You are on page 1of 7

Journal of Alloys and Compounds 457 (2008) 279–285

Grain refinement under multi-axial forging


in Fe–32%Ni alloy
Baojun Han ∗ , Zhou Xu
Key Laboratory for High Temperature Materials and Tests of Ministry of Education, Shanghai Jiao
Tong University, Shanghai 200240, PR China
Received 15 January 2007; received in revised form 11 March 2007; accepted 12 March 2007
Available online 20 March 2007

Abstract
The deformation behavior and the grain refinement mechanism of Fe–32%Ni alloy under multi-axial forging were studied by the flow stress
curve analysis and the microstructure observation. The multi-axial forging was carried out at a strain rate of 10−2 s−1 under temperature of 773
and 1073 K, respectively. The results show that the general character of flow stress curve and the microstructure of Fe–32%Ni alloy sensitively
depended on the deformation temperature. At the temperature of 773 K, the flow stress increased to a single peak with work hardening and then
showed a steady-state-like one during further deformation. The new grains are formed in the process that the initial deformation leads to the
formation of subgrain boundaries and further deformation causes the misorientations of subgrain boundaries to increase by usual intrinsic slip and
the absorption of accommodated dislocations, in which way the new ultra-fine grains are formed homogeneously. The grain refinement mechanism
is classified as continuous dynamic recrystallization (CDRX). At the temperature of 1073 K, the flow stress curve exhibited a single peak stress
followed by work softening and then a steady state. The grain refinement process is that new grains that are almost defect-free are formed by
nucleating at the bulged boundaries during deformation, then they grow and consume the deformed microstructure until the material is completely
recrystallized, which process is classified as discontinuous dynamic recrystallization (DDRX).
© 2007 Elsevier B.V. All rights reserved.

Keywords: Metals and alloys; Grain boundaries; Transmission electron microscopy (TEM); Electron backscattered diffraction (EBSD)

1. Introduction pler industrial alloys and lead to economic benefits as well as


improved recyclability [1]. It has been reported that ultra-fine
The properties of metals are strongly influenced by its grain grained materials could be produced by severe plastic deforma-
size, and the well-known Hall–Patch equation, which predicts tion (SPD) [2–4]. Several special SPD methods such as equal
an increase in yield strength (σ) with a decrease in grain size (d), channel angular pressing (ECAP) [5–7], high pressure torsion
has been shown to be applicable to a wide variety of materials: (HPT) [8,9] and mechanical milling (MM) [10–12] have suc-
ceeded in producing the ultra-fine grained materials. Whereas
σ = σ0 + Kd −1/2 the disadvantage of these processes is that they are not applicable
where σ 0 is the friction stress below which dislocations will to large bulk materials. Not small functional materials but large
not move in a single crystal and K a constant. The benefits of structural materials do require high strength and toughness. The
grain refinement are the improvements in fracture resistance and accumulative roll bonding (ARB) had been used successfully
the phenomenon of superplasticity in metals with grain sizes to produce ultra-fine grains in a number of large bulk materials
less than about 10 ␮m. Another reason for the current interest [13–15], however, the ARB was easy to introduce impurities
in ultra-fine grained materials is that it would be more desir- into the deformed materials.
able to control its mechanical properties by processing than Multi-axial forging (MF) is also one method to exert large
by alloying, since it would result in the use of fewer and sim- strain on bulk materials and may be the most effective way [16].
However, much work remained to be done about this kind of
processing method and the microstructure evolution mechanism
∗ Corresponding author. Tel.: +86 21 54742618; fax: +86 21 62829641. is not clear yet. In the present work, the deformation behavior
E-mail address: Hanbaojun80@tom.com (B. Han). and the microstructure evolution during MF were studied by the

0925-8388/$ – see front matter © 2007 Elsevier B.V. All rights reserved.
doi:10.1016/j.jallcom.2007.03.067
280 B. Han, Z. Xu / Journal of Alloys and Compounds 457 (2008) 279–285

Fig. 1. The optical micrographs of Fe–32%Ni alloy before multi-axial forging


(MF). Fig. 3. The flow stress curves of Fe–32%Ni alloy compressed at a strain rate of
10−2 s−1 and temperature of 773 and 1073 K, respectively.

