7

Magnetic nanowires and submicron wires prepared by the quenching and drawing technique

T.-A. Óvári; N. Lupu; H. Chiriac    National Institute of Research and Development for Technical Physics, Iasi, Romania

Abstract

Glass-coated amorphous ferromagnetic nanowires and submicron wires prepared by means of rapid solidification techniques are a novel class of ultrathin, magnetically soft materials with great potential to be used in novel micro- and nanosensing devices and domain wall logic-based applications. The main characteristics of their magnetic bistable behavior, along with the specific aspects of the domain wall propagation, are analyzed. These phenomena are likely to be among the first operating principles for applications based on these new materials. New characterization methods that have been specifically developed for such ultrathin wires are also discussed, together with future work to be carried out on this subject.

Keywords

Amorphous magnetic nanowires

Domain wall propagation

Ferromagnetic glass-coated wires

Magnetic bistability

Magnetic hysteresis.

Acknowledgements

This work was supported by the Romanian Ministry of Education and Scientific Research through project no. PN-II-ID-PCE-2012-4-0424 (contract no. 46/2013) and by the European Commission through the FP7-REGPOT-2012-2013-1 project entitled “Upgrading the capacity of NIRDTP to develop sensing applications for biomedicine using magnetic nanomaterials and nanostructured materials”—NANOSENS (grant agreement no. 316194).

7.1 Introduction

Composite amorphous glass-coated microwires prepared using the drawing and quenching method specifically known as “glass-coated melt spinning” have metallic nuclei with diameters between 1 and 50 μm and a glass coating thickness within the same range (Chiriac and Óvári, 1996). These rapidly solidified microwires have been extensively studied in the past 20 years, mainly in view of developing various sensing applications (Chiriac et al., 1999; Vázquez, 2001; Zhukov, 2006). On the other hand, recent advances in sensor miniaturization, as well as novel developments in magnetic logic devices (Allwood et al., 2005) and spintronic applications (Parkin et al., 2008), have brought to the attention of scientists and engineers various types of magnetic nanowires with much smaller dimensions when compared with microwires. Among these, nanowires prepared using electrochemical methods (Medina et al., 2009) and those prepared using lithographic techniques (Hayward et al., 2010) represent the mainstream preoccupations with respect to the study of materials aimed at such applications. Many newly sought applications are associated with the motion of domain walls within magnetic nanowires (Negoita et al., 2013), which is also the most encountered mechanism of their magnetization reversal (Pereira et al., 2013), being the origin of their main magnetic characteristics.

Domain wall propagation has been a topic of growing interest in rapidly solidified amorphous glass-coated microwires (e.g., Zhukov, 2001; Varga et al., 2005; Chizhik et al., 2008; Chiriac et al., 2008; Óvári et al., 2011a.). The results of numerous studies of domain wall propagation in amorphous glass-coated microwires have pointed to some unique aspects of this phenomenon, which have been clearly linked to the special magnetic structures formed in positive and nearly zero magnetostrictive microwires. Thus, the domain wall velocities reported in amorphous microwires are very large, of the order of 1000 to well over 2000 m/s (Olivera et al., 2008), and in some particular situations are extremely large, for example, 18.5 km/s (Varga et al., 2007). These values are generally larger than those encountered in the typical planar nanowires used in domain wall logic-related experiments, and, most importantly, they are obtained at much smaller applied magnetic fields—for example, a few amperes per meter for nearly zero magnetostrictive microwires (Chiriac et al., 2009)—than in the case of lithographically patterned nanowires, in which thousands of amperes per meter are required (Atkinson et al., 2003). The wall mobility values are also quite large in microwires, and both wall velocity and mobility depend on the type of magnetic structure from the near-surface region of the microwires.

Given the advantages of the domain wall propagation characteristics in microwires, there has been an increasing interest in creating a new class of nanowires with reduced dimensions that are suitable for novel domain wall logic applications and for the development of new, miniaturized sensing devices, such as microsensors and nanosensors, with the added benefits of the fast domain wall motion observed in amorphous microwires. Therefore, the idea of making magnetically soft amorphous nanowires by means of rapid solidification methods has arisen. If successful, such materials would be a cheaper alternative to nanowires prepared using various lithographic techniques.

The main idea was to improve the glass-coated melt spinning method to allow the preparation of wire-shaped magnetic materials with much smaller dimensions. Thus, in this well-known technique, special attention has been paid to increasing the drawing speed and attenuating the associated vibrations to keep the wire-forming process continuous. The most important parameter is the surface tension of the molten alloy, which has to be optimal to ensure its continuous flow through the softened glass capillary. Therefore, the alloy has to be overheated more than in the case of the glass-coated microwires prepared using the same method. Additional changes refer to the cooling of the nanowire, which has been adapted to the novel, reduced transverse dimensions.

The successful results of the first attempts to prepare such ultrathin wires by means of glass-coated melt spinning were reported a few years ago (Chiriac et al., 2010a), when nearly zero magnetostrictive glass-coated submicron amorphous wires with a Co68.15Fe4.35Si12.5B15 metallic nucleus with a diameter of 800 nm and a 6-μm-thick Pyrex glass coating were made and studied from the point of view of their magnetization reversal mechanism and magnetic anisotropy distribution. Figure 7.1 shows a scanning electron microscopy image of the first submicron amorphous glass-coated wire prepared by means of rapid solidification. The preparation of iron (Fe)-rich submicron amorphous wires with similar nucleus diameters has been recently reported (Chizhik et al., 2013).

f07-01-9780081001646
Figure 7.1 Scanning electron microscopy image of a glass-coated amorphous submicron wire with a metallic nucleus diameter of 800 nm.

