Nuclear Instruments and Methods in Physics Research B: C. Djebbari, S. Alleg, J.M. Greneche

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 5

Nuclear Instruments and Methods in Physics Research B 268 (2010) 306–310

Contents lists available at ScienceDirect

Nuclear Instruments and Methods in Physics Research B


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

Effect of aluminium addition on the structural properties of nanostructured


Fe50Co50 alloy
C. Djebbari a,b,*, S. Alleg a, J.M. Greneche b
a
Laboratoire de Magnétisme et de Spectroscopie des Solides, Département de Physique, Faculté des Sciences, Université de Annaba. B.P. 12, Annaba 23000, Algeria
b
Laboratoire de Physique de l’Etat Condensé, UMR CNRS 6087, Université du Maine, Faculté des Sciences, F-72085, Le Mans Cedex 9, France

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

Article history: Nanostructured Fe50Co50 and (Fe50Co50)80Al20 powders were obtained by mechanical alloying in a high-
Received 2 June 2009 energy planetary ball mill under argon atmosphere. Structural and microstructural changes during the
Received in revised form 3 August 2009 milling process were followed by X-ray diffraction using the Maud program which is based on the Riet-
Available online 12 September 2009
veld method. For the Fe50Co50 powders, the non-homogeneity of the a-Fe(Co) solid solution is evidenced
by the existence of two bcc structures having nearly the same lattice parameter but different crystallite
Keywords: sizes, microstrains and proportions. The Al addition leads, above 6 h of milling, to the formation of a dis-
Nanostructured materials
ordered Fe(Co, Al) solid solution and B2 AlCo nanophase (CsCl type).
Fe–Co alloys
Mechanical alloying
Ó 2009 Published by Elsevier B.V.
X-ray diffraction

1. Introduction tion until the recent decades when certain of these alloys attracted
interest for ultrasonic applications [7].
Nanocrystalline (NC) materials have been intensively investi- For the FeAl binary alloy, a typical ordering system, two types of
gated during the last decades because of their remarkable physical, phase separation A2 + B2 (FeAl) and A2 + DO3 (Fe3Al) have been
magnetic and mechanical properties in comparison to coarse- found experimentally [8]. The FeCo alloy is also a typical ordering
grained polycrystalline materials. Those features arise from the system with a coexistent region of A2 + B2 phases. In contrast to
crystallite size refinement down to the nanometer scale and the abundance of studies in Fe–Al and Fe–Co alloys, those in the ter-
important fraction of atoms residing in the grain boundaries. NC nary Fe–Co–Al system are scarce. According to the Fe–Co–Al phase
materials have been prepared by different methods such as elec- diagram, the a-Fe disordered region is continuously connected to
trodeposition, inert gas consolidation, physical vapour deposition, the B2 ordered region without any phase boundary [9]. It has been
chemical vapour deposition, mechanical alloying (MA), etc. This la- reported that the phase separation during ageing of Fe-rich Fe–Co–
ter is described as a high-energy milling process in which the pow- Al ordering alloys [10] can be classified into: (i) a coexistent region
der particles are subjected to heavy plastic deformation, repeated of two kinds of B2 phases at low temperature with an identical
fracturing, cold welding and rewelding. MA has been used to syn- structure (CsCl type superlattice) but different compositions. These
thesize amorphous alloys, intermetallic compounds, NC alloys and two phases merge into a B2 single phase at high temperature; and
solid solutions [1–4]. (ii) a coexistent region of a-Fe disordered and B2 ordered phases at
The Fe–Co system has received special attention due to its soft low temperature, and an A2 or B2 single phase at high tempera-
magnetic properties which are superior in the nanostructured al- ture. However, it is important to mention that there is no study
loys [5]. The interest of Fe–Al system is based on several applica- on the structural changes during the alloying process in the Fe–
tions, particularly at elevated temperature in hostile environment Co–Al system to our knowledge.
because of its excellent oxidation and corrosion resistance [6]. The present work reports the first results on the effect of Al
The Fe–Co–Al system, which is one of the most important ternary addition on the structural and microstructural properties of nano-
systems for magnetic materials, has received only sporadic atten- structured Fe50Co50 powders studied by X-ray diffraction.

2. Experimental procedures
* Corresponding author. Address: Laboratoire de Physique de l’Etat Condensé,
UMR CNRS 6087, Université du Maine, Faculté des Sciences, F-72085, Le Mans
Cedex 9, France. Elemental iron (6–8 lm, 99.7%), cobalt (45 lm, 99.5%) and alu-
E-mail address: djebbaritbs@yahoo.fr (C. Djebbari). minium (74 lm, >99%) powders were used as starting materials.

