Sergej O. Demokritov - Spin Wave Confinement - Propagating Waves, Second Edition-Pan Stanford (2017)
Sergej O. Demokritov - Spin Wave Confinement - Propagating Waves, Second Edition-Pan Stanford (2017)
Confinement
1BO4UBOGPSE4FSJFTPO3FOFXBCMF&OFSHZ7PMVNF
Spin Wave
Confinement
Second Edition
Propagating Waves
editors
Preben Maegaard edited by
Anna Krenz
Wolfgang PalzSergej O. Demokritov
Wind Power
for the World
Published by
Pan Stanford Publishing Pte. Ltd.
Penthouse Level, Suntec Tower 3
8 Temasek Boulevard
Singapore 038988
Email: [email protected]
Web: www.panstanford.com
Introduction 1
1.1 Introduction 11
1.2 Spin Wave Dispersion 15
1.3 Spin Wave Excitation 18
1.4 Spin Wave Steering 25
1.5 Spin Wave Output 31
1.6 Spin Wave Control and Magnonic Devices 32
1.7 Conclusions and Outlook 37
2.1 Introduction 47
2.2 Theoretical Approach 48
2.3 Spin Waves in Coupled Magnetic Stripes 51
2.4 Nonlinear Spin Wave Coupling in Magnonic Crystals 56
2.5 Multilayer Magnonic Crystals 60
2.6 Frequency-Selective Tunable Spin Wave Channeling 67
2.7 Conclusions 72
vi Contents
3.1 Introduction 78
3.2 Sample Fabrication and Brillouin Light Scattering
Measurements 80
3.3 Micromagnetic Simulations 83
3.4 Results and Discussion 84
3.4.1 Spin Wave Band Structure and Mode
Spatial Profiles for the NWs with
Rectangular Cross Section 84
3.4.2 Spin Wave Band Structure and Mode
Spatial Profiles for the NWs with L-Shaped
Cross Section 88
3.5 Conclusions 91
Index 427
Introduction
Sergej O. Demokritov
Yekateringburg
January 2017
References
13. Hoffmann, A. (2013). Spin Hall effects in metals, IEEE Trans. Magn., 49,
5172–5193.
14. Kajiwara, Y., Harii, K., Takahashi, S., Ohe, J., Uchida, K., Mizuguchi, M.,
Umezawa, H., Kawai, H., Ando, K., Takanashi, K., Maekawa, S., and
Saitoh, E. (2010). Transmission of electrical signals by spin-wave
interconversion in a magnetic insulator, Nature, 464, 262–266.
15. Wang, Z., Sun, Y., Wu, M., Tiberkevich, V., and Slavin, A. (2011).
Control of spin waves in a thin film ferromagnetic insulator through
interfacial spin scattering, Phys. Rev. Lett., 107, 146602.
16. Demidov, V. E., Urazhdin, S., Rinkevich, A. B., Reiss, G., and Demokritov,
S. O. (2014). Spin Hall controlled magnonic microwaveguides, Appl.
Phys. Lett., 104, 152402.
17. An, K., Daniel, R., Birt, D. R., Pai, C.-F., Olsson, K., Ralph, D. C.,
Buhrman, R. A., and Li, X. (2014). Control of propagating spin
waves via spin transfer torque in a metallic bilayer waveguide, Phys.
Rev. B, 89, 140405(R).
18. Urazhdin, S., Demidov, V. E., Ulrichs, H., Kendziorczyk, T., Kuhn, T.,
Leuthold, J., Wilde, G., and Demokritov, S. O. (2014). Nanomagnonic
devices based on the spin-transfer torque, Nat. Nanotechnol., 9,
509–513.
19. Vlaminck, V., and Bailleul, M. (2008). Current-induced spin-wave
Doppler shift, Science, 322, 410–413.
20. Dzyaloshinsky, I. (1958). A thermodynamic theory of ‘weak’
ferromagnetism of antiferromagnetics, J. Phys. Chem. Solids, 4,
241–255.
21. Moriya, T. (1960). New mechanism of anisotropic superexchange
interaction, Phys. Rev. Lett., 4, 228–230.
22. Bode, M., Heide, M., von Bergmann, K., Ferriani, P., Heinze, S.,
Bihlmayer, G., Kubetzka, A., Pietzsch, O., Blügel, S., and Wiesendanger, R.
(2007). Chiral magnetic order at surfaces driven by inversion
asymmetry, Nature, 447, 190–193.
23. Kittel, C. (1958). Excitation of spin waves in a ferromagnet by
uniform RF field, Phys. Rev., 110, 1295–1297.
24. Seavey Jr., M. H., and Tannenwald, E. (1958). Direct observation of
spin wave resonance, Phys. Rev. Lett., 1, 168–169.
25. Damon, R. W., and Eshbach, J. R. (1961). Magnetostatic modes of a
ferromagnet slab, J. Phys. Chem. Solids, 19, 308–320.
26. Grunberg, P., and Metawe, F. (1977). Light scattering from bulk and
surface spin waves in EuO, Phys. Rev. Lett., 39, 1561–1565.
References
27. Grunberg, P., Schreiber, R., Pang, Y., Brodsky, M. B., and Sowers, H.
(1986). Layered magnetic structures: evidence for antiferromagnetic
coupling of Fe layers across Cr interlayers, Phys. Rev. Lett., 57,
2442–2445.
28. Mathieu, C., Jorzick, J., Frank, A., Demokritov, S. O., Slavin, A. N.,
Hillebrands, B., Bartenlian, B., Chappert, C., Decanini, D., Rousseaux, F.,
and Cambrill, E. (1998). Lateral quantization of spin waves in micron
size magnetic wires, Phys. Rev. Lett., 81, 3968–3971.
29. Jorzick, J., Demokritov, S. O., Hillebrands, B., Berkov, D., Gorn, N. L.,
Guslienko, K., and Slavin, A. N. (2002). Spin wave wells in nonellipsoidal
micrometer size magnetic elements, Phys. Rev. Lett., 88, 047204.
30. Park, J. P., Eames, P., Engebretson, D. M., Berezovsky, J., and Crowell, P. A.
(2002). Spatially resolved dynamics of localized spin-wave modes
in ferromagnetic wires, Phys. Rev. Lett., 89, 277201.
31. Gubbiotti, G., Tacchi, S., Carlotti, G. Singh, N., Goolaup, S., Adeyeye, A. O.,
and Kostylev, M. (2007). Collective spin modes in monodimensional
magnonic crystal consisting of dipolarly coupled nanowires, Appl.
Phys. Lett., 90, 092503.
32. Wang, K., Zhang, V. L., Lim, H. S., Ng, S. C., Kuok, M. H., Jain, S., and
Adeyeye, A. O. (2009). Observation of frequency band gaps in a one-
dimensional nanostructured magnonic crystal, Appl. Phys. Lett., 94,
083112.
33. Krawczyk, M., Mamica, S., Mruczkiewicz, M., Klos, J. W., Tacchi,
S., Madami, M., Gubbiotti, G., Duerr, G., and Grundler, D. (2013).
Magnonic band structures in two-dimensional bi-component
magnonic crystals with in-plane magnetization, J. Phys. D: Appl. Phys.,
46, 495003.
34. Tsoi, M., Jansen, A. G. M., Bass, J., Chiang, W.-C., Seck, M., Tsoi, V., and
Wyder, P. (1998). Excitation of a magnetic multilayer by an electric
current, Phys. Rev. Lett., 80, 4281–4284.
35. Tsoi, M., Jansen, A. G. M., Bass, J., Chiang, W.-C., Seck, M., Tsoi, V.,
and Wyder, P. (2000). Generation and detection of phase-coherent
current-driven magnons in magnetic multilayers, Nature, 406, 46–48.
36. Kiselev, S. I., Sankey, J. C., Krivorotov, I. N., Emley, N. C., Schoelkopf, R. J.,
Buhrman, R. A., and Ralph, D. C. (2003). Microwave oscillations of
a nanomagnet driven by a spin-polarized current, Nature, 425,
380–383.
37. Rippard, W. H., Pufall, M. R., Kaka, S., Russek, S. E., and Silva, T. J. (2004).
Direct-current induced dynamics in Co90Fe10/Ni80Fe20 point
contacts, Phys. Rev. Lett., 92, 027201.
10 Introduction
38. Krivorotov, I. N., Emley, N. C., Sankey, J. C., Kiselev, S. I., Ralph, D. C., and
Buhrman, R. A. (2005). Time-domain measurements of nanomagnet
dynamics driven by spin-transfer torques, Science, 307, 228–232.
39. Demidov, V. E., Urazhdin, S., and Demokritov, S. O. (2010). Direct
observation and mapping of spin waves emitted by spin-torque
nano-oscillators, Nat. Mater., 9, 984–988.
40. Demidov, V. E., Urazhdin, S., Ulrichs, H., Tiberkevich, V., Slavin, A.,
Baither, D., Schmitz, G., and Demokritov, S. O. (2012). Magnetic nano-
oscillator driven by pure spin current, Nat. Mater., 11, 1028–1031.
Chapter 1
1.1 Introduction
The most general definition of magnonics [1, 2], as the study
of spin waves [3, 4], leaves a lot of freedom for interpretation
Figure 1.1 (a) The relationship between magnonics and its sister fields
of research and technology. (b) A block diagram of a generic magnonic
device and its constituents. Adapted from Davies et al. [46].
and so are admittedly rather slow. However, this also means that,
at the same frequency, they have a much shorter wavelength
compared to, e.g., electromagnetic waves, paving the way toward
device miniaturization.
Figure 1.2 (a) The magnetostatic spin wave dispersion plotted in one
quadrant of reciprocal space for a film in the yz plane with in-plane
magnetization in the z direction. Above and below the FMR frequency
fFMR, there is a single and infinite manifold of dispersion branches,
referred to as the surface (Damon–Eshbach) mode and backward volume
modes, respectively. Only the first branch of the volume spin wave modes
is shown, corresponding to the uniform precession across the film
thickness. (b) The isofrequency curves characterizing the propagation
of magnetostatic spin waves, where f2 > fFMR > f1. (c) The isofrequency
curves characterizing the propagation of dipole exchange spin waves,
where f2 > f1 > fFMR.
