Amorphous nickel nanophases inducing ferromagnetism in equiatomic NiTi alloy

Amorphous nickel nanophases inducing ferromagnetism in equiatomic NiTi alloy

Acta Materialia 161 (2018) 47e53 Contents lists available at ScienceDirect Acta Materialia journal homepage: www.elsevier.com/locate/actamat Full l...

2MB Sizes 0 Downloads 29 Views

Acta Materialia 161 (2018) 47e53

Contents lists available at ScienceDirect

Acta Materialia journal homepage: www.elsevier.com/locate/actamat

Full length article

Amorphous nickel nanophases inducing ferromagnetism in equiatomic NieTi alloy M.R. Chellali a, *, S.H. Nandam a, S. Li a, b, M.H. Fawey a, E. Moreno-Pineda a, L. Velasco a, T. Boll b, c, L. Pastewka b, d, R. Kruk a, P. Gumbsch a, b, e, H. Hahn a, f, g a

Institute of Nanotechnology, Karlsruhe Institute of Technology, 76344, Eggenstein-Leopoldshafen, Germany Institute for Applied Materials, Karlsruhe Institute of Technology, 76344, Eggenstein-Leopoldshafen, Germany Karlsruhe Nano Micro Facility, Karlsruhe Institute of Technology, 76344, Eggenstein-Leopoldshafen, Germany d Department of Microsystems Engineering, University of Freiburg, 79110, Freiburg, Germany e Fraunhofer Institute for Mechanics of Materials IWM, 79108, Freiburg, Germany f KIT-TUD Joint Research Laboratory Nanomaterials, Institute of Materials Science, Technical University Darmstadt, 64287, Darmstadt, Germany g Herbert Gleiter Institute of Nanoscience, Nanjing University of Science and Technology, Nanjing, China b c

a r t i c l e i n f o

a b s t r a c t

Article history: Received 8 June 2018 Received in revised form 11 September 2018 Accepted 11 September 2018 Available online 15 September 2018

Ni50Ti50 nm-sized amorphous particles are prepared using inert-gas condensation followed by in situ compaction. Elemental segregation of Ni and Ti is observed in the consolidated nanostructured material. Amorphous, nearly pure Nickel (96%) nanophases form within the amorphous Ni50Ti50 alloy. Combining atom probe tomography and scanning transmission electron microscopy with computer modelling, we explore the formation process of such amorphous nanophase structure. It is shown that the Ni rich amorphous phase in the consolidated nanostructured material is responsible for the ferromagnetic behavior of the sample whereas the rapidly quenched amorphous and crystalline samples with the same chemical composition (Ni50Ti50) were found to be paramagnetic. Due to the high cooling rate obtained using the inert gas condensation technique, an exceptional control over the crystallization processes is possible, promoting the formation of various amorphous phases, which are not obtained by standard rapid quenching techniques. Our findings demonstrate the potential of amorphous metallic nanostructures as advanced technological materials, and useful magnetic compounds. © 2018 Published by Elsevier Ltd on behalf of Acta Materialia Inc.

Keywords: Amorphous nickel Molecular dynamics Magnetic Atom probe tomography

1. Introduction In many aspects modern technologies are focused on crystalline materials because of their beneficial magnetic [1], electrical [2], mechanical [3], and chemical [4] properties. However, the recent advances in experimental and modelling methods have opened new areas of investigation in the field of amorphous solids [5,6]. In this regard, a new family of non-crystalline solids termed as “nanoglasses” has been proposed by Gleiter and co-workers [6], which are fabricated by consolidating quenched nanometer-sized amorphous or glassy particles obtained with either thermal evaporation [7] or magnetron sputtering [8] in an inert gas atmosphere. The consolidated nanosized particles consist of amorphous regions, which are connected by interfaces characterized by a lower density

* Corresponding author. E-mail address: [email protected] (M.R. Chellali). https://doi.org/10.1016/j.actamat.2018.09.019 1359-6454/© 2018 Published by Elsevier Ltd on behalf of Acta Materialia Inc.

and higher free volume compared to the amorphous nanoparticle cores [6]. Molecular dynamics (MD) simulations have shown that these interfaces are characterized by a short- and medium-range order and an excess free volume different from the bulk [5,9]. Evidence to support the existence of such interfaces in consolidated nanostructured materials has been gathered from the combination of small angle X-ray diffraction and positron annihilation lifetime spectroscopy [10]. The findings have been further confirmed by €ssbauer spectroscopy (MS) investigations of Fe-Sc nanoglasses, Mo which clearly show interfacial regions that are a few atomic layers in thickness with a reduced density and a larger distribution of distances between nearest neighbor atoms [11]. This structural model is further supported by the fact that the volume fraction of the interfaces was found to vary inversely with the size of the amorphous nanoparticles and confirmed by observing the isomer €ssbauer spectra of PdeFeeSi alloy [12]. Although the shift in Mo structural studies have confirmed the high free volume in the interfacial regions, it is not clear whether the free volume is a

