Mirtha Pillaca Quispe‡
*a,
Carlos V. Landauroa,
Milida Z. Pinto Vergaraa,
Justiniano Quispe-Marcatomaa,
Chachi Rojas-Ayalaa,
Víctor A. Peña-Rodrígueza and
Elisa Baggio-Saitovitchb
aFaculty of Physical Sciences, National University of San Marcos, P.O. Box 14-0149, Lima 14, Peru. E-mail: mpillacaq@unmsm.edu.pe
bCentro Brasileiro de Pesquisas Físicas (CBPF), Rua Dr Xavier Sigaud 150, Urca, Rio de Janeiro, 22290-180, Brazil
First published on 5th January 2016
The effect of mechanical milling in i-Al64Cu23Fe13 quasicrystalline and ω-Al70Cu20Fe10 crystalline phases is systematically investigated in the present work. The Al–Cu–Fe samples were obtained by arc furnace technique and then nanostructured by means of mechanical milling. The results indicate that the solid samples present a weak ferromagnetic behavior at 300 K, showing a saturation magnetization of 0.124 emu g−1 for the icosahedral phase (i-phase) and 0.449 emu g−1 for the tetragonal phase (ω-phase). These small values could be an indication that only a few percentage of Fe atoms carry magnetic moment. The magnetic response in the nanostructured ω-phase increases up to 3.5 times higher than its corresponding solid counterpart. Whereas for the i-phase this increment is about 16 times higher. Moreover, the speed of the variation of the studied physical parameters after reducing the average grain size has been obtained from the exponent (α) of a power law fit of the experimental data. The values of α, corresponding to the magnetic response, are slightly different in each phase, which should be related to the different chemical composition and/or the type of long range order. Additionally, we also search for a critical grain size. However, this critical value has not been observed in the studied samples.
The first studies of magnetism in the stable icosahedral Al–Cu–Fe13 phase have shown that local atomic environments (i.e., the nearest-neighbors coordination shell of the Fe atoms in i-Al–Cu–Fe) play an important role in the magnetic nature of these alloys.8,12 Thus, these works have indicated the non-magnetic character of the Al–Cu–Fe quasicrystal. Furthermore, a recent work14 reported that single-quasicrystalline icosahedral Al64Cu23Fe13 samples, considered to be in the ideal icosahedral composition, are diamagnetic in the whole temperature range (2–300 K) with a Curie–Weiss temperature near to θ = −2.3 K. Taking into account that the studied sample was a mono-quasicrystal of high structural quality, they considered this behavior as an intrinsic property of the i-Al–Cu–Fe quasicrystal. Accordingly, one important conclusion is that the presence of defects, strain, grain size, and grain boundaries may influence the physical properties of the i-Al–Cu–Fe QC in such a way that they can overlap the true intrinsic properties of quasicrystals.14
In particular, the grain boundary volume is a relevant parameter in nanostructured materials (NM). This is because the grain boundary volume fraction increases after reducing the grain size due to the nanostructuration process. Consequently, grain boundaries growth leads to significant changes in the physical properties of the system under study. However, these effects have been poorly studied. Despite the fact that there are several methods to synthesize NM, among them, mechanical milling (MM) technique has been extensively used for the nanostructuration of materials.15,16 This is not only due to the low production costs at the industrial level but also for producing NM with a high density of interstitial regions. Thus, MM techniques are a good tool if our purpose is to study the influence of the nanostructuration process on the physical properties of the system. Moreover, it would be interesting to manipulate these properties, in a controlled way, for specific technological applications. In this regard, in a recent work17 it has been shown that a power law type fit of the simulations can be successfully used to describe the physical properties in function of the size of the system. Another important point to be considered in this adjustment is that these nanostructuration processes also depend on specific characteristics of the system such as: (i) periodicity or aperiodicity, and (ii) chemical composition. In this sense, the study of both the i-Al64Cu23Fe13 quasicrystal and its crystalline counterpart ω-Al70Cu20Fe10 (ref. 18) phases are a suitable option to perform such researches.
In the context described above, the present work reports a systematic comparative study of the influence of mechanical milling on the microstructure and magnetic properties of the i-Al64Cu23Fe13 quasicrystal and the ω-Al70Cu20Fe10 crystalline phases. The synthesis and the milling process were performed at least twice to verify the reproducibility of the results. Additionally, the dependence of the physical parameters in function of the average grain is analyzed. The paper is organized as follows. The experimental details are described in the next section. Section 3 is devoted to the results and Section 4 to the discussions. Finally, the conclusions are given in Section 5.