TEM and the EBSD observation. The microstructure evolution 3. Results and discussions
mechanism was discussed in detail in this paper.
3.1. The stress–strain (σ–ε) curves analysis
2. Experimental procedures
Fig. 3 shows the typical stress–strain (σ–ε) curves for
The material used in the present experiment was Fe–32%Ni alloy (com-
position in wt.%: 0.007C, 0.01Si, 0.04Mn, 0.005P, 0.0006S, 32.4Ni, 0.022Al, Fe–32%Ni alloy tested at a strain rate of 10−2 s−1 under tem-
0.00074N, 0.020O, and the balance Fe). The alloy fits for the investigation of the perature of 773 and 1073 K, respectively. It is clearly seen that
microstructure evolution during deformation because its martensite starting tem- the general character of flow curves sensitively depended on the
perature (Ms) is below the room temperature. The alloy was annealed at 1473 K deformation temperature. The flow stress curve of 1073 K exhib-
for 60 min before MF, and the initial optical micrograph was shown in Fig. 1.
ited a typical type of dynamic recrystallization (DRX) one, i.e.
The samples were machined into cubic with starting dimensions of
16 mm × 16 mm × 16 mm. Preheated at the deformation temperature for 600 s, a single peak flow stress followed by work softening, and then a
the samples were alternately forged with loading direction changed through 90◦ steady state at large strain [17]. On the other hand, any softening
(see Fig. 2) by Gleeble3500, using graphite as lubricant to make the specimens did not appear during forging at the temperature of 773 K, and at
deformed homogeneously. After each stage of forging, the curved surfaces were which temperature, the flow stress approached a saturation value
cut by electric discharge machining. The forgings were carried out at a strain
of about 300 MPa gradually and then showed a steady-state-like
rate of 10−2 s−1 under temperature of 773 and 1073 K, respectively. The strain
achieved in each stage of forging was about 0.5. The samples were taken out one during further deformation. The stress–strain curve in such
quickly and quenched into cold water after each stage of forging. a shape is similar to that appears under dynamic recovery (DRV)
Microstructure observations were carried out on sections parallel to the com- [18].
pression axis using EBSD and Hitachi H800 TEM. Thin foils for the TEM
investigations were twin-jet polished and samples for the EBSD investigations
were grounded and then electro-polished in a solution of 5% perchloric acid and 3.2. The microstructural evolution
95% ethanol at 223 K. The EBSD data was analyzed using orientation imag-
ing microscopy (OIM) software of TSL Lab. Boundaries with misorientations According to the flow stress behavior of Fe–32%Ni alloy
between 2◦ and 15◦ were defined as low angle boundaries (LABs) and those of
higher than 15◦ as high angle boundaries (HABs). In the EBSD maps presented in
(Fig. 3), it can be deduced that different microstructure evolution
this paper, the grain boundaries were colored according to their misorientations. mechanism may operate under different deformation tempera-
The LABs were depicted as thin grey lines, and the HABs as thick black lines. ture.

Fig. 2. Diagrammatic representation of the multi-axial forging (MF) process.


B. Han, Z. Xu / Journal of Alloys and Compounds 457 (2008) 279–285 281

Fig. 4. The typical micrographs of Fe–32%Ni alloy evolved at 1073 K under multi-axial forging with cumulative strain of (a) 0.5, (b) 1.0, and (c) 1.5.

3.2.1. Microstructure evolution at 1073 K ture was almost completely replaced by new grains when the
A series of typical micrographs of Fe–32%Ni alloy evolved at cumulative strain was 1.5 (Fig. 4c). It should be pointed out that
1073 K with different cumulative strains are shown in Fig. 4. The the microstructure was not homogeneous eventually.
main feature of the structure changes during MF at 1073 K was Fig. 5 shows the typical EBSD image quality maps of
the grain refinement. And it could be recognized in Fig. 4a that Fe–32%Ni alloy under different deformation conditions. It can
some new small grains emerged at the original grain boundaries be seen in Fig. 5a that the grain boundaries became serrated,
after a single pass deformation. The mean grain size decreased and small grains (indicated by arrows) evolved at the origi-
with increasing cumulative strain and more new grains were nal serrated grain boundaries and triple junctions where there
formed after two passes compression (Fig. 4b). The microstruc- were high-density dislocations. The initial grain structure disap-

Fig. 5. The typical EBSD image quality maps of Fe–32%Ni alloy after MF with cumulative strain of (a) 0.5 and (b) 1.5.
282 B. Han, Z. Xu / Journal of Alloys and Compounds 457 (2008) 279–285