Spin Wave Dispersion 17
distance from the end increases (Fig. 1.4a). The different values
of fFMR(r) correspond to different propagating spin wave modes,
the dispersion of which is shown in Fig. 1.4b. Hence, when the
stripe is excited by the uniform microwave magnetic field with a
frequency matching fFMR(r) at a particular distance from the end,
the magnetization at the distance gets resonantly excited and
becomes a source of spin waves propagating into the stripe.
The spin waves are imaged using the time-resolved scanning
Kerr microscope (TRSKM) [26, 61], with the acquired images
presented in Fig. 1.4c,d for the excitation at frequencies of 5.76
and 7.52 GHz, respectively. Further measurements (not shown
here) have demonstrated successful excitation of spin waves
across a frequency range of more than 4 GHz, which far surpasses
in bandwidth the Fano resonance–assisted mechanism described
by Au et al. [61, 62].
A peculiar example of a graded magnonic landscape is given
by magnetic domain walls, i.e., boundaries between regions of
different magnetization orientation. In particular, it has been
recently shown that a domain wall can generate spin waves when
excited by an external magnetic field or a spin-polarized current
[69–74], while arrays of domain walls were proposed as spin
wave grating couplers [75]. The spin wave emission has traditionally
been attributed to the effect of the domain wall oscillations.
However, at least in two of the studies referenced above, the
spin waves were observed to have the frequency of the driving
stimuli, rather than twice its value (as one would expect for
a nonlinear process of interaction between two oscillatory
modes). So, to uncover the mechanism of the emission, we
have developed a linear analytical theory [108] in the exchange
approximation using the formalism from Ref. [58].
Figure 1.5a shows a domain wall excited by a uniform
microwave magnetic field oriented perpendicular to the
magnetization in both the domain wall and domains. Our
calculations show that propagating spin waves at the frequency of
the incident field are emitted from the domain wall (Fig. 1.5b), as
a result of a linear process. The underlying mechanism is similar to
that proposed by Schlömann [55], except the domain wall creates
a natural graded magnonic index landscape with a reduction of
Spin Wave Excitation 23
Figure 1.4 (a) A cross section of the calculated fFMR(r) profile along
the length of the stripe is shown for the first 5 μm from its left end
and y = 5 μm. (b) The spin wave dispersion calculated from simulations
of an infinitely long stripe is shown in grayscale, with the overlaid red
circles showing the points deduced from the measurements of the
finite length stripe. For the dispersion calculation, the results of the
simulations were spatially smoothed so as to mimic the experimental
resolution. The horizontal dashed lines extending from panel (a) to
(b) illustrate the correspondence between the source region and wave
number of propagating spin waves excited at frequencies of 5.76 and
7.52 GHz, the Kerr images of which are shown in panels (c) and
(d) respectively. After Mushenok et al. [107].
24 Graded Magnonic Index and Spin Wave Fano Resonances in Magnetic Structures
same time, the energy flow along the arms dictates the direction
of the group velocity, which is therefore bound to be parallel to
their horizontal symmetry axis. Hence, the anisotropic dispersion
of the magnetostatic spin waves (Fig. 1.2) leads to a small,
non-zero angle between the directions of the group velocity and
wave vector (phase velocity), explaining the tilt of the phase fronts.
The same explains the tilts of the spin wave phase fronts observed
in Fig. 1.4c,d. When the bias magnetic field is rotated from the
symmetry axis by just ±15°, we observe only one spin wave beam
propagating into one of the arms (Fig. 1.6b,c). The direction of
the propagation is “switched” between the two arms by the sign
of the tilt angle. In each case, the spin wave beam (emitted from
the leg-arm junction) propagates at an oblique angle to the arm’s
axis, hits its edge, and is reflected into a much broader beam,
propagating approximately along the arm’s length. The incidence
and reflection angles are different, again resulting from the
anisotropy of the magnetostatic dispersion relation and the tilt of
the static magnetisation.
We interpret our observations in terms of the graded magnonic
index induced by the spatial variations of the orientations of
the magnetisation and the value of the internal magnetic field in
the sample. Using the convincing agreement between the
measured and numerically simulated results, we apply the theory
from Vashkovsky and Lock [96, 97] to the numerically computed
static magnetization and field distributions (Fig. 1.7a) to derive
the local directions of the wave vectors and group velocities of
the propagating spin waves (shown Fig. 1.6c). The confinement of
the precessing magnetization to the width of the T junction’s
leg results in a broad kx spectrum (Fig. 1.7c). For each kx value,
the isofrequency curve corresponding to the frequency of the
incident microwave field returns allowed (by the magnetostatic
dispersion relation) values of ky, while the normals to the
isofrequency curves show the group velocity directions
(Fig. 1.7d–h). The field and magnetization distributions in the
arms are quite uniform along the x axis starting from about 1 µm
from the leg-arm boundary, which ensures conservation of the
kx value of the spin wave propagating across the arm’s width.
In contrast, the values of ky and the group velocity adjust
adiabatically to the variation of the internal field magnitude and
direction of the magnetization [98]. The non-uniformity also
28 Graded Magnonic Index and Spin Wave Fano Resonances in Magnetic Structures
Apart from the small region at small wave vectors, the spin
wave isofrequency curves depicted in Fig. 1.7d–h consist of
nearly straight lines. This leads to virtually the same direction of
group velocity for a wide range of wave vectors, giving rise to the
formation of spin wave caustic beams [47, 96, 97]. This explains
the strongly directional beam emitted from the leg-arm boundary
for the tilted-bias magnetic field, but not the absence of the
other beam. Due to the inhomogeneities of the internal field and
magnetization (Fig. 1.7a), the beam curves slightly and experiences
distributed scattering, with the group velocities of the scattered
waves being roughly aligned with the arm’s length (Fig. 1.7d,e).
The group velocity of the reflected beam switches direction near
the far edge of the arm (Fig. 1.7f), leading to the phenomenon
of “back reflection” [96]. The reflected beam is confined by the
non-uniform demagnetizing field and magnetization near the
arm’s edge. In addition to this, some spin waves with small
(negative or positive) kx values are not supported in parts of
the magnetic landscape at all. In Fig. 1.7h, for example, there
is no intersection between the line kx = 0.94 μm–1 and the
isofrequency curves, giving rise to a “forbidden” path for spin waves
of certain wave vector. Finally, and quite surprisingly, we find
that the beam formed from spin waves with negative kx values
cannot possibly propagate into the left arm of the junction (for
the bias field direction in Fig. 1.7). Indeed, the beam is curved
into the nearest edge of the left arm from which it is then
scattered backwards into the right arm (Fig. 1.7g). The observed
complete disappearance of one of the two beams in favor of
the other one would be impossible without the graded
distribution of the magnonic index. The anisotropic dispersion is
also required for the effects to take place, but on its own it could
only lead to a tilt of both beams and asymmetry of their
intensities [92].
Figure 1.6 clearly shows that the spin wave beam initiated
near the leg-arm boundary then propagates along the arms of
the structure as prescribed by the direction of the bias magnetic
field. However, the beam is also quite wide and not as distinct
as one could wish. So, Fig. 1.8 presents results of simulations
for a Permalloy T junction that has a narrower (1 µm wide)
leg, which leads to a better-defined spin wave caustic beam
propagating into one arm of the structure. The smaller width
30 Graded Magnonic Index and Spin Wave Fano Resonances in Magnetic Structures
Figure 1.9 (a) The Fano control element is positioned 5 nm above the
waveguide. A spin wave (SW) excited elsewhere propagates along the
waveguide in the negative x direction. (b) The first (top) snapshot shows
the out-of-plane component of magnetization (mz) when the control
element is absent. The second (middle) and third (bottom) snapshots
show mz with the control element present and magnetized parallel and
antiparallel to the y axis, respectively. (c, d) The same as (a, b) but for
the control element positioned 20 nm above the waveguide. Adapted
from Davies et al. [46].
Spin Wave Control and Magnonic Devices 35
Figure 1.11 (a) Magnonic NAND gate realized as a pair of two transducers
positioned on a shared waveguide. (b) Snapshots of mz corresponding to
the same moment of time are shown for the four possible NAND gate input
combinations. Adapted from Davies et al. [46].
Conclusions and Outlook 37
wave propagation in Fig. 1.11b, spin waves are only absent from
the center of the waveguide when an input of (1,1) is used. The
NAND gate functionality is therefore obtained, as demonstrated
in the truth table shown in the inset of Fig. 1.11a.
Acknowledgments
The research leading to these results has received funding from
the Engineering and Physical Sciences Research Council of the
United Kingdom (Project Nos. EP/L019876/1, EP/L020696,
and EP/P505526/1), and from the European Union’s Horizon
2020 research and innovation program under Marie Skłodowska-
Curie Grant Agreement No. 644348 (MagIC).
References
9. Gubbiotti, G., Tacchi, S., Madami, M., Carlotti, G., Adeyeye, A. O., and
Kostylev, M. (2010). Brillouin light scattering studies of planar
metallic magnonic crystals, J. Phys. D: Appl. Phys., 43, 264003.
10. Lenk, B., Garbs, F., Ulrichs, H., Abeling, N., and Munzenberg, M. (2013).
Photo-magnonics, Top. Appl. Phys., 125, 71.
11. Kalashnikova, A. M., Kimel, A. V., and Pisarev, R. V. (2015). Ultrafast
opto-magnetism, Phys. Usp., 58, 969.
12. Bauer, G. E. W., Saitoh, E., and van Wees, B. J. (2012). Spin caloritronics,
Nat. Mater., 11, 391.
