Open access peer-reviewed chapter

Waveguides in Magnetism and Spintronics

Written By

Rebeca Díaz Pardo and Jorge Martínez Garfias

Reviewed: 29 April 2024 Published: 16 August 2024

DOI: 10.5772/intechopen.115052

From the Edited Volume

Optical Waveguide Technology and and Applications

Edited by Ki Young Kim

Chapter metrics overview

40 Chapter Downloads

View Full Metrics

Abstract

In this chapter, we propose to review and discuss the use of waveguides in spintronic devices and other magnetic systems as well as some emergent phenomena linked to ultrafast magnetic dynamics. Spintronic devices are expected to replace the recent nanoelectronic memories and sensors due to their efficiency in energy consumption and functionality with scalability. In the field of spintronic devices, it is required the development of magnetic thin films with a wide range of magnetic properties. To achieve the characterization of magnetic thin films, several techniques are very useful, particularly in studying magnetization dynamics described by the phenomenological Landau-Lifshitz-Gilbert equation. These techniques are developed based on key phenomena such as spin pumping, ferromagnetic resonance, the recently observed terahertz oscillations and ultrafast switching. The coupling of waveguides and microwave generators to produce radio-frequency magnetic fields and integrate them with magnetic thin films is crucial to characterize the magnetization dynamics.

Keywords

  • spintronics
  • coplanar waveguides
  • spin current
  • magnetization dynamics
  • magnetism

1. Introduction

This chapter shows the state of the art of the use of waveguides in different experiments in magnetism, spintronics and magnonics and gives a general view of the relevance of waveguides in the development of the areas of spintronics and magnonics. We also discuss perspectives on new technologies related to these elements.

The chapter is divided as follows: we give a brief introduction to magnetization dynamics, the phenomena of resonance and the basic characteristics of spinwaves. Afterward, we describe the type of waveguides that are commonly used in in order to excite spinwaves and generate resonance conditions in magnetic materials. Then, we describe some phenomena in spintronics where oscillating fields give rise to novel and interesting phenomena. Finally, we briefly give a non-exhaustive view of potential applications and concluding remarks.

Advertisement

2. Magnetic dynamics

In this section, we give a brief introduction to the principles of magnetization dynamics. We start by explaining the Landau-Lifshitz-Gilbert (LLG) equation since this expression describes the dynamics of magnetic moments generated by a magnetic field and in general, by a driving force.

We suppose that we have a ferromagnetic layer with magnetic moment per unit volume M and it is under an effective magnetic field Heff. The effective field has several components: the Zeeman energy HZ, exchange interaction Hexch, magnetocrystalline anisotropy energy Hanis, and the demagnetization energy Hdemag, among others, depending on the system [1]. All these terms sum the energy interactions of magnetic materials per unit volume. We find the lowest energy state or the equilibrium state when the magnetic moment points in the same direction as the effective field, Heff. On the other hand, if the magnetization vector M deviates from the equilibrium state, then a torque is exerted on the magnetic moment [2]:

τ=m×μ0HeffE1

where m=M/M is a unitary vector in direction of M. Using the semiclassical approximation, we have that the relation between the magnetic moment m and the quantum angular momentum L is m=γL where γ=gμB/ is the gyromagnetic ratio, with g is a dimensionless quantity that expresses the proportionality between the quantum angular momentum and the magnetic moment and μB is the Bohr magneton. Due to the conservation of angular momentum, the ratio between L and time t is equal to the torque: dLdt=τ. Therefore, from Eq. (1) and the definition of torque, we have [2]:

dmdt=γμ0m×HeffE2

which is the unitary version of the Landau Lifschitz (LL) equation with no damping [2]. The complete version of the dynamic description includes a damping term therefore the complete description is called the Landau-Lifschitz-Gilbert (LLG) equation:

dmdt=γμ0m×Heff+αm×dmdtE3

The second term describes the damping in the magnetization, where the precession decays continuously. This decaying depends on the α term, this is the dissipation term that encompasses all the energy losses [3]. Figure 1 shows the magnetization trajectory for different values of α. Without this extra damping term, m would precess indefinitely. The dissipation causes m to align parallel to Heff its state of minimum energy.

Figure 1.

Sketch of the trajectory followed by m over time under Heff with (a) α=0 and (b) α0.

2.1 Ferromagnetic resonance

In this section, we describe the phenomenology of ferromagnetic resonance, which is a widely used technique to characterize ferromagnetic materials. The magnetic moments of a ferromagnetic material present an intrinsic magnetic moment that precesses at a given frequency (Larmor frequency) and couples with an electromagnetic wave, absorbing its power if the frequency of the wave is the same as the Larmor frequency. This phenomenon is known as ferromagnetic resonance (FMR) [4]. This effect has various applications such as in spectroscopic techniques and in the conception of microwave devices.

Let us consider a ferromagnetic sample under a uniform external magnetic field H0 with an arbitrary shape. We can express the internal field Hint as [3]:

Hint=H0NME4

with N the tensor of the demagnetizing factors. If we choose the Cartesian coordinate system with the x, y, and z directions along the symmetry axes of the sample as shown in Figure 2a, we have that the tensor is diagonal, and that its components satisfy [3]:

Figure 2.

General coordinate system xyz and local coordinate system xyz where the precessing magnetization m is located. (a) Precession around the applied field Hext perpendicular to the sample. (b) The transformation is performed by first rotating φ around z=z and then rotating by (π2θ) around the new y axis [4]. Representation for the case when the external applied field in the sample plane, where we have the component of the magnetization mx [5].

Nx+Ny+Nz=1E5

If we consider an external uniform field Hext applied along the z axis and the magnetization precessing around it as shown in Figure 2a. From Eq. (4), we consider mx and my, the components of the magnetization equation of motion. Solving these equations, we obtain the general Kittel resonance frequency [3]:

ω0=γμ0H0+NxNzM1/2H0+NyNzM1/2E6

The above equation shows that the FMR frequency depends on the shape of the sample and the direction of the applied field. For thin films we take the cases that are represented in Figure 2. For the case of the configuration of Figure 2a, the field Hext is perpendicular to the sample plane, and the magnetization Mext only acts on the z axis, so the component Nz=1. Since it does not act on the x,y plane, we have that Nx=Ny=0. For the configuration in Figure 2b, the magnetization acts on the plane of the thin film; for simplicity, we consider it acts on the direction of the x axis, that is, Nx=1, and since it only acts on this axis, we have that for the axes x and y Ny=Nz=0. The solutions of these two configurations are represented by the following equations, respectively [3]:

fres=γ/2πH0μ0MeffE7
fres=γ/2πH0H0+μ0Meff1/2E8

With these resonance conditions, we can describe the magnetization motion by taking as reference the field with local coordinates already transformed, as in Figure 2b. We assume that mx is the saturated component along the external field and therefore mx=MxMs=1. If this is the case when we change the parametric description of m in the local coordinate system xyz we perform two rotations [6]:

mxmymz=senθcosϕsenϕcosθcosϕsenθsenϕcosϕcosθsenϕcosθ0senθmxmymzE9

We have that in Figure 2b, the dynamics are limited to the yz plane, in this case z=z. If we rewrite the external field as the sum of the static field and the alternating field Hext=H+ht, and replace this in the LLG equation, we can describe the dynamics of the magnetization. Moreover, we can write the magnetization components in the local coordinate system as a function of ht through the susceptibility tensor that expresses the shape of the curve close to the resonance [5]:

mymz=χyyχyzχzyχzzhyhzE10

The shape of absorbed power during the resonance is in general a combination of symmetric and antisymmetric Lorentzian functions. The components of the susceptibility tensor in Eq. (10) depend on the parameters of these Lorentzian functions [5].

