusual and cleavage-like fracture occurs (Müllner et al. 1994).
work hardening behavior is dynamic strain aging pro- In the present study single crystals were employed
duced by the reorientation of carbon in C–Mn couples to exclude the contributions of the grain boundaries
in the cores of dislocations (Dastur and Leslie 1981; to work hardening and fracture. Moreover, the choice
Owen and Grujicic 1999), which favors fast dislocation of specimens with the tensile axes oriented along the
accumulation. [011], [1̄44], and [1̄11] directions allowed to study
Deformation twinning in Hadfield steel occurs by strain hardening and fracture under conditions where
discontinuous plain splitting of a/2 < 110 > perfect slip mechanical twinning dominates, as the Schmid factors
dislocations into two a/6 < 211 > Shockley partials joint for twinning in these orientations are higher than those
by stacking faults on {111} slip planes, which are also for slip (Table 1) (Berner and Kronmüller 1965). More-
the habit planes. This results in a diffusionless shear of over, the variation of the initial orientations from [1̄11]
a plate-shaped region of the parent crystal, which dif- to [011] and [1̄44] also changes the number of twinning
ferentiates it from the surrounding. Mechanical twin- systems (Table 1), which in turn provides information
ning in FCC metals is less common compared to BCC on the influence of multiplicity of twinning on the frac-
or HCP materials (Christian and Mahajan 1995). In ture mechanism.
Hadfield steel, however, the high solute concentrations Aluminium increases the stacking fault energy of
simultaneously raise the shear stress for slip (mainly Hadfield steel to a value of 0.05 J m−2 and reduces
due carbon) and reduce the stacking fault energy (man- the mobility of carbon in solid solution (Zuidema et
ganese) to γSF = 0.03 J m−2 (Adler et al. 1986). The al. 1987). Thus, alloying of Hadfield steel with alu-
stacking fault width is inversely proportional to the minium suppresses both twinning and dynamic strain
stacking fault energy, which makes dislocation cross aging (Zuidema et al. 1987; Astafurova et al. 2007a;
slip more difficult. Moreover, the stress for nucleation Canadinc et al. 2005). The unusually high strain hard-
of deformation twins is also proportional to the stack- ening coefficient of G/23 (G-shear modulus) observed
ing fault energy, and thus, a lower stacking fault energy in the aluminum-alloyed Hadfield steel single crystals
increases the tendency for twinning. (2.6% Al, in wt.) is attributed to the formation of high