13. Bonetti, S., and Akerman, J. (2013). Nano-contact spin-torque
oscillators as magnonic building blocks, Top. Appl. Phys., 125, 177.
14. Locatelli, N., Cros, V., and Grollier, J. (2014). Spin-torque building
blocks, Nat. Mater., 13, 11.
15. Chumak, A. V., Vasyuchka, V. I., Serga, A. A., and Hillebrands, B. (2015).
Magnon spintronics, Nat. Phys., 11, 453.
16. Adeyeye, A. O., and Jain, S. (2011). Coupled periodic magnetic
nanostructures (invited), J. Appl. Phys., 109, 07B903.
17. Barman, A., and Haldar, A. (2014). Time-domain study of magnetization
dynamics in magnetic thin films and micro- and nanostructures,
Solid State Phys., 65, 1.
18. Demidov, V. E., and Demokritov, S. O. (2015). Magnonic waveguides
studied by microfocus Brillouin light scattering, IEEE Trans. Magn.,
51, 0800215.
19. Grundler, D. (2016). Nanomagnonics, J. Phys. D: Appl. Phys., 49,
391002.
20. Serga, A. A., Chumak, A. V., and Hillebrands, B. (2010). YIG magnonics,
J. Phys. D: Appl. Phys., 43, 264002.
21. Khitun, A., Bao, M. Q., and Wang, K. L. (2010). Magnonic logic circuits,
J. Phys. D: Appl. Phys., 43, 264005.
22. Chumak, A. V., Karenowska, A. D., Serga, A. A., and Hillebrands, B. (2013).
The dynamic magnonic crystal: new horizons in artificial crystal
based signal processing, Top. Appl. Phys., 125, 243.
23. Landau, L. D., and Lifshitz, E. M. (1977). Quantum Mechanics:
Non-Relativistic Theory (Pergamon Press, Oxford).
24. Marchand, E. W. (1978). Gradient Index Optics (Academic Press,
London).
25. Leonhardt, U., and Philbin, T. G. (2010). Geometry and Light: The
Science of Invisibility (Dover Publications Inc, New York).
40 Graded Magnonic Index and Spin Wave Fano Resonances in Magnetic Structures
26. Davies, C. S., Francis, A., Sadovnikov, A. V., Chertopalov, S. V., Bryan,
M. T., Grishin, S. V., Allwood, D. A., Sharaevskii, Y. P., Nikitov, S. A., and
Kruglyak, V. V. (2015). Towards graded-index magnonics: steering
spin waves in magnonic networks, Phys. Rev. B, 92, 020408(R).
27. Davies, C. S., and Kruglyak, V. V. (2015). Graded-index magnonics,
Low Temp. Phys., 41, 976.
28. Gorobets, Y. I., and Reshetnyak, S. A. (1998). Reflection and refraction
of spin waves in uniaxial magnets in the geometrical-optics
approximation, Tech. Phys., 43, 188.
29. Reshetnyak, S. A. (2004). The approximation of geometrical optics
for bulk spin waves in spatially inhomogeneous ferromagnetic
insulators with an exchange defect, Low Temp. Phys., 30, 398.
30. Jeong, D.-E., Han, D.-S., Choi, S., and Kim, S.-K. (2011). Refractive index
and Snell’s law for dipole-exchange spin-waves in a confined planar
structure, SPIN, 1, 27.
31. Gruszecki, P., Romero-Vivas, J., Dadoenkova, Y. S., Dadoenkova, N. N.,
Lyubchanskii, I. L., and Krawczyk, M. (2014). Goos-Hänchen effect
and bending of spin wave beams in thin magnetic films, Appl. Phys.
Lett., 105, 242406.
32. Elyasi, M., Bhatia, C. S., Qiu, C. W., and Yang, H. (2016). Cloaking the
magnons, Phys. Rev. B, 93, 104418.
33. Stigloher, J., Decker, M., Korner, H. S., Tanabe, K., Moriyama, T., Taniguchi,
T., Hata, H., Madami, M., Gubbiotti, G., Kobayashi, K., Ono, T., and
Back, C. H. (2016). Snell’s law for spin waves, Phys. Rev. Lett., 117,
037204.
34. Xing, X. J., and Zhou, Y. (2016). Fiber optics for spin waves, NPG
Asia Mater., 8, e246.
35. Yu, W. C., Lan, J., Wu, R. Q., and Xiao, J. (2016). Magnetic Snell’s law
and spin-wave fiber with Dzyaloshinskii-Moriya interaction, Phys.
Rev. B, 94, 140410.
36. Demokritov, S. O. (2008). Spin Wave Confinement (Pan Stanford,
Singapore).
37. Fano, U. (1961). Effects of configuration interaction on intensities
and phase shifts, Phys. Rev., 124, 1866.
38. Miroshnichenko, A. E., Flach, S., and Kivshar, Y. S. (2010). Fano
resonances in nanoscale structures, Rev. Mod. Phys., 82, 2257.
39. Al-Wahsh, H., El Boudouti, E. H., Djafari-Rouhani, B., Akjouj, A., Mrabti,
T., and Dobrzynski, L. (2008). Evidence of Fano-like resonances
in mono-mode magnetic circuits, Phys. Rev. B, 78, 075401.
References 41
67. Yu, H. M., d’Allivy Kelly, O., Cros, V., Bernard, R., Bortolotti, P., Anane,
A., Brandl, F., Heimbach, F., and Grundler, D. (2016). Approaching
soft X-ray wavelengths in nanomagnet-based microwave technology,
Nat. Commun., 7, 11255.
68. Hermsdoerfer, S. J., Schultheiss, H., Rausch, C., Schafer, S., Leven, B.,
Kim, S. K., and Hillebrands, B. (2009). A spin-wave frequency doubler
by domain wall oscillation, Appl. Phys. Lett., 94, 223510.
69. Marchenko, A. N., and Krivoruchko, V. N. (2012). Magnetic structure
and resonance properties of a hexagonal lattice of antidots, Low
Temp. Phys., 38, 157.
70. Roy, P. E., Trypiniotis, T., and Barnes, C. H. W. (2010). Micromagnetic
simulations of spin-wave normal modes and the resonant
field-driven magnetization dynamics of a 360 degrees domain wall
in a soft magnetic stripe, Phys. Rev. B, 82, 134411.
71. Boone, C. T., and Krivorotov, I. N. (2010). Magnetic domain wall
pumping by spin transfer torque, Phys. Rev. Lett., 104, 167205.
72. Mozooni, B., and McCord, J. (2015). Direct observation of closure
domain wall mediated spin waves, Appl. Phys. Lett., 107, 042402.
73. Van de Wiele, B., Hamalainen, S. J., Balaz, P., Montoncello, F., and van
Dijken, S. (2016). Tunable short-wavelength spin wave excitation
from pinned magnetic domain walls, Sci. Rep., 6, 21330.
74. Sluka, V., Weigand, M., Kakay, A., Erbe, A., Tyberkevych, V., Slavin, A.,
Deac, A., Lindner, J., Fassbender, J., Raabe, J., and Wintz, S. (2015).
Stacked topological spin textures as emitters for multidimensional
spin wave modes, Abstract DE-03 in the Book of Abstracts of the
2015 IEEE Intermag Conference (May 11–15, 2015, Beijing, China).
75. Truetzschler, J., Sentosun, K., Mozooni, B., Mattheis, R., and McCord,
J. (2016). Magnetic domain wall gratings for magnetization reversal
tuning and confined dynamic mode localization, Sci. Rep., 6, 30761.
76. Pöschl, G., and Teller, E. (1933). Bemerkungen zur Quantenmechanik
des anharmonischen Oszillators, Z. Phys., 83, 143.
77. Demidov, V. E., Jersch, J., Demokritov, S. O., Rott, K., Krzysteczko, P.,
and Reiss, G. (2009). Transformation of propagating spin-wave modes
in microscopic waveguides with variable width, Phys. Rev. B, 79,
054417.
78. Demidov, V. E., Urazhdin, S., Zholud, A., Sadovnikov, A. V., and
Demokritov, S. O. (2015). Dipolar field-induced spin-wave waveguides
for spin-torque magnonics, Appl. Phys. Lett., 106, 022403.
79. Lan, J., Yu, W. C., Wu, R. Q., and Xiao, J. (2015). Spin-wave diode, Phys.
Rev. X, 5, 041049.
44 Graded Magnonic Index and Spin Wave Fano Resonances in Magnetic Structures
80. Garcia-Sanchez, F., Borys, P., Soucaille, R., Adam, J. P., Stamps, R.
L., and Kim, J. V. (2015). Narrow magnonic waveguides based on
domain walls, Phys. Rev. Lett., 114, 247206.
81. Perez, N., and Lopez-Diaz, L. (2015). Magnetic field induced spin-
wave energy focusing, Phys. Rev. B, 92, 014408.
82. Dzyapko, O., Borisenko, I. V., Demidov, V. E., Pernice, W., and
Demokritov, S. O. (2016). Reconfigurable heat-induced spin wave
lenses, Appl. Phys. Lett., 109, 232407.
83. Hertel, R., Wulfhekel, W., and Kirschner, J. (2004). Domain-wall
induced phase shifts in spin waves, Phys. Rev. Lett., 93, 257202.
84. Vasiliev, S. V., Kruglyak, V. V., Sokolovskii, M. L., and Kuchko, A. N.
(2007). Spin wave interferometer employing a local nonuniformity
of the effective magnetic field, J. Appl. Phys., 101, 113919.
85. Kanazawa, N., Goto, T., Sekiguchi, K., Granovsky, A. B., Ross, C. A.,
Takagi, H., Nakamura, Y., and Inoue, M. (2016). Demonstration of a
robust magnonic spin wave interferometer, Sci. Rep., 6, 30268.
86. Lee, K. S., and Kim, S. K. (2008). Conceptual design of spin wave
logic gates based on a Mach-Zehnder-type spin wave interferometer
for universal logic functions, J. Appl. Phys., 104, 053909.