The most significant breakthrough was reported soon after the preparation of the first submicron amorphous wire (Chiriac et al., 2011a), when the preparation of glass-coated amorphous nanowires with metallic nucleus diameters between 90 and 180 nm were reported for the first time. Samples with both nearly zero magnetostriction (Co68.15Fe4.35Si12.5B15) and highly positive magnetostriction (Fe77.5Si7.5B15) have been made and studied experimentally. A scanning electron microscopy image of a rapidly solidified, Fe77.5Si7.5B15 amorphous glass-coated nanowire with a nucleus diameter of 100 nm is shown in Figure 7.2. Obviously, amorphous glass-coated nanowires and submicron wires with intermediate values of nucleus diameter also have been prepared and investigated, so that the entire range of diameters between 90 and 950 nm has been fully covered (Chiriac et al., 2011b). Highly magnetostrictive and nearly zero magnetostrictive alloy compositions have been studied for the whole range of dimensions to understand the peculiarities of the nanowires’ magnetic behavior, which were believed to originate in the magnetomechanical coupling between magnetostriction and internal stresses induced by the quenching and drawing preparation technique.

f07-02-9780081001646
Figure 7.2 Scanning electron microscopy image of a glass-coated amorphous nanowire with a metallic nucleus diameter of 100 nm.

The main characteristic of the magnetic behavior of amorphous glass-coated nanowires and submicron wires is the presence of magnetic bistability. The magnetic bistable behavior implies a one-step axial magnetization reversal through the displacement of a single magnetic domain wall along the entire wire length when the axially applied field reaches to a certain threshold value, called the switching field, H* (Vázquez et al., 1994). It occurs either through the nucleation of a new end domain with reversed magnetization followed by its propagation, in which case the switching field is the nucleation field (H* = Hn), or through the propagation of a preexisting end domain, in which case the switching field is the propagation field (H* = Hp) (Chiriac et al., 1998). The experimental proof of magnetic bistability is a rectangular hysteresis loop measured at low fields.

Magnetic bistability is a general feature of highly positive, magnetostrictive amorphous microwires, and in some particular cases it can also occur in nearly zero magnetostrictive ones (Chiriac et al., 2007). Nevertheless, nearly zero magnetostrictive amorphous glass-coated microwires generally exhibit an anhysteretic magnetic behavior (Zhukov et al., 2003).

Taking into account the need for cheaper and more versatile magnetic nanowires for sensing and magnetic domain wall logic applications, as well as the basic scientific interest in magnetic bistable behavior as one of the main sources of application-related phenomena in magnetic wires, microwires, submicron wires, and nanowires, there has recently been an increased focus on understanding the changes that take place in the parameters of magnetic switching with a reduction in the diameter of amorphous, wire-shaped magnetic materials. Purely scientific interest has been doubled by the practical requirements to control and tailor such parameters for the various envisaged applications. Hence, we present an overview of the main results concerning magnetic behavior and domain wall propagation in glass-coated amorphous nanowires and submicron wires prepared by means of the drawing and quenching technique, as well as the most significant aspects of their fundamental and practical importance.

7.2 Magnetic behavior

7.2.1 Characterization methods

One of the main sources of information with respect to the magnetization processes in ultrathin magnetic nanowires and submicron wires prepared by rapid solidification is represented by experimental hysteresis loops. There are two types of hysteresis loop measurements that give an overall picture of the magnetic behavior of glass-coated amorphous wires in general: bulk hysteresis loops, usually measured by means of inductive methods (Zhukov et al., 1995), and surface hysteresis loops, typically measured using magneto-optical Kerr effect (MOKE) techniques (Chizhik et al., 2006). In the case of the ultrathin wires, such as glass-coated amorphous nanowires and submicron wires, the typical inductive experimental setup is not suitable because of the very small signal induced in the pickup coils given by the very small transverse dimensions of the samples.

To be able to measure the bulk hysteresis loops of amorphous nanowires and submicron wires, a new method has been recently developed and used (Corodeanu et al., 2011a). The novel method is based on the magnetic bistability of the wires, which always results in rectangular hysteresis loops, and uses a digital integration technique. Figure 7.3 shows a schematic of the experimental setup. The main components are

f07-03-9780081001646
Figure 7.3 Schematic of the experimental setup for hysteresis loop measurements in amorphous glass-coated nanowires and submicron wires.

 a magnetizing solenoid;

 a system of pickup coils;

 a low-noise preamplifier;

 a function generator; and

 a data acquisition board.

The length of the samples measured with this system is about 4 cm, whereas the diameter can take any value between 90 and 1000 nm. The technique, which uses a so-called window method (Butta et al., 2009), works for any magnetically bistable ultrathin wire, and the resulting hysteresis loop is free of noise, as illustrated in Figure 7.4.

f07-04-9780081001646
Figure 7.4 Rectangular hysteresis loop of a glass-coated amorphous Fe77.5Si7.5B15 nanowire measured using the digital integration technique.

MOKE, on the other hand, is an effective tool for studying the surface magnetization of magnetic materials. It has been successfully used for wire-shaped materials, starting from conventional amorphous wires quenched in rotating water with diameters larger than 100 μm (e.g., Chizhik et al., 2002; Chiriac et al., 2011c), continuing with thinner glass-coated amorphous microwires (Chizhik et al., 2012; Chiriac et al., 2010b), and finishing with glass-coated, ultrathin amorphous submicron wires (Chiriac et al., 2011b) and nanowires (Țibu et al., 2012). Given that the magnetic properties of the near-surface region are particularly interesting in amorphous microwires (Óvári et al., 2009), investigation of these properties in much thinner submicron wires and nanowires is essential. For instance, the MOKE axial surface hysteresis loop of a Co68.15Fe4.35Si12.5B15 amorphous glass-coated submicron wire with a nucleus diameter of 800 nm is shown in Figure 7.5.

f07-05-9780081001646
Figure 7.5 Axial magneto-optical Kerr effect hysteresis loop for a Co68.15Fe4.35Si12.5B15 submicron amorphous glass-coated wire with a metallic nucleus diameter of 800 nm.