0168-583X/$ - see front matter Ó 2009 Published by Elsevier B.V.


doi:10.1016/j.nimb.2009.09.008
C. Djebbari et al. / Nuclear Instruments and Methods in Physics Research B 268 (2010) 306–310 307

First, Fe and Co were mixed, in appropriate weight ratio, to give crystallite size, lattice parameters, phase fraction and residual mi-
nominal composition of Fe50Co50 and milled up to 6 h. The Al cro strains. For the Fe50Co50 powders, the local Fe environment
was added to the Fe50Co50 powders milled for 6 h in order to obtain changes during the milling process was investigated by Mössbauer
the (Fe50Co50)80Al20 mixture (wt.%) to be milled for different times spectrometry in a transmission geometry with a conventional con-
(1 h, 3 h, 6 h, and 48 h). The milling process was performed in a stant acceleration spectrometer using a 57Co source diffused into a
planetary ball mill Fritch P7, at room temperature, using hardened Rh matrix. The spectrum was computer fitted with the least-
steel balls and vials. The ball-to-powder weight ratio was 35/3 and squares MOSFIT program [12]. The isomer shift, IS, values are given
the rotation speed was 1125 rpm. The hardened steel vials and with respect to a-Fe at 300 K.
balls were sealed under argon atmosphere in a glove box to pre-
vent from oxidation. In order to avoid the local temperature rise in- Table 1
Lattice parameter, a, average crystallite size, <L>, microstrains, <r2>1/2, and phase
side the vials, milling was interrupted after 1=4 h for 1=4 h. The phase
proportion deduced from the Rietveld refinement of the Fe50Co50 powders milled for
identification during the milling process was followed by X-ray dif- 6 h.
fraction (XRD) by means of a Siemens D501 diffractometer in a
Phases a (nm) <L> (nm) <r2 > 1/2 Proportion
(h 2h) geometry using Cu Ka radiation (k = 0.154056 nm). A
±10 4 ±1 ±10 3 (%) ± 10
structural refinement of X-ray powder data according to the Riet-
veld method [11] has been used to follow the evolution of the a-Fe(Co)1 0.2859 28.6 0.967 65
a-Fe(Co)2 0.2861 8.2 1.77 35
milled powders. The refined structural parameters were essentially

Fig. 1. Rietveld refinement of the XRD pattern of the Fe50Co50 powders milled for 6 h with one bcc a-FeCo phase.

Fig. 2. The best Rietveld refinement of the XRD pattern of the Fe50Co50 powders milled for 6 h with two bcc a-FeCo phases.
308 C. Djebbari et al. / Nuclear Instruments and Methods in Physics Research B 268 (2010) 306–310

B2 of the diffraction peaks and the decrease of their intensities with


B1 increasing milling time are due to the crystallite size refinement
down to the nanometer scale and an increase of the atomic level
strain because of the heavy plastic deformation. The high speed
Transmission (a. u.)

milling conditions (high milling energy) of Fe50Co50 powders lead,


after 6 h of milling, to the formation of a non-homogeneous Fe(Co)
solid solution. In fact, the best Rietveld refinement of the XRD pat-
tern cannot be obtained by means of a single bcc a-Fe(Co) phase as
shown in Fig. 1. Therefore, to well describe the inhomogeneous
broadening of Bragg peaks two bcc a-Fe(Co) phases, designed here
by a-Fe(Co)1 and a-Fe(Co)2 in order to distinguish them, have to be
considered (Fig. 2). These two phases have nearly the same lattice
parameter but different crystallite size, microstrains and relative
proportions. The refined values of these parameters are summa-
-8 -4 0 4 8 rized in Table 1. The lattice parameter change with grain size in
V(mm/s) the two bcc a-Fe(Co) phases is in good agreement with those re-
ported in other NC materials [4,13]. It has been recently reported
Fig. 3. Mössbauer spectrum, taken at 300 K, of the Fe50Co50 powders milled for 6 h.
that the low speed ball-milling conditions of the Fe–12Co (wt.%)
powders lead, after 24 h of milling, to the formation of a single
bcc Fe(Co) solid solution whereas, the high speed ball-milling con-
ditions lead, after 3 h of milling, to the formation of an heteroge-
Table 2
Hyperfine magnetic field, B, isomer shift, IS, quadrupole shift, 2e, relative area and
neous two-bcc Fe(Co) structures with different lattice
magnetic moment, l, of the Fe50Co50 powders milled for 6 h. parameters, crystallite sizes, microstrains and hyperfine parame-
ters [4]. It has been also reported that XRD pattern of the ball
Phases B (T) IS (mm/s) 2e (mm/s) Relative area l (lB)
±0.2 ±0.01 ±0.01 (%) ± 5
milled Fe67Co33 shows a single crystalline phase while Mössbauer
spectrometry indicates the existence of two components: one com-
a-Fe(Co)1 33.1 0.025 0.05 19 2.20
ponent is linked to large crystallite size and the second is attrib-
a-Fe(Co)2 36.36 0.036 0.007 79 2.42
uted to the contribution from very small crystallites [14].
According to the fact that the lattice parameter decreases with
increasing Co content [3], one can suppose that the a-Fe(Co)1
phase with a lattice parameter a = 0.2859 nm can be attributed to
3. Results and discussion a Co-rich FeCo environment, while the a-Fe(Co)2 phase
(a = 0.2861 nm) can be ascribed to an Fe-rich FeCo environment.
The structural changes during the alloying process from Fe, Co This assumption can be supported by the Mössbauer spectrometry
and Al elemental powders into the nanostructured Fe50Co50 and results. Indeed, the Mössbauer spectrum of the milled Fe50Co50
(Fe50Co50)80 Al20 powders were followed by XRD. The broadening powders for 6 h has to be computer fitted with two components