87. Csaba, G., Papp, A., and Porod, W. (2014). Spin-wave based realization
of optical computing primitives, J. Appl. Phys., 115, 17C741.
88. Klingler, S., Pirro, P., Bracher, T., Leven, B., Hillebrands, B., and
Chumak, A. V. (2015). Spin-wave logic devices based on isotropic
forward volume magnetostatic waves, Appl. Phys. Lett., 106, 212406.
89. Gertz, F., Kozhevnikov, A., Khivintsev, Y., Dudko, G., Ranjbar, M.,
Gutierrez, D., Chiang, H., Filimonov, Y., and Khitun, A. (2016).
Parallel read-out and database search with magnonic holographic
memory, IEEE Trans. Magn., 52, 3401304.
90. Vogt, K., Fradin, F. Y., Pearson, J. E., Sebastian, T., Bader, S. D.,
Hillebrands, B., Hoffmann, A., and Schultheiss, H. (2014). Realization
of a spin-wave multiplexer, Nat. Commun., 5, 3727.
91. Sadovnikov, A. V., Davies, C. S., Grishin, S. V., Kruglyak, V. V., Romanenko,
D. V., Sharaevskii, Y. P., and Nikitov, S. A. (2015). Magnonic beam
splitter: the building block of parallel magnonic circuitry, Appl. Phys.
Lett., 106, 192406.
92. Davies, C. S., Sadovnikov, A. V., Grishin, S. V., Sharaevskii, Y. P., Nikitov,
S. A., and Kruglyak, V. V. (2015). Field-controlled phase-rectified
magnonic multiplexer, IEEE Trans. Magn., 51, 3401904.
93. Braecher, T., Pirro, P., Westermann, J., Sebastian, T., Lagel, B., Van de
Wiele, B., Vansteenkiste, A., and Hillebrands, B. (2013). Generation
References 45
107. Mushenok, F. B., Dost, R., Davies, C. S., Allwood, D. A., Inkson, B. J.,
Hrkac, G., and Kruglyak, V. V. (2017). Broadband conversion of
microwaves into propagating spin waves in patterned magnetic
structures, arXiv:1706.04409.
108. Whitehead, N. J., Horsley, S. A. R., Philbin, T. G., Kuchko, A. N.,
and Kruglyak, V. V. (2017). Theory of linear spin wave emission from
a Bloch domain wall, arXiv:1705.01852.
109. Davies, C. S., Poimanov, V. D., and Kruglyak, V. V. (2017). Mapping
the magnonic landscape in patterned magnetic structures,
arXiv:1706.03212.
Chapter 2
2.1 Introduction
Recent developments in magnetic thin-film technology led to
the fabrication of basic functional elements for spin wave (SW)
signal transmission and processing in the microwave range in
micro- and nanoscales: SW waveguides, delay lines, resonators,
and oscillators [1–5]. A magnonic waveguide, formed from a
magnetic stripe, is a building block of any complex integral
reconfigurable magnonic network [6, 7]. Control over the dispersion
of SWs can be achieved, for example, by periodic patterning of
thin magnetic films. Periodic variation of the magnetic materials’
parameters allows fabrication of magnonic crystals (MCs) [3, 4, 8],
where
H1,2(t )= H0 + h1,2(is
t )+the dynamic
Kh2,1 (t ) magnetic microwave field in each layer
and K is the coupling coefficient that determines the coupling
between microwave magnetic fields of each waveguide. The
external magnetic (t )= H0 +ish1,2directed
H1,2field (t )+ Kh2,1along
(t ) the x axis. We can
obtain the following set of wave equations for magnetic waveguides
using the equation of motion for high-frequency components
of magnetization and the equation for the magnetostatic potential
in each film, as well as the appropriate boundary conditions in
the long-wave approximation for magnetostatic surface waves
(MSSWs) [12] propagating along the y axis [13, 14]:
∂ 2m1,2 w2M d ∂
2 + wH ( wH + wM )m1,2 + (m + Km2,1)= 0 , (2.2)
∂t 2 ∂y 1,2
where m1,2 = my1,2/M0 are dynamic magnetization components in
each waveguide, M0 is the saturation magnetization, d is the
magnetic film thickness, wM = g4pM0, wH = gH0, and g is the
gyromagnetic ratio. This approach can be adopted to a periodic
structure using expansion into Fourier series for dynamic
magnetization components in each waveguide. We consider
further two coupled 1D MCs separated by a dielectric layer with
thickness D (for a layered structure) or width D (for lateral
magnetic stripes). The surfaces of each MC (MC 1 and MC 2) are
periodically corrugated with the period L. In general, it is assumed
that the MCs’ periods are shifted relatively to each other in the
direction of the y axis by the value of q.
The solution of the wave equations in each MC can be
represented as a sum of spatial harmonics [15]. We shall consider
only the zero-order harmonic of forward waves and the –1
harmonic of backward waves in the first Brillouin zone. In this
case the solution of Eq. 2.2 for each of the crystals is the
superposition of forward and backward waves:
where A1,2 and B1,2 are the amplitudes of the forward and
backward waves [14], respectively; k0 is the propagation constant
of the zero-order harmonic; k– = k0 – 2p/L refers to the −1
50 Coupled Spin Waves in Magnonic Waveguides
–
where D1,2 = – w2 + w2H + wM wH + 1 ,2k– , 1,2 = w2Md01,02 /2, and
± i ( y1 ,2 ) ± ±
±
q1,2 = e d1,2; and d1,2 = k± dd1,2/2, y1 = 0, and y2 = 2pq/L.
If dd1,2 = 0 and B = 0 Eq. 2.5 describes a structure consisting
of two coupled homogeneous film (HFs). If K = 0 then Eq. 2.5
describes SW dynamics in separated MCs.
We determine the dispersion relation for waves in a coupled
structure by equating the determinant of the set of equations in
Eq. 2.5 to zero:
the phase shifts y1 and y2 between the MCs. Note that when K ≠ 0
±
and d1,2 = 0, Eq. 2.6 describes the dispersion relation for MMSWs
±
in the structure of the two coupled HFs. When K = 0 and d1,2 ≠0
in Eq. 2.6 we obtain the dispersion equations for MC 1 and MC 2
separately [17]. If d2± = 0 in Eq. 2.6 for MC 2, the second layer is
an HF.
The main problem in the coupled wave approach is the
lack of definition of the coupling coefficient K and geometry of
the magnetic structures. Moreover, it is necessary to take into
account the multimode coupling between the transverse width
modes of each finite-width magnetic waveguide [18]. Thus the
micromagnetic numerical simulations reveal the properties
of modes coupling and can be used for geometry design of the
planar and layered directional multimode SW coupler. Using the
finite element method (FEM) [19] the spectra of eigenmodes of
the coupled magnetic structures and the coupling parameter K
can be calculated, whereas both the time- and space-dependent
magnetization evolution and SW transmission can be numerically
simulated by means of the finite-difference method [20].
Figure 2.1 (a) Schematic of the experiment. (b) Profile of the static
internal magnetic field Hint along the x axis.
Figure 2.2 (a) Dispersion characteristics for first three symmetric (solid
lines) and antisymmetric (dashed lines) modes. (b) Transmission
characteristics measured with a signal network analyzer (solid line)
and calculated with only first symmetric and antisymmetric modes
(dashed lines): the results of the calculation for first and third modes
are plotted with a dotted line. (c) Frequency dependence of the coupling
length obtained with FEM (solid lines), micromagnetic simulation
(closed circles), and BLS technique (open squares). (d) Coupling length
versus dimensionless ratio d/w at a frequency of 3.125 GHz.
(e–h) Profiles for first (n = 1) and second (n = 2) transverse symmetric
and antisymmetric modes.
56 Coupled Spin Waves in Magnonic Waveguides
Figure 2.3 (a) Normalized color-coded BLS intensity map. (b) Calculated
intensity (FEM) obtained by taking into account first three symmetric
and antisymmetric modes. (c) Micromagnetic simulation of SW amplitude.
Figure 2.5 (a) Transmission (blue solid curve) and dispersion (red
dash-dotted curve) characteristics, measured with the VNA. Calculated
transmission is shown with the dashed green curve; (b) shows the dynamic
magnetization along the z-coordinate of ports C1 and C2 as a function of
frequency. Yellow area is the guide for the eye to show the frequency and
wavenumber region of the magnonic forbidden zone; vertical dashed-
dotted line shows the central frequency of magnonic forbidden gap.
(c) Power exchange coefficient as a function of frequency detuning from
the center of magnonic forbidden zone. The dotted curve shows the
numerical results. Open circles, squares, and triangles show the
experimental data for power levels 26, 10, and−10 dBm, respectively.
Vertical dashed lines show the frequency detuning at frequencies f1, f2
and f3.
As it is well known [1, 9], the dispersion curve for the MSSW
splits into two normal modes, which correspond to fast and
slow waves in coupled HFs (dd1,2 =0 and K ≠ 0). Dispersion
characteristics for these waves are shown as dashed lines in
62 Coupled Spin Waves in Magnonic Waveguides
Fig. 2.7a–c: red lines 2 for the fast incident wave and blue
lines 3 for the slow incident wave. Corresponding dispersion
characteristics of the reflected waves are denoted by lines 2¢ and
3¢. Interaction between the waves of four described types leads
to the formation of BGs (shaded region in Fig. 2.7) at frequencies
of the phase synchronism. As can be seen from Fig. 2.7а, in the
symmetric structure at y = 0 the two BGs G-1 and G-2 at k = kB can
be formed. BG G-1 is observed at frequencies higher than the
frequency of the Bragg resonance for a single MC, fMC (red region
in Fig. 2.7a). BG G-1 is formed because of the interaction of
the incident and reflected fast waves (red dashed lines 2
and 2¢ in Fig. 2.7a). BG G-2 is observed at Re(k) = kB and at
frequencies lower than the frequency of the Bragg resonance for
a single MC, fMC (blue region in Fig. 2.7a). BG G-2 is formed by the
interaction of the incident and reflected slow waves (blue
dashed lines 3 and 3 in Fig. 2.7a). The presence of only two BGs
in Fig. 2.7a is a symmetric degenerate case when the structure
consists of two identical MCs placed symmetrically.