123
The effect of aluminum alloying on ductile-to-brittle transition in Hadfield steel single crystal 145

density dislocation walls due to planar dislocation slip. 3 Results and discussion
However, twinning is also present as a secondary
deformation mechanism in aluminum-alloyed Hadfield The effect of temperature on uniform elongation εp and
steel, although the volume fraction is small (Astafurova critical resolved shear stress τcr for steels (I) and (II) is
et al. 2007b; Canadinc et al. 2005). Thus, using the presented in Fig. 1, which demonstrates that the DBT
[011], [1̄44], and [1̄11] single crystals of aluminum- temperature interval does not depend on the single crys-
alloyed Hadfield steel allowed for investigation of the tal orientation.
deformation behaviour and fracture behaviour under The strong temperature dependences of critical
conditions where the activity of mechanical twinning resolved shear stress (Fig. 1) is normally not observed
and dynamic strain aging are reduced as compared to in conventional FCC materials. The rapid increase
the aluminium-free steel. in τcr below 200 K is a result of the high content
of interstitials in Hadfield steel (1.3 wt.% carbon) as
thermal activation processes become less likely at low
2 Methods temperatures (Berner and Kronmüller 1965). Simi-
lar strengthening mechanisms have been reported for
The materials used in this study were Hadfield steel nitrogen-alloyed in austenitic steels with low stacking
(Fe-13 Mn-1.3C wt.%, referred to as steel I) and alu- fault energy (Chumlyakov et al. 1996).
minium-alloyed Hadfield steel (Fe-13 Mn-2.7 Al-1.3C,
steel II). Single crystals were grown by the Bridgman
technique in an inert gas atmosphere. All crystals were
then homogenized in an argon atmosphere at 1373 K for
24 h. Electro-discharge machining was utilized to cut
regular dog-bone shaped flat tensile specimens with
nominal dimensions of 12 mm × 3 mm × 1.5 mm in
the gauge section. The tensile specimens were solu-
tion-treated at 1373 K for 1 h and then water-quenched.
Finally, grinding and an electrochemical polish (50 g
CrO3 , in 200 g H3 PO4 ) were employed to remove the
entire processing-affected surface layer.
The orientation of the crystal axis was determined
using a DRON-3 M X-ray diffractometer. Tensile tests
were conducted in the temperature interval of 77–673 K
at a strain rate of 4 × 10−4 s−1 . The critical resolved
shear stresses were determined as τcr = mσ0.2 , where
σ0.2 is the 0.2% offset yield stress and m is the Schmid
factor for slip (assuming that slip precedes twinning
for all test temperatures). The uniform elongation was
determined as the maximum uniform plastic strain (εp )
before fracture. Experiments were repeated on five
companion specimens to check repeatability. For frac-
ture analysis a Philips SEM 515 scanning electron
microscope (SEM) was utilised.
In order to characterize the microstructural changes
occurring during deformation, tensile tests were inter-
rupted at given strains, and electron transparent foils
Fig. 1 The effect of temperature on critical resolved shear
were prepared from these samples by conventional stresses τcr (curves 1) and uniform elongation εp (curves 2, 3)
twin-jet electropolishing. For microstructural analysis, for steel (I) (a) and steel (II) (b). a Curves 1, 2–[1̄11] orientation,
a Philips CM 200 transmission electron microscope steel (I); 3–[1̄44] orientation, steel (I), b 1, 2–[1̄11] orientation,
(TEM) operated at 200 kV was employed. steel (II); 3–[011] orientation, steel (II)

123
146 E. G. Astafurova et al.

Fig. 2 SEM micrographs


for the [1̄11]-oriented single
crystals of steel (I) (a, c, e)
and steel (II) (b, d, f): a,
b–T = 77 K; c, d–T = 295 K;
e, f–T = 673 K

Both Hadfield steels show a non-linear increase in the plane facets of cleavage lie along crystallographic
plastic strain with temperature (Fig. 1). For steel (I), the planes of the {111}-type, which are slip and twinning
temperature interval can be divided into two regimes. planes for FCC crystals. Moreover, steel (II) not only
For T > 200 K temperature has only a minor effect on demonstrates brittle fracture in tension at T = 77 K,
εp and τcr , whereas below 200 K, an abrupt embrit- but fracture surface is divided into small cleavage-like
(I) fragments, which are misoriented with respect to each
tlement occurs (Fig. 1a). The DBT temperature TDBT1
of ∼ 140 K for steel (I) was estimated as the mean other (Fig. 2b). Macroscopically the fracture surface
value between the temperature corresponding to the of steel (II) is oriented normal to the applied tensile
minimum crystal ductility (77 K) and the temperature stress, whereas on the micro scale the small flat regions
(∼200 K) corresponding to the onset of the plateau are close to {111} cleavage planes. As temperatures is
observed in the εp (T)-curve (Fig. 1a). The DBT regime increased, the fracture surfaces of both steels start to
for steel (II) is not so well defined, but can be recog- develop a more complex morphology revealing a com-
nized using the change in the dεp /dT slope. Alloying bination of ductile and brittle regions (Fig. 2c, d) up to
with aluminum widens the DBT interval and moves the about 473 K. Finally, at T = 673 K the fracture mode
(II)
DBT temperature to TDBT1 ∼ 180 K (Fig. 1b). is completely ductile for both steels (Fig. 2e, f).
Typical SEM images of fracture surfaces for [1̄11]- With the increase in test temperature, the transition
oriented single crystals of steels (I) and (II) at temper- to ductile fracture sets in a bit earlier for steel (II) as
atures of 77, 295, and 673 K are shown in Fig. 2. For compared to steel (I). This is demonstrated in Fig. 3a,
[011], [1̄44], and [1̄11]-oriented single crystals of steel which shows the amount of brittle fracture components
(I) at T = 77 K, a cleavage-type fracture takes place measured on the fracture surfaces of failed specimens.
(Fig. 2a). As revealed by X-ray diffraction and TEM, A commonly used estimate for the DBT temperature