48

M.R. Chellali et al. / Acta Materialia 161 (2018) 47e53

chemical or structural effect. Recently, Cu and Fe enriched regions have been observed in Cu50Zr50 and Fe90Sc10 systems, respectively [13,14]. By comparing Pd80Si20 and Cu64Zr36 amorphous nanoparticles, Adjaoud et al. [15] demonstrated that palladium segregates to the shell and silicon segregates to the core in the PdeSi system, while copper migrates to the shell and zirconium to the core in CueZr nanospheres. As a result of such novel nanoamorphous structure, new properties of non-crystalline materials can be expected. In fact, they exhibit spectacular properties, such as high catalytic activity [16], improved mechanical performance [10], and enhanced magnetic characteristics [11], which differ significantly from any known amorphous material. Specifically, it has been shown by means of micro-compaction and tensile tests that such consolidated nanostructured materials have excellent plastic deformation ability compared to the melt spun ribbon with identical chemical composition. Such an improvement in ductility is probably due to less dense interfaces, which accommodate deformation thereby avoiding the fatal nucleation of system-spanning shear bands. A change in magnetic properties of the consolidated nanometer-sized amorphous particles in comparison to the reference ribbon has been observed in Fe90Sc10 alloy [11,17]. It has been demonstrated that the contribution of the itinerant electrons to the magnetization is different in nano-amorphous structure because of the high free volume in the interfaces. In this regard, we synthesized amorphous nanoparticles of equiatomic Nickel-Titanium (NieTi) similarly by inert gas condensation and thoroughly characterized their structural and magnetic properties. Ni50Ti50 alloys are in widespread use in robotic, engineering, medical, and aerospace applications [18e20], due to their favorable corrosion resistance, high hardness, and good electrical conductivity. While the alloys show unique properties as shape memory alloys and for bio-compatibility properties, to date, no ferromagnetic state has been reported for this material. In this study, we report on the finding of ferromagnetism in Ni50Ti50 consolidated nanostructured materials, and trace this to the effect of the interfacial regions. These are found to consist of amorphous Nickel phases with distinct ferromagnetic properties differing from those of the conventional paramagnetic crystalline and amorphous materials with the same chemical composition.

Philips and Bruker X-ray diffractometer equipped with a Mo-Ka and Cu-Ka radiation sources, respectively. Transmission electron microscopy (TEM) was carried out using FEI Tecnai F20-ST and FEI Titan 80e300 aberration corrected microscopes. The powder sample used for TEM analysis was collected directly from the cold finger in the IGC chamber using a glass test tube, which was transferred to the microscope in a special vacuum holder to prevent exposing the sample to air. Preparation of TEM lamellae and tips for atom probe tomography (APT) samples was performed by employing both FEI Strata 400 and Zeiss Auriga 60 focused ion beam systems (FIB). Prior to the lift-out, a platinum protection layer (150 nm) was deposited over the area of interest to protect the sample surface from gallium ion beam milling damage and sample degradation. To produce the required atom probe specimen geometry, annular milling was used to create needle-shaped morphology with a tip diameter smaller than 100 nm. More details about the FIB milling process can be found in Ref. [21]. The APT measurements were carried out using a Cameca-LEAP 4000 HR instrument in laser pulse mode (wavelength 355 nm, pulse frequency 100 kHz, pulse energy 60 pJ, evaporation rate 0.50%) at 50 K. Data processing was achieved with the CAMECA integrated visualization and analysis software (IVAS-version 3.6.1), incorporating standard reconstruction algorithms, allowing the extraction of three-dimensional nanoscale chemical distribution of all detected atoms in the analysis volume. The magnetic properties were investigated using the Quantum Design MPMS 5XL (SQUID) magnetometer. MD simulations were performed using the large-scale atomic/ molecular massively parallel simulator code [22]. A modified embedded-atom method potential was used to describe the interatomic interactions in the NieTi system [23]. The chosen potential gives a good fit to basic properties of pure crystalline Ni, pure crystalline Ti, and NieTi alloys. To replicate the synthesis procedure in IGC, we prepared the formation of nanoparticles by repeatedly adding 100 atoms (50 Ni atoms and 50 Ti atoms) with velocity on the order of kilometer per second in a virtual reaction chamber. The system was then equilibrated at 800 K for 100 ps. This process was repeated until a total of 20000 atoms were added to the chamber. XRD line profiles from the simulations were obtained by using the virtual diffraction algorithm [24].