The absorbed power in the resonance condition can be expressed as:

pabsHH0ΔH=AFSHcosε+FAHsinεE11

Where A is an amplitude factor, ε is an angle that mixes the symmetric and antisymmetric Lorentzian functions FS and FA, respectively [6]:

FSH=ΔH2HH02+ΔH2E12
FAH=ΔHHH0HH02+ΔH2E13

The line width ΔH is half the width at half maximum in the symmetric Lorentzian curve, and the terms of the type χ denote the susceptibilities at resonance. The term ΔH is the result only of the Gilbert damping parameter αeff which is composed of the interaction of Heff with the resonance frequency given by the equation of Kittel as in the text [5, 6]:

ΔH=αeffωμ0γE14

Thus we have that our resonance condition FMR expresses the magnetization movement where the parameter of αeff corresponds to the speed at which the spin angular momentum dissipates [6].

In order to detect the FMR signal, a microwave resonant cavity with an electromagnet is required. The resonant cavity is fixed at a frequency in the super GHz range, and a detector is placed at the end of the cavity to detect the microwaves. To achieve a full characterization of thin film materials, one needs to carry out FMR measurements at a wide range of microwave frequencies, and for this end, a cavity at a fixed frequency is not convenient. It has been shown that using a broadband microwave signal generator, a coplanar waveguide (CPW), and a broadband microwave detector, it is possible to obtain a good signal-to-noise ratio when a GHz current is injected into the CPW from one end [5, 7].

2.2 Spin waves

So far, we have considered only the case where the magnetization dynamics are homogeneous throughout all the samples. However, it is possible to find varying magnetization through the sample; the elementary excitations are known as spin waves (SW). The quanta of these excitations are known as magnons.

To describe this phenomenon, we consider a linear chain of N classical spins, each with magnitude S coupled with nearest-neighbor exchange interaction, uniformly spaced, and separated by a distance a. We consider only the Heisenberg interaction and the ground state to be where all the spins are parallel, therefore, SiSi+1=S2 and the exchange energy becomes U02JNS2 [3].

In 1930 Felix Bloch showed that the low-lying excitations of the spin system consisted of non-localized, collective spin deviations that he called SW. This was a further conceptualization from the Pierre Weiss model that considered the first excited state to be one spin reversed in the spin chain described before [8]. The existence of low-energy excitations is related to the spin system having an infinite number of degenerate ground states with infinitesimally different spin orientations in the absence of an external field.

In the linear spin chain, the effective field arising from the exchange interaction is:

Hexceff=2JgμBμ0Si1+Si+1E15

Using the first term of the LLG equation and considering that the spins are subject to an applied static magnetic field H, we write the total field acting over the spins as BT=μ0H+Hexceff. Therefore, the spin equation of motion is written as follows [3]:

dSidt=γμ0Si×H+HexceffE16

where γ=gμB/ is the gyromagnetic ratio. Developing these equations for each component and taking into account that the magnetization points in the opposite direction of the spin and therefore taking H=zH, we find that the possible solutions are in the form of traveling waves:

Six=Axekxiωt,Sixy=AyekxiωtE17

with ω the angular frequency and k the wave number. Substituting these solutions in the equations of spin motion for each component, we obtain the dispersion relation, in other words, the dependence of the spin wave frequency on the wave number:

ωk=γμ0H+4JS1coskaE18

From here, we get the relation between the amplitudes of the spin components, and we can write the real parts of the transverse components as:

Six=A0coskxiωkt,Siy=A0sinkxiωkt,E19

From these equations, we can represent the classical picture of a SW in one dimension where the spins precess circularly around the equilibrium direction, as shown in Figure 3. If the wave travels along the +x shortest distance between two spins that precess with the same phase is the wavelength corresponding the wave number λ=2π/k.

Figure 3.

Representation of a spin wave in a 1D chain propagating in the x direction. In the upper part of the figure, we show the top view of the spins. The wavelength corresponds to the distance between two spins pointing in the same direction.

When k=0, the magnon energy is determined only by the magnetic field intensity ω0=γμ0H, however, the general expression for the energy of one magnon in a 1D chain is:

ωk=γμ0H+4JS1coskaE20

As the wave number increases, the phase difference of precession for neighboring spins. Typically, the Zeeman contributions (ω0=γμ0H) of the applied magnetic fields are small in comparison with the exchange energy. For ka=π, the exchange energy is maximum and its value is 8JS, and it is the energy of a linear chain when two consecutive spins are reversed.

At low temperatures, the modes with small wave numbers are dominant, and they are important for many magnonic phenomena. For ka1, we can rewrite Eq. (20) as:

ωkγμ0H+2JSa2k2E21

which is a quadratic dispersion relation with an energy gap. The frequencies of SWs with ka1 are in the microwave range. When k1/a the spin wave frequencies are in terahertz range.

There are three types of SWs based on the propagation direction (k) with respect to the external magnetic field (Hext) and the magnetization (M): Damon-Eshbach (DE) or surface waves when k is perpendicular to the field H , but both are on the sample plane, backward volume waves (BV) when k is parallel to the field H and both are on the sample plane and forward volume waves (FV), when k is perpendicular to the field H but the field is perpendicular to the sample plane [9].

Advertisement

3. Coplanar waveguides in magnetic resonance experiments

Waveguides are important in the study of any high-frequency signal and since we have described the magnetization dynamics in previous sections, we continue to describe examples of applications in magnetism and spintronics. In ferromagnetic materials, the Larmor frequency usually falls in the range of a few tens of GHz. Thus, transmission lines, such as coplanar waveguides, provide a large frequency band in which the impedance is relevant in FMR-related experiments where ferromagnetic films are studied [10].

Coplanar waveguides are composed of a thin metal film layer that is the conductor with a width Wsl with two ground planes parallel to the conductor with a width Wgap separated by a distance Wgl which in this case are composed of copper and have a thickness tTL where they are placed on a dielectric substrate with a thickness tsubstrate as shown in Figure 4 [11].

Figure 4.

Coplanar waveguide. The current j flows inside and outside along the direction y, generating the transverse magnetic field lines ht. The most important parameters of a coplanar waveguide are indicated: the metal thickness tTL, the substrate thickness tsubstrate, Wsl, the width of the signal line, the width of the gap between the signal line and the ground lines Wgap, and the width of the ground line Wgl [10].

One of the main features of coplanar waveguides is the different possible combinations of Wgap and Wsl to design circuits with a specific impedance and modulation of the electromagnetic field generated by the signal line and its two ground lines. This is of great interest to us because we use this property to control the magnetic field generated by the CPW at frequencies from 0GHz to 20GHz, which corresponds to the study of the properties of ferromagnetic resonance (FMR) [12].