The obtained products were characterized by different methods and techniques. The phases and the grain size of the samples were analyzed by means of X-ray diffraction (XRD) technique employing a Bruker-D8 Focus diffractometer at 40 kV and 40 mA with Cu-Kα radiation (λ = 1.5406 Å). The local structure around the Fe sites was analyzed by a Transmission Mössbauer Spectrometer (TMS) with a 25 mCi 57Co/Rh radioactive source. Mössbauer spectra were fitted using the NORMOS program.21 The isomer shift values are given relative to the α-Fe. Finally, the magnetic properties were measured with a vibrating sample magnetometer (VSM) using a Quantum Design equipment with a magnetic field, H, in the range of ±20 kOe. All measurements were recorded at room temperature (RT).
![]() | ||
Fig. 1 X-ray diffraction pattern of the Al–Cu–Fe samples obtained after heat treatment at 700 °C. (a) Al64Cu23Fe13 icosahedral quasicrystalline phase, the corresponding N/M (reduced) indices23 are also indicated. (b) Al70Cu20Fe10 tetragonal crystalline phase including the Miller indices (hkl), similar as reported in the literature.25 The insets show their corresponding as cast samples where some metastable crystalline phases (ε-Al2Cu3, λ-Al13Fe4, θ-Al2Cu, η-AlCu and fcc-Al) co-exist. |
![]() | (1) |
![]() | ||
Fig. 3 Evolution of the average grain size 〈D〉 as function of the milling time for both (a) the i-phase and (b) the ω-phase (○ series A, □ series B). Inset: log–log plot of 〈D〉 versus t (t > 0). The slopes corresponding to the i-phase and ω-phase (−0.59 and −0.19, respectively) were obtained from a linear fit of the experimental data. The error bars are smaller than or equal to the symbol sizes and correspond to the accuracy determining the line-width of the diffraction peaks employed in eqn (1). |
It is worth mentioning that although the Scherrer method is not the best way to calculate 〈D〉, it can be considered as a good approximation to carry out a systematic study of the effect of the nanostructuration process on the physical properties of the studied samples. Moreover, according to the Rietveld refinement, due to the small value of the lattice strain (∼0.19) produced in the Al70Cu20Fe10 phase after five hours milling,18 it is possible to neglect this effect in the present analysis for both kind of samples: the i-phase and the ω-phase.
In order to compare the variation of the average grain size during the MM of both phases we determine the slope of the log–log plot (base 10) of 〈D〉 versus t (see inset of Fig. 3). Thus, the slope value for the i-phase is −0.59 which is very close to the theoretical value of −2/3 (see ref. 28), as was also observed in other phases.28 However, the corresponding slope value for the ω-phase is −0.19, which is very different to the expected theoretical value. This is an indication that the dependence of the mechanical properties on the grain size for both phases are very different.
Milling time (h) | i-Al64Cu23Fe13 | ω-Al70Cu20Fe10 | ||||||||
---|---|---|---|---|---|---|---|---|---|---|
〈D〉 (nm) | 〈δ〉 (mm s−1) | 〈Δ〉 (mm s−1) | Ms (emu g−1) | Hc (Oe) | 〈D〉 (nm) | 〈δ〉 (mm s−1) | 〈Δ〉 (mm s−1) | Ms (emu g−1) | Hc (Oe) | |
0.0 | 110.88(2) | 0.241(1) | 0.390(1) | 0.124(4) | 17.41(1) | 76(3) | 0.170(3) | 0.198(2) | 0.449(6) | 20.12(9) |
0.5 | 41.30(2) | 0.244(1) | 0.410(1) | 0.784(4) | 27.72(6) | 28(6) | 0.179(6) | 0.229(2) | 0.854(6) | 30.11(5) |
1.0 | 25.92(2) | 0.244(1) | 0.434(2) | 0.727(5) | 80.32(7) | 21(5) | 0.184(7) | 0.277(2) | 1.063(4) | 57.08(9) |
1.5 | 16.76(2) | 0.248(0) | 0.437(1) | 0.969(5) | 80.68(4) | — | — | — | — | — |
2.0 | 13.88(2) | 0.250(0) | 0.440(2) | 1.166(4) | 94.46(1) | 19(2) | 0.197(7) | 0.317(2) | 1.176(6) | 78.33(7) |
2.5 | 12.55(2) | 0.252(0) | 0.454(1) | 1.108(4) | 111.13(9) | 18(3) | 0.191(8) | 0.317(3) | 1.389(4) | 49.12(8) |
3.0 | 11.14(2) | 0.254(1) | 0.464(1) | 1.045(4) | 127.45(6) | 21(5) | 0.188(8) | 0.294(2) | 1.368(5) | 74.98(7) |
4.0 | 10.68(2) | 0.249(1) | 0.457(2) | 1.465(4) | 81.85(8) | 18(1) | 0.200(7) | 0.324(2) | — | — |
5.0 | 8.40(2) | 0.258(1) | 0.460(2) | 1.994(5) | 117.22(9) | 17(4) | 0.208(8) | 0.370(2) | 1.602(5) | 94.78(2) |
The magnetic parameters (saturation magnetization, Ms, and coercive field, Hc) are given in Table 1. It can be seen that Ms in the nanostructured ω-Al70Cu20Fe10 phase increases up to 3.5 times higher than its corresponding solid counterpart. Surprisingly, for the case of the i-Al64Cu23Fe13 phase this increment is about 16 times. The origin of this behavior will be discussed below.