2. Methods 3. Results and discussion Ni50Ti50 amorphous nanoparticles were prepared by using magnetron sputtering in inert gas condensation system (IGC), which has been described elsewhere [14]. The main difference to conventional thin film sputtering process is the much higher inert gas pressure in the chamber on the order of 10 1 mbar. Pellets with a diameter of ~8 mm, and a thickness of ~300 mm were then fabricated by in-situ compaction at a uniaxial pressure of 3 GPa, followed by ex-situ compaction in a high pressure torsion machine at 6 GPa. Such consolidated structure will be referred to as the amorphous nanophase materials (ANM's). The reference sample, amorphous metallic thin film (AMTF), was synthesized by Direct Current magnetron sputtering at a working pressure of 8 10 3 mbar. The base pressure of the magnetron sputtering chamber was ~10 8 mbar. The thickness of the films was ~900 nm. In order to avoid contamination with oxygen, a 5 nm Gold (Au) capping layer was deposited in-situ on top of the NieTi thin film. Crystallization of the thin film samples was initiated by annealing in an ultra-high vacuum furnace with a residual gas pressure of 1  10 8 mbar at 650  C for 60 min. The compositions of the ANM and AMTF were measured using an energy dispersive X-ray spectroscopy (EDX) (Oxford Instruments X-MaxN 50 mm2 Silicon Drift Detector) in a LEO 1530 scanning electron microscope. The structural properties of the samples were determined by X-ray diffraction (XRD) using a

The nature of the amorphous nature of ANM and AMTF is confirmed by X-ray diffraction and Transmission electron microscopy (see Fig. 1(aee)). A completely crystallized AMTF after annealing at 650  C for 1 h is also shown in Fig. 1(a). XRD peak profile shows only the B2 phase without any other phases. In addition, X-ray diffraction profiles were calculated based in the atomistic simulations. The corresponding profiles show a good match to the experimental ones, which will be discussed later. Fig. 1(b) shows a representative TEM image of the as prepared powder with sizes varying between 3 and 15 nm. The noncrystalline nature of the prepared nanoparticles is confirmed by the selected area electron diffraction (SAED) in Fig. 1(c), which is composed of broad diffraction halos without any crystalline diffraction spots. All these observations clearly indicate that the nanometer sized particles are completely amorphous. After high pressure compaction, the amorphous structure is still retained as can be seen in the XRD and high resolution TEM images as well as the electron diffraction patterns (see Fig. 1(a) and (d)). The elemental distribution mapping of the ANM sample obtained from the scanning transmission electron microscopy using an EDX shows Ni-rich and Ti-rich regions separated from each other (see Fig. 1(f)) on the length scale of about 10 nm. No segregation was observed in

M.R. Chellali et al. / Acta Materialia 161 (2018) 47e53

49

Fig. 1. (a) Experimental and simulated XRD patterns of the Ni50Ti50 amorphous nanophase material, amorphous metallic thin film and crystalline samples. TEM images of (b) Ni50Ti50 amorphous powder scraped from the cold finger, (c), and insets in (d) and (e) show the SAEDs confirming the amorphous nature of the samples, (d) and (e) HRTEM images showing amorphous nature of ANM and AMTF. (f) and (g) the compositional mappings of Ti in ANM and AMTF, respectively. The amorphous nature of Ni-rich zone (indicated by the arrows in d and f) is confirmed by the FFT in the inset (f).

the case of AMTF specimen (see Fig. 1(g)). The amorphous nature of Ni-rich zone in ANM was confirmed by the diffuse ring in the fast Fourier transform in the inset (f). Fig. 2(a) shows a full 3D chemical reconstruction of the ANM tip with sub-nanometer resolution. Individual atoms are represented by color-coded dots within the volume. By extracting a thin slice in the perpendicular direction (top to bottom) and using a color map of the atomic concentration, one can offer a suitable way to visualize the spatial and compositional Ni and Ti rich nanophase regions (see Fig. 2(b)), which are difficult to observe in 3D atomic reconstruction or isoconcentration surfaces because of the varied heterogeneity of the sample. Note that the nanophases are in the size range of 5e10 nm, which agrees closely with the average size of the primary nanoparticles. An isolated subvolume of the ANM sample is shown in the inset of Fig. 2(c). An iso-concentration surface at 74 at.% Ni is drawn into the picture. To give a better statistical picture of the NieTi segregation, a composition profile for each iso-concentration surface (see the inset of Fig. 2(c)) was then calculated by employing the so-called proxigram method [25]. This process was used to remove the effect of the elongated shapes of the nanoparticles caused from local magnification effects and the complicated morphology of nanoclusters that are in direct interaction with each other. Proxigram analysis reveals the chemical compositional variations at fixed distances from an isoconcentration surface, regardless of the actual shape and particle size. Positive and negative proximities correspond to the concentrations outside and inside of the selected surface (rose surface in the inset of Fig. 2(c)), respectively, which has the same sign as the Ni concentration gradient at the surface. The resulting concentration profile averaged over the surfaces is shown in Fig. 2(c). Ni is