3.1 Theoretical model coplanar waveguide in magnetism

We give a brief theoretical description of the physics involved in the operation of CPWs. We start by assuming that electromagnetic waves propagate along the y direction of the CPW, as shown in Figure 4. In this CPW system, transverse electromagnetic modes (TEM) are admitted, and the local electric E and magnetic field H are perpendicular to each other. The direction of propagation and the direction of the magnetic field H are indicated in Figure 4. Applying Maxwell’s equations to a general homogeneous lossy medium [10]:

×E=iωμHE22
×H=iωεE+σEE23

where ω/2π is the frequency of the electromagnetic wave, μ=μ+iμ is the complex magnetic permeability, ε=εiε is the complex electrical permittivity and σ is the conductivity of the guide wave [10]. For the electric field, we have to [10, 13]:

2E+ω2με1iσωσE=0E24

We define the complex propagation constant for the medium as follows [10]:

γ=με1iσωεE25

If we assume that the electric field acts only in the x direction and, at the same time, remains uniform in the xy plane, Eq. (24) reduces to [10]:

2Exz2γ2Ex=0E26

The general solution of this differential equation is given by [10]:

Exz=E+expγz+EexpγzE27

where the term on the right side describes a wave that travels toward the positive side of the z axis and the left side of a wave that travels in the opposite direction in the z direction. The time-dependent propagation factor of the positively moving wave (in this case γz) is written as [10]:

expγzexpiωt=expαzexpiβzexpiωt=expαzcosωtβzE28

For good conductors ω<<σ, this means that the dielectric losses are small compared to the conductivity, and when the conductivity loss predominates, we have that Eq. (25) reduces to [10]:

γ=μεσjωε=1+iωμσ2E29

Where the real part of the above equation corresponds to α and this defines the skin depth δs of the electric field in the material with the relationship [10]:

δs=1α=2ωμσE30

where α=με. As we can see from the above equation, the amplitude of the current density decreases exponentially on the scale of the characteristic length δs within conductive transmission lines [10].

This result is related to the geometric parameters of CPW. If tTl<<δs we approximate the current density to be homogeneously distributed along the z axis in the CPW. In contrast, along the x direction and due to Wsl, Wgl>>δs we cannot assume that the current density is homogeneous. Thus, to determine the local in-plane and out-of-plane amplitudes of the excitation fields, one has to consider the distribution of the current density layers as a function of frequency along the x direction [10, 13].

Thus, we have that the change in the increase in current densities that goes toward the limits of the transmission lines is proportional to the frequency, where they are related through the depth of the skin δ. Therefore, the frequency-dependent magnetic field within the CPW can now be calculated by summing all contributions from the current density paths using the Biot-Savart law. Thus, the field can be calculated at each point of the CPW [10].

3.2 Example of CPW compatible with FMR experiment

As explained in the previous section, a coplanar waveguide (CPW), as shown in Figure 4, consists of ground lines and conducting lines. The ground lines must be at ground potential, and the signal line must have relative voltage amplitude [10, 13]. We notice in Figure 4 that the magnetic field lines are parallel to axis x and if we put a thin ferromagnetic film on top, these field lines would lie on the sample plane [10].

Now, we will discuss an example of a CPW design, starting with the choice of materials for an FMR experiment. The conductive material that was used is copper. This metal is commonly chosen due to its low cost, good electrical conductivity and low dielectric loss. It is also compatible with the transmission of the wave frequencies studied in an adequate way. The substrate we use is Rogers RO4350B, which is made of glass-reinforced woven hydrocarbon/ceramic with a dielectric constant of εr=3.48±0.05 [14]. According to the data we have, this material has a high frequency for applications of up to 18–22 GHz with a loss of ±0.05. Moreover, its dielectric constant does not change abruptly with temperature therefore is a suitable material for temperature variation experiments [14].

To design the waveguide, we use the Advancing the Wireless Revolution Design Environment (AWR) program, which is provided under license by Cadence Design System, Inc. [15].

The process to use this software is to delimit the parameters. For this purpose, we use the AWR transmission line calculator (TX-LINE) to get a first approximation of realistic parameters (considering an impedance of 50Ω and a loss close to 0) [15]. Then, in circuit schematics, using the predefined designs and the dimensions of the CPW, we create an electromagnetic (EM) structure. At this stage, we design the CPW in three-dimensions (3D) and analyze how the electromagnetic fields act on the waveguide and how they affect the loss and power in it. What this procedure is broadly about is comparing the design in Figure 5b with the ideal dimensions that we consider in Figure 5a, finding the best configuration, following this procedure.

Figure 5.

The coplanar waveguide layout of our example designed in AWR is shown [15]. (a) Circuit diagram for the designed coplanar waveguide. (b) Return loss and power vs. frequency. The pink squares represent the loss of the CPW modeled in the first approximation with the TXline parameters, the brown diamonds represent the loss of the AWR-designed CPW, the blue triangles represent the power of the approximate CPW, and the red double triangles represent the power of the AWR-designed CPW.

Furthermore, the CPW is designed in 3D, taking into account the conductor and substrate mentioned above in the Axiem EM section of AWR [15]. This software simulates the electromagnetic fields in the CPW in order to calculate the loss and power of our reference CPW, which is shown in Figure 5b and compared to the loss and power in the CPW with the parameters obtained by TX-LINE. Thus, by the aforementioned procedure, we refine our design until we reach the one we have graphed in Figure 5a, which minimizes the loss and maximizes the power (Figure 5a).

We see how the loss of both the approximate CPW and that of the CPW designed in 3D is very close to zero, and the power of the CPW designed in 3D (EM Structure) is delimited below the CPW that we take as a base (Coplanar), this tells us that we have a suitable design for our purposes. Thus, after this process, we have the dimensions found for this example of CPW using the characteristics of Figure 4 are: Wgl=3mm, Wsl=1mm, Wgap=0.2mm, tTL=0.2mm.

Advertisement

4. Spintronics and magnonics phenomena with waveguides

In this section, we describe experiments and physical phenomena that are intensely studied in the subareas of spintronics and magnetization dynamics. We will focus on potential applications based on the production of spin currents, phenomena related to spin-orbit torques and spin transport via spinwaves.

4.1 FMR, spin pumping and spin currents

The magnetization dynamics in thin magnetic multilayers, specifically in ferromagnetic films in contact with paramagnetic conductors, can produce a moving magnetization vector that causes “pumping” of spins into adjacent nonmagnetic layers [16]. This spin transfer affects the magnetization dynamics, similar to the Landau-Lifshitz-Gilbert phenomenology. This phenomenon is known as spin pumping, and it is the Onsager reciprocal to spin-transfer torque (STT) [17]. In both cases, there is transfer of spin angular momentum between a ferromagnet (F) and a metallic reservoir.

In the first one, a spin (polarized) current injected from the metal into the ferromagnet transfers angular momentum onto the magnetization M of the latter, thereby inducing magnetization dynamics [18]. In contrast, for spin pumping, the nonequilibrium magnetization dynamics (magnons) in the ferromagnet act as the source for an angular momentum flow from the ferromagnet into the metal, inducing a spin current [19].