Fig. 4. XRD patterns of the Fe50Co50 milled for 6 h (a), Co80Al20 for 6 h (b), and (Fe50Co50)80Al20 powders for 1 h (c), 3 h (d), 6 h (e) and 48 h (f).
C. Djebbari et al. / Nuclear Instruments and Methods in Physics Research B 268 (2010) 306–310 309

(Fig. 3) related to different iron environments and therefore, to dif- atoms into the a-Fe(Co) matrix and consequently, to the formation
ferent compositions. The deduced hyperfine parameters are given of the a-Fe(Co, Al) solid solution. This is evidenced by the slight
in Table 2. The first component with B1 = 33.1 T can be assigned shift in the XRD peaks towards smaller 2h angles since the Al radius
to an Fe-rich FeCo environment where the Fe atoms are sur- (rAl = 0.143 nm) is larger than those of Fe (rFe = 0.124 nm) and Co
rounded preferentially by Fe atoms as first nearest neighbours. (rCo = 0.125 nm) ones. The appearance of new diffraction peaks,
This is confirmed by the magnetic moment value. The second com- after 3 h of milling, can be linked to the formation of a B2 structure
ponent with B2 = 36.4 T can be assigned to a Co-rich FeCo environ- (CoAl and/or Fe/Al). Such a reaction can be attributed to the heat
ment since the presence of one Co atom as first or second nearest generated by the MA process which involves the transfer of a ki-
neighbour of the Fe atoms increases the hyperfine magnetic field netic energy of a highly energetic ball to the powder. Additionally,
by about 0.8 T. fine grain structure and structural defects generated by the re-
The XRD patterns of the milled (Fe50Co50)80Al20 powders for se- peated mechanical deformation during milling would favour the
lected milling times are shown in Fig. 4 with those of the milled reaction kinetics. On the basis of the XRD pattern of the milled
Fe50Co50 and Co80Al20 powders for 6 h. The broadening of the Al Co80Al20 powders for 6 h in the same conditions (Fig. 5), one can
diffraction peaks and the decrease in their heights are due to the conclude that the obtained B2 structure is the CoAl phase. Indeed,
crystallite size refinement and the introduction of internal strains according to the thermodynamic data, the CoAl phase has a rela-
and structural defects because of the induced heavy plastic defor- tively large negative enthalpy of formation ( 43 kJ/mol) compared
mation during the milling process. The total vanishing of the Al to that of the FeAl phase ( 32 kJ/mol) [13]. Milling for 48 h gives
Bragg peaks above 6 h of milling is related to the diffusion of Al rise to a mixture of B2 AlCo nanophase and two bcc a-Fe(Co, Al)

Fig. 5. Rietveld refinement of the XRD pattern of the Co80Al20 powders milled for 6 h.

Fig. 6. Rietveld refinement of the XRD pattern of the (Fe50Co50)80Al20 powders milled for 48 h.
310 C. Djebbari et al. / Nuclear Instruments and Methods in Physics Research B 268 (2010) 306–310

Table 3 thus the formation of bimetallic compounds [18]. It has been re-
Lattice parameter, a, average crystallite size, <L>, microstrains, <r2>1/2, and phase ported that the addition of Al content greater than 10 at.% to the
proportion deduced from the Rietveld refinement of the (Fe50Co50)80Al20 powders
milled for 48 h.
Fe–Co alloys leads to the formation of two phase types having dif-
ferent Al concentrations: the phases poor in Al have a small atomic
Phase a (nm) <L> (nm) <r2>1/2 Proportion volume, whereas those rich in Al are B2 type, with a greater lattice
±10 4 ±1 ±10 3 (%) ± 5
constants [18]. The observed plateau above 24 h of milling for both
a-Fe(Co, Al)1 0.2876 13.4 0.943 79 structures can be attributed to a solubility limit and therefore, to a
a-Fe(Co, Al)2 0.2877 4 2.35 11
B2 AlCo 0.2884 9 2.18 10
steady state.