highly enriched and is monotonically increasing from less than 50 at.% outside to ~96 at.% inside Ni-rich region. However, the APT analysis also reveals that no segregation is observed in AMTF (see Fig. 2(d)). In the present study, ANM consists of two regions, representing the core and the interface (see movie in the Supporting Information). It has been unambiguously identified that the oxygen within a range of 1e2 at.% segregates to the Ti rich core, while no oxygen segregation was found in the Ni-rich interface. It was surprising to identify a highly Ni rich (96 at.%) amorphous phase because there has been no clear and comprehensive study on pure or nearly pure amorphous nickel segregation in the literature. Murty et al. [26] investigated the amorphous transition in the composition ranges in Ti100-xNix system by high energy ball milling. The report indicates a broader amorphous-forming composition range from 10 to 70 at.% Ni, while crystallinity is found above 70 at.% Ni. Similarly, a broad amorphous-forming composition range was also observed by few authors in TieNi system in the content range of 23e72 at.% Ni27, 28. Moreover, Levit et al. [29] investigated Ni90Ti10 alloy thin films with a thickness of 100 nm using XRD, and analytical TEM. Their results revealed that the system has a crystalline nature corresponding to randomly oriented nanocrystalline Ni grains. Using the sonochemical processing, Ramesh et al. [30] have attempted to synthesize nanophasic amorphous Nickel on silica microspheres. Although this mechanism was effective to prepare fine powder with homogeneous particle size distribution, the prepared nanoparticles have either chemical purity or stability issues. Furthermore, Kim et al. [31] discovered pure Fe amorphous phases in submicron spheres of Iron-Nickel produced in high vacuum by electro-hydrodynamic atomization technique. Due to the small size

50

M.R. Chellali et al. / Acta Materialia 161 (2018) 47e53

Fig. 2. (a) A three-dimensional atom-by-atom LEAP tomographic reconstruction of an amorphous nanophase material in an analyzed volume of 20  20  300 nm3 (Ni atoms in blue and Ti atoms in magenta). (b) Section perpendicular to the specimen axis, alternating Ni-rich concentration gradient map. (c) Proxigram showing Ni and Ti concentrations as a function of distance to interface shown in the inset, using Ni 74 at.% isosurface. (d) The composition profile of Ni and Ti in the AMTF sample. (For interpretation of the references to color in this figure legend, the reader is referred to the Web version of this article.)

of the droplets, which makes the surface area-to-volume ratio very high, the droplets could cool at about 106 K/s. This prevents nucleation to occur even in highly under-cooled melts. In our study, amorphous Ni phases with content up to 96 at.% were present in the entire sample. Broad haloes and peaks were clearly visible in the SAED and XRD pattern of the Ni rich regions, and the entire sample, respectively (see Fig. 1). Due to the high cooling rates in IGC of around 108e1010 K/s [32], which is very close to the critical cooling rate (109e1010 K/s), required to freeze metals into metallic glasses [33], the amorphization of almost pure Nickel appears plausible. Supplementary video related to this article can be found at https://doi.org/10.1016/j.actamat.2018.09.019. To further investigate the atomic structure of amorphous nanoparticles synthesized from the sputtered atoms in IGC, we performed the atomistic simulations to prepare nanoparticles formed with a composition of Ni50Ti50 (see the method section). By the gradual addition of atoms in the ratio of 1:1 of Ni and Ti into the simulation box, small clusters were formed. As more and more atoms were added, these small clusters grew larger resulting in the coalescence of the nanoparticles. To check the possibility of surface segregation in the synthesized nanoparticles (see top inset in Fig. 3(a)), the compositional variation of species from the core to the surface was computed. The atoms were classified by their shortest distance to a free surface d. The maximum value of d in our simulated nanoparticles is approximately 3 nm. The bottom inset of Fig. 3(a) shows a cross section through the final cluster with atoms color-coded by d. Fig. 3(a) shows the computed distribution of Ni