There are two different detection schemes for spin pumping in ferromagnetic/metallic (F/M) bilayers: the first one detects spin pumping spectroscopically by the increase of the (Gilbert) damping constant of coherent magnetization precession in the ferromagnet, in this case, the metal acts like an angular momentum sink. The second is the electrical detection of the spin current electrically via the inverse spin Hall effect in the metal, which acts as a spin current detector. The excedent of angular momentum can also flow out or be “pumped” across the interface from the ferromagnet into the normal metal, where both longitudinal and transversal angular momentum are pumped. The latter is the main contributor to damping for small precession angles θ. This is schematically shown in Figure 6d. Since spin pumping increases the damping α, one way to detect spin pumping is through the increase in the line width in FMR measurements, as compared in Figure 6a and b.

Figure 6.

(a) The absorption in the resonance condition corresponding to (b). The linewidth ΔH is determined by all the relaxation mechanisms, (c) the magnetization vector of a ferromagnet out of its equilibrium position precessing around the magnetic field with an angle θ and excited via resonance produced by a microwave magnetic field hRF, and (d) ferromagnet/normal metal bilayer. In this structure, there is an additional component to the damping mechanism due to the excess of angular momentum that broads the FMR line width. This can be understood as the generation of pure spin current js with longitudinal spin-polarization; that is, the spin current is pumped from the ferromagnet in its resonance condition into the normal metal.

The spin current density (the spin current per interface area) pumped across the F/N interface by the spin-pumping mechanism is [20, 21, 22]:

jspump=4πRegm×dmdtImgdmdt.E31

where g is the spin mixing conductance that can be understood as the available spin transport channels at the interface, m=M/M is the magnetization unit vector and is the reduced Planck constant.

We can see from Eq. (31) that the current generated as a result of spin pumping across the F/N interface is equivalent to an additional Gilbert-like damping channel, increasing the FMR line width. The additional Gilbert-like damping arising from spin pumping can thus be expressed as an increase of α by αSP [23, 24]:

αSP=gμBReg4πMS1tFηE32

where g Landé g-value, the Bohr magneton μB, the saturation magnetization of the ferromagnet MS, the thickness of the ferromagnet tF, and the backflow parameter 0η1, account for a possible spin current backflow into the ferromagnet.

Another way to detect spin currents arising from spin pumping is via electrical measurements through the inverse spin Hall effect (ISHE) [25, 26]. The charge current density can be written as:

jc=2eΘSHjs×sE33

where js is the spin current in conductors with nonzero spin-orbit coupling and e is the electronic charge, s is the spin current polarization and represents the orientation of the spin flow in the direction of jS and ΘSH is the spin Hall angle, which represents the spin-to-charge current conversion efficiency; for heavy metals, we expect large spin-to-charge conversion.

From Eq. (31), we can see that the spin current jS has AC and DC components. The DC spin current component arises from the longitudinal nonequilibrium magnetization components and the AC spin current components from the transverse components.

From Eq. (31), we can see that the spin current jS has AC and DC components. The DC spin current component arises from the longitudinal nonequilibrium magnetization components and the AC spin current components from the transverse components.

For the electrically detected spin pumping, we will consider only the DC part of js. The DC js is pumped across the surface and, depending on the direction of M and according to Eq. (33), jc lies along the N plane, as shown in Figure 7. The pumped DC spin current js propagates across the F/N interface and is normal to it; this is the physical origin of the ISHE. From Eq. (33), we see that jc lies in the plane of the N layer and perpendicular to s. The pumped spin current into N decays at the same scale as the spin diffusion length λsp, usually between 1 and 100 nm.

Figure 7.

Schematic of the DC electrically detected spin-pumping experiments, the DC spin current js pumped into the normal metal is converted into a charge current jc via the inverse spin Hall effect and then detected via the corresponding open-circuit voltage VDC.

It is important to notice that js will be pumped into N within the resonance condition. Under these conditions, it is possible to detect jc=jc or Ec=VDC/L=jc/σtot where L is the distance between the electrical contacts and σtot is the total conductivity of the F/N bilayer. The spin-pumping spin Hall voltage VDC is given by [27, 28]:

VDCL=2eΘSHjsηλsdtanhtN2λsdσFtF+σNtNE34

The DC electric field can also be generated by microwave rectification phenomena apart from ISHE. In this context, the resonant magnetization precession in FMR generates a modulation of the sample’s resistivity tensor at the microwave frequency. The resistivity modulations rectify microwave-frequency charge currents induced in the sample by the microwave signal. Since different components of the resistivity tensor (anisotropic magnetoresistance, anomalous Hall effect, etc.) can contribute to the rectification voltages, based on symmetry or line shape can be ambiguous the presence of an additional, magnetically ordered layer may provide an additional magnetization damping channel [29]. Furthermore, there are other effects that should be taken into account in microwave rectification or magneto-caloric effects such as the anomalous Nernst effect, which may then play a role in the proximity of polarized metal.

Note that the ferromagnetic layer in this type of experiment does not need to be metallic in order to produce a spin current in the adjacent layer. There has been experimental evidence both for and against static magnetic polarization in the Pt in F/Pt hybrids made from an electrically insulating ferromagnet such as yttrium iron garnet (YIG) [30, 31, 32].

Spin pumping, as well as other interesting phenomena well known in the spintronics areas such as spin Hall effect (SHE) [25, 33], Rashba effect [34, 35], and Dyaloshinskii-Moriya Interaction (DMI) are all known to have origin in spin-orbit coupling (SOC) [36].

Spin-orbit torques (SOT) originating from pure spin current are expected to be crucial to control the magnetization as well as SW much more efficiently than for instance conventional spin-transfer torque, since SOT spin-polarization arises from the carrier velocity difference, and there is no need for a magnetized layer that plays as a spin filter. SOT is capable of compensating Gilbert damping over extended regions, which has opened up a new avenue for on-chip SW communication devices. These applications will be discussed further in Section 5.

We have discussed spin-pumping experiments in which the magnetization was driven out of equilibrium via conventional FMR. In other words, the spin current across the F/N interface was excited by microwave photons. However, magnetization dynamics can be caused by other mechanisms: thermal gradients, acoustic strain fields, charge currents, and high-frequency elastic deformations.

4.2 Microwave techniques to study spin waves

In magnonic device technology, the main mechanisms under study are the ones implicated in three major blocks: generation, detection, and manipulation of the spin waves, which correspond to input, gate, and output. The magnetic waveguide in a magnonic device guides the spin wave signal of a given frequency , wave vector, phase, and amplitude, which are the key ingredients for wave-based computing [37].

An important breakthrough was the understanding of the charge current to spin current conversion and vice versa. The charge-to-spin current has been achieved via SHE, where a charge current is injected into a heavy metal with a large SOC, leading to an orthogonal spin current, as explained in the previous section. The inverse mechanism ISHE is useful to detect spin waves (Figure 7) [33, 38].

In order to propagate a magnon current, we need to be able to regulate the input, output, and gate. To do so, it is necessary to manipulate and control regimes of frequency, wavevector, and phase of the spin wave. We also need to be able to control the decay length (λ), which, as explained before, is the length over which the spin wave intensity drops by a factor of e. For this, we need materials with low damping as YIG, as we already have pointed out; other examples include Permalloy (Ni80Fe20) [39], CoFeB [40], and Heusler Alloys [41].