MOKE measurements in ultrathin wire-shaped samples have been typically performed using a NanoMOKE II magnetometer produced by Durham Magneto Optics Ltd. In a longitudinal configuration, the rotation of the plane of polarization is proportional to the magnetization component parallel to the plane of incidence. The plane of incidence is parallel to the wire axis. Polarized light from a helium–neon laser, with a wavelength of 635 nm, is reflected from the wire to the detector. The diameter of the light beam is about 2 μm, whereas the penetration depth of the laser light is about 9 nm.

To fill the eventual gaps in the experimental study of the magnetization process of rapidly solidified amorphous nanowires and submicron wires, ferromagnetic resonance (FMR) measurements have been performed. These investigations are aimed at studying the magnetic anisotropy distribution within the near-surface region of the nanowires and submicron wires (Chiriac et al., 2010a). The correlation between FMR and MOKE results can offer comprehensive information about the magnetic characteristics within the wire surface region. FMR spectra are typically measured using an X-band spectrometer (Chiriac et al., 2000). A modulation technique is applied, in which the direct current magnetic field is modulated with an alternating field with a low frequency (1 kHz) and an amplitude of about 800 A/m. Changing the working frequencies, for example, from 8.5 to 10.5 GHz, results in the modification of the magnetic penetration depth.

7.2.2 Main aspects of the magnetic behavior

As was pointed out in the Introduction section, magnetic bistability is the main characteristic of glass-coated amorphous nanowires and submicron wires prepared by means of rapid solidification. It has been emphasized in highly positive magnetostrictive compositions, such as Fe77.5Si7.5B15, in nearly zero magnetostrictive ones, for example, Co68.15Fe4.35Si12.5B15, as well as in FINEMET- and NANOMET-type compositions.

Figure 7.6 illustrates the inductive hysteresis loop of a glass-coated, amorphous Fe77.5Si7.5B15 submicron wire with a nucleus diameter of 500 nm, whereas Figure 7.7 shows the inductive loop of a Co68.15Fe4.35Si12.5B15 glass-coated, amorphous submicron wire with the same nucleus diameter. The coercivities are one order of magnitude larger in the Fe-rich sample compared with the cobalt (Co)-rich one. This shows the importance of the magnetomechanical coupling between internal stresses induced during preparation and the magnetostriction of the alloy in the case of highly magnetostrictive samples. Based on the results obtained in the case of amorphous glass-coated microwires, one can expect to have two main contributions that decide the overall characteristics of the magnetic behavior in the case of ferromagnetic, amorphous glass-coated nanowires and submicron wires prepared using a similar technique: (1) the magnetoelastic term, which is the result of the above-mentioned coupling between mechanical internal stresses and magnetostriction, and (2) the magnetostatic term, which gives rise to so-called magnetic shape anisotropy. The magnetoelastic term is expected to be preponderant in the case of highly magnetostrictive alloy compositions, whereas the magnetostatic one should be more important in samples with low magnetostriction. Hysteresis loop and coercivity data are not sufficient to clearly understand the impact of each contribution. Therefore, further investigations were required to obtain additional information on the magnetization reversal characteristics in these new wire-shaped materials.

f07-06-9780081001646
Figure 7.6 Inductive hysteresis loop of a glass-coated, amorphous Fe77.5Si7.5B15 submicron wire with a nucleus diameter of 500 nm.
f07-07-9780081001646
Figure 7.7 Inductive hysteresis loop of a glass-coated, amorphous Co68.15Fe4.35Si12.5B15 submicron wire with a nucleus diameter of 500 nm.

The study of nanowires and submicron wires made from FINEMET-type alloys (Fe73.5Cu1Nb3Si13.5B9) is important for trying to clarify such aspects. In the as-cast amorphous state, they exhibit a typical amorphous structure with a large and positive magnetostriction constant, as the main alloy group from which they originate (Fe–silicon–boron). However, the copper and niobium additions lead to the controlled formation and growth of α-FeSi nanosized grains, which are randomly spread through the remaining amorphous matrix. The fine balance between the two phases, which also implies certain limits to the size of the grains and certain distances among them, leads to the formation of a special structure called the nanocrystalline structure, whose main attribute is a vanishing magnetostriction (Herzer, 1997). This structure is obtained after well-defined stages of annealing to achieve the right density of grains and suitable grain dimensions to average out the overall magnetostriction. This recipe applies to FINEMET materials of any shape. Therefore, it was expected to work for rapidly solidified nanowires and submicron wires as well. Figure 7.8 illustrates the dependence of the maximum relative magnetic permeability on the annealing temperature for FINEMET nanowires and submicron wires with various diameters. The peak value corresponds to the formation of the nanocrystalline structure, whereas the subsequent decrease is determined by the growth of the crystalline grains, which alters the zero magnetostriction state. It is important to emphasize that the FINEMET nanowires and submicron wires are magnetically bistable in the as-cast state, and bistability is maintained after the various stages of annealing (1 h at temperatures between 500 and 650 °C). The nanocrystalline structure appears after annealing at 600 °C. This means that in this case, magnetic bistability is the result of the magnetostatic contribution (shape anisotropy), rather than of the magnetoelastic term. Thus, both terms are able to yield a magnetically bistable behavior.

f07-08-9780081001646
Figure 7.8 Maximum relative magnetic permeability versus annealing temperature for FINEMET glass-coated nanowires and submicron wires with various diameters.