and Ti concentration as the function of the minimum distance to the surface atoms d. In the inner core region (d > 1.6 nm), the data displays a large fluctuation due to the relatively limited number of atoms. In the middle core region (0.4 nm < d < 1.6 nm), Ni is depleted and the stoichiometry is roughly Ni46Ti54. The Ni content increases towards the surface. At around d z 0.4 nm, the concentration of Ni almost equals Ti. Ni is enriched at the surface with a stoichiometry of Ni63Ti37. The outermost atoms on the surface are almost pure Ni as seen on the top inset of Fig. 3(a). This is in agreement with the simulation results obtained on amorphous nanophase materials using Lennard Jones potential [5]. To simulate the structure of the ANM's synthesized experimentally, compaction of the nanoparticles was performed. Specifically, eight nanoparticles with a radius of 2.3 nm are placed in the simulation box. The nanoparticles were arranged into two layers with hexagonal-closed packed stacking. Periodic boundary conditions were applied in all directions. Subsequently, a hydrostatic pressure of 2.5 GPa was applied to the system and kept constant for 4 ns while equilibrating the system at 300 K in order to promote relaxation. The pressure was then increased to 6 GPa and the system was further relaxed by another 8 ns to fully relax the atoms. Finally, the pressure was released to zero and the stress-free state configuration was used for the following analysis. Fig. 3(b) shows the distribution of concentration of Ni atoms in a slice cut from the ANM. A strong segregation is found in the interfaces (red regions) where the maximum concentration of Ni can reach an atomic concentration of almost 80%. Correspondingly, the Ni concentration is low in the interiors of nanograins. To characterize the structure of

M.R. Chellali et al. / Acta Materialia 161 (2018) 47e53

51

Fig. 3. Atomistic simulations on the structural analysis of Ni50Ti50 an amorphous nanophase material. (a) Concentration of two species as a function of the minimum distance to surface d (blue for Ni and red for Ti) in Ni50Ti50 nanoparticle. Top inset: The configuration of nanoparticle. Atoms in blue and red refer to Ni and Ti. Bottom inset: The nanoparticle is cut in half with the color-coded by d. The atoms in black are close to the surface and the atoms in white are in the core region. The distance unit is in nanometer. (b) The distributions of Ni concentration in ANM, which was prepared by consolidating nanoparticles. The colors are shown according to the Ni concentration. The unit is in atomic ratio. (c) The radial distribution functions in core region (red dashed curve) and interface region (black solid curve) in ANM. (d) The dependence of maximum Ni concentration on nanograins of ANM. The black and red data were obtained from MD simulation and experiments separately. (For interpretation of the references to color in this figure legend, the reader is referred to the Web version of this article.)

the ANM's, we first distinguished the interface and core regions according to the Ni concentration (>0.6 for interface and <0.6 for core). The radial distribution function (RDF) g(r) was then computed in both regions. The black and red curve in Fig. 3(c) represent the g(r) for interface and core, respectively. For both curves, we can see that there is one broad peak in the short range (r ~ 0.26 nm) and no sharp peaks in the long range. This typical profile confirms that we obtained non-crystalline structure in both regions in accordance with experimental results (see inset in Fig. 1(g)). To characterize the structure, we carried out X-ray diffraction (XRD) calculations for B2 crystal, bulk amorphous material and ANM with composition Ni50Ti50 with both Cu and Mo radiation sources. The line profiles match quite well with the experimental measurements (Fig. 1(a)), which further confirms the validity for characterizing the structure from simulation. Furthermore, ANM have been prepared with different grain sizes. Fig. 3(d) shows the maximum Ni concentration at the interface region within ANMs (solid black squares) as a function of nanograin size. Here we can see an obvious size dependency that the concentration undergoes an almost linear decrease as radius of the core-interface like structure decreases. The experimental results obtained by APT (red circles) support the MD studies that the source for an increase in concentration is due to an increase in cluster size. As can be seen from experimental and computational results, there is a net segregation of Ti atoms to the cores and Ni to the interfaces. This can be explained by one of the two major physical

mechanisms [15,34]: (1) surface energy effect: the atoms with lower surface energy will segregate to the surface; and (2) the atomic sizes: when the atomic sizes of the two elements are extremely different, one component will segregate to the surface to reduce the strain energy in the system. As shown in Fig. 1, the particles are spherical in shape; there must be enough surface diffusion for the coalesced particles to form a sphere during synthesis. The preparation of nanoparticles in MD simulations were also performed at high temperature (800 K) leading to spherical nanoparticles. Therefore, we can assume that the temperature of the particles is above 800 K during the synthesis in IGC, from which the particles are cooled by collisions with the inert gas. This means that temperature simply enables diffusion of atoms, leading to elemental segregation in order to minimize the total Gibb's free energy. This observation is in agreement with results obtained by Ruban et al. [35], who showed that Ni, which has a negative segregation energy segregate towards the surface, and Ti, which has a positive segregation energy, tend to remain in the interior of the system. Since Ni rich amorphous regions were observed in ANM, their magnetic properties were studied using Superconducting Quantum Interference Device (SQUID) magnetometer. Fig. 4 illustrates the magnetization loops, M (magnetization) versus H (external magnetic field) of a ANM, a AMTF and a crystalline sample, all with the same overall chemical composition (Ni50Ti50). It is obvious that the crystalline and thin film specimens are paramagnetic at room