As having been found by Sakai, etc. [19], the hot deformed


microstructures are roughly similar to each other, but are
characterized by heterogeneous on a substructural level. The
substructure of multi-axially forged Fe–32%Ni alloy could be
divided into three types due to the dislocation distribution fea-
ture (Fig. 6): (A) fine grains with low dislocation density (the
nuclei of dynamic recrystallization); (B) grains with disloca-
tion density in a gradient distribution and low density near the
boundaries and high density within the grains (growing dynamic
recrystallized grains); (C) dynamic recrystallized grains with
high dislocation density (work-hardened dynamic recrystallized
grains). The three kinds of structures were distributed uniformly
in a proportion during deformation, and such substructures are
frequently developed in hot-worked metallic materials. Obvi-
ously, the present substructure was related to the microstructure
evolution process. The new grains were formed continuously
during deformation, and were deformed and work-hardened dur-
ing further deformation, so there were always three kinds of
Fig. 6. The TEM microstructures of Fe–32%Ni alloy under MF with cumulative substructures in the TEM micrographs. The nucleation and the
of 1.5. growth of these new grains caused the softening of the deformed
materials.
Many materials undergo dynamic recrystallization (DRX)
peared and the microstructure was consisted of almost equiaxed during hot working, and the microstructure evolution during
grains (Fig. 5b) when the strain summed up to 1.5. Although hot working has been examined extensively. Ponge etc. [20]
the original grains were almost completely replaced by the new found that DRX was set off by strain induced bulging of prior
grains, it could be recognized that some gains were free of crys- grain boundaries and preceded by the diminishing of the original
tal defects and some were severely distorted, especially at the deformed microstructures. Wusatowska–Sarnek [21] elucidated
grain boundaries. the microstructure evolution of copper, it was found that the new

Fig. 7. The typical micrographs of Fe–32%Ni alloy evolved at 773 K under MF with cumulative strain of (a) 0.5, (b) 1.0, (c) 1.5, and (d) 3.0.
B. Han, Z. Xu / Journal of Alloys and Compounds 457 (2008) 279–285 283

grains were evolved by bulging of the serrated grain boundaries, There was no nucleation phenomenon as that in Fig. 4. It could
which was companied either by rotation of a bulged portion be clearly seen that the dynamic changes in grain size during
or twining at the back of the migrating boundary. The present deformation strongly depended on the cumulative strain, and
results indicated that the fine-grained microstructures could be the grains were just elongated after the first pass compression
produced by MF at 1073 K, and the new grains were promoted (Fig. 7a). The grains were greatly fined after two passes com-
by the large stored energy and the large density nucleation sites pression, and the breaking-up characteristic of the original grains
(Fig. 5a). The grain refinement process is that the new grains that could be found (Fig. 7b). The grain size was about 1 ␮m after
are almost defect-free are formed by nucleating at the serrated six passes MF (Fig. 7d). Although no strain softening appeared
grain boundaries where there is large stored energy resulting at high strain (Fig. 3), the present deformation was still able
from deformation; then they grow and consume the deformed to produce almost full development of new fined-grains, which
microstructure until the material is completely recrystallized suggested that other microstructure evolution mechanism may
(Fig. 5b). In such a process, the microstructure is heterogeneous take place during the present deformation conditions.
and can be divided into those regions that have undergone recrys- Fig. 8 shows the typical EBSD image quality maps of
tallization and that have not. Generally, the grain refinement Fe–32%Ni alloy under MF at the temperature of 773 K. It could
process is classified as discontinuous dynamic recrystallization be found that large amounts of LABs were formed in the grain
(DDRX). interior after one pass forging (Fig. 8a), and some dislocation
walls and deformation bands could be recognized. More HABs
3.2.2. Microstructure evolution at 773 K were formed as the cumulative strain increased (Fig. 8b). The
A series of typical micrographs of Fe–32%Ni alloy evolved grain size was about 300 nm when the cumulative strain reached
at 773 K with different cumulative strain are shown in Fig. 7. about 7.5 (Fig. 8c), and it was obvious that the new grains were