When y = 0.6 p (see. Fig. 2.7c) three BGs (G-1, G-2, G-3) within
the dispersion curve are visible. An additional band G-3 (green
Multilayer Magnonic Crystals 63
(green area G-3). Note that a similar situation exists for the
structure based on MCs with a shift.
Figure 2.8 (a) Dependence of the width and position of bandgaps from
the phase shift between MCs. (b) Dependence of the width of bandgaps
(G-1, G-2, G-3) (shaded areas) and central frequencies (dashed lines) on
values of coupling K for the structure MC-HF.
2.7 Conclusions
In this chapter we considered the concept of SW coupling in
lateral and vertical magnonic structures. We showed that the
functionalities of simple magnetic stripes can be extended by
lateral or vertical coupling with adjacent stripes or MCs. We
demonstrated nonlinear SW switching in two side-coupled MCs.
We showed that in a fabricated 2D MCA the control over a
propagation distance in SW channels can be implemented by the
orientation of an external magnetic field. Collimated SW beams can
be used as signal carriers in the magnonic platform for applications
such as signal multiplexing. Thus, coupled magnonic stripes and
crystals can act both as a magnonic splitter and a multimode
References 73
Acknowledgments
This work was supported, in part, by the grant from the Russian
Science Foundation (Project No. 16-19-10283, 14-19-00760), the
Russian Foundation for Basic Research (Project No. 15-07-05901),
and the grant from the president of the Russian Federation (No. MK-
5837.2016.9).
References
3.1 Introduction
Magnonic crystals (MCs) consist of periodically modulated
magnetic material and make use of fundamental properties of
waves, such as scattering and interference, to create “bandgaps”—
ranges of wavelength (or frequency) within which spin waves
(SWs) cannot propagate through the structure. The concept of
MCs followed by two decades the analogous concept of photonic
crystals for the propagation of electromagnetic waves [1, 2]
where the band gap is caused by a periodic variation in the
refractive index of an artificially structured material. In this
respect, MCs offer opportunities to tailor-made novel properties
for SW propagation such as high group velocity [3], magnetic
field tunability, and controllable bandgap opening [4–15].
Interesting phenomena such as anisotropic damping [16], Bragg
diffraction [17, 18], and SW mode conversion [19] have also
been demonstrated. Besides, bi-component MCs (BMCs), consisting
of two different periodically arranged magnetic materials, have
additional degrees of freedom thanks to the contrasting properties
of the ferromagnetic (FM) materials [20–22].
The progress in the field of MCs goes in parallel with the
capability to realize magnetic nanostructures over a large area by
Introduction 79
(a)
(b)
Table 3.1 Geometric and magnetic parameters for the single- and bilayered
NWs
sin[2pf0 (t – t 0 )]
h(t )= h0
2pf0 (t – t 0 )
k= k= k= 0 k= 0
Figure 3.6 Measured (points) and calculated (grey scale) dispersion
curves for spin waves propagating perpendicularly to the NWs length
(x direction) for H = 500 Oe applied along the NW length for
90 Tuning of the Spin Wave Band Structure in Nanostructured Iron/Permalloy NW Arrays
(d) 10.0 GHz (d) 12.3 GHz (d) 13.2 GHz (d) 14.9 GHz (d) 15.7 GHz
Amplitude (arb. units)
(c) 8.9 GHz (c) 11.4 GHz (c) 11.6 GHz (c) 13.7 GHz (c) 13.8 GHz
(d) 7.9 GHz (b) 10.2 GHz (b) 10.5 GHz (b) 11.7 GHz (b) 11.4 GHz
(a) 6.3 GHz (a) 7.1 GHz (a) 7.7 GHz (a) 8.3 GHz (a) 8.2 GHz
0 50 100 150 200 250 300 0 50 100 150 200 250 300 0 50 100 150 200 250 300 0 50 100 150 200 250 300
Width, x (nm) Width, x (nm) Width, x (nm) Width, x (nm) Width, x (nm)
3.5 Conclusions
In this chapter, we analyzed the dispersion of collective spin waves
in arrays of iron/permalloy nanowires with either rectangular
92 Tuning of the Spin Wave Band Structure in Nanostructured Iron/Permalloy NW Arrays
Acknowledgments
A. O. A. was supported by National Research Foundation, Prime
Minister’s Office, Singapore, under its Competitive Research
Programme (CRP Award No. NRF-CRP 10-2012-03).
References
4. Ma, F. S., Lim, H. S., Wang, Z. K., Piramanayagam, S. N., Ng, S. C., and
Kuok, M. H. (2011). Micromagnetic study of spin wave propagation
in bicomponent magnonic crystal waveguides, Appl. Phys. Lett., 98,
153107.
5. Neusser, S., Bauer, H. G., Duerr, G., Huber, R., Mamica, S., Woltersdorf, G.,
Krawczyk, M., Back, C. H., and Grundler, D. (2001). Tunable
metamaterial response of a Ni80Fe20 antidot lattice for spin waves,
Phys. Rev. B, 84, 184411.
6. Topp, J., Heitmann, D., Kostylev, M. P., and Grundler, D. (2010). Making
a reconfigurable artificial crystal by ordering bistable magnetic
nanowires, Phys. Rev. Lett., 104, 207205.
7. Schwarze, T., Huber, R., Duerr, G., and Grundler, D. (2012). Complete
band gaps for magnetostatic forward volume waves in a two-
dimensional magnonic crystal, Phys. Rev. B, 85, 134448.
8. Wang, Z. K., Zhang, V. L., Lim, H. S., Ng, S. C., Kuok, M. H., Jain, S., and
Adeyeye, A. O. (2010). Nanostructured magnonic crystals with
size-tunable bandgaps, ACS Nano, 4, 643–648.
9. Tacchi, S., Duerr, G., Klos, J. W., Madami, M., Neusser, S., Gubbiotti, G.,
Carlotti, G., Krawczyk, M., and Grundler, D. (2012). Forbidden band
gaps in the spin-wave spectrum of a two-dimensional bicomponent
magnonic crystal, Phys. Rev. Lett., 109, 137202.
10. Duerr, G., Madami, M., Neusser, S., Tacchi, S., Gubbiotti, G., Carlotti,
G., and Grundler, D. (2011). Spatial control of spin-wave modes in
Ni80Fe20 antidot lattices by embedded Co nanodisks, Appl. Phys. Lett.,
99, 202502.
11. Krawczyk, M., Mamica, S., Mruczkiewicz, M., Klos, J. W., Tacchi, S.,
Madami, M., Gubbiotti, G., Duerr, D., and Grundler, D. (2013). Magnonic
band structures in two-dimensional bi-component magnonic
crystals with in-plane magnetization, J. Phys. D: Appl. Phys., 46,
495003.
12. Ma, F. S., Lim, H. S., Zhang, V. L., Wang, Z. K., Piramanayagam, S. N., Ng,
S. C., and Kuok, M. H. (2012). Materials optimization of the magnonic
bandgap in two-dimensional bi-component magnonic crystal
waveguides, Nanosci. Nanotechnol. Lett., 4, 663–666.
13. Mamica, S., Krawczyk, M., and Klos, J. W. (2012). Spin-wave band
structure in 2D magnonic crystals with elliptically shaped scattering
centres, Adv. Condens. Matter Phys., 2012, 161387.
14. Rychły, J., Kłos, J. W., Mruczkiewicz, M., and Krawczyk, M. (2015). Spin
waves in one-dimensional bicomponent magnonic quasicrystals,
Phys. Rev. B, 92, 054414.
94 Tuning of the Spin Wave Band Structure in Nanostructured Iron/Permalloy NW Arrays
15. Nikitov, S. A., Kalyabin, D. V., Lisenkov, I. V., Slavin, A. N., Barabanenkov,
Yu N., Osokin, S. A., Sadovnikov, A. V., Beginin, E. N., Morozova,
M. A., Sharaevsky, Yu P., Filimonov, Yu A., Khivintsev, Yu V.,
Vysotsky, S. L., Sakharov, V. K., and Pavlov, E. S. (2015). Magnonics:
a new research area in spintronics and spin wave electronics,
Phys. Usp., 58, 1002–1028.
16. Neusser, S., Duerr, G., Bauer, H. G., Tacchi, S., Madami, M., Woltersdorf,
G., Gubbiotti, G., Back, C. H., and Grundler D. (2010). Anisotropic
propagation and damping of spin waves in a nanopatterned
antidot lattice, Phys. Rev. Lett., 105, 067208.
17. Zivieri, R., Tacchi, S., Montoncello, F., Giovannini, L., Nizzoli, F., Madami,
M., Gubbiotti, G., Carlotti, G., Neusser, S., Duerr, G., and Grundler, D.
(2012). Bragg diffraction of spin waves from a two-dimensional
antidot lattice, Phys. Rev. B, 85, 012403.
18. Yu, H., Duerr, G., Huber, R., Bahr, M., Schwarze, T., Brandl F., and
Grundler, D. (2013). Omnidirectional spin-wave nanograting coupler,
Nat. Commun., 4, 2702.
19. Tacchi, S., Botters, B., Madami, M., Klos, J. W., Sokolovskyy, M. L.,
Krawczyk, M., Gubbiotti, G., Carlotti, G., Adeyeye, A. O., Neusser, S., and
Grundler, D. (2012). Mode conversion from quantized to propagating
spin waves in a rhombic antidot lattice supporting spin wave
nanochannels, Phys. Rev. B, 86, 014417.