52

M.R. Chellali et al. / Acta Materialia 161 (2018) 47e53

Fig. 4. Magnetization curves for an amorphous nanophase materials, amorphous metallic thin film and crystalline sample at 300 K and 10 K. (a) and (b) Crystalline and AMTF samples exhibit a paramagnetic behavior, while the ANM shows a ferromagnetic shape reaching a magnetization of 0.22 mB and per 0.25 mB per Ni atom at RT and 10 K, respectively. (c) The magnetization of as prepared powder was not fully saturated at 5T, indicating a superparamagnetic behavior. The small sizes of the amorphous nanoparticles induce a thermal fluctuation in the magnetization.

temperature (RT). In contrast, the magnetization curve of ANM's display ferromagnetic characteristics and it achieves an average magnetization of about 0.26 mB per Ni atom at RT. This is about 40% of the total magnetization of pure crystalline nickel [36]. To further explore the magnetic ground state of the samples, magnetic measurements have been performed at low temperature. At 10 K, the AMTF and crystalline samples again indicate paramagnetic behavior. In contrast, the ANM reveals a typical hysteresis behavior and a small coercive field, with a magnetization of about 0.28 mB per Ni atom at 5 T. Although the nano-amorphous powder was ferromagnetic at RT and low temperature, the magnetization of these nanoparticles was not fully saturated. This could result from the small size of the as-prepared primary particles (leading to superparamagnetism) and from the partial segregation of Ni. In the latter case, only partially segregated Ni would result in weaker and more complicated ferromagnetic interactions (for example creation of canted ferromagnetism) showing, on average, smaller magnetic moment per Ni atom. The increase of the magnetic moment of Ni in ANMs compared to as-prepared particles can originate from an enhancement of the elemental segregation of amorphous Nickel at the interfaces. Such chemical disorder in magnetic nanostructures could have considerable impact on the magnetic properties of system [37,38]. Firstprinciple density functional theory calculations also show that segregation of platinum (Pt) to the surface of FeePt nanoparticles causes a decrease in their total magnetic moments, and a reduction in their magnetic anisotropy [37]. Both the magnetic moment and the Curie temperature are found to increase in the Iron-Chromuim

(FeeCr) system [38]. It was suggested that this enhancement is due to the alloy segregating into Cr-rich and Iron-rich phases. From elemental mapping and APT profile of the Ni50Ti50 amorphous nanophase materials specimen (see Figs. 1 and 2), the maximum composition of the Ni-rich region was found to be Ni96Ti4. To explore the Ni behavior at the above composition, a thin film of the similar composition of Ni90Ti10 was prepared using magnetron sputtering. XRD revealed the crystalline nature of the Ni90Ti10 thin film with no signature of any amorphous phase (See supplementary Fig. 1), which is in agreement with results reported in references [26e29]. This is another proof that the synthesis of Ni rich structure in binary NieTi alloys is difficult to make in amorphous form, however a nearly pure amorphous Nickel phase was successfully produced by using IGC. The supplementary fig. (2), presents the room temperature (RT) and low temperature magnetization loops, M (magnetization) vs. H (external magnetic field), of a Ni90Ti10 crystalline sample. The film shows a ferromagnetic behavior, with a magnetization of about 0.48 (1) mB and 0.59 (1) mB per Ni atoms at RT and low temperature, respectively. For a rough comparison of the net Ni magnetic moments in both cases (NG and sputtered thin film), we assume that the composition at the interfaces is Ni90Ti10, and the volume fraction is about 30%. Considering the saturation magnetization of Ni90Ti10 crystalline thin film as 0.59 mB per Ni atom, then the calculated magnetic moment per Ni atom in the segregated amorphous phase would be 0.2 mB, which is close to the value obtained in the amorphous nanophase materials (0.28 mB). The above estimate of the magnetic moment is performed under the assumption that the rest of Ni located in the core is not