A very common technique to study long spin waves in ferromagnets is the use of microwave radiation. For a monochromatic driving field with frequency ω, the absorbed microwave power by a spin wave when ωk=ω is given by [3]:

P=12μ0ωℜhmdrE35

where h and m are the driving field and the small signal magnetization with frequency ω, they are both complex quantities. If we consider the driving field along the x direction, we can write the power absorbed by a spin wave with magnetization mxkexpikr due to a driving field with a spatial variation hxexpiqr as:

P=12μ0ωVℜhxqmxkδq,k=12μ0ωVχxxhxk2E36

where δq,k is the Kronecker delta, equal to 1 for q=k and zero for other values, χxx is the susceptibility tensor. This expression shows that a spin wave with a certain wave vector k can only be excited by a driving magnetic field with a Fourier component at the same wave vector.

In bulk materials with uniform internal magnetizing fields, it is only possible to excite the k0 magnon, that is, the case of ferromagnetic resonance.

In order to excite higher-order magnons, we need microwave radiation, fine wire antennas, and thin magnetic films. For instance, a metallic multielement antenna is deposited by lithography on the surface of the magnetic film [42].

Another possibility is to use the boundary conditions at the film surface in order to generate a nonuniform internal driving field. In this scheme, the driven spin waves are comparable to the film thickness [43]. In films, the first observations of spin wave resonances were measured in a Permalloy film, with the static field perpendicular to the film surface. In this experiment, plane spin waves with wave vectors perpendicular to the film plane generated standing waves between the two film surfaces [43].

The wavelength is given by d=nλ/2 where d is the film thickness and n is an integer. As the field increases, the wave number varies, and the spin wave resonances produce absorption peaks corresponding to a different wavelength, with index number n.

The first studies of spin waves were done in YIG bulk single crystals in the shape of disks, rods, or slabs. When the sample has these shapes, the applied uniform field produces a nonuniform internal field due to demagnetizing effects on the surfaces. For static fields, the wave frequency remains unchanged; however, the wave number varies since there is no momentum conservation. The experiment described in ref. [44], uses a YIG displaced in a cylindrical microwave cavity with a 9.4 GHz microwave input. In the experiment performed by Eshbach et al., the oscilloscope signal showed magneto-elastic wave echoes at various magnetic field values. This indicated that due to the nonuniform dc field, the microwave pulse drives the spin wave in the region where it has a wave number k0.

The wave packet with circular wavefront propagates toward the center of the disk with constant frequency. As the field decreases, the wave number increases and approaches a crossover region. In a uniform field, the two branches of magnetoelastic dispersion would correspond to normal modes. Since the field varies in space, the excitation can hop into another branch [45]. The hopping probability depends on the combination of magnetic field values at a given position and if, in that position, the magnon and phonon wave numbers are equal. On the other hand, if the field gradient is much smaller, the probability for branch hopping decreases, and in this case, the spin wave packet is converted very efficiently into an elastic wave, just like in the experiment performed in Ref. [44].

Magnon–phonon conversion in a time-varying field can be experimentally demonstrated in a YIG rod using a setup consisting of the microwave source, that produces rf bursts with frequency in the range of 1–2 GHz and duration 0.2μs, and the receiving system, which consists of a broadband microwave amplifier, a diode detector or a narrowband superheterodyne system, composed of a local oscillator, mixer, amplifier and diode detector.

In this type of experiment, one can tune the amplifier and measure the pulse frequency by scanning the oscillator frequency in the narrowband. Then, an oscilloscope displays the pulses and the waveform of the current pulses produced by the time-varying field [46].

We have discussed the first experiments in order to measure spin wave frequency shift and magnon–phonon conversion under a time-varying magnetic field; however, there are other measuring techniques that have recently become very important in the study of spin waves, such as inelastic light scattering techniques that nowadays are crucial to study elementary excitations in solids, among which one of the most used ones is Brillouin light scattering (Figure 8) [3].

Figure 8.

Diagram of the setup used to observe spin wave frequency shift and magnon-phonon conversion in a magnetic field varying with time.

Advertisement

5. Potential applications

Today’s computing systems rely on information read and written by electric charge or voltage, and therefore, computation is performed by charge movements. The fundamental circuit element is the transistor, which plays the role of both a switch and an amplifier. Nowadays, large-scale integrated circuits are based on complementary metal–oxide–semiconductor (CMOS) field-effect transistors due to their high density, low power consumption, and low fabrication cost [47].

Logic gates can be built with CMOS transistors, such that they represent a full set of Boolean algebraic operations. The device density and the performance of the CMOS have been steadily improved in the last decades according to Moore’s law [48]. CMOS technology has driven and adapted to an exploding information technology market and therefore CMOS has consolidated as a preferred mechanism in the digital domain.

Spintronics, which uses the spin degree of freedom for information coding, is particularly promising as an alternative CMOS paradigm in computing technology due to the low intrinsic energies of magnetic excitations as well as their collective nature [49, 50, 51].

Several implementations of spintronic Boolean logic devices have been investigated such as magnetic semiconductors [52], spin currents [53], nanomagnets [54], domain walls [55, 56], and skyrmions [57]. Beyond logic gates, interconnects are the key elements of any circuit. Usually more energy is typically dissipated in interconnects by moving data around. The fraction of the power dissipated in interconnects may be even larger for spin-based logic if communication happens via spin currents, domain wall propagation, or by converting back-and-forth spin signals to electrical signals. In spin waves, there is no charge transported; therefore, using them as information carriers is a potential solution for energy dissipation [58].

For this reason, of the most novel spintronic logic device concepts have been based on spin waves as information carriers, and several spin-wave-based logic devices have been proposed [47] and experimentally demonstrated, for instance, via interferometer-based logic gates [59] and spin-wave majority gates [60]. However, in order to develop practical applications of such logic devices, a combination of circuits and computing systems containing both logic and memory is necessary.

The current challenge in magnonic circuits is to keep losses under control. If this is achieved, communication may not require extra energy costs. Magnonic circuits can scale to nanoscale dimensions, unlike optical waveguides. For now, some realistic proposals point toward a hybrid systems concept with local spin-wave islands embedded in a CMOS periphery.

Another advantage of spin waves is their relatively large coherence length lpl, even at room temperature (up to 1 cm in YIG of micrometer thicknesses) as well as their broad frequency range. These characteristics make spin wave devices promising candidates for applications related to data transfer and processing for rf applications. Additionally, the recent miniaturization of spin-wave wavelength λ below 100 nm that, for now, is limited to the lattice constant of the material gives further perspective to enable access to magnonics in the THz range [61].

Current technology covers the frequency range of up to 100 GHz, which is of special interest for the upcoming 5G communication systems. As a result, typical components for wireless communication (antennas, amplifiers, filters, mixers, modulators, and frequency synthesizers) will have to adapt to new frequency bands and communication protocols.

The development of ultrahigh-speed beyond–CMOS spin-wave-based technology for signal processing requires, as said before, long-range enhancement of spin waves, that is, enhanced spin wave coherence and propagation length. These can be addressed by spin-orbit torques (SOT) since, through these torques, it is possible to electrically control damping, resulting in significantly enhanced spin-wave propagation.