The variations of the Mössbauer and magnetic parameters for un-milled and milled Al–Cu–Fe samples suggest that the chemical order, type of long range order, and the average grain size play an important role in the physical properties of Al–Cu–Fe systems. Thus, despite of having very close chemical compositions, the i-phase (with lattice parameter a = 6.308(2) Å) and ω-phase (a = 6.336(1) Å and c = 14.870(2) Å) present a quite different long range order, as can be seen in Fig. 2 for the corresponding un-milled samples. Hence, we expect that their behavior during the mechanical milling process should be different. In fact, in Table 1 can be seen that the values of the hyperfine parameters for the un-milled samples (i-Al64Cu23Fe13 and ω-Al70Cu20Fe10) are completely different. In their corresponding Mössbauer spectra can be observed that the Fe site in the ω-phase have a more symmetric environment than in the quasicrystalline phase, which is expressed in their corresponding quadrupole splittings. In this regard, we can mention some important points considering a structural model for the 1/1 approximant of the Al–Cu–Fe phase (Cockayne model) employed to describe the icosahedral phase.36 We note two Fe sites (Fe1 sites surrounded exclusively by aluminum atoms and Fe2 sites surrounded by Cu and predominantly by Al atoms), while in the ω-phase there is just one Fe site surrounded only by Al atoms. Therefore, it can be seen that these slight differences in the atomic local environment will play an important role in the nanostructuration process by mechanical milling, which is characterized by generating disordered sites in the interstitial region. Thus, it is expected that this might influence considerably in the physical properties of the studied samples. Also, the differences in the isomer shift values could be related to the amount of Fe atoms associated to each phase. According to this, the quasicrystalline phase with 13 at% and the ω-phase with 10 at% have 〈δ〉 = 0.241 mm s−1 and 0.170 mm s−1, respectively. These results are in agreement with ref. 37 and 38. However, the nanostructuration process does not affect significantly the isomer shift of the milled samples, as happened for Ms (see Table 1). This is expected because 〈δ〉 is related to both the atomic radius and the electronic charge distribution in the nucleus, which remain almost constant during the MM process.
Furthermore, it is worth noting that the Mössbauer spectra during the nanostructuration process broadens. For the i-Al64Cu23Fe13 samples, the Mössbauer spectra become wider, without changing their shape, after reducing the average grain size which is evidenced in the slight increment of the quadrupole splitting values (see Table 1). However, for the ω-Al70Cu20Fe10 phase these variations are more pronounced. This remarkable difference could be explained considering the structural stability of the phases. It is known that the Bergman cluster is considered the building block of the icosahedral Al–Cu–Fe alloys (i-phase). This kind of atomic arrangement, consisting mainly of two atomic shells with icosahedral symmetry, is more stable than other kind of local order such as the tetragonal order observed in the ω-phase.39 Additionally, according to the indicated above, the changes in the shape of the Mössbauer spectra (mainly in the ω-phase) means that the environment of some Fe sites are partially distorted. In fact, when the average grain size decreases, the shape of the quadrupole distribution, centered on 〈Δ〉 = 0.415 mm s−1, becomes slightly broader which is associated to the slight distortion of atomic environment in the icosahedral structure, as can be seen in the right panel of Fig. 2. The same behavior is observed in the ω-phase, but this distribution changes from a narrow peak (centered at 〈Δ〉 = 0.152 mm s−1) to a distribution with two peaks (centered at 0.168 mm s−1 and 0.587 mm s−1). The second peak could be associated to small disorder in the grain boundary due to the employed technique for the nanostructuration process. In this sense, the same criterion is applied to the i-phase where such small quadrupole distribution peak would be overlapped by the full distribution curve (whose width is approximately the sum of the two peak widths in the ω-phase).
The Ms obtained for the 0 h quasicrystal (0.124 emu g−1) is smaller than that reported by M. Roy (Ms = 6.4 emu g−1).9 This difference could be attributed to the presence of a remnant of the bcc-Fe phase in the samples synthesized in ref. 9 because those samples were obtained by mechanical alloying of elemental powders of Al, Cu, Fe in the composition Al65Cu20Fe15. In our case the QC sample is obtained by arc furnace and then mechanically milled to obtain the nanostructured QCs. Furthermore, the mechanical milling process contributes to a systematic increment of the magnetic parameters (Ms and Hc) for both compositions, as indicated before (see Table 1). This increment can be associated to the reduction of the average grain size or, in other words, to the growth of a magnetic interstitial region.
On the other hand, an important observation in these experiments is the absence of new phases during the nanostructuring process. Therefore, this ensures that the changes in the physical properties, in particular the magnetic ones, of the Al–Cu–Fe system should be related to the reduction of the average grain size of the system.
Hence, following ref. 40 we estimate the magnetic fraction in the Al–Cu–Fe samples (i-Al64Cu23Fe13 and ω-Al70Cu20Fe10). For this purpose, we assume that the full magnetic moment of each iron atom is 2.2 μB. Then, if all the Fe ions were magnetic, the saturation magnetization should be Msi = 13 × 2.2 μB f.u.−1 = 28.6 μB f.u.−1 for i-phase and Msω = 10 × 2.2 μB f.u.−1 = 22 μB f.u.−1 for ω-phase (f.u. standing for the “formula unit”). Now, knowing that for the un-milled i-phase the experimental value is Mexps = 0.124 emu g−1 = 0.087 μB f.u.−1 (see Table 1) we determine that only ∼0.30% of all the Fe atoms in this sample carry a magnetic moment. Analogously, for the un-milled ω-phase we obtain that ∼1.36% of the Fe atoms carry a magnetic moment. In similar way, we obtain the percentage of magnetic Fe atoms for the milled samples (nanostructured samples) and observe that these percentages increase after reducing the average grain size of the samples, as can be seen in Fig. 6(c) and (f). Thus, after five hours milling % Fe atoms with magnetic moments reaches ∼4.89%, and ∼4.85% for the i-phase and ω-phase, respectively. It is worth mentioning that a similar amount of magnetic atoms, for solid samples, have been reported in the Al–Cr–Fe approximant phase (0.8% of Fe atoms40) and the Al–Pd–Mn quasicrystal (1.0% of Mn atoms30,41).
However, according to Mössbauer spectroscopy measurements, no magnetic behavior is observed. In the other hand, the magnetization curves show a weak ferromagnetic behavior. This apparent disagreement between the VSM measurements, showing a magnetic behavior of the samples, and the Mössbauer spectra that do not show a magnetic signal (sextet lines) can be understood considering the following: (i) from XRD measurements there are no evidences of secondary phases, this implies that the observed magnetic signal cannot be due to the formation of spurious phases during the MM process; (ii) it is known that VSM is much more sensitive to detect a magnetic signal than TMS; consequently, can be argued that the percentage of magnetic Fe sites is below the detection limit of TMS, as was observed in several systems.42,43 To test this idea, we mixed bcc-Fe with the QC so that the amount of the Fe sites in the bcc-Fe phase represents ∼1.9% of the total amount of Fe (bcc-Fe and QC together). As suggested, the resulting spectrum did not show any magnetic signal with TMS (sextet lines). We also carried out a chemical analysis by energy dispersive X-ray spectroscopy (EDX) of the samples. The results indicate a low presence of chromium for the five hours milled samples (of the order of 0.25 wt% and 0.16 wt% for the i-phase and the ω-phase, respectively), which is due to contamination by the milling tools (hardened steel vial and balls). Thus, one could argues that the magnetic signal is due to Cr. However, Cr is not found in the samples milled by 0.5 hours. Contrary, their corresponding Ms values increase from 0.124 emu g−1 to 0.784 emu g−1 (6 times higher) in the i-phase and from 0.449 emu g−1 to 0.854 emu g−1 (2 times higher) in the ω-phase. Thus, the variation of Ms cannot be attributed to Cr. In general, Cr is present during the full milling process but its variation is not in the same proportion as the variations of Ms, as indicated above (see Table 1). Analogously, oxygen is also present during the milling process. However, for 0.5 hours milling the increment of O is of ∼1.9 while Ms increases 6 times (i-phase). Hence, the presence of oxygen could be associated to a very thin aluminum oxide layer44 without magnetic signal. In summary, we do not neglect the possibility that impurities, originated from the milling tool debris (formation of iron oxides or external elements like Cr), could contribute to the observed magnetism. However, the noticeable difference in the variation of the magnetic response for both milled samples (16 and 3.5 times higher than its corresponding solid counterpart for i-phase and ω-phase, respectively) is a clear indication that the magnetic signal is not only due to the employed experimental procedure, which is of the same type and order in both samples, but mainly due to a physical origin. In this regard, Jagličić et al. presented a study of the Mn magnetism in the i-Al–Pd–Mn quasicrystal subjected to different thermal annealing sequences. They reported that the degree of the disorder induced in the i-Al–Pd–Mn quasicrystalline structure, during the heat treatments or cooling modes, was decisive to generate the magnetic moment formation in this sample.45 In similar way, we consider that in the case of the milled Al–Cu–Fe samples, the impact of the balls produces lattice defects and grain boundaries in the system. For longer milling times the grain border becomes an interstitial disordered region that grows after increasing the milling of the samples. This disorder could be correlated with the observed magnetic response. However, since the i-Al64Cu23Fe13 phase is generally not perfect in the structure (which can contain twin domains or nucleation of approximants locally) the origin of the magnetic properties may be more complex than only disorder. Further measurements on atomic scale are required to understanding and clarify this magnetic behavior, which is beyond the scope of this study.
F = b〈D〉α, | (2) |
![]() | ||
Fig. 5 Variation of hyperfine parameters, 〈δ〉 and 〈Δ〉, versus the average grain size 〈D〉 plotted on a log–log scale (○ series A, □ series B). The dashed lines indicate a fit according to eqn (2). From the slope of the fit we obtain the exponent α. The error bars are smaller than or equal to the symbol sizes. |
![]() | ||
Fig. 6 log–log plot of the variation of the saturation magnetization (Ms), coercivity (Hc) and % Fe atoms carry magnetic moment with the average grain size 〈D〉 for (a–c) quasicrystalline i-Al64Cu23Fe13 and (d–f) crystalline ω-Al70Cu20Fe10 samples (○ series A, □ series B). The dashed lines indicate a fit according to eqn (2). From the slope of the fit we obtain the exponent α. The error bars are smaller than symbol size. |
F | α | |
---|---|---|
i-phase | ω-phase | |
〈δ〉 | −0.018 | −0.096 |
〈Δ〉 | −0.063 | −0.402 |
Ms | −1.059 | −0.757 |
Hc | −0.540 | −0.926 |
The difference in these changes is associated to the larger content of Fe atoms and the smaller average grain size (13 at% Fe, ∼9 nm) in the i-phase, as compared to the ω-phase (10 at% Fe, ∼19 nm). Furthermore, the slight difference of the exponent α, corresponding to Ms (see Table 2), should be related to the type of long range order.
Considering that the interstitial region depends inversely of the volume (∼〈D〉−3) of the nano-grains, we can express (according to the scaling theory) the dependence of the magnetic response with the average grain size as a power law function where the exponent α is the relevant quantity. In this way, we can determine the speed of the variation of the studied physical parameters after reducing the average grain size during the nanostructuration process. Thus, the values of α, corresponding to the magnetic response, are slightly different in each phase, which could be related to the different chemical composition and/or the type of long range order. Additionally, we also search for a critical grain size. However, this critical value has not been observed in the studied samples.
It is worth mentioning that a similar behavior in terms of grain size reduction has also been found in other alloys. Hence, the study of the parameter α is required to explore its possible universal behavior as well as to find a critical grain size in the nanostructured samples. In particular, Al-rich based alloys offers interesting perspectives due to their similarities to the Al–Cu–Fe system. Furthermore, similar studies in other quasicrystals with greater magnetism (such as Al–Pd–Mn or Al–Cu–Co) could be interesting to find if its magnetic response increases in the same proportion as in the case of the i-Al64Cu23Fe13. Finally, it is necessary to have a clear understanding of the magnetic behavior of both Al–Cu–Fe phases during the mechanical milling process. Thus, further HRTEM or STEM images at atomic resolutions are required. Additional studies to quantify and/or discard completely the possible contamination of the milling tool debris are being planned. Hence, a study of the nanostructuration of the samples employing other type of containers (e.g. agate and/or zirconia tools), which do not contain magnetic elements, is also required.
Footnotes |
† Electronic supplementary information (ESI) available. See DOI: 10.1039/c5ra21093c |
‡ Present address: Crystallography Section, Department of Earth and Environmental Sciences, Ludwig-Maximilians-Universität München, Theresienstraβe 41, 80333 München, Germany. E-mail: E-mail: mirtha.pillaca@campus.lmu.de. |
This journal is © The Royal Society of Chemistry 2016 |