M.R. Chellali et al. / Acta Materialia 161 (2018) 47e53

ferromagnetic because of the excessive amount of paramagnetic Ti atoms in the core. It is apparent that the ferromagnetic behavior of the ANM, compared to paramagnetic behavior of the AMTF and crystalline samples, comes from the internal magnetic nature of the amorphous ANM specimen. It is not caused by impurities such as oxygen or other metals. Possible impurities such as Ni-oxide (NiO) and Tidioxide (TiO2) give magnetic behaviors that are dissimilar from that of crystalline or amorphous nickel. It should also be noted that NiO and TiO are antiferromagnetic [39] and diamagnetic [40] compounds, respectively, which confirms that even a slight oxidation cannot be responsible for the observed ferromagnetism. Moreover, paramagnetic response has also been reported in pure Ti [41]. The results of the XRD diffraction and SAED pattern shown above seem to exclude crystallites of fcc Nickel or hcp Titanium as the source of the ferromagnetism. Therefore, the observed ferromagnetic behavior can be safely related to the amorphous-Ni rich interfaces. 4. Conclusion In the present study, Ni50Ti50 nm-sized amorphous particles were synthesized by means of inert gas condensation. The cooling rate reached in this technique is high enough to successfully freeze pure Nickel atoms to form amorphous nanophases within ANM. The presence of Ni- and Ti-rich regions at interface and core, respectively, is confirmed by detailed analysis of composition fluctuations in both atom probe tomography and atomistic simulation. The radial distribution function confirms the amorphous nature of both regions, keeping in accordance with experimental results obtained by XRD and TEM. Due to the unique atomic and electronic structures at the interfacial regions, the magnetic properties of amorphous nanophase materials were found to differ significantly from the corresponding properties of metallic amorphous metallic thin film and crystalline samples with the same chemical compositions. Besides the remarkable magnetic properties reported in this communication, introducing interfaces into metallic, ionic, or covalent glasses on a nanometer scale may provide an alternative dimension for tailoring properties and performance of nanostructured materials. Acknowledgements The authors are grateful to the Karlsruhe Nano Micro Facility (KNMF) for support and access to APT, FIB and TEM facilities. MRC acknowledges the financial support received from the Karlsruhe Institute of Technology (KIT). HH and SHN are grateful for the financial support by the Deutsche Forschungsgemeinschaft (DFG) under contract HA 1344/30-1. LP acknowledges financial support by the Deutsche Forschungsgemeinschaft (DFG) under contract PA 2023/2. Calculations were carried out at the Jülich Supercomputing Center (project hfr13). Appendix A. Supplementary data Supplementary data to this article can be found online at https://doi.org/10.1016/j.actamat.2018.09.019.