Fig. 8. The typical EBSD image quality maps of Fe–32%Ni alloy under MF at the temperature of 773 K with cumulative strain of (a) 0.5, (b) 1.5, and (c) 7.5.
284 B. Han, Z. Xu / Journal of Alloys and Compounds 457 (2008) 279–285

grains, so the grain refinement during MF deformation should


be considered as a strain induced phenomenon [25].
Different phenomenological models for the development of
such structural refinement have been presented. Some of them
extended the continuous evolution of dislocation structure by
the crystallographic glide from low and moderate strain to very
large strain [26]. An alternative approach described severe plas-
tic deformation as discontinuous evolution due to localized
flow inside shear bands of non-crystallographic orientations
[27–29]. The present experimental results suggested that grain
subdivision and rotation were the mechanism of the present
structural evolution. The initial deformation led to the formation
of dislocation substructures such as cells and dense disloca-
tion walls (DDWs) (Fig. 8a). Subdividing the grains interiors
into several blocks, the DDWs are generally called subgrain
boundaries (Fig. 9a). Due to the increase in lattice dislocation
density evolved by the usual intrinsic slip and the absorption of
accommodated dislocations coming from strain incompatibility
of subgrains, further deformation caused the misorientations of
subgrain boundaries to increase. Under repetitive deformation
with changing compression axis, the DDWs were developed
in various directions and intersected with each other, in which
way the original grains were broken up into several subgrains.
Then the subgrains were gradually reoriented to new inde-
pendent grains by rotation, and the new grains were formed
homogeneously and continuously during deformation [30,31].
Moreover, as noted in Ref. [32], the difference in the orientations
of adjacent grains caused differences in the strain each grain suf-
fered, and consequently led to local stresses in the vicinity of the
boundaries. The relaxation of the internal stresses was realized
by the subgrain rotation and formation of new strain-induced
boundaries. Due to the heterogeneity of internal stresses within
each grain, different parts of each grain were rotated to various
crystallographic orientations at last.
The grain refinement took place gradually and homo-
geneously throughout the whole microstructure, and the
‘recrystallized’ and ‘unrecrystallized’ regions could not be dis-
Fig. 9. The TEM microstructure of Fe–32%Ni alloy after MF at 773 K with tinguished from each other. Thus the present microstructure
cumulative of 7.5. The number in indicates the grain boundary misorientations. evolution process is considered to be a kind of continuous
reaction and should be phenomenally classified as continuous
distributed homogeneously relative to those in Fig. 5b. Large dynamic recrystallization (CDRX).
proportion HABs was formed in Fig. 8c compared with that in
Fig. 8a. 4. Conclusions
The typical TEM microstructure of multi-axially forged
Fe–32%Ni alloy was shown in Fig. 9. It revealed that the early The deformation behavior and the grain refinement mecha-
multiple deformations brought about the evolution of subgrains nism of Fe–32%Ni alloy during MF were studied by the flow
that were crossed by dense dislocation walls (Fig. 9a). The stress testing and the microstructure observation. The following
multi-axially forged microstructure was characterized by the conclusions could be obtained:
full development of well-defined grains with high misorienta-
tions throughout the whole area, and the high-density dislocation (1) The flow stress of Fe–32%Ni alloy gradually increased to
and non-equilibrium grain boundaries were also could be found saturation with work hardening and showed a steady-state-
(Fig. 9b). Such kind of microstructure is generally similar to like one without any strain softening when multi-axially
those developed under cold deformation [22–24]. It is generally forged at 773 K; while when multi-axially forged at 1073 K,
agreed that DRV controls the flow behavior in athermal region, the flow stress curve displayed a typical type of dynamic
but dose not lead to new grain evolution. In the present study, recrystallization (DRX) flow one, i.e. a single peak flow
the effect of strain accumulation under MF became very impor- stress followed by a work softening and then a steady
tant for the evolution of dislocation structures as well as new state.
B. Han, Z. Xu / Journal of Alloys and Compounds 457 (2008) 279–285 285