The theoretical concepts of SOC and spinwaves were both formulated separately. However, the interplay between these two topics has drawn attention recently due to its various advantages [61, 62]. There are several phenomena that originate from SOT, for instance, perpendicular magnetic anisotropy (PMA), SHE, Rashba effect, spin pumping and Dyaloshinskii-Moriya Interaction (DMI) [63, 64]. At the same time, it has recently and experimentally realized that SOT originates from pure spin current, as it is explained in Section 4.1.

Observation of chiral spin textures is caused, among other factors, by the presence of DMI in thin film heterostructures where a ferromagnetic layer is placed adjacent to a heavy metal layer or 2D material. The influence of DMI or antisymmetrical interaction in the spin wave dispersion has become a topic of interest. It has been theoretically predicted that spin wave asymmetry can be induced by DMI, and this can be measured by the frequency difference of counter-propagating spin waves in the Brillouin light scattering technique [62, 65].

SOTs provide a versatile tool to electrically manipulate the magnetization of all classes of magnetic materials, that is, metals, semiconductors, or insulators, as well as different types of magnetic order, including ferrimagnetic and antiferromagnetic structures. This versatility has led to a variety of experimental and conceptual results, which have greatly expanded the scope of spintronics in the last decade [50].

The demonstration of the current-induced switching of a single-layer ferromagnet, by injecting a current density 108A/cm2 into a Pt layer a few nm thick, adjacent to a Co dot with perpendicular magnetization [36]. Further experiments showed switching of three terminal magnetic tunnel junctions (MTJs) based on Ta/CoFeB/MgO or W/CoFeB/MgO stacks with either perpendicular or in-plane magnetization [66, 67], using ns current pulses with extremely low error rates and without external fields. In MTJs, SOTs allow for the switching of the free layer without passing a current through the tunnel barrier. The separation of the write and read current paths further avoids write errors during readout and voltage breakdowns. For these reasons, SOTs are attracting increasing attention as a replacement for or addition to STT in magnetic random access memories (MRAMs).

The goal of the scientific challenges is to combine a number of spin-orbit effects with the magnonic crystals and therefore to the magnonic devices. For example, the simultaneous existence of PMA and DMI as well as patterning of 2D magnonic crystals may stabilize topological magnetic textures like skyrmions and might improve the control of their dynamics.

A new path for novel physical phenomena may be paved by the combined effect of current flow and magnon on skyrmions due to the competition between thermally driven magnon current, transverse magnon current, and electron flow. One of the main challenges of magnonics for applications is the large damping of spinwaves in metallic ferromagnets, which limits its propagation distance and efficiency. The properties of SOT may indeed dramatically reduce damping.

Advertisement

6. Conclusions

Waveguides in spintronic phenomena are crucial due to the nature of magnetization dynamics, which makes magnetic moments sensitive to time-varying fields, producing a wide variety of spin phenomena from spin waves to resonance phenomena which, with proper adjacent layers, produce pure spin currents due to SHE or Rashba effect. The optimization of waveguides so that the magnetic moments reach the resonant condition is a key ingredient in generating spin currents and observing phenomena like spin pumping, SHE, and ISHE, which arise from SOC and produce SOTs.

SOTs provide a variety of tools to electrically manipulate the magnetization of all classes of magnetic materials: metals, semiconductors, or insulators, as well as different types of magnetic order (ferrimagnetic and antiferromagnetic structures). This has produced an enormous amount of experimental results, which have greatly expanded the scope of spintronics in the last decade.

Spin waves are considered to be promising as information carriers in low-power-consuming computing and logic. Since, in the case of spin waves, there is no charge transport but purely spin transport, and it presents interesting advantages such as the absence of Joule heating and clock frequencies up to the THz frequency regime.

On the other hand, the wide variety of physical phenomena and properties of materials linked to SOTs open the door for innovative technology that can leverage the topological properties of certain materials in spintronic devices, spin logic, efficient reading and writing mechanisms, etc.

One of the current challenges in the field of spintronics-magnonics lies in implementing spin-orbit effects to modify the magnon properties in tailored magnetic materials. This achievement is very promising for energy-efficient high-frequency nanoelectronic devices.

Advertisement

Conflict of interest

The authors declare no conflict of interest.