20. Vasseur, J. O., Dobrzynski, L., Djafari-Rouhani, B., and Puszkarski, H.
(1996). Magnon band structure of periodic composites, Phys. Rev. B,
54, 1043.
21. Puszkarski, H., and Krawczyk, M. (2003). Magnonic crystals—the
magnetic counterpart of photonic crystals, Solid State Phenom., 94,
125–134.
22. Nikitov, S. A., Tailhades, Ph., and Tsai, C. S. (2001). Spin waves in
periodic magnetic structures-magnonic crystals, J. Magn. Magn.
Mater., 236, 320–330.
23. Zheng, M., Yu, M., Liu, Y., Skomski, R., Liou, S. H. Sellmyer, D. J.,
Petryakov, V. N. Verevkin, Y. K. Polushkin, N. I., and Salashchenko,
N. N. (2001). Magnetic nanodot arrays produced by direct laser
interference lithography, Appl. Phys. Lett., 79, 2606.
24. Heyderman, L. J., Solak, H. H., David, C., Atkinson, D., Cowburn, R. P.,
and Nolting, F. (2004). Arrays of nanoscale magnetic dots: fabrication
by x-ray interference lithography and characterization, Appl. Phys.
Lett., 85, 4989–4991.
References 95
25. Jensen, T. R., Malinsky, M. D., Haynes, C. L., and Van Duyne, R. P.
(2000). Nanosphere lithography: tunable localized surface plasmon
resonance spectra of silver nanoparticles, J. Phys. Chem. B, 104,
10549–10556.
26. Haynes, C. L., and Van Duyne, R. P. (2001). Nanosphere lithography:
a versatile nanofabrication tool for studies of size-dependent
nanoparticle optics, J. Phys. Chem. B, 105, 5599–5611.
27. Ding, J., and Adeyeye, A. O. (2013). Binary ferromagnetic nano-
structures: fabrication, static and dynamic properties, Adv. Funct.
Mater., 23, 1684–1691.
28. Navas, D., Bi, B., Adeyeye, A. O., and Ross, C. A. (2015). Templates as
shadow masks to tune the magnetic anisotropy in nanostructured
CoCrPt/Ti bilayer films, Adv. Mater. Interfaces, 1400551.
29. Shimon, G., Ross, C. A., and Adeyeye, A. O. (2015). Self-aligned
Ni/NiFe/Fe magnetic lateral heterostructures, J. Appl. Phys., 118,
153901.
30. Gubbiotti, G., Malagò, P., Fin, S., Tacchi, S., Giovannini, L., Bisero,
D., Madami, M., Carlotti, G., Ding, J., Adeyeye, A. O., and Zivieri, R.
(2014). Magnetic normal modes of bicomponent Permalloy/cobalt
structures in the parallel and antiparallel ground state, Phys. Rev. B,
90, 024419.
31. Wang, Z. K., Zhang, V. L., Lim, H. S., Ng, S. C., Kuok, M. H., Jain, S., and
Adeyeye, A. O. (2009). Observation of frequency band gaps in a
one-dimensional nanostructured magnonic crystal, Appl. Phys. Lett.,
94, 083112.
32. Sokolovskyy, M. L., and Krawczyk, M. (2011). The magnetostatic
modes in planar one-dimensional magnonic crystals with nanoscale
sizes, J. Nanopart. Res., 13, 6085–6091.
33. Zhang, V. L., Lim, H. S., Lin, C. S., Wang, Z. K., Ng, S. C., Kuok, M. H.,
Jain, S., Adeyeye, A. O., and Cottam, M. G. (2011). Ferromagnetic
and antiferromagnetic spin-wave dispersions in a dipole-exchange
coupled bi-component magnonic crystal, Appl. Phys. Lett., 99, 143118.
34. Livesey, K. L., Ding, J., Anderson, N., Camley, R. E., Adeyeye, A. O.,
Kostylev, M. P., and Samarin, S. (2013). Resonant frequencies of a
binary magnetic nanowire, Phys. Rev. B, 87, 064424.
35. Lin, C. S., Lim, H. S., Zhang, V. L., Wang, Z. K., Ng, S. C., Kuok, M. H.,
Cottam, M. G., Jain, S., and Adeyeye, A. O. (2012). Interfacial
magnetization dynamics of a bi-component magnonic crystal
comprising contacting ferromagnetic nanostripes, J. Appl. Phys., 111,
033920.
96 Tuning of the Spin Wave Band Structure in Nanostructured Iron/Permalloy NW Arrays
36. Gubbiotti, G., Tacchi, S., Madami, M., Carlotti, G., Jain, S., Adeyeye,
A. O., and Kostylev, M. P. (2012). Collective spin waves in a
bicomponent two-dimensional Magnonic Crystal, Appl. Phys. Lett.,
100, 162407.
37. Krawczyk, M., Mamica, S., Mruczkiewicz, M., Klos, J. W., Tacchi,
S., Madami, M., Gubbiotti, G., Duerr, D., and Grundler, D. (2013).
Magnonic band structure in tangentially magnetized thin film of
two-dimensional bi-component magnonic crystals, J. Phys. D: Appl.
Phys., 46, 495003.
38. Zivieri, R. (2014). Bandgaps and demagnetizing effects in a Py/Co
magnonic crystal, IEEE Trans. Magn., 50, 1100304.
39. Duerr, D., Tacchi, S., Gubbiotti, G., and Grundler, D. (2014). Field-
controlled rotation of spin-wave nanochannels in bi-component
magnonic crystals, J. Phys. D: Appl. Phys., 47, 325001.
40. Malagò, P., Giovannini, L., Zivieri, R., Gruszecki, P., and Krawczyk, M.
(2015). Spin-wave dynamics in Permalloy/cobalt magnonic crystals
in the presence of a nonmagnetic spacer, Phys. Rev. B, 92, 064416.
41. Rychły, J., Gruszecki, P., Mruczkiewicz, M., Kłos, J. W., Mamica, S., and
Krawczyk, M. (2015). Magnonic crystals—prospective structures for
shaping spin waves in nanoscale, Low Temp. Phys., 41, 745.
42. Gubbiotti, G., Kostylev, M., Tacchi, S., Madami, M., Carlotti, G., Ding, J.,
Adeyeye, A. O., Zighem, F., Stashkevich, A. A., Ivanov, E., and Samarin, S.
(2014). Collective spin waves on a nanowire array with step-
modulated thickness, J. Phys. D: Appl. Phys., 47, 105003.
43. Gubbiotti, G., Tacchi, S., Madami, M., Carlotti, G., Yang, Z., Ding, J.,
Adeyeye, A. O., and Kostylev, M. (2016). Collective spin excitations in
bi-component magnonic crystals consisting of bi-layer Permalloy/Fe
nanowires, Phys. Rev. B, 93, 184411.
44. Gubbiotti, G., Silvani, R., Tacchi, S., Madami, M., Carlotti, G., Yang,
Z., Adeyeye, A. O., and Kostylev, M. (2017). Tailoring the spin waves
band structure of 1D magnonic crystals consisting of L-shaped
iron/permalloy nanowires, J. Phys. D: Appl. Phys., 50, 105002.
45. Kostylev, M., Yang, Z., Maksymov, I. S., Ding, J., Samarin, S., and
Adeyeye, A. O. (2016). Microwave magnetic dynamics in ferromagnetic
metallic nanostructures lacking inversion symmetry, J. Appl. Phys.,
119, 103903.
46. Sandercock, J. R. (1982). In Light Scattering in Solids III, eds. Cardona,
M., and Güntherodt, G., Springer Series in Topics in Applied Physics
Vol. 51 (Springer-Verlag, Berlin), p. 173.
References 97
Magnetization Dynamics of
Reconfigurable 2D Magnonic Crystals
4.1 Introduction
Nanomagnet arrays and clusters are considered vital building
blocks for many technological applications such as ultrahigh-
density media [1–3], magnetic random access memory [4, 5], logic
[6–10], microwave signal processing devices [11–13], magnonics
[14–21], and spin torque nano-oscillators (STOs) [22–26]. Some
of these applications require nanomagnets to be packed more
closely together in order to achieve higher-areal-density devices.
Beyond the challenges for device fabrication and integration,
the strength of dipolar interactions between neighboring
nanomagnets is known to be significantly enhanced when
interelement spacing becomes much smaller than their lateral
dimensions. Such enhanced dipolar interaction has been found
to largely modify magnetization reversal [27–29], switching field
distribution [30–32], and dynamic magnetization reversal processes
[33–36]. Dipolar interaction can also be engineered to realize
functional magnetic systems. For example, magnetic quantum
cellular automata (MQCA) have been utilized to perform logic
operations and propagate magnetic information [6, 9, 37].
Similarly, artificial spin ice has been shown to produce localized
spin wave (SW) or selective SW propagation based on defects
or change of magnetization states [38–40]. Additionally, flavors
of tunable dynamic behaviors were made possible by exploiting
their dipolar interaction [11, 33–36].
Magnonic crystals (MCs), magnetic analogues of photonic
crystals [41, 42], hold the possibility to tailor SW propagation and
frequency on the basis of geometrical confinement and periodic
Introduction 101
g
f res = (Hz +( N y – N z)4 pMz ) (Hz +( N x – N z )4 pMz ) (4.1)
2p
yield accurate results, for example, lex of NiFe is ~5.7 nm. Typically,
a unit cell size of 5 nm × 5 nm × 5 nm was chosen for our
simulations.
There are two common simulation routines used in our work.
First was the basic quasi-static simulation, which simulates the
steady-state response of magnetic samples to Heff. For quasi-static
simulation, a damping coefficient of a = 0.5 was chosen to obtain
rapid convergence. Second was the dynamic simulation that
simulates the dynamic response and quantifies the spatial
characteristics of the SW mode. Time-dependent simulations
are performed using a = 0.008 for NiFe. A time-varying sinc
sin(2pft )
wave excitation field hsinc = h0 is used to yield uniform
t
excitation in the frequency domain, where h0 is the initial
amplitude of the sinc wave (within the linear excitation regime,
typically up to few tens of oersteds). The dynamic simulation
results are analyzed in the frequency domain by performing fast
Fourier transform (FFT) processing.
