Paula G. Bercoff*ab,
Soledad Apreaab,
Eva Céspedesc,
José Luis Martínezc,
Silvia E. Urretaa and
Manuel Vázquezc
aUniversidad Nacional de Córdoba. Facultad de Matemática, Astronomía, Física y Computación (FAMAF). Ciudad Universitaria, 5000 Córdoba, Argentina. E-mail: paula.bercoff@unc.edu.ar
bCONICET, Instituto de Física Enrique Gaviola (IFEG). Ciudad Universitaria, 5000 Córdoba, Argentina
cInstituto de Ciencia de Materiales de Madrid, CSIC, E-28049 Madrid, Spain
First published on 28th August 2024
Arrays of 50 nm diameter Fe85Pd15 cylindrical nanowires were electrochemically grown, crystallizing in a metastable γ-Fe(Pd) fcc A1 disordered solid solution. After performing a heating–cooling thermal cycle between 300 K and 1000 K, the γ-Fe(Pd) fcc metastable phase still predominates (97%), coexisting with a not-fully-identified minority phase. The thermal cycling induces a moderate increase in the crystallite size and a reduction of the lattice parameter although leading to a significant heating–cooling magnetic hysteresis. No further changes in temperature-dependent magnetization, M(T), are observed during subsequent cycling. The full-range (5 K to 800 K) saturation magnetization Ms(T) curve is quite accurately described by a phenomenological expression, which provides a Bloch-type contribution as T → 0 and undergoes the critical behavior near the Curie temperature TC. An upturn in Ms(T) is observed below 100 K which is described by a spin-glass-like second contribution, with freezing temperature Tf = (80 ± 2) K, and kBTf comparable to the exchange interactions in Fe–Pd systems. A Curie temperature of TC = 830 K, and a critical exponent value β = 0.42 ± 0.05 are estimated. These regimes (below and above 100 K) are also observed in the magnetization process. The temperature dependence of coercivity between 100 K and 800 K is consistent with a nucleation/propagation remagnetization mechanism, with activation energy of (320 ± 20) kJ mol−1 and critical field for magnetization reversal of (65 ± 1) mT, at 0 K. The analysis of the effective magnetic anisotropy as a function of temperature allows us to conclude that it essentially arises from the balance between different magnetostatic contributions.
In the last decades, low-dimensional nanostructures based on this alloy series have been intensively investigated, such as nanoparticles12 or Fe–Pd 1D (nanowires)13–15 and 2D (thin films)16,17 with different compositions produced by electrodeposition, a relatively low-cost and practical technique.
The bimetallic Fe–Pd system has a quite rich structural phase diagram;18 at high temperatures, below the solidus line, a thermodynamically stable disordered γ-FexPd100−x fcc type solid solution exists in the whole composition range. However, in the intermediate- and low-T regimes the Fe–Pd system becomes more complex, with many different stable and even metastable phases which are strongly dependent on the alloy composition as well as on the previous heat-treatment history details.
At room temperature, for iron concentration x ≥ 70, α-FexPd100−x alloys with a bcc-type structure are found, while for x ≤ 60 phases with fcc-type lattice are observed.19,20 The high temperature disordered γ-FexPd100−x alloys having compositions around x = 50 at% Fe and 25 at% Fe undergo a disorder–order transformation on cooling. Alloys with 40 ≤ x ≤ 50, order to L10 type superstructure via a thermodynamically order–disorder transformation of first order, to achieve a tetragonal structure, by alternating stacking of Fe and Pd planes along the [001] principal axis with respect to the fcc-lattice. In the composition range 14 ≤ x ≤ 38, the ordered phase is the cubic FePd3 phase, with L12-type structure. Atomic rearrangements at low T are often hindered by the sluggish kinetics of the equilibrium approach, leading to metastability. In these cases, slow diffusion processes limit the reaction pathway making difficult the transformation to more stable structures.
Fe–Pd low dimensional nanostructures synthesized by electrodeposition near room temperature are found to consist of a disordered, metastable γ-FexPd100−x phase in all the composition range, as also found in bulk samples quenched from high temperature, where solubility is complete. The as-deposited nanostructures are then heat-treated at temperatures above 830 K to promote the stable phase separation corresponding to the given composition. As soon as atomic mobility is enabled via a vacancy-mediated atomic diffusion process during heating, two simultaneous processes begin: defect recovery and long-range ordering.
While most of the investigations to date have been performed within an intermediate range of compositions in the Fe–Pd system, little has been reported about the hysteresis and structural transformations in high Fe content Fe–Pd alloys, and specifically on metastable Fe-rich γ-FexPd100−x alloys.
In this article, we report for the first time that the fcc disordered metastable γ-Fe85Pd15 phase initially observed in as-electrodeposited nanowires (NWs) remains almost unchanged, without transforming to the expected equilibrium bcc α-phase nor undergoing ordering, even after thermal cycling between 350 K and 800 K. This interesting fact has allowed us to completely characterize the magnetic properties of this disordered solid solution in a quite broad temperature range, describing the main magnetization mechanism, the effective anisotropy controlling coercivity and also the critical exponent associated to the ferro-paramagnetic transition. In addition to these fundamental physical properties, a spin-glass-like contribution to overall magnetization is detected below 100 K, associated to the NW's large area/volume ratio. The large resilience of this phase to structural (phase) transformations is crucial for applications where the material is exposed to high temperatures for extended periods. As this fcc phase also retains good mechanical properties and oxidation resistance, it is also suitable for small components operating under high thermal stress.
The porous alumina membranes used as templates were imaged by scanning electron microscopy, SEM (not shown), to estimate the pore diameter D, the center-to-center interpore distance dcc histograms, and the porosity P (= π/2√3(D/dcc)2).22 The templates exhibit self-assembled pores in a hexagonal lattice with D = (50 ± 2) nm; dcc = (110 ± 5) nm and P = (0.20 ± 0.05).
Microstructural details were determined by TEM-STEM techniques in a TEM TALOS F200X device, equipped with 4 windowless SDD Super-X detector system. For TEM characterization, the alumina template was dissolved in chromic acid during a few days, and the released nanowires were washed several times in ethanol. Then, a drop of the solution was deposited on a carbon-coated copper grid.
The low-temperature magnetic properties were measured in a Cryogenic vibrating sample magnetometer (VSM) from 5 K to 350 K. A SQUID magnetometer (MPMS-3), from Quantum Design (San Diego, USA), supplying a maximum magnetic field of 7 T and a furnace option allowing sensitive magnetic measurements at controlled high temperatures, was used to measure the magnetic properties between 300 K and 1000 K. A special ceramic holder was utilized to measure between 300 and 1000 K and a special alumina cement was used to fix the samples to the holder, before covering the ensemble with a non-magnetic Cu foil. Hysteresis loops were measured with the magnetic field applied both parallel (PA) and perpendicular (PE) to the NW's axis. All the magnetic measurements were performed while keeping the NWs within the template.
Fig. 1 SEM image corresponding to Fe85 NWs, after being partially released from the alumina template. Histograms of the NWs' diameter and length were obtained by measuring several images. |
Fig. 2 displays (upper panel) the room temperature X-ray diffractogram corresponding to the as-deposited Fe85 nanowire array. Measurements were performed with the X-ray beam incident on the array's bottom surface, after removing most of the gold layer. In the as-deposited sample, peaks corresponding to a majority (97%) γ-Fe(Pd) fcc disordered (A1) phase and to the α-Fe3Pd bcc solid solution phase are indexed, together with extra peaks arising from the remnant of the sputtered gold.
Fig. 2 X-ray diffractograms of Fe85 in the as-deposited state (before annealing, upper pattern) and after a heating–cooling thermal cycle, between 350 K and 1000 K at 10 K min−1 (lower pattern). |
The crystallographic phases and grain substructure in the as-deposited NWs were further characterized with TEM and STEM and the results are summarized in Fig. 3. Composition values and Fe/Pd maps obtained from HAADF imaging are depicted in Fig. 3a and b for an individual NW, confirming that these elements are alloyed with a composition of about 85 at% Fe, close to that obtained with SEM-EDS. TEM (Fig. 3c) and HRTEM (Fig. 3d) images confirm that the nanowires are polycrystalline, with mean grain size of about 16 nm, a value which is statistically indistinguishable from that estimated for the mean value of crystallite size with XRD data and the Williamson–Hall formula (s = (15 ± 5) nm). The electron diffraction pattern (FFT from the indicated zone) shown in the inset of Fig. 3d is consistent with a fcc lattice, as observed in the X-ray data.
These metastable atomic phases are expected to remain unchanged during measuring the magnetic properties at low temperature, while above room temperature (up to 1000 K), phase and/or ordering transformations can be promoted in addition to magnetic phase transitions, leading to complex effects during thermal cycling. Then, before measuring the magnetic hysteresis properties above room temperature, a thermal treatment consisting in a heating–cooling thermal cycle between 350 K and 1000 K, at 10 K min−1 was performed, which proved to be enough to reach a stable atomic microstructure during further high temperature measurements.
The XRD pattern corresponding to the final microstructure after the thermal cycle annealing is shown in the lower panel of Fig. 2. A similar volume fraction of the original γ-Fe(Pd) fcc disordered phase is still detected together with a quite small fraction (∼3%) of an ordered fcc phase, replacing the original α phase, which may be the L10 FePd equiatomic phase or the L12 Fe3Pd iron-rich one. As the majority phase peaks shift to higher angles after the thermal cycle (Δθ ∼ 0.2°), they overlap with the peaks of the minority phase, making it difficult to perform a confident identification of this secondary phase. Summarizing, from the data in Fig. 2 and 3, the crystallite size of the γ-Fe(Pd) phase is (15 ± 5) nm in the as-deposited state and (27 ± 5) nm after the thermal treatment, while the crystallite size of the minority phase is ∼9 nm in both conditions. The reduction in the lattice constant of the γ-Fe(Pd) phase observed after the high temperature cycle (from a = 3.8980 Å before the thermal cycle to a = 3.8873 Å afterwards) indicates that some atomic rearrangements take place, likely leading to a more relaxed structure and a more uniform matrix composition, but not large enough to achieve a phase transition to the equilibrium phases.
The changes in atomic structure and magnetic configurations during the thermal cycle were explored by measuring the evolution of the as-deposited array magnetization during a thermal cycle between 350 K to 1000 K, at a constant rate of 10 K min−1. The temperature dependence of the NW array magnetic moment was measured under an external magnetic field of 100 mT (about 1/10 of the saturation field), applied parallel (PA orientation) to the NW's axis. The obtained curves—shown in Fig. 4— display heating–cooling hysteresis indicating that some changes in the initial crystalline phases take place.
The initial heating curve in Fig. 4 (indicated with red arrows) exhibits nearly constant magnetization values up to about 750 K (increasing only by 1.8%); then, it notably decreases displaying a marked shoulder above 830 K, a temperature quite close to 823 K, reported for the ordering debut in these metastable γ-Fe(Pd) alloys.23 This indicates that above this temperature, magnetic and atomic processes overlap in the measured magnetization. After cooling and completing the thermal cycle, the magnetization is higher than at the beginning, confirming that some atomic rearrangements took place in the microstructure. All the magnetic measurements above room temperature were performed after a thermal cycle as the one just described.
These diffusion-controlled rearrangements in the γ-Fe(Pd) matrix at high temperature (which are also detected by XRD measurements) are responsible for the heating–cooling hysteresis observed in Fig. 4. It may be then concluded that the annealing tends to stabilize the initial main phase and changes the precipitated phase, leading to a microstructure that remains unchanged during subsequent thermal cycles below 1000 K. Then, magnetic measurements performed in the as-deposited condition at low temperature (below room temperature) should be well correlated to those performed at temperatures above 300 K and up to 1000 K, because the initial atomic microstructure in the as-deposited array is quite similar to that obtained after the high temperature thermal treatment. The changes underwent by the small precipitates are not detected, likely because of their small volume fraction in both conditions.
In what follows, the magnetic hysteresis properties at low temperature (below 400 K) are first characterized and then the same magnetic properties are measured at high temperature in the thermally treated array. Finally, hysteresis properties are analyzed in the whole temperature range.
Assuming that this maximum represents the blocking temperature distribution of the small α-Fe(Pd) precipitates, d = 9 nm in diameter, with a magnetocrystalline energy of about KCα ∼ 1 × 104 J m−3 (ref. 24) the mean blocking temperature TB may be estimated as , with kB the Boltzmann constant. This temperature is close to that estimated from the inset of Fig. 5.
In the FC curve, an increase in the magnetic moment below about 80 K is measured, which is observed for different values of applied field (not shown), suggesting a spin glass-like behavior at low temperature. This phenomenon is frequently reported in magnetic nano-objects,25–28 being ascribed to surface effects promoting ‘spin canting’, ‘spin pinning’ or ‘broken exchange bonds’. This effect will be addressed in more detail below.
The hysteresis mechanisms, leading to changes in the array magnetization were further investigated by measuring the hysteresis loops at different temperatures between 5 K and 800 K, for two field configurations: with the applied field parallel (PA) to the NWs axis and with the field perpendicular (PE) to this direction. The results for the PA and PE configurations shown in Fig. 6, between 5 K and 800 K, are consistent with a magnetization easy axis close to the NWs axis, as expected because of the array's shape anisotropy. Both, coercivity and remanence decrease with temperature in PA and PE configurations, as illustrated in Fig. 6c and d.
Fig. 7 shows the temperature dependence of the coercive field μ0HC, and the relative remanent magnetization (or squareness S = Mr/MS) in the range 5 K to 800 K, where a quite smooth transition between measurements at low and high temperature regimes—taken in different equipments—is observed.
Fig. 7 Temperature dependence of the coercive field (left axis) and the relative remanence or squareness ratio S (right axis), taken from the hysteresis loops shown in Fig. 6 for the PA field configuration. The yellow and the green solid lines over coercivity data correspond to linear fits described in the text. |
Both magnitudes undergo a similar transition at about 100 K, where two regimes can be clearly identified. The temperature dependence of coercivity below 100 K is well fitted by a linear law Hc(T) = Hc(0) – γT, where Hc(0) is the coercivity at T = 0 K, and γ is a fitting parameter; in our case, Hc(0) = (80 ± 2) mT, and γ = (0.32 ± 0.03) mT K−1; the best fit is shown as a solid yellow line in Fig. 7. At these low temperatures, frozen spin-glass-like surface spins are strongly pinned by exchange interaction with ordered core spins and hence extra energy is needed to switch the core spins. This “core–shell” interaction in the individual NWs is likely the responsible of the observed increase in coercive field at low temperatures.
Above 100 K, a linear coercivity decrease with temperature is also observed. In this case it is possible to correlate this dependence with a thermally activated, nucleation-controlled magnetization reversal mechanism. Hence, the temperature and field dependence of the activation energy barrier and the coercive field may be described as:29,30
(1) |
(2) |
Here, parameters μ0H0 and E0 are respectively, the critical field for magnetization reversal and the activation barrier height at T = 0 K; the slope of the line described by eqn (2) is then given by . The fitting of experimental data to eqn (1) and (2) lead to activation energy values of (320 ± 20) kJ mol−1 (3.3 ± 0.2 eV) and a critical reversal field at 0 K of (65 ± 1) mT. The activation energy values are comparable to those reported for a mechanism of reverse domain nucleation in Fe nanowires31 (2.4–5.1 eV), and Fe–Rh nanowires32 (2.1–4 eV) and larger by a factor two for Fe–Pd biphasic NWs above 300 K (ref. 15) (1.7 eV).
The temperature dependence of saturation magnetization MS is shown in Fig. 8 for the full range between 5 K and 800 K. The two regimes noticed for coercivity and squareness in Fig. 7 are also identified in MS, with an upturn below 100 K—see Fig. 8—likely arising from the spin glass-like contribution from magnetic moments in surface/interface zones in the NWs, which can be described by:33
(3) |
Fig. 8 Temperature dependence of the saturation magnetization relative to its value at 5 K. The solid line corresponds to the fitting of eqn (5) to experimental data. |
In addition to this low temperature spin-glass phase contribution to saturation magnetization, another one exists in the whole temperature range, between 5 K and 800 K, arising from the main phase inside the nanowires. This contribution can be well described by the phenomenological expression:
(4) |
Fig. 8 shows the relative saturation magnetization Ms(T)/Ms(5 K) (assuming that Ms(0) ≈ Ms(5 K)), together with the curve resulting from the best fit of the sum of eqn (3) and (4) to experimental data:
(5) |
Eqn (5) provides a good description of the data points in the whole temperature range, with the parameters values given in Table 1.
Fe85 | |
---|---|
A | 0.17 ± 0.05 |
Tf [K] | 80 ± 2 |
TC [K] | 830 ± 20 |
a | 4.8 ± 0.4 |
β (= b−1) | 0.42 ± 0.05 |
The low-temperature magnetization increase described by eqn (3) has also been observed in small-sized nanoparticle systems, in the range from 2.5 nm to 25 nm, being explained by considering the quantization of the spin-wave spectrum due to the small finite size of the particles,35,36 or in different systems by a surface effect.37–39 In our case, the NWs are ∼50 nm in diameter so such strong confinement effects are not expected. On the other hand, a differential contribution arising from the relatively large amount of spins in surface (misaligned) configurations, leading to a non-uniform magnetization through the nanostructures cannot be excluded. It is likely that the wires exhibit a magnetic ordered core, surrounded by a surface shell of disordered spins with a large number of broken exchange bonds, which can result in frustration and spin disorder. These spins are expected to freely fluctuate at high temperatures, more than those in the core, but they freeze at low temperatures in a disordered “spin-glass like” structure, being the energy barrier for freezing kBTf intimately related to the exchange interactions. In the present case, we obtain an energy kBTf = (11 ± 1) × 10−22 J, which is of the same order of magnitude as the exchange coupling constant between first neighbors (JFe–Fe(fcc) = 3.52 × 10−22 J and JFe–Pd = 17.3 × 10−22 J (ref. 40)) inside the nanomaterial.
The resulting value of parameter a = 4.8 is three times larger than that of the Bloch law in bulk ferromagnets (3/2), indicating that in the present conditions some hypotheses might not be valid. Recently, based on experimental results, an extension of Bloch law has been proposed,41,42 which also considers a temperature power law but with the exponent dependent on three fundamental conditions: dimensionality, predominant spin and anisotropy. Köbler et al.43 find that for integer spin 3-D exchange interactions, the T9/2 power function holds, which agrees with our result.
Concerning the critical exponent β = 0.42 ± 0.05 (= b−1) resulting from fitting the data in Fig. 8, it is close to that obtained with the Monte Carlo method, β = 0.5,44 and practically indistinguishable from the value β = 0.44 reported for FePd3.45 Note that a quite reasonable critical exponent β = 0.42 ± 0.05 is derived from the whole temperature range fitting to eqn (5).
Since the magnetocrystalline anisotropy of the major phase in the NWs is quite low (∼2 × 104 J m−3 (ref. 25)) then, only contributions arising from magnetostatic energy—the shape anisotropy KS of a single NW and the magnetostatic anisotropy KI, associated with inter-wire magnetostatic interaction in the array—are considered. Then, in this simple scenario we have:
Keff = KS + KI. | (6) |
The total magnetostatic effective anisotropy constant (the one arising from the shape anisotropy of individual nanowires added to that arising from dipolar inter-wire interactions) may be estimated as proposed by Carignan et al.,47 so the effective constant can be written as:
(7) |
In Fig. 9, the temperature dependence of KS and KI are shown; their respective positive and negative values in the whole temperature range correspond to the field configuration parallel to the NW magnetization easy axis. That is, for the present geometrical values of the NWs length and diameter and the inter-wire distance, the shape anisotropy of the individual cylindrical nanowires is stronger than that arising from the magnetostatic interactions’ energy among NWs. Thus, the effective anisotropy estimated by only considering magnetostatic contributions describes quite well the experimental data particularly above 100 K, supporting the dependence on MS2 proposed.
In the range 5 K–100 K, even when experimental data are statistically indistinguishable from the calculated values, errors are too large and other effects cannot be excluded. As commented before, the observed temperature dependence of the coercive field, the relative remanence and the saturation magnetization are consistent with a spin glass-like behavior associated with surface spin configurations.
After a heating–cooling thermal cycle between 350 K and 1000 K, the γ-Fe(Pd) fcc metastable phase still predominates (97%). The corresponding diffraction maxima are shifted to higher angles, indicating a slight contraction in the lattice parameter, likely due to the annealing of atomic defects and the debut of structural ordering at high temperature.
The saturation magnetization curve between 5 K and 800 K is well described by the superposition of two contributions: one arising from surface spins behaving as a “shell magnetic phase” in the NWs and a second one, a “core magnetic phase”, involving the magnetic moments inside the wires. The core contribution is accurately described by a phenomenological expression, which reproduces a Bloch-like law as T → 0, with an exponent larger than 3/2, and near the Curie temperature a critical exponent β = 0.42 ± 0.05 could be estimated. On the other hand, the spin-glass-like phase contribution becomes detectable below 100 K, being the freezing temperature Tf = (80 ± 2) K, and the energy kBTf comparable to the exchange interactions in FePd systems. These regimes (below and above 100 K) are also observed in the coercivity and the relative remanence curves as functions of temperature.
The temperature dependence of the coercive field in the array between 100 K and 800 K is consistent with a mechanism involving the nucleation of reverse domains and the subsequent propagation of the domain walls, with an activation energy of (320 ± 20) kJ mol−1 (3.3 ± 0.2 eV) and a critical field for magnetization reversal of (65 ± 1) mT, at 0 K.
The effective magnetic anisotropy estimated from the hysteresis loops between 5 K and 800 K shows a good concordance with an effective anisotropy arising only from magnetostatic effects, that is the shape anisotropy of individual NWs and dipolar inter-wire interaction in the array.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d4nr03119a |
This journal is © The Royal Society of Chemistry 2024 |