Initial hysteresis studies have been extended to samples with different dimensions to analyze the influence of wire dimensions on coercivity values. The dependence of coercivity on nanowire diameter for a fixed glass coating thickness is represented in Figure 7.9 for Fe77.5Si7.5B15 amorphous samples (Óvári and Chiriac, 2014). Figure 7.10 shows the normalized coercivity and normalized magnetoelastic anisotropy constant values (calculated based on the internal stress distributions) versus wire diameter. The two curves display a similar shape. They are basically overlapped in the range of 100–300 nm, proving the magnetoelastic nature of the magnetization reversal process in the case of the thinnest samples. As the diameter increases, however, there is a continuously widening gap between the two curves, with the switching field versus diameter curve dropping below the other one. This clearly indicates that the magnetoelastic term is not solely responsible for magnetic switching, even in highly magnetostrictive amorphous nanowires and submicron wires. The magnetostatic term is expected to play a significant role as well. This means that shape anisotropy produces various effects on the magnetization process of amorphous nanowires and submicron wires prepared by rapid solidification, depending on their dimensions and composition.

f07-09-9780081001646
Figure 7.9 Coercivity versus metallic nucleus diameter for Fe77.5Si7.5B15 amorphous glass-coated nanowires and submicron wires prepared by rapid solidification.
f07-10-9780081001646
Figure 7.10 Dependence of normalized coercivity and normalized magnetoelastic anisotropy constant on wire diameter for Fe77.5Si7.5B15 amorphous submicron wires and nanowires.

To extend the investigations and to explain these aspects, the actual magnetization reversal mechanism has been brought into focus, that is, the domain wall displacement along the wire length. Understanding the peculiarities of domain wall propagation allows one to understand the basic aspects of magnetization reversal and constitutes the basis for future applications, mainly domain wall logic applications.

7.3 Domain wall propagation

Domain wall propagation is the mechanism through which the magnetization reversal takes place in ultrathin amorphous wires. It is actually the only magnetization reversal mechanism encountered in glass-coated amorphous nanowires and submicron wires prepared by drawing and quenching, and therefore its study is expected to shed some light on the insights of their basic magnetic behavior. Moreover, domain wall displacement is among the main future application principles envisaged for this new class of rapidly solidified nanometric materials. As was briefly mentioned in Section 7.1, there are two ways in which domain wall propagation occurs: nucleation and propagation or just propagation. Because of the demagnetizing effects, there usually are end domains with reversed magnetization at the wire ends (see Figure 7.11).

f07-11-9780081001646
Figure 7.11 Schematic drawing of an end domain with reversed magnetization that preexists at one end of the wire.

Therefore, propagation of the preexistent wall between such an end domain and the rest of the sample is the most encountered situation. Once the applied magnetic field reaches the critical switching field value, which is the propagation field Hp, the wall moves towards the opposite end, in the direction of the applied field. Wall displacement is characterized by the domain wall velocity v and by the domain wall mobility S, defined as dv/dH. Both parameters should be related to the nature of the preponderant axial magnetic anisotropy—be it magnetoelastic or shape anisotropy—and their values should be influenced by any misalignments of the magnetization (magnetic moments) within the near-surface region (Óvári et al., 2011b). Their values and dependence on various nanowire characteristics, such as dimensions and composition, should therefore give essential information about the basic magnetic properties and behavior. In addition, the wall velocity determines the speed of operation in devices and applications based on domain wall movements. Consequently, the study of magnetic domain wall propagation in these novel amorphous wires is crucial because it should be able to clarify many of their basic properties and to offer ways of controlling and tailoring the domain wall velocity and mobility for practical reasons.

7.3.1 Experimental techniques

Similar to the case of the specific experimental setup required for studying the hysteresis loops of such ultrathin wires, a special setup was designed to measure the domain wall propagation characteristics of glass-coated amorphous nanowires and submicron wires (Corodeanu et al., 2011b). The method is based on the classical Sixtus–Tonks method (Sixtus and Tonks, 1932); however, it addresses a couple of shortcomings, one in particular that is related to the small transverse dimensions of the rapidly solidified nanowires and submicron wires:

1. it picks up very small signals associated with the ultrathin wires of interest; and

2. it allows the detection of additional domain walls, which appear as a result of newly nucleated domains anywhere along the wire, to avoid having false readings of unrealistic and very large domain wall velocities.

The issue of small signals has been addressed using two identical systems of pickup coils that are connected in series-opposition to obtain a compensated system that provides only the signal because of the propagating domain walls. The detection of additional domain walls—those due to the generated domains with reversed magnetization anywhere else on the wire length during the propagation of the end domain walls—has been addressed using a number of four pickup coils in each of the compensated systems. Moreover, to easily detect the direction of the propagating domain walls, each of the four pickup coils has had three windings; the middle one is wound in the opposite direction with respect to the other two. Figure 7.12 shows a schematic of the improved system used to measure domain wall velocity in amorphous glass-coated nanowires and submicron wires.

f07-12-9780081001646
Figure 7.12 Schematic of the improved system used to measure domain wall velocity in amorphous glass-coated nanowires and submicron wires prepared by means of glass-coated melt spinning.

More recently, a significant development has allowed the experimentally investigation of the shape of the propagating domain wall in rapidly solidified amorphous glass-coated nanowires and submicron wires (Țibu et al., 2012). This investigation is based on simultaneously using MOKE and the Sixtus–Tonks method in the same setup, as illustrated in Figure 7.13.

f07-13-9780081001646
Figure 7.13 Schematic of the experimental setup used to investigate the shape of the propagating domain walls in glass-coated amorphous submicron wires and nanowires.

7.3.2 Domain wall velocity and mobility

Figure 7.14 shows the field dependence of the domain wall velocity for Fe77.5Si7.5B15 amorphous glass-coated submicron wires prepared by rapid solidification, whereas Figure 7.15 shows the same dependence in Co68.15Fe4.35Si12.5B15 samples.

f07-14-9780081001646
Figure 7.14 Field dependence of the domain wall velocity for Fe77.5Si7.5B15 amorphous glass-coated submicron wires prepared by rapid solidification.
f07-15-9780081001646
Figure 7.15 Field dependence of the domain wall velocity for Co68.15Fe4.35Si12.5B15 amorphous glass-coated submicron wires prepared by rapid solidification.