(2) At the temperature of 1073 K, the grain refinement mecha- [3] R.Z. Valev, R.K. Islamgaliev, I.V. Alexandrov, Prog. Mater. Sci. 45 (2000)
nism is classified as discontinuous dynamic recrystallization 103–189.
(DDRX). The microstructure evolution process is that the [4] C. Xu, M. Furukawa, Z. Horita, J. Alloys Compd. 378 (2004) 27–34.
[5] Y. Lwahashi, Z. Horita, M. Nemoto, Acta Mater. 46 (1998) 3317–3325.
new grains that are almost defect-free are formed by nucle- [6] S. Ferrasse, V.M. Segal, K.T. Hartwig, Metall. Mater. Trans. A 29 (1997)
ating at grain boundaries, then they grow and consume the 1047–1057.
deformed microstructure until the material is completely [7] V.M. Segal, Mater. Sci. Eng. A 197 (1995) 157–164.
recrystallized. [8] T. Hebesberger, H.P. Stüwe, A. Vorhauer, Acta Mater. 53 (2005) 393–402.
(3) At the temperature of 773 K, the grain refinement mecha- [9] A. Vorhauer, T. Hebesberger, R. Pippan, Acta Mater. 51 (2003) 677–686.
[10] Y. Kimura, S. Takaki, Mater. Trans. JIM 36 (1995) 289–296.
nism is classified as continuous dynamic recrystallization [11] K. Ameyama, M. Hiromitsu, N. Imai, in: T. Chandra, S.R. Leclair, J.A.
(CDRX). The new microstructure formation process is that Meech (Eds.), Proceedings of Australia-Pacific From on Intelligent Pro-
the initial deformation leads to the formation of subgrain cessing and Manufacturing of Materials (IPMM 97), 1997, p. 982.
boundaries, then further deformation causes the misorien- [12] K. Ameyama, M. Hiromitsu, N. Imai, Tetsu-to-Hagane 84 (1998) 357–362.
tations of subgrain boundaries to increase by usual intrinsic [13] N. Tsuji, Y. Saito, H. Utsnomiya, Scrip. Mater. 40 (1999) 795–800.
[14] Y. Saito, N. Tsuji, H. Utsnomiya, Scrip. Mater. 39 (1998) 1221–1227.
slip and the absorption of accomodated dislocations com- [15] M.T. Pérez-Prado, J.A. del Valle, O.A. Ruano, Scrip. Mater. 51 (2004)
ing from strain incompatibility of subgrains. At last, the new 1093–1097.
grains are formed homogeneously. [16] A. Belyakov, T. Sakai, H. Miura, Mater Trans. JIM 41 (2000) 476–484.
(4) Multi-axial forging (MF) is an effective SPD method to pro- [17] F.J. Humphreys, M. Hantherly, Recrystallization and Related Annealing
duce ultra-fine grained materials, and the grains obtained Phenomena, Pergamon, Oxford, 1996.
[18] A. Belyakov, W. Gao, H. Miura, Metall. Trans. A 24 (1998) 2957–2965.
through CDRX by warm deformation are much smaller and [19] T. Sakai, M. Ohashi, Mater. Sci. Tech. 6 (1990) 1251–1257.
more homogeneous than those obtained through DDRX by [20] D. Ponge, G. Gottstein, Acta Mater. 46 (1998) 69–80.
hot deformation. [21] A.M. Wusatowska-Sarnek, H. Miura, T. Sakai, Mater. Sci. Eng. A 323
(2002) 177–186.
Acknowledgements [22] A. Galiyev, R. Kaibyshev, G. Gottstein, Acta Mater. 49 (2001) 1199–1207.
[23] S.Yu. Mironov, G.A. Salishchev, M.M. Myshlyev, R. Pippan, Mater. Sci.
Eng. A 418 (2006) 257–267.
The authors would like to acknowledge the financial sup- [24] A. Belyakov, T. Skai, H. Miura, R. Kaibyshev, Scrip. Mater. 42 (2000)
port of the National Natural Science Foundation of China 319–325.
under granted number 50471017. The authors would also like [25] A. Belyakov, T. Sakai, H. Miura, ISIJ Inter. 39 (1999) 592–599.
to express thanks to professor Xuyue Yang of the Electro- [26] D.A. Hughes, N. Hansen, Acta Mater. 45 (1997) 3871–3886.
[27] T.R. MacNalley, D.L. Swisher, M.T. Perez-Prado, Met. Mater. Trans. A 33
Communications University of Japan for help in the EBSD test. (2002) 279–290.
[28] P.C. Wu, C.P. Chang, P.W. Kao, Mater. Sci. Eng. A 374 (2004) 196–203.
References [29] V.M. Segal, Mater. Sci. Eng. A 406 (2005) 205–216.
[30] D. Kuhlmann-Wilsdorf, N. Hansen, Scrip. Metall. Mater. 25 (1991)
[1] F.J. Humphreys, P.B. Prangnell, R. Priestner, Curr. Opin. Solid State Mater. 1557–1562.
Sci. 5 (2001) 15–21. [31] J. Huang, Z. Xu, Mater. Lett. 60 (2006) 1854–1858.
[2] P.B. Prangnell, J.R. Bowen, P.J. Apps, Mater. Sci. Eng. A 375–377 (2004) [32] V.V. Rybin, Large Plastic Strains and Fracure of Metals, Metallurgy,
178–185. Moscow, 1986.

You might also like