(a)
(b)
(c)
Figure 4.3 (a–c) SEM images of disks with s = 500, 200, and 50 nm.
(d) Normalized BLS spectra and (e) the corresponding simulated spectra
for different s and orientations. Plots with filled (open) symbols represent
the BLS spectra in parallel (perpendicular) orientations between
the interdisk coupling direction and Happ. Insets in (d): Schematic
diagram of stray field models for the disk pair in the parallel (left)
and perpendicular (right) orientations. Adapted with permission from
Shimon and Adeyeye [58]. Copyright 2015 WILEY-VCH Verlag GmbH & Co.
KGaA, Weinheim.
state. For each s (except for s = 500 nm), there are two BLS spectra
plotted, representing parallel (filled symbols) and perpendicular
(open symbols) coupling orientations with respect to the Happ
direction. At Happ = 1 kOe, only one prominent mode was observed
in the disk for various s. We named the mode observed for the
isolated disk as mode C1, while those in the coupled disks were
modes C2A and C2B when their coupling orientation with respect
to the Happ direction is parallel and perpendicular, respectively.
For the isolated disk, the resonant frequency of mode C1 (fC1) was
detected at 9.2 GHz. As s is reduced, we observed a systematic
shift of fres with respect to that of the isolated disk. In addition,
we observed opposing trends of this fres shift, depending on the
coupling orientation (two dotted lines in Fig. 4.3d). For parallel
orientation, fC2A increases by 300 MHz as s is reduced from
200 nm to 50 nm. In contrast, for perpendicular orientation,
fC2B decreases by 300 MHz as s is reduced from 200 nm to 50 nm.
To understand the observed fres shift, we consider two stray
field models (see insets in Fig. 4.3d). The first model (right), for
coupling in parallel orientation, illustrates how a dipolar field
along +x (+Hdx, blue arrow) is exerted on the disk under the probe
by its neighbor from the left. Conversely, the second model (left),
for coupling in perpendicular orientation, shows how the
dipolar field along –x (–Hdx, red arrow) is exerted on the disk
under the probe by its neighbor from the top. On the basis of this
model, we analyze how additional Hd influences the fres of the
disk using the modified Kittel resonant formulation as Eq. 4.2:
g
f res = (H x +( N y – N x )4 pMx ) (H x +( N y – N x )4 pMx ) , (4.2)
2p
where Nx, Ny, and Nz are the demagnetization factors for x, y,
and thickness (z) directions, respectively; g is the gyromagnetic
ratio of the materials; and 4pMx and Hx are the magnetization
of the sample and the effective field along x, respectively.
In the parallel orientation, +Hd–x adds to Hx such that
Hx = Happ + Hd–x, and as a result fres will increase as compared
to the isolated case, where Hx = Happ. As a larger Hd–x is expected
with a reduction of s, Dfres will consequently increase. In the
perpendicular orientation, –Hd–x reduces Hx following
Hx = Happ – Hd–x and as a result fres will decrease with s (opposite
to that with parallel orientation). Note that the dipolar field along
Coupled Nanodisks 111
nm
dipolar field in modifying the
s=500 nm
Simulated Mode Profiles at 1 kOe (norm.)
dynamic behavior in a2D ȝ-BLS images of
disk.
s=500
resonant mode at 1 kOe
(e) (j) (k)
(a)
nm
s=500 nm
(f) (l)
(g)
s=500
(b) s = 200 nm
s=200 nm
(f)
(c) (l)
s=100 nm
(m)
(g)
(b)
s=200 nm
(h)
s = 200 nm
(l) (m)
(c) (g) s = 100 nm
s=100 nm
(i) Y
s=50 nm
800 nm
1 μm
(c)
s=100 nm
s = 100 nm
(h) hrf
(d) s = 50 nm
s=50 nm
(i) Y
800 nms=50 nm
s = 100 nm
800 nm
1 μm
(i) s = 50 nm Y
s=50 nm
1 μm
+Happ
1.7 μm 1.5Z μm
s = 50 nm
X
1.7 μm 1.5 μm
Figure 4.5 (a–c) Frequency versus Happ plot for (a) C1, (b) C2A, and
(c) C2B. Symbols are experimentally extracted fres values, lines are fitted
curves based on Eq. 4.2. Adapted with permission from Shimon and
Adeyeye [58]. Copyright (2015) WILEY-VCH Verlag GmbH & Co. KGaA,
Weinheim.
d=500 nm
C5A 9.8GHz 9.38GHz
(b) C2A s
8.98GHz 9.77GHz
(e) C5A 9GHz 9.5GHz
C3
s=50 nm 8.59GHz
C2B 8.9GHz
(c) C2B
9.5GHz 9.77GHz
s
C2A
(f) s C5B
1 9.2GHz 9.18GHz
y
G hrf C1
s
S 0
G
Happ 6 8 10 12 6 8 10 12
x 500 nm Frequency (GHz) Frequency (GHz)
Figure 4.6 (a–f) SEM images of disks with various cluster configurations
(D = 500 nm, s = 50 nm). Green arrows in (a–f) label the disk under the
probe. (g) Measured BLS spectra and corresponding 2D mode profiles
of C1, C2A, C2B, C3, C5A, and C5B with Happ = 1 kOe. (h) Corresponding
simulated spectra and 2D mode profiles. Color scale bar in (g) represents
the normalized intensity of the resonant mode profile. Adapted with
permission from Shimon and Adeyeye [59]. Copyright (2016) American
Institute of Physics.
at fC5A = 9.38 GHz, which comes from the coupled resonant of the
center disk with its four neighboring disks (inset of Fig. 4.6h).
14
(a) C3 (b) Higher f
Frequency (GHz)
12 Hd-x = +99 Oe
10
+Hd-x
-Hd-x 8 Lower f C3
Hd-x = -22 Oe
6
14
(c) C5A (d)
Frequency (GHz)
12
+Hd-x
10 Hd-x = +150 Oe
8
C5A
+Hd-x
6
14
(e) C5B (f)
Higher f
Frequency (GHz)
12 Hd-x = +132 Oe
+Hd-x 10
C5B
8
Lower f
-Hd-x 6
Hd-x = -22 Oe
0.5 1.0 1.5 2.0
Field (kOe)
Figure 4.7 Models of static dipolar interaction for (a) C3, (c) C5A, and
(e) C5B. Frequency versus Happ plot for (b) C3, (d) C2A, and (f) C2B.
Symbols are experimentally extracted fres values, and lines are fitted
curves based on Eq. 4.2. Adapted with permission from Shimon and
Adeyeye [59]. Copyright (2016) American Institute of Physics.
Figure 4.8 (a) Remanence state of P-, Q-, and R-type nanomagnets
after field initialization, HI: 2000 Oe 0 along the x axis. (b) Energy
landscape illustrating shift in energy minima and maxima in P magnets
as compared to R magnets. (c, d) Hysteresis loops in P and R magnets
along geometrical short and long axis, respectively. (e, f) SEM images
of isolated PQ and reference R magnet array. (g, h) MFM contrasts for
isolated PQ and R magnet array after initialization with HI: 2000 Oe 0
along the short axis. Adapted with permission from Haldar and Adeyeye
[74]. Copyright (2016) American Chemical Society.
Reconfigurable Magnetization Dynamics 121
Figure 4.9 (a–c) SEM images of PQ, PQP, and PPQ networks. (d–f) MFM
contrasts for all the samples mentioned before after initialization with
HI: 2000 Oe 0 along the short axis. (g) Easy axis MOKE hysteresis
loops for all the above-mentioned samples. The results are shown for
d = 100 nm for the coupled networks. (h) Simulated magnetic hysteresis
loops for PQ network (d = 100 nm). Simulated spin configurations for the
hard axis loop at field points (i to vi) are shown on the right. Adapted
with permission from Haldar and Adeyeye [74]. Copyright (2016)
American Chemical Society.
Figure 4.11 (a) SEM images for five different types of metamaterials
denoted by S1–S5. One unit cell of these crystals is indicated by a
dotted box. MFM images for S1–S5 MCs in (b) AFM and (c) FM magnetic
ground states. AFM and FM states were obtained by field initialization
(HI: 1000 Oe 0) along the x axis and the y axis, respectively.
(d) MOKE and (e) simulated hysteresis loops for samples S1–S5 when
the field was applied along the y axis, that is, the easy axis of the RNM.
In addition, a simulated hysteresis loop for S2 is shown when the field
was applied along the x axis, that is, the hard axis of the RNM. Adapted
with permission from Haldar and Adeyeye [75]. Copyright (2016)
American Institute of Physics.
Figure 4.13 (a) Experimental and (b) simulated FMR responses at the FM
and FiM ground states. (c) Simulated spatial profiles of the SW modes
for f1 to f5.
4.5 Summary
In the first part of this chapter, we used the µ-BLS spectroscopy
technique to investigate the dynamic response of a single magnetic
disk as a function of the neighboring disks’ separation and
coupling configurations. A simple method has been developed to
130 Magnetization Dynamics of Reconfigurable 2D Magnonic Crystals
Acknowledgments
This work was supported by the National Research Foundation,
Prime Minister’s Office, Singapore, under its Competitive Research
Programme (CRP Award No. NRF-CRP 10-2012-03), SMF-NUS
New Horizon Awards, and Ministry of Education, Singapore, AcRF
Tier 2 (Grant No. R-263-000-A19-112). A.O.A is a member of
the Singapore Spintronics Consortium (SG-SPIN).
References 131
References
1. Shiroishi, Y., Fukuda, K., Tagawa, I., Iwasaki, H., Takenoiri, S., Tanaka, H.,
Mutoh, H., and Yoshikawa, N. (2009). Future options for HDD
storage, IEEE Trans. Magn., 45, 3816–3822.