Advertisement

Abbreviations

LLG

Landau-Lifshitz-Gilbert

FMR

ferromagnetic resonance

CPW

coplanar waveguide

DE

Damon-Eshbach

TEM

transverse electromagnetic

EM

electromagnetic

STT

spin-transfer torque

ISHE

inverse spin Hall effect

YIG

ytrium iron garnet

SHE

spin Hall effect

DMI

Dyaloshinskii-Moriya interaction

SOC

spin-orbit coupling

SOT

spin-orbit torque

SW

spin wave

CMOS

complementary metal-oxide-semiconductor

PMA

perpendicular magnetic anisotropy

MTJ

magnetic tunnel junction

MRAM

magnetic random access memory

References

  1. 1. Coey J. Magnetism and Magnetic Materials. The Edinburgh Building, Cambridge CB2, UK: Cambridge University Press; 2009. DOI: 10.1017/CBO9780511845000
  2. 2. Xu J, Jin L, Liao Z, Wang Q, Tang X, Zhong Z, et al. Quantum spin-wave materials, interface effects and functional devices for information applications. Frontiers in Materials. 2020;7:2-6. Available from: https://www.frontiersin.org/articles/10.3389/fmats.2020.594386
  3. 3. Rezende SM. Fundamentals of Magnonics. Cham: Springer; 2020. DOI: 10.1088/1361-6463/aac7b5
  4. 4. Moura A. Application of magnon coherent states in spintronics. Journal of Magnetism and Magnetic Materials. 2020;498:166091. Available from: https://www.sciencedirect.com/science/article/pii/S0304885319329920
  5. 5. Decker M. Spin current induced control of magnetization dynamics [Ph.D. dissertation]. Universität Regensburg; 2017. DOI: 10.5283/epub.33381. Available from: https://epub.uni-regensburg.de/37166/1/PHD_Decker_Martin_2017.pdf
  6. 6. Pietanesi L. Spin-to-charge conversion in Bi2Se3/NiFe heterostructures [master thesis]. Politecnico di Milano; 2020. Available from: https://www.politesi.polimi.it/handle/10589/151301
  7. 7. Montoya E, McKinnon T, Zamani A, Girt E, Heinrich B. Broadband ferromagnetic resonance system and methods for ultrathin magnetic films. Journal of Magnetism and Magnetic Materials. 2014;356:12-20. Available from: https://www.sciencedirect.com/science/article/pii/S0304885313009256
  8. 8. Bloch F. Zur theorie des ferromagnetismus. Zeitschrift für Physik. 1930;61:206-219
  9. 9. Prabhakar A, Stancil DD. Spin Waves: Theory and Applications. Springer; 2010
  10. 10. Obstbaum M. Inverse spin Hall effect in metallic heterostructures [Ph.D dissertation]. Regensburg Universität; 2015. Available from: https://epub.uni-regensburg.de/33381/1/Doktorarbeit%20Martin%20Obstbaum.pdf
  11. 11. Simons RN. Coplanar Wavegude Circuits, Components, and Systems. John Wiley and Sons, Inc.; 2001. ISBN: 978-0-471-46393-1
  12. 12. Pizarro DCP. Estudio de la respuesta en frecuencia de líneas de guía de onda coplanar en función de diferentes tipos de estructuras periódicas ranuradas en el plano de tierra. Centro de Investigación CIentífica y de Educación Superior De Ensenada (CICESE); 2008. Available from: https://cicese.repositorioinstitucional.mx/jspui/bitstream/1007/184/1/179031.pdf
  13. 13. Pozar DM. Microwave Engineering. John Wiley and Sons Inc.; 2012. ISBN: 978-0-470-63155-3
  14. 14. RO4359B Laminates. Rogers Corporation; 2024. Available from: https://www.rogerscorp.com/advanced-electronics-solutions/ro4000-series-laminates/ro4350b-laminates
  15. 15. AWR Design Environment. Cadence Design Systems, Inc., version 17, 2024. Available from: https://www.cadence.com/en_US/home/tools/system-analysis/rf-microwave-design.html
  16. 16. Althammer MS, Weiler M. Spin Pumping. John Wiley and Sons, Ltd; 2015;6:111-124. DOI: 10.5772/intechopen.115052. Available from: https://onlinelibrary.wiley.com/doi/abs/10.1002/9781118751886.ch6
  17. 17. Ralph D, Stiles M. Spin transfer torques. Journal of Magnetism and Magnetic Materials. 2008;320(7):1190-1216. Available from: https://www.sciencedirect.com/science/article/pii/S0304885307010116
  18. 18. Liu L, Moriyama T, Ralph DC, Buhrman RA. Spin-torque ferromagnetic resonance induced by the spin hall effect. Physical Review Letters. 2011;106:036601. DOI: 10.1103/PhysRevLett.106.036601
  19. 19. Wang Y, Ramaswamy R, Yang H. FMR-related phenomena in spintronic devices. Journal of Physics D: Applied Physics. 2018;51:273. DOI: 10.1088/1361-6463/aac7b5
  20. 20. Tserkovnyak Y, Brataas A, Bauer GEW. Enhanced gilbert damping in thin ferromagnetic films. Physical Review Letters. 2002;88:117601. DOI: 10.1103/PhysRevLett.88.117601
  21. 21. Tserkovnyak Y, Brataas A, Bauer WB, Gerrit E. Spin pumping and magnetization dynamics in metallic multilayers. Physical Review B. 2002;66:224403. DOI: 10.1103/PhysRevB.66.224403
  22. 22. Tserkovnyak Y, Brataas A, Bauer GEW, Halperin BI. Nonlocal magnetization dynamics in ferromagnetic heterostructures. Reviews of Modern Physics. 2005;77:1375-1421. DOI: 10.1103/RevModPhys.77.1375
  23. 23. Jiao H, Bauer GEW. Spin backflow and ac voltage generation by spin pumping and the inverse spin hall effect. Physical Review Letters. 2013;110:217602. DOI: 10.1103/PhysRevLett.110.217602
  24. 24. Chen YT, Takahashi S, Nakayama H, Althammer M, Goennenwein STB, Saitoh E, et al. Theory of spin hall magnetoresistance. Physical Review B. 2013;87:144411. DOI: 10.1103/PhysRevB.87.144411
  25. 25. Dyakonov M, Perel V. Current-induced spin orientation of electrons in semiconductors. Physics Letters A. 1971;35(6):459-460. Available from: https://www.sciencedirect.com/science/article/pii/0375960171901964
  26. 26. Valenzuela S, Tinkham M. Direct electronic measurement of the spin hall effect. Nature. 2006;442:176-179. Available from: https://www.nature.com/articles/nature04937
  27. 27. Mosendz O, Pearson JE, Fradin FY, Bauer GEW, Bader SD, Hoffmann A. Quantifying spin hall angles from spin pumping: Experiments and theory. Physical Review Letters. 2010;104:046601. DOI: 10.1103/PhysRevLett.104.046601
  28. 28. Mosendz O, Vlaminck V, Pearson JE, Fradin FY, Bauer GEW, Bader SD, et al. Detection and quantification of inverse spin hall effect from spin pumping in permalloy/normal metal bilayers. Physical Review B. 2010;82:214403. DOI: 10.1103/PhysRevB.82.214403
  29. 29. Sun KY, Chang H, Kabatek M, Song YY, Wang Z, Jantz M, et al. Damping in yttrium iron garnet nanoscale films capped by platinum. Physical Review Letters. 2013;111:106601. DOI: 10.1103/PhysRevLett.111.106601
  30. 30. Geprägs S, Meyer S, Altmannshofer S, Opel M, Wilhelm F, Rogalev A, et al. Investigation of induced Pt magnetic polarization in Pt/Y3Fe5O12 bilayers. Applied Physics Letters. 2012;101(26):262407. DOI: 10.1063/1.4773509
  31. 31. Huang SY, Fan X, Qu D, Chen YP, Wang WG, Wu J, et al. Transport magnetic proximity effects in platinum. Physical Review Letters. 2012;109:107204. DOI: 10.1103/PhysRevLett.109.107204
  32. 32. Lu YM, Choi Y, Ortega CM, Cheng XM, Cai JW, Huang SY, et al. Pt magnetic polarization on Y3Fe5O12 and magnetotransport characteristics. Physical Review Letters. 2013;110:147207. DOI: 10.1103/PhysRevLett.110.147207
  33. 33. Hirsch JE. Spin Hall effect. Physical Review Letters. 1999;83:1834-1837. DOI: 10.1103/PhysRevLett.83.1834