4. Conclusion

0.289 Nanostructured Fe50Co50 and (Fe50Co50)80Al20 alloys were ob-


α−Fe(Co.Al)1
tained by MA. The milling of the Fe50Co50 mixture leads to a
α−Fe(Co.Al)2 non-homogeneous solid solution with two bcc phases having
Lattice parametrers (nm)

nearly the same lattice parameter but different hyperfine parame-


0.288 ters, crystallite sizes, microstrains and percentages. The dissolution
of Al atoms into the bcc a-Fe(Co) solid solution gives rise, after 48 h
of milling, to a mixture of two bcc a-Fe(Co, Al) structures and or-
dered B2 AlCo nanophase (CsCl type).
0.287
Acknowledgments

This work was supported by the Ministère de l’Enseignement


0.286 Supérieur et de la Recherche Scientifique – Algeria. The authors
are very grateful to Anne Marie Mercier from the Laboratoire des
Fluorures de l’Université du Maine UMR CNRS 6010, Le Mans –
0 6 12 18 24 30 36 42 48
France for performing XRD measurements.
Milling time (h)
References
Fig. 7. Milling time dependence of the a-Fe(Co, Al)1 and a-Fe(Co, Al)2 lattice
parameters.
[1] S. Alleg, S. Azzaza, R. Bensalem, J.J. Suñol, S. Khene, G. Fillion, J. Alloys Compd.
482 (2009) 86.
[2] Hongwei Shi, Debo Guo, Yifang Ouyang, J. Alloys Compd. 455 (2008) 207.
structures having the same lattice parameter but different crystal- [3] H. Moumeni, S. Alleg, C. Djebbari, F.Z. Bentayeb, J.M. Greneche, J. Mater. Sci. 39
(2004) 5441.
lite sizes, microstrains and proportions (Fig. 6). The refined param- [4] S. Alleg, F.Z. Bentayeb, R. Bensalem, C. Djebbari, L. Bessais, J.M. Greneche, Phys.
eters are given in Table 3. Stat. Sol. (a) 205 (7) (2008) 1641.
The relative deviation (Da) of the lattice parameter of the AlCo [5] C. Suryanarayana, Bull. Mater. Sci. 17 (1994) 307.
[6] M.M. Rico, J.M. Greneche, G.A. Pérez Alcazar, J. Alloys Compd. 398 (2005) 26.
nanophase (a) with respect to the perfect crystal (a0), which is gi- [7] I. Jacobs, IEEE Trans. Magn. 21 (4) (1985) 13.
ven by Da = (a a0)/a0, reaches as much as 0.73% after 48 h with a [8] S.M. Allen, J.W. Cahn, Acta Metall. 24 (1976) 425.
crystallite size of about 9 nm. The lattice distortion which is char- [9] O.S. Iwanov, Izv. Sekt. Fiz. Khim. Anal. 19 (1949) 503.
[10] T. Miyazaki, K. Isobe, T. Kozakai, M. Doi, Acta Metall. 35 (1987) 317.
acterised by lattice expansion or compression and increase/de- [11] L. Lutterotti, MAUD Program, CPD, Newsletter (IUCr), 24, 2000.
crease interatomic spacing with respect to the corresponding [12] F. Varret, J. Teillet, Unpublished Mosfit Program, Université du Maine, France.
perfect crystal has been reported for many mechanically alloyed [13] N.C. Abhik, R. Vivek, V. Udhayabanu, B.S. Murty, J. Alloys Compd. 465 (2008)
106.
systems [15–17]. [14] E.C. Passamani, C. Lartica, V.P. Nascimento, J. Mater. Sci. 37 (2002) 819.
The variation of the a-Fe(Co, Al)1 and a-Fe(Co, Al)2 lattice [15] H. Moumeni, S. Alleg, J.M. Greneche, J. Alloys Compd. 419 (2006) 140.
parameters is plotted against milling time in Fig. 7. The increase [16] W. Tebib, S. Alleg, R. Bensalem, N. Bensebaa, F.Z. Bentayeb, J.J. Suñol, J.M.
Greneche, J. Nanosci. Nanotechnol. 8 (2008) 2029.
of the lattice parameters can be attributed to the Al diffusion into
[17] K. Lu, M.L. Sui, J. Mater. Sci. Technol. 9 (1993) 419.
the a-Fe(Co) solid solution. Indeed, the presence of Al atoms in the [18] S.C. Chadjivasiliou, K.G. Efthimiadis, I.A. Tsoukalas, J. Hesse, Mater. Res. Bull. 34
transition alloy lattice causes an expansion of the cell, and favours (4) (1999) 581.

You might also like