123
The effect of aluminum alloying on ductile-to-brittle transition in Hadfield steel single crystal 147

is the temperature at which the fraction of the brittle


component on the fracture surfaces reaches 50%. These
(I)
temperatures were determined as TDBT2 ∼ 350 K and
(II)
TDBT2 ∼ 300 K for steels (I) and (II), respectively
(Fig. 3a).
Clearly, DBT temperatures determined based on the
change in ductility (Fig. 1) are different from those
calculated based on fractographic appearance (Fig. 3).
Consequently, the temperature interval studied was
divided into three regimes as demonstrated in Fig. 3b
for the example of a [1̄11]-oriented single crystals of
steel (I). The boundaries of these regions are not exactly
defined but were chosen according to the following
considerations. Regime A corresponds to brittle frac-
ture in a general sense that is confirmed both by the
cleavage-like fracture mode and a low elongation to
failure. Regime B is a wide transition range, where
ductility can be as high as for the pure ductile mode,
but fracture surfaces demonstrate brittle behavior. For
steel (II), regime B is significantly narrower than for
steel (I), but, nevertheless, exists. Finally, Regime C is
the temperature interval where both data for elonga-
tion to failure as well as fractography indicate ductile
behavior. Fig. 3 a Temperature dependence of the amount of brittle com-
ponent on fracture surface for [1̄11]-oriented specimens of steels
The reason for the discrepancy between the brittle (I) (curve 1) and (II) (curve 2), and the mean twin thickness for
fractographic appearance and high ductility in Regime steel (I) (curve 3), b the amount of brittle component on fracture
B appears to be the temperature dependence of mor- surface (curve 1), uniform elongation ε (curve 2) depending on
phology of mechanical twinning, i.e. twin thickness test temperature for [1̄11]-oriented specimens of steel (I)
and density. As expected, twinning is active in the tem-
perature interval of 77–473 K for steel (I) (Fig. 4).
By contrast, alloying with aluminum should suppress main reason for the shift of the onset of twinning to
twinning in steel (II). Still, the basic features of the larger plastic strains and to lower temperatures. In fact,
DBT remained the same for both steels. The similar- extensive TEM investigations of dislocation arrange-
ities observed in the fracture modes of both steel can ments in single crystals of steel (II) demonstrated that
be rationalized if one assumes that aluminum alloying (i) independent of the initial tensile axis, no twins were
initially decreases the intensity of mechanical twinning observed at T > 473 K and (ii) at T = 300 K thin
in Hadfield steel, but does not fully suppress it (Asta- twins of one system were the distinctive feature of
furova et al. 2007b). the microstructure in [1̄11] crystals at ε ≥10% and
As revealed by TEM (Fig. 4a, b), mechanical in [011] crystals at ε ≥15% (Fig. 4c). Moreover, a low
twinning sets in from the very beginning of plastic temperature test on [011]-oriented samples of steel (II)
deformation (ε > 0.5%) for the [011]-, [1̄44]-, and demonstrated that a decrease in the temperature moves
[1̄11]-oriented crystals of steel (I) in the temperature the onset of twinning to earlier stages of plastic defor-
interval of 77–293 K. At higher temperatures (293– mation, i.e. twinning was observed at ε = 3–5% for
673 K) twinning also is also observed but it only sets T = 200 K (Fig. 4d) and at ε = 1% for T = 77 K.
in after substantial slip deformation. Returning to the fracture peculiarities for both steels,
As outlined before, alloying with aluminum the following assumption can be made: The twins tend
increases the stacking fault energy of Hadfield steel to be narrower in the lower the temperature (Christian
γSF from 0.03 J m−2 to 0.05 J m−2 . The rise of γSF in and Mahajan 1995). As demonstrated by TEM, twins
steel (II) as compared to steel (I) is assumed to be the are very thin at 77 K < T < 200 K in steel (I) (Fig. 3a),