  34. 34. Edelstein V. Spin polarization of conduction electrons induced by electric current in two-dimensional asymmetric electron systems. Solid State Communications. 1990;73(3):233-235. Available from: https://www.sciencedirect.com/science/article/pii/003810989090963C
  35. 35. Bel’kov VV, Ganichev SD. Magneto-gyrotropic effects in semiconductor quantum wells. Semiconductor Science and Technology. 2008;23(11):114003. DOI: 10.1088/0268-1242/23/11/114003
  36. 36. Miron I, Garello K, et al. Perpendicular switching of a single ferromagnetic layer induced by in-plane current injection. Nature. 2011;476:189-193
  37. 37. Haldar A, Adeyeye AO. Functional magnetic waveguides for magnonics. Applied Physics Letters. 2021;119(6):060501. DOI: 10.1063/5.0061528
  38. 38. Sinova J, Culcer D, Niu Q, Sinitsyn NA, Jungwirth T, MacDonald AH. Universal intrinsic spin hall effect. Physical Review Letters. 2004;92:126603. DOI: 10.1103/PhysRevLett.92.126603
  39. 39. Demidov VE, Demokritov SO, Rott K, Krzysteczko P, Reiss G. Linear and nonlinear spin-wave dynamics in macro- and microscopic magnetic confined structures. Journal of Physics D: Applied Physics. 2008;41(16):164012. DOI: 10.1088/0022-3727/41/16/164012
  40. 40. Yu H, Huber R, Schwarze T, Brandl F, Rapp T, Berberich P, et al. High propagating velocity of spin waves and temperature dependent damping in a CoFeB thin film. Applied Physics Letters. 2012;100(26):262412. DOI: 10.1063/1.4731273
  41. 41. Sebastian T, Ohdaira Y, Kubota T, Pirro P, Brächer T, Vogt K, et al. Low damping spin-wave propagation in a micro-structured Co2Mn0.6Fe0.4Si heusler waveguide. Applied Physics Letters. 2012;100:112402. Available from: https://api.semanticscholar.org/CorpusID:121064753
  42. 42. Lim J, Bang W, Trossman J, Kreisel A, Jungfleisch MB, Hoffmann A, et al. Direct detection of multiple backward volume modes in yttrium iron garnet at micron scale wavelengths. Physical Review B. 2019;99:014435. DOI: 10.1103/PhysRevB.99.014435
  43. 43. Seavey MH, Tannenwald PE. Direct observation of spin-wave resonance. Physical Review Letters. 1958;1:168-169. DOI: 10.1103/PhysRevLett.1.168
  44. 44. Eshbach JR. Spin-wave propagation and the magnetoelastic interaction in yttrium iron garnet. Physical Review Letters. 1962;8:357-359. DOI: 10.1103/PhysRevLett.8.357
  45. 45. Rezende SM, Morgenthaler FR. Magnetoelastic waves in time-varying magnetic fields. I. Theory. Journal of Applied Physics. 1969;40(2):524-536
  46. 46. Rezende SM. Magnetoelastic and magnetostatic waves in time-varying magnetic fields [Ph.D. dissertation]. Massachusetts Institute of Technology; 1967. Available from: https://apps.dtic.mil/sti/citations/AD0664333
  47. 47. Mahmoud A, Ciubotaru F, Vanderveken F, Chumak AV, Hamdioui S, Adelmann C, et al. Introduction to spin wave computing. Journal of Applied Physics. 2020;128(16):161101. DOI: 10.1063/5.0019328
  48. 48. Moore G. Cramming more components onto integrated circuits. Proceedings of the IEEE. 1998;86(1):82-85
  49. 49. Wolf SA, Awschalom DD, Buhrman RA, Daughton JM, von Molnár S, Roukes ML, et al. Spintronics: A spin-based electronics vision for the future. Science. 2001;294(5546):1488-1495. Available from: https://www.science.org/doi/abs/10.1126/science.1065389
  50. 50. Sander D, Valenzuela SO, Makarov D, Marrows CH, Fullerton EE, Fischer P, et al. The 2017 magnetism roadmap. Journal of Physics D: Applied Physics. 2017;50(36):363001. DOI: 10.1088/1361-6463/aa81a1
  51. 51. Bernstein K, Cavin RK, Porod W, Seabaugh A, Welser J. Device and architecture outlook for beyond cmos switches. Proceedings of the IEEE. 2010;98(12):2169-2184
  52. 52. Žutić I, Fabian J, Das Sarma S. Spintronics: Fundamentals and applications. Reviews of Modern Physics. 2004;76:323-410. Available from: https://link.aps.org/doi/10.1103/RevModPhys.76.323
  53. 53. Hall KC, Flatté ME. Performance of a spin-based insulated gate field effect transistor. Applied Physics Letters. 2006;88(16):162503. DOI: 10.1063/1.2192152
  54. 54. Imre A, Csaba G, Ji L, Orlov A, Bernstein GH, Porod W. Majority logic gate for magnetic quantum-dot cellular automata. Science. 2006;311(5758):205-208. Available from: https://www.science.org/doi/abs/10.1126/science.1120506
  55. 55. Allwood DA, Xiong G, Faulkner CC, Atkinson D, Petit D, Cowburn RP. Magnetic domain-wall logic. Science. 2005;309(5741):1688-1692. Available from: https://www.science.org/doi/abs/10.1126/science.1108813
  56. 56. Nikonov DE, Bourianoff GI, Ghani T. Proposal of a spin torque majority gate logic. IEEE Electron Device Letters. 2011;32(8):1128-1130
  57. 57. Koumpouras K, Yudin D, Adelmann C, Bergman A, Eriksson O, Pereiro M. A majority gate with chiral magnetic solitons. Journal of Physics: Condensed Matter. 2018;30(37):375801. DOI: 10.1088/1361-648X/aad82f
  58. 58. Barman A, Gubbiotti G, Ladak S, et al. The 2021 magnonics roadmap. Journal of Physics: Condensed Matter. 2021;33(41):413001. DOI: 10.1088/1361-648X/abec1a
  59. 59. Kostylev MP, Serga AA, Schneider T, Leven B, Hillebrands B. Spin-wave logical gates. Applied Physics Letters. 2005;87(15):153501. DOI: 10.1063/1.2089147
  60. 60. Khitun A, Wang KL. Non-volatile magnonic logic circuits engineering. Journal of Applied Physics. 2011;110(3):034306. DOI: 10.1063/1.3609062
  61. 61. Chumak AV. Spintronics Handbook, Second Edition: Spin Transport and Magnetism. CRC Press; 2019. ch. Magnonic Spintronics. DOI: 10.1201/9780429441189
  62. 62. Hellman F, Hoffmann A, Tserkovnyak Y, Beach G, et al. Interface-induced phenomena in magnetism. Reviews of Modern Physics. 2017;89:025006. Available from: https://link.aps.org/doi/10.1103/RevModPhys.89.025006
  63. 63. Ganguly A, Rowan-Robinson RM, Haldar A, Jaiswal S, Sinha J, Hindmarch AT, et al. Time-domain detection of current controlled magnetization damping in Pt/Ni81Fe19 bilayer and determination of Pt spin Hall angle. Applied Physics Letters. 2014;105(11):112409. DOI: 10.1063/1.4896277
  64. 64. Panda SN, Mondal S, Sinha J, Choudhury S, Barman A. All-optical detection of interfacial spin transparency from spin pumping in Ta/CoBeB thin films. Science Advances. 2019;5(4):eaav7200 https://www.science.org/doi/abs/10.1126/sciadv.aav7200
  65. 65. Chaurasiya AK, Banerjee C, Pan SSS, Choudhury S, Sinha ABJ. Direct observation of interfacial Dzyaloshinskii-Moriya interaction from asymmetric spin-wave propagation in W/CoFeB/SiO2 heterostructures down to sub-nanometer CoFeB thickness. Scientific Reports. 2016;6(32592):1-8
  66. 66. Cubukcu M, Boulle O, Mikuszeit N, Hamelin C, Brcher T, Lamard N, et al. Ultra-fast perpendicular spin–orbit torque MRAM. IEEE Transactions on Magnetics. 2018;54(4):1-4
  67. 67. Aradhya SV, Rowlands GE, Oh J, Ralph DC, Buhrman RA. Nanosecond-timescale low energy switching of in-plane magnetic tunnel junctions through dynamic oersted-field-assisted spin hall effect. Nano Letters. 2016;16(10):5987-5992

Written By

Rebeca Díaz Pardo and Jorge Martínez Garfias

Reviewed: 29 April 2024 Published: 16 August 2024