In general, a higher wall velocity corresponds to a thinner wire, which also exhibits higher coercivity because of larger internal stresses. This is also valid for the thinnest Fe77.5Si7.5B15 submicron wire, shown in Figure 7.14, although in this case a significantly different slope is observed. A slight increment in slope with the decrease in the wire diameter is also observed for Co68.15Fe4.35Si12.5B15 submicron wires (Figure 7.15). A decrease in the wire diameter is equivalent to an increase in shape anisotropy. On the other hand, in the nearly zero magnetostrictive samples (Co68.15Fe4.35Si12.5B15), shape anisotropy is expected to be more influential with respect to magnetostrictive samples. Therefore, the slope of the velocity versus field dependence curve seems to be a key element in understanding the effect of each of the two main types of magnetic anisotropy encountered in ultrathin amorphous nanowires and submicron wires prepared by rapid solidification.

Maximum domain wall velocity values are usually larger in nearly zero magnetostrictive submicron wires. For even thinner samples, for example, Co68.15Fe4.35Si12.5B15 and Fe77.5Si7.5B15 amorphous nanowires, the situation is somewhat similar, as illustrated in Figures 7.16 and 7.17.

f07-16-9780081001646
Figure 7.16 Field dependence of the domain wall velocity for a Co68.15Fe4.35Si12.5B15 amorphous glass-coated nanowire.
f07-17-9780081001646
Figure 7.17 Field dependence of the domain wall velocity for a glass-coated amorphous Fe77.5Si7.5B15 nanowire.

It is important to note that compared with the highly magnetostrictive wires, nearly zero magnetostrictive wires require a much smaller axially applied magnetic field to reach high values of domain wall velocity. This is also clearly visible in the case of as-cast and nanocrystallized FINEMET nanowires (Figure 7.18).

f07-18-9780081001646
Figure 7.18 Field dependence of the domain wall velocity for glass-coated FINEMET submicron wires in the as-quenched and nanocrystalline states.

As mentioned in the Introduction section, the maximum wall velocity values in rapidly solidified amorphous glass-coated nanowires and submicron wires are quite large—in most cases larger than those measured in planar nanowires. Moreover, the applied fields required to propagate the walls are generally smaller, and they can be tailored in a wide range to respond to the specifications of the envisaged applications. These propagation fields are related to the depinning of the preexistent domain walls and, therefore, to the preponderant type of magnetic anisotropy within the wires.

Based on the discussions above, a close connection between the domain wall propagation-related parameters and the type of preponderant magnetic anisotropy can be inferred. Therefore, the next step was to analyze the domain wall mobility in these ultrathin amorphous wires and its connection with the magnetic anisotropy.

Figures 7.19 and 7.20 show the domain wall mobility versus the diameter of the metallic nucleus for Fe77.5Si7.5B15 and Co68.15Fe4.35Si12.5B15 glass-coated amorphous nanowires and submicron wires, respectively. In the case of positive magnetostrictive samples, the mobility values are small, but the dependence is highly nonlinear, with a maximum at a nucleus diameter of about 350 nm. In the case of nearly zero magnetostrictive wires, the domain wall mobility values are significantly larger (about two orders of magnitude), and there is a monotonic increase in mobility with the diameter, with a tendency of saturation for diameters larger than 600 nm.

f07-19-9780081001646
Figure 7.19 Domain wall mobility versus metallic nucleus diameter for Fe77.5Si7.5B15 glass-coated amorphous nanowires and submicron wires.
f07-20-9780081001646
Figure 7.20 Domain wall mobility versus metallic nucleus diameter for Co68.15Fe4.35Si12.5B15 glass-coated amorphous nanowires and submicron wires.

Let us discuss first the case of the Co68.15Fe4.35Si12.5B15 nearly zero magnetostrictive nanowires and submicron wires. Because of the very small magnetostriction constant (λ ≅ − 1 × 10− 7), the magnetoelastic anisotropy Kme, which is proportional to the product between λ and the preponderant component of the internal stress tensor σ (Velázquez et al., 1996), is expected to be negligible compared with the shape anisotropy. This assumption is well supported by the small propagation fields when compared with the case of highly magnetostrictive wires and by the much larger values of the domain wall mobility. Therefore, one can assume that the monotonic increase in domain wall mobility with the wire diameter is an indication of a preponderant shape anisotropy, the strength of which decreases monotonically with an increase in the wire diameter. In brief, a positive slope in the mobility (S) versus wire diameter (Φ) curve indicates a preponderant magnetic shape anisotropy.

In the case of Fe77.5Si7.5B15 amorphous glass-coated nanowires and submicron wires with large and positive magnetostriction (λ = 25 × 10− 6), the S versus Φ curve also exhibits a region with a positive slope. This region corresponds to the smallest diameters, that is, nanowires with Φ < 350 nm. This indicates that shape anisotropy is more important at smaller diameters, which is very likely. Nevertheless, the situation is not so straightforward in the sense that large values of the applied field are still required to propagate the domain walls, and therefore the magnetomechanical coupling between internal stresses and magnetostriction cannot be dismissed altogether. It plays an important role in the pinning of the domain walls in highly magnetostrictive amorphous nanowires and submicron wires. Thus, a positive slope of the S(Φ) curve and large values of S show a domain wall propagation controlled mostly by shape anisotropy, with a negligible effect of magnetoelastic anisotropy—as in Co68.15Fe4.35Si12.5B15 wires—whereas a positive slope and small values of S seem to indicate domain wall displacement controlled by shape anisotropy but pinning that is determined by the magnetomechanical coupling, as in the case of Fe77.5Si7.5B15 nanowires with diameters < 350 nm.

For Fe77.5Si7.5B15 samples with diameters larger than 350 nm, the slope of the S versus Φ dependence is negative, and the values of the mobility are small. On the other hand, one would expect in this case an increased role of magnetoelastic anisotropy, enhanced by the reduction of shape anisotropy as the diameter increases and by the large frozen-in stresses. Obviously, domain wall pinning remains controlled by the magnetoelastic term. The important role of magnetoelastic anisotropy in highly magnetostrictive Fe77.5Si7.5B15 amorphous submicron wires is supported by the results of FMR investigations. Thus, Figure 7.21 illustrates the FMR spectra of a glass-coated Fe77.5Si7.5B15 amorphous submicron wire with a nucleus diameter of 500 nm at three frequencies: 8.5, 9.5, and 10.5 GHz. Irrespective of the working frequency, one can observe a small split of the resonance curves, indicating a slightly different anisotropy direction in the near-surface region of the wire. This is consistent with the effect of the magnetoelastic term, since it is the only possible cause of magnetization misalignment with respect to the wire axis. It is clearly not a well-defined domain in the sense of the radial outer shell encountered in microwires with the same composition; it is more likely a magnetization ripple, which nevertheless produces visible effects on the FMR spectra (Chiriac et al., 2011b). To summarize, domain wall mobility is a good indicator of the type of preponderant magnetic anisotropy and of the nature of the pinning of preexistent end domain walls.

f07-21-9780081001646
Figure 7.21 Ferromagnetic resonance spectra for a glass-coated amorphous Fe77.5Si7.5B15 submicron wire with a metallic nucleus diameter of 500 nm at frequencies of 8.5, 9.5 and 10.5 GHz.

7.3.3 Shape of the propagating domain walls

The combined MOKE and Sixtus–Tonks method allowed the shape of the moving domain walls within glass-coated amorphous submicron wires and nanowires prepared by means of rapid solidification to be analyzed. The main idea is that there is a delay Δt between the simultaneously recorded MOKE and Sixtus–Tonks signals generated by a moving magnetic domain wall. It is important to emphasize that the MOKE signal comes from the surface layer of the wall, within the near-surface region of the samples, whereas the Sixtus–Tonks signal is determined by the entire volume of the moving domain wall. Therefore, the synchronization of the two signals under various circumstances is key in interpreting the results.

Figure 7.22 sums up the results obtained through the simultaneous MOKE–Sixtus–Tonks method in Fe77.5Si7.5B15 amorphous submicron wires and nanowires. For these samples, the inductive peaks of the Sixtus–Tonks signal precede the Kerr transition, which means that the edges of the moving domain wall follow the bulk of the wall in its movement. Thus the wall is not planar but most likely has a parabolic or conical shape. The delay Δt is actually the delay between the top of the wall and its edges form the near-surface region of the wires. As the diameter of the sample decreases, so does Δt, until it vanishes completely for the thinnest wires (< 300 nm for Fe77.5Si7.5B15). Hence, the wall becomes planar only for very thin nanowires.

f07-22-9780081001646
Figure 7.22 Interpretation of the domain wall curvature as a function of the delay between the magneto-optical Kerr effect and Sixtus–Tonks signals in amorphous glass-coated Fe77.5Si7.5B15 submicron wires and nanowires.

For Co68.15Fe4.35Si12.5B15 wires, the situation is exactly the opposite: the Kerr transition precedes the Sixtus–Tonks signal. Again, Δt vanishes for very thin nanowires (< 400 nm in this case). In summary, highly magnetostrictive submicron wires and nanowires exhibit domain walls shaped as positive parabolas, whereas nearly zero magnetostrictive ones display walls shaped as negative parabolas. The shape of the propagating walls is linked to the preponderant type of magnetic anisotropy in a way that is more or less similar to the link between anisotropy and domain wall mobility. The delay of the surface for highly magnetostrictive wires is most likely the result of the magnetization misalignment caused by the above-mentioned magnetization ripple. On the other hand, the surface mobility is enhanced in nearly zero magnetostrictive samples, in which no such ripple caused by the magnetoelastic term has been emphasized.

7.4 Final remarks and future work

An overview of the most recent results on the magnetic behavior and domain wall propagation in amorphous magnetic glass-coated nanowires and submicron wires prepared by rapid solidification methods has been presented.

The main feature of their magnetic behavior is the one-step magnetization reversal that takes place through the propagation of a 180° domain wall along the entire wire length. This process is the only mechanism of magnetization reversal emphasized in this novel type of ultrathin magnetic wires, and it occurs irrespective of sample diameter (metallic nucleus diameter within the characteristic range of 90–950 nm) and magnetostriction; magnetization reversal is observed in both highly magnetostrictive and nearly zero magnetostrictive wires.

In terms of hysteresis, this behavior is characterized by a rectangular loop. There is a much larger coercivity in highly magnetostrictive wires compared with nearly zero magnetostrictive ones, which indicates an essential contribution of the magnetoelastic term in the former.

However, in typical amorphous microwires with nucleus diameters larger than 1 μm and up to several tens of micrometers prepared through a similar quenching and drawing technique, rectangular loops appear mainly in highly positive magnetostrictive wires, whereas typical nearly zero and negative magnetostrictive wires exhibit anhysteretic behavior. This difference offers a wide range of tailoring opportunities with respect to the magnetic bistable behavior of nanowires and submicron wires, depending on their size and composition, given that the switching field in their case has a huge range of values that can be used for various applications, from nanosensors and microsensors to magnetic domain wall logic devices.

Thus the decrease in diameter from the range of micrometers to hundreds of nanometers and even further down to tens of nanometers results in significant changes in the distribution of the magnetic anisotropy and the associated domain structure, which both affect magnetization mechanisms. At the nanoscale, irrespective of composition, sample dimensions no longer allow the formation of a complex core–shell magnetic domain structure as a result of magnetoelastic energy minimization. Hence, despite the larger values of the internal stresses in these ultrathin, rapidly solidified materials, the shape anisotropy is preponderant at the nanoscale. The large internal stresses give rise to quite large switching fields.

The study of the magnetization reversal mechanism has led to a thorough investigation of domain wall dynamics, starting with wall velocity and mobility, continuing with the specific aspects of the domain wall pinning, and ending with the correlation between wall mobility and the preponderant type of magnetic anisotropy, which can be either magnetoelastic or shape anisotropy.

To complete these studies, novel measuring techniques have been developed and implemented, for example, a method for investigating the hysteresis loops in nanowires and submicron wires, a method for measuring the domain wall velocity in these ultrathin amorphous wires, and a method for studying the shape of the propagating walls within such novel materials.

Given the extent of the research dedicated to ferromagnetic nanowires prepared either by lithographic methods or by electrodeposition techniques in the past 20 years, we anticipate a fresh impetus in this field driven by the development and study of this new category of rapidly solidified amorphous nanowires and submicron wires.

Some advantages of nanowires and submicron wires prepared by drawing and quenching over lithographically or electrochemically prepared ones are listed below:

 The preparation technique is inexpensive and straightforward in comparison with any type of lithography and electrodeposition.

 They are prepared as single nanowires at practically any length – from tens of centimeters to ten of meters or longer – as opposed to electrodeposited nanowires, which are usually short (the length is determined by the thickness of the membrane in which they are deposited, usually from several micrometers to several tens of micrometers).

 Their magnetic properties can be tailored in a wide range, that is, their coercivity/switching field, magnetization and domain wall velocity and mobility, as well as the magnetic anisotropy, are easily tailored through composition, magnetostriction, and sample dimensions (nucleus diameter and glass coating thickness).

 Additional tailoring possibilities are offered by postproduction annealing of nanowires and submicron wires with the suitable composition, for example, FINEMET-type wires.

There are extensive studies of domain wall propagation in lithographically prepared nanowires, which confer an apparent advantage for these materials over rapidly solidified nanowires. However, taking into consideration the most recent results in the control of domain wall motion in rapidly solidified ferromagnetic wires (Vázquez et al., 2012), which open new ways of trapping and injecting domain walls in these materials, one can expect new developments in this field in the near future.

In spite of the significant progress made in the few years since the first reports of the preparation and properties of these ultrathin wires, there are still issues that need to be addressed and understood before fully exploiting the application potential of these new materials. Future work refers to the further investigation of both basic and application-related properties and phenomena of rapidly solidified amorphous nanowires and submicron wires. Some of the envisaged basic aspects include the interactions among magnetic domain walls in adjacent nanowires and submicron wires, the effects of in situ removal of the glass coating on their magnetic behavior, as well as the theoretical and experimental study of the correlation among magnetic anisotropy distribution, hysteretic behavior and domain wall dynamics. Application-related studies involve the preparation of novel compositions for new applications, for example, electromagnetic shielding, controlled motion of functionalized particles and other medical applications.

References

Allwood DA, Xiong G, Faulkner CC, Atkinson D, Petit D, Cowburn RP. Magnetic domain-wall logic. Science. 2005;309:1688–1692.

Atkinson D, Allwood DA, Faulkner CC, Xiong G, Cooke MD, Cowburn RP. Magnetic domain wall dynamics in a permalloy nanowire. IEEE Trans. Magn. 2003;39:2663–2665.

Butta M, Infante G, Ripka P, Badini-Confalonieri GA, Vázquez M. M-H loop tracer based on digital signal processing for low frequency characterization of extremely thin magnetic wires. Rev. Sci. Instrum. 2009;80:083906.

Chiriac H, Óvári T-A. Amorphous glass-covered magnetic wires: preparation, properties, applications. Prog. Mater. Sci. 1996;40:333–407.

Chiriac H, Óvári T-A, Marinescu CS, Barariu F, Vázquez M, Hernando A. Non-linear effect in the switching mechanism of Fe-based amorphous wires’. In: Kose V, Sievert J, eds. Non-Linear Electromagnetic Systems: Advanced Techniques and Mathematical Methods. Amsterdam: IOS Press; 1998:285–288.

Chiriac H, Marinescu CS, Óvári T-A, Neagu M. Sensor applications of amorphous glass-covered wires. Sensors Actuators A Phys. 1999;76:208–212.

Chiriac H, Colesniuc CN, Óvári T-A, Castano FJ. Ferromagnetic resonance investigation of surface anisotropy distribution in amorphous glass-covered wires. J. Appl. Phys. 2000;87:4816–4818.

Chiriac H, Corodeanu S, Țibu M, Óvári T-A. Size triggered change in the magnetization mechanism of nearly zero magnetostrictive amorphous glass-coated microwires. J. Appl. Phys. 2007;101:09N116.

Chiriac H, Óvári T-A, Țibu M. Domain wall propagation in nearly zero magnetostrictive amorphous microwires. IEEE Trans. Magn. 2008;44:3931–3933.

Chiriac H, Óvári T-A, Tibu M. Effect of surface domain structure on wall mobility in amorphous microwires. J. Appl. Phys. 2009;105:07A310.

Chiriac H, Corodeanu S, Lostun M, Ababei G, Óvári T-A. Magnetic behavior of rapidly quenched submicron amorphous wires. J. Appl. Phys. 2010a;107:09A301.

Chiriac H, Lostun M, Óvári T-A. Surface magnetization processes in amorphous microwires. IEEE Trans. Magn. 2010b;46:383–386.

Chiriac H, Corodeanu S, Lostun M, Stoian G, Ababei G, Óvári T-A. Rapidly solidified amorphous nanowires. J. Appl. Phys. 2011a;109:063902.

Chiriac H, Lostun M, Ababei G, Óvári T-A. Comparative study of the magnetic properties of positive and nearly zero magnetostrictive submicron amorphous wires. J. Appl. Phys. 2011b;109:07B501.

Chiriac H, Lostun M, Óvári T-A. Magnetization process and domain structure in the near-surface region of conventional amorphous wires. J. Appl. Phys. 2011c;109:07B504.

Chizhik A, Gonzalez J, Zhukov A, Blanco JM. Magnetization reversal of Co-rich wires in circular magnetic field. J. Appl. Phys. 2002;91:537–539.

Chizhik A, Garcia C, Gonzalez J, Zhukov A, Blanco JM. Study of surface magnetic properties in Co-rich amorphous microwires. J. Magn. Magn. Mater. 2006;300:e93–e97.

Chizhik A, Varga R, Zhukov A, Gonzalez J, Blanco JM. Kerr-effect based Sixtus-Tonks experiment for measuring the single domain wall dynamics. J. Appl. Phys. 2008;103:07E707.

Chizhik A, Zhukov A, Blanco JM, Gonzalez J. Magneto-optical study of domain wall dynamics and giant Barkhausen jump in magnetic microwires. J. Magn. Magn. Mater. 2012;324:3563–3565.

Chizhik A, Zhukov A, Gonzalez J. Magnetic properties of sub-micrometric Fe-rich wires. Thin Solid Films. 2013;543:130–132.

Corodeanu S, Chiriac H, Lupu N, Óvári T-A. Magnetic characterization of submicron wires and nanowires using digital integration techniques. IEEE Trans. Magn. 2011a;47:3513–3515.

Corodeanu S, Chiriac H, Óvári T-A. Accurate measurement of domain wall velocity in amorphous microwires, submicron wires, and nanowires. Rev. Sci. Instrum. 2011b;82:094701.

Hayward TJ, Bryan MT, Fry PW, Fundi PM, Gibbs MRJ, Allwood DA, Im M-Y, Fischer P. Direct imaging of domain-wall interactions in Ni80Fe20 planar nanowires. Phys. Rev. B. 2010;81:020410.

Herzer G. Amorphous and nanocrystalline soft magnets. In: Hadjipanayis GC, ed. Magnetic Hysteresis in Novel Magnetic Materials. Dordrecht: Kluwer Academic Publishers; 1997:711–730.

Medina JD, Darques M, Piraux L. Exchange bias anisotropy in Co nanowires electrodeposited into polycarbonate membranes. J. Appl. Phys. 2009;106:023921.

Negoita M, Hodges MPP, Bryan MT, Fry PW, Im M-Y, Fischer P, Allwood DA, Hayward TJ. Linear transport of domain walls confined to propagating 1-D potential wells. J. Appl. Phys. 2013;114:163901.

Olivera J, Varga R, Vojtanik P, Prida VM, Sánchez ML, Hernando B, Zhukov A. Fast domain wall dynamics in amorphous glass-coated microwires. J. Magn. Magn. Mater. 2008;320:2534–2537.

Óvári and Chiriac, 2014 Óvári T-A, Chiriac H. Intrinsic domain wall pinning in rapidly solidified amorphous nanowires. J. Appl. Phys. 2014;115:17A329.

Óvári et al., 2009 Óvári T-A, Corodeanu S, Chiriac H. Near-surface magnetic structure and GMI response in amorphous microwires. IEEE Trans. Magn. 2009;45:4282–4285.

Óvári et al., 2011a Óvári T-A, Țibu M, Chiriac H. Controlled manipulation of domain walls in amorphous microwires. IEEE Trans. Magn. 2011a;47:2838–2840.

Óvári et al., 2011b Óvári T-A, Corodeanu S, Chiriac H. Domain wall velocity in submicron amorphous wires. J. Appl. Phys. 2011b;109:07D502.

Parkin SSP, Hayashi M, Thomas L. Magnetic domain-wall racetrack memory. Science. 2008;320:190–194.

Pereira A, Gallardo C, Espejo AP, Briones J, Vivas LG, Vázquez M, et al. Tailoring the magnetic properties of ordered 50-nm-diameter CoNi nanowire arrays. J. Nanopart. Res. 2013;15:2041.

Sixtus KJ, Tonks L. Propagation of large Barkhausen discontinuities. II. Phys. Rev. 1932;42:419–435.

Țibu et al., 2012 Țibu M, Lostun M, Óvári T-A, Chiriac H. Simultaneous magneto-optical Kerr effect and Sixtus-Tonks method for analyzing the shape of propagating domain walls in ultrathin magnetic wires. Rev. Sci. Instrum. 2012;83:064708.

Varga R, Garcia KL, Vázquez M, Vojtanik P. Single-domain wall propagation and damping mechanism during magnetic switching of bistable amorphous microwires. Phys. Rev. Lett. 2005;94:017201.

Varga R, Zhukov A, Zhukova V, Blanco JM, Gonzalez J. Supersonic domain wall in magnetic microwires. Phys. Rev. B. 2007;76:132406.

Vázquez M. Soft magnetic wires. Phys. B. 2001;299:302–313.

Vázquez M, Gomez-Polo C, Chen D-X, Hernando A. Magnetic bistability of amorphous wires and sensor applications. IEEE Trans. Magn. 1994;30:907–912.

Vázquez M, Basheed GA, Infante G, Del Real RP. Trapping and injecting single domain walls in magnetic wire by local fields. Phys. Rev. Lett. 2012;108:037201.

Velázquez J, Vázquez M, Zhukov AP. Magnetoelastic anisotropy distribution in glass-coated microwires. J. Mater. Res. 1996;11:2499–2505.

Zhukov A. Domain wall propagation in a Fe-rich glass-coated amorphous microwire. Appl. Phys. Lett. 2001;78:3106–3108.

Zhukov A. Design of the magnetic properties of Fe-rich, glass-coated microwires for technical applications. Adv. Funct. Mater. 2006;16:675–680.

Zhukov AP, Vázquez M, Velazquez J, Chiriac H, Larin V. The remagnetization process in thin and ultra-thin Fe-rich amorphous wires. J. Magn. Magn. Mater. 1995;151:132–138.

Zhukov A, Zhukova V, Blanco JM, Cobeno AF, Vázquez M. Magnetostriction in glass-coated magnetic microwires. J. Magn. Magn. Mater. 2003;258–259:151–157.

..................Content has been hidden....................

You can't read the all page of ebook, please click here login for view all page.
Reset