123
148 E. G. Astafurova et al.

Fig. 4 TEM images shown


twinning development in
steels (I) and (II). a Steel
(I), [1̄11], ε = 2%,
T = 77 K, b steel (I),
[1̄11], ε = 5%, T = 295 K,
c steel (II), [011], ε = 20%,
T = 295 K, d steel (II),
[011], ε = 10%, T = 200 K;
arrows show twin
reflections in SAD patterns

and thus, shear transition over twin boundaries is com- (I) (Dastur and Leslie 1981; Astafurova et al. 2007a).
plicated, as the distance between twin boundaries is The high mobility of carbon in steel (I) also leads to
small for dislocation loop nucleation. In addition, the rapid recovery of short range order upon dislocation
stress level is high in this temperature regime and movement (Dastur and Leslie 1981), and thus, to fast
accommodation of stresses by slip is also suppressed dislocation accumulation. The latter provides for addi-
(Fig. 1a). The combination of both effects leads to tional barriers to the motion of dislocations especially
cleavage along {111} planes in regime A. at the twin boundaries during twin–twin or twin–slip
At higher temperatures (regime B) the mutual pen- interactions.
etration of twins and slip dislocations is much easier Despite the fact that strain hardening of the alu-
due to the increase in twin thickness and decrease in minium-alloyed steel (II) is basically governed by the
resolved critical shear stresses (Fig. 1a). Müllner and interaction of pile-ups and multipoles (Astafurova et al.
Solenthaler (1997) have also shown an effect of strain 2007b; Canadinc et al. 2005), twinning still influences
on dislocation and twin densities in austenitic steels. the fracture mode of steel (II). Alloying with alumi-
Plastic strain alters the twin density and the twin thick- num only suppresses twinning initially, and thus, the
ness. Slip and twinning in conjugate systems at temper- trends in fracture mode are similar. The differences in
atures where the intersection with twin boundaries is the DBT temperatures of steel (II) and steel (I), are thus,
possible, results in noticeable homogeneous strain, lin- attributed to the differences in temperature dependence
ear stress-strain dependence and thickening of twins of of twinning deformation, twin morphology and dislo-
the primary twin system in Hadfield steel. After con- cation arrangement.
siderable plastic strain, when twins become thin and Finally, as the twin volume fraction becomes low in
stresses are high due to strain hardening, fracture finally steel (I) or negligible in steel (II) at high test tempera-
occurs along twin boundaries. In this case, the total tures, ductile failure results for both steels in regime C.
elongation of the single crystals is high yet the frac-
ture surfaces demonstrate a brittle character (regime B).
Thus, the contribution of the conjugate twin systems to 4 Conclusions
overall strain might be small, but it is significant with
respect to the hardening behavior and fracture. Using single crystals of austenitic Fe-13 Mn-1.3C (steel
In Hadfield steel, dynamic strain aging and the high I) and Fe-13 Mn-2.7 Al-1.3C (steel II) (wt.%) Had-
mobility of carbon make the strain-rate dependence field steels oriented along [011]-, [1̄44]-, and [1̄11]
β=σ/ln ε̇ near room temperature negative in steel directions, a ductile-to-brittle transition (DBT) has

123
The effect of aluminum alloying on ductile-to-brittle transition in Hadfield steel single crystal 149

been experimentally established. The DBT temperature Canadinc D, Sehitoglu H, Maier HJ, Chumlyakov YI (2005)
interval was found to be independent of the orientation Strain hardening behavior of aluminum alloyed Hadfield
steel single crystals. Acta Mat 53:1831–1842
of the single crystals and to be influenced by the tem- Christian JW, Mahajan S (1995) Deformation twinning. Prog
perature dependence and morphology of mechanical Mater Sci 39:1–157
twinning. Chumlyakov YuI, Kireeva IV, Korotaev AD, Litvinova EI, Zuev
Twinning was revealed to be the primary deforma- YuL (1996) Mechanisms of plastic deformation, hardening,
and fracture in single crystals of nitrogen-containing austen-
tion mechanism of the [011]-, [1̄44]-, and [1̄11]-ori- itic stainless steels. Russ Phys J 39(3):189–210
ented Hadfield steel single crystals in the temperature Dastur YN, Leslie WC (1981) Mechanism of work hardening in
interval of 77 to 293 K. Alloying with aluminum sup- Hadfield manganese steel. Metall Trans 12A:749–759
presses twinning initially. After a significant amount of Müllner P, Solenthaler S, Uggowitzer PJ, Speidel MO (1994)
Brittle fracture in austenitic steel. Acta Metal Mater
slip twinning still set in and has a pronounced influence 42:2211–2217
on the fracture mode of steel II. Consequently, frac- Müllner P, Solenthaler S (1997) On the effect of deformation
ture surfaces appear brittle at intermediate temperature twinning on defect densities. Mater Sci Ing A 230:107–115
despite the high elongation to failure demonstrated by Müllner P (1997) On the ductile to brittle transition in austenitic
steel. Mater Sci Ing A 234(236):94–97
the samples. Owen WS, Grujicic M (1999) Strain aging of austenitic Hadfield
manganese steel. Acta Mater 47:111–126
Acknowledgments The work was partially supported by Rus- Panfilov P, Yermakov A (2004) On brittle fracture in polycrys-
sian Foundation for Basic Researches (project # 07-08-00064). talline iridium. J Mater Sci 39:4543–4552
Raghavan KS, Sastri AS, Marcinkowski MJ (1969) Nature of
work-hardening behavior in Hadfield’s manganese steel.
References Trans Metall Soc AIME 245:1569–1575
Riedel H (1993) Fracture mechanisms. In: Cahn RW, Haasen P,
Kramer EJ (eds) Materials science and technology: a com-
Adler PH, Olson GB, Owen WS (1986) Strain hardening of Had- prehensive treatment. Weinheim, New York, pp 565–634
field manganese steel. Metall Trans A 17:1725–1737 Tomota Y, Xia Y, Inoue K (1998a) Mechanism of low tempera-
Astafurova EG, Tukeeva MS, Chumlyakov YI (2007a) The ture brittle fracture in high nitrogen bearing austenitic steel.
effect of aluminum alloying on strength properties and Acta Mater 46:1577–1587
deformation mechanisms of the <123> Hadfield steel sin- Tomota Y, Nakano J, Xia Y, Inoue K (1998b) Unusual strain rate
gle crystals. Russ Phys J 50:959–963 dependence of low temperature fracture behavior in high
Astafurova EG, Kireeva IV, Chumlyakov YI, Maier HJ, Sehito- nitrogen bearing austenitic steels. Acta Mater 46:3099–
glu H (2007b) The influence of orientation and aluminum 3108
content on the deformation mechanisms of Hadfield steel Zuidema BK, Subramanyam DK, Leslie WC (1987) Effect of
single crystals. Int J Mat Res 98:144–149 aluminum on the work hardening and wear resistance of
Berner R, Kronmüller H (1965) Plastische verformung von eink- Hadfield manganese steel. Metall Trans 18:1629–1639
ristallen, moderne probleme der metallphysik. Springer,
Berlin

123

You might also like