2. Park, K. S., Park, Y. P., and Park, N. C. (2011). Prospect of recording
technologies for higher storage performance, IEEE Trans. Magn., 47,
539–545.
3. Joel, K. W. Y., Yunjie, C., Tianli, H., Huigao, D., Naganivetha, T.,
Hui Kim, H., Siang Huei, L., and Vivian, N. (2011). Fabrication and
characterization of bit-patterned media beyond 1.5 Tbit/in2,
Nanotechnology, 22, 385301.
4. Kawahara, T., Ito, K., Takemura, R., and Ohno, H. (2012). Spin-transfer
torque RAM technology: review and prospect, Microelectron. Reliab.,
52, 613–627.
5. Khvalkovskiy, A. V., Apalkov, D., Watts, S., Chepulskii, R., Beach, R. S.,
Ong, A., Tang, X., Driskill-Smith, A., Butler, W. H., Visscher, P. B., Lottis,
D., Chen, E., Nikitin, V., and Krounbi, M. (2013). Basic principles
of STT-MRAM cell operation in memory arrays, J. Phys. D: Appl.
Phys., 46, 074001.
6. Cowburn, R. P., and Welland, M. E. (2000). Room temperature
magnetic quantum cellular automata, Science, 287, 1466–1468.
7. Csaba, G., Imre, A., Bernstein, G. H., Porod, W., and Metlushko, V.
(2002). Nanocomputing by field-coupled nanomagnets, IEEE Trans.
Nanotechnol., 1, 209–213.
8. Kostylev, M. P., Serga, A. A., Schneider, T., Leven, B., and Hillebrands, B.
(2005). Spin-wave logical gates, Appl. Phys. Lett., 87, 153501-3.
9. Imre, A., Csaba, G., Ji, L., Orlov, A., Bernstein, G. H., and Porod, W. (2006).
Majority logic gate for magnetic quantum-dot cellular automata,
Science, 311, 205–208.
10. Schneider, T., Serga, A. A., Leven, B., Hillebrands, B., Stamps, R. L., and
Kostylev, M. P. (2008). Realization of spin-wave logic gates, Appl. Phys.
Lett., 92, 022505-3.
11. Ustinov, A. B., and Kalinikos, B. A. (2007). Ferrite-film nonlinear
spin wave interferometer and its application for power-
selective suppression of pulsed microwave signals, Appl. Phys. Lett.,
90, 252510-3.
12. Phuoc, N. N., Xu, F., and Ong, C. K. (2009). Ultrawideband microwave
noise filter: hybrid antiferromagnet/ferromagnet exchange-coupled
multilayers, Appl. Phys. Lett., 94, 092505-3.
132 Magnetization Dynamics of Reconfigurable 2D Magnonic Crystals
50. Verba, R., Melkov, G., Tiberkevich, V., and Slavin, A. (2012). Fast
switching of a ground state of a reconfigurable array of magnetic
nano-dots, Appl. Phys. Lett., 100, 192412.
51. Kittel, C. (1947). Interpretation of anomalous larmor frequencies
in ferromagnetic resonance experiment, Phys. Rev., 71, 270–271.
52. Kittel, C. (1948). On the theory of ferromagnetic resonance
absorption, Phys. Rev., 73, 155–161.
53. Demokritov, S. O., Hillebrands, B., and Slavin, A. N. (2001). Brillouin
light scattering studies of confined spin waves: linear and nonlinear
confinement, Phys. Rep., 348, 441–489.
54. Sandercock, J. R. (2014). Tandem Fabry-Perot interferometer TFP-
1 operator manual, Available at: http://tablestable.com/uploads/
ckeditor/TFP-1/Manual_TFP_1.pdf
55. Demidov, V. E., Demokritov, S. O., Hillebrands, B., Laufenberg, M.,
and Freitas, P. P. (2004). Radiation of spin waves by a single
micrometer-sized magnetic element, Appl. Phys. Lett., 85, 2866–2868.
56. Donahue, M. J., and Porter, D. G. (1999). OOMMF User’s Guide,
Version 1.0. Interagency Report NISTIR 6376, National Institute of
Standards and Technology, Gaithersburg, MD.
57. Gilbert, T. L. (1955). A Lagrangian formulation of the gyromagnetic
equation of the magnetization field. [Abstract only; full report,
Armor Research Foundation Project No. A059, Supplementary
Report, May 1, 1956] (unpublished), Phys. Rev., 100, 1243.
58. Shimon, G., and Adeyeye, A. O. (2015). Direct detection of static
dipolar interaction on a single nanodisk using microfocused
Brillouin light scattering spectroscopy, Adv. Electron. Mater., 1,
1500070.
59. Shimon, G., and Adeyeye, A. O. (2016). Direct observation of
configurational anisotropy in coupled magnetic disk cluster using
micro-focused Brillouin light scattering spectroscopy, Appl. Phys.
Lett., 109, 032407.
60. Pigeau, B., Hahn, C., de Loubens, G., Naletov, V., Klein, O., Mitsuzuka,
K., Lacour, D., Hehn, M., Andrieu, S., and Montaigne, F. (2012).
Measurement of the dynamical dipolar coupling in a pair of magnetic
nanodisks using a ferromagnetic resonance force microscope, Phys.
Rev. Lett., 109, 247602.
61. Dvornik, M., Bondarenko, P. V., Ivanov, B. A., and Kruglyak, V. V. (2011).
Collective magnonic modes of pairs of closely spaced magnetic
nano-elements, J. Appl. Phys., 109, 07B912.
136 Magnetization Dynamics of Reconfigurable 2D Magnonic Crystals
62. Sugimoto, S., Fukuma, Y., Kasai, S., Kimura, T., Barman, A., and Otani,
Y. (2011). Dynamics of coupled vortices in a pair of ferromagnetic
disks, Phys. Rev. Lett., 106, 197203.
63. Au, Y.-Y., and Ingvarsson, S. (2009). Ferromagnetic resonance of
individual magnetic double layer microwires, J. Appl. Phys., 106,
083906-4.
64. Ding, J., Kostylev, M., and Adeyeye, A. O. (2011). Magnetic hysteresis
of dynamic response of one-dimensional magnonic crystals consisting
of homogenous and alternating width nanowires observed with
broadband ferromagnetic resonance, Phys. Rev. B, 84, 054425.
65. Keatley, P. S., Gangmei, P., Dvornik, M., Hicken, R. J., Grollier, J., Ulysse,
C., Childress, J. R., and Katine, J. A. (2013). Bottom up magnonics:
magnetization dynamics of individual nanomagnets, in Magnonics.
Vol. 125, eds. Demokritov, S. O., and Slavin, A. N. (Springer, Berlin,
Heidelberg), pp. 17–28.
66. Shimon, G., and Adeyeye, A. O. (2015). Size-dependent magnetization
dynamics in individual Ni80Fe20 disk using micro-focused
Brillouin light scattering spectroscopy, AIP Adv., 5, 097124.
67. Liu, X. M., Ding, J., Singh, N., Shimon, G., and Adeyeye, A. O. (2014).
Magnetization dynamics of coupled Ni80Fe20 dots: effects of
configurational anisotropy and dipolar coupling, Appl. Phys. Lett., 105,
052413.
68. Gubbiotti, G., Carlotti, G., Zivieri, R., Nizzoli, F., Okuno, T., and Shinjo,
T. (2003). Spin wave modes in submicron cylindrical dots, J. Appl.
Phys., 93, 7607–7609.
69. Gubbiotti, G., Carlotti, G., Okuno, T., Shinjo, T., Nizzoli, F., and Zivieri,
R. (2003). Brillouin light scattering investigation of dynamic spin
modes confined in cylindrical Permalloy dots, Phys. Rev. B, 68,
184409.
70. Gubbiotti, G., Madami, M., Tacchi, S., Carlotti, G., and Okuno, T.
(2006). Normal mode splitting in interacting arrays of cylindrical
Permalloy dots, J. Appl. Phys., 99, 08C701.
71. Giovannini, L., Montoncello, F., Nizzoli, F., Gubbiotti, G., Carlotti, G.,
Okuno, T., Shinjo, T., and Grimsditch, M. (2004). Spin excitations
of nanometric cylindrical dots in vortex and saturated magnetic
states, Phys. Rev. B, 70, 1–4.
72. Shaw, J., Silva, T., Schneider, M., and McMichael, R. (2009). Spin
dynamics and mode structure in nanomagnet arrays: effects of size
and thickness on linewidth and damping, Phys. Rev. B, 79, 184404.
References 137
73. Keatley, P., Kruglyak, V., Neudert, A., Galaktionov, E., Hicken, R.,
Childress, J., and Katine, J. (2008). Time-resolved investigation of
magnetization dynamics of arrays of nonellipsoidal nanomagnets
with nonuniform ground states, Phys. Rev. B, 78, 214412.
74. Haldar, A., and Adeyeye, A. O. (2016). Deterministic control of
magnetization dynamics in reconfigurable nanomagnetic networks
for logic applications, ACS Nano, 10, 1690–1698.
75. Haldar, A., and Adeyeye, A. O. (2016). Artificial metamaterials for
reprogrammable magnetic and microwave properties, Appl. Phys.
Lett., 108, 022405.
76. Haldar, A., and Adeyeye, A. O. (2015). Vortex chirality control in
circular disks using dipole-coupled nanomagnets, Appl. Phys. Lett.,
106, 032404.
77. Ding, J., Kostylev, M., and Adeyeye, A. O. (2012). Realization of a
mesoscopic reprogrammable magnetic logic based on a nanoscale
reconfigurable magnonic crystal, Appl. Phys. Lett., 100, 073114.
78. Alexander, K., Mingqiang, B., and Kang, L. W. (2010). Magnonic logic
circuits, J. Phys. D: Appl. Phys., 43, 264005