53

References [1] M.E. McHenry, D.E. Laughlin, Acta Mater. 48 (1) (2000) 223e238. n ~ ez, Z. Luo, A. Genç, L. Piveteau, S. Ortega, D. Cadavid, O. Dobrozhan, [2] M. Iba Y. Liu, M. Nachtegaal, M. Zebarjadi, J. Arbiol, M.V. Kovalenko, A. Cabot, Nat. Commun. 7 (2016), 10766. [3] R. Valiev, Nature 419 (2002) 887. [4] M.P. Stewart, J.M. Buriak, Adv. Mater. 12 (12) (2000) 859e869. [5] D. Danilov, H. Hahn, H. Gleiter, W. Wenzel, ACS Nano. 10 (3) (2016) 3241e3247. [6] H. Gleiter, Acta Mater. 56 (19) (2008) 5875e5893. € fler, J. Logas, Appl. Phys. Lett. 57 (17) (1990) [7] R. Averback, H. Hahn, H. Ho 1745e1747.  [8] Z. Sniadecki, D. Wang, Y. Ivanisenko, V.S.K. Chakravadhanula, C. Kübel, H. Hahn, H. Gleiter, Mater. Char. 113 (Supplement C) (2016) 26e33. [9] D. S¸opu, K. Albe, Y. Ritter, H. Gleiter, Appl. Phys. Lett. 94 (19) (2009), 191911. [10] J.X. Fang, U. Vainio, W. Puff, R. Würschum, X.L. Wang, D. Wang, M. Ghafari, F. Jiang, J. Sun, H. Hahn, H. Gleiter, Nano Lett. 12 (1) (2012) 458e463. [11] M. Ghafari, H. Hahn, H. Gleiter, Y. Sakurai, M. Itou, S. Kamali, Appl. Phys. Lett. 101 (24) (2012) 243104. €mer, R. Birringer, H. Gleiter, U. Gonser, J. Non-Cryst. Solids 113 (2) [12] J. Jing, A. Kra (1989) 167e170. [13] C. Wang, D. Wang, X. Mu, S. Goel, T. Feng, Y. Ivanisenko, H. Hahn, H. Gleiter, Mater. Lett. 181 (Supplement C) (2016) 248e252.  [14] S.H. Nandam, Y. Ivanisenko, R. Schwaiger, Z. Sniadecki, X. Mu, D. Wang, R. Chellali, T. Boll, A. Kilmametov, T. Bergfeldt, H. Gleiter, H. Hahn, Acta Mater. 136 (Supplement C) (2017) 181e189. [15] O. Adjaoud, K. Albe, Acta Mater. 113 (Supplement C) (2016) 284e292. [16] N. Chen, R. Frank, N. Asao, D.V. Louzguine-Luzgin, P. Sharma, J.Q. Wang, G.Q. Xie, Y. Ishikawa, N. Hatakeyama, Y.C. Lin, M. Esashi, Y. Yamamoto, A. Inoue, Acta Mater. 59 (16) (2011) 6433e6440. [17] R. Witte, T. Feng, J.X. Fang, A. Fischer, M. Ghafari, R. Kruk, R.A. Brand, D. Wang, H. Hahn, H. Gleiter, Appl. Phys. Lett. 103 (7) (2013), 073106. [18] Y. Furuya, H. Shimada, Mater. Des. 12 (1) (1991) 21e28. [19] N. Sharma, K.K. Jangra, T. Raj, Proc. IME J. Mater. Des. Appl. 232 (3) (2015) 250e269. [20] S.S. Abdullah, J.H. Taha, S.M. Hasan, J. Genet. Environ. Resour. Conserv. 4 (2) (2016) 75e79. [21] M.K. Miller, K.F. Russell, K. Thompson, R. Alvis, D.J. Larson, Microsc. Microanal. 13 (6) (2007) 428e436. [22] S. Plimpton, J. Comput. Phys. 117 (1) (1995) 1e19. [23] W.-S. Ko, B. Grabowski, J. Neugebauer, Phys. Rev. B 92 (13) (2015) 134107. [24] S. Coleman, D. Spearot, L. Capolungo, Model. Simulat. Mater. Sci. Eng. 21 (5) (2013), 055020. [25] O.C. Hellman, J.A. Vandenbroucke, J. Rüsing, D. Isheim, D.N. Seidman, Microsc. Microanal. 6 (5) (2002) 437e444. [26] B.S. Murty, S. Ranganathan, M. Mohan Rao, Mater. Sci. Eng. 149 (2) (1992) 231e240. [27] R.B. Schwarz, R.R. Petrich, C.K. Saw, J. Non-Cryst. Solids 76 (2) (1985) 281e302. [28] K.H.J. Buschow, J. Appl. Phys. 56 (2) (1984) 304e306. [29] M. Levit, I. Grimberg, B.Z. Weiss, J. Appl. Phys. 80 (1) (1996) 167e173. [30] S. Ramesh, Y. Koltypin, R. Prozorov, A. Gedanken, Chem. Mater. 9 (2) (1997) 546e551. [31] Y.-W. Kim, H.-M. Lin, T.F. Kelly, Acta Metall. 36 (9) (1988) 2525e2536. [32] C.C. Koch, Nanostructured Materials: Processing, Properties and Applications, William Andrew, 2006. [33] S. Banerjee, P. Mukhopadhyay, Phase Transformations: Examples from Titanium and Zirconium Alloys, vol. 12, Elsevier, 2010. [34] M. Müller, K. Albe, Phys. Rev. B 72 (9) (2005), 094203. [35] A.V. Ruban, H.L. Skriver, J.K. Nørskov, Phys. Rev. B 59 (24) (1999) 15990e16000. [36] L. Sacconi, J. Am. Chem. Soc. 74 (18) (1952) 4503e4505. [37] H. Lv, Y. Lei, A. Datta, G. Wang, Appl. Phys. Lett. 103 (13) (2013), 132405. [38] Y. Ishikawa, R. Tournier, J. Filippi, J. Phys. Chem. Solid. 26 (12) (1965) 1727e1745. [39] E. Winkler, R.D. Zysler, M.V. Mansilla, D. Fiorani, D. Rinaldi, M. Vasilakaki, K.N. Trohidou, Nanotechnology 19 (18) (2008) 185702. rez, F. Jime nez-Vil[40] A. Serrano, E.F. Pinel, A. Quesada, I. Lorite, M. Plaza, L. Pe €mer, J.F. Fernandez, lacorta, J. de la Venta, M.S. Martín-Gonz alez, J.L. Costa-Kra J. Llopis, M.A. García, Phys. Rev. B 79 (14) (2009), 144405. [41] H. Helten, B. Dutta, J.R. Vance, M.E. Sloan, M.F. Haddow, S. Sproules, D. Collison, G.R. Whittell, G.C. Lloyd-Jones, I. Manners, Angew. Chem. Int. Ed. 52 (1) (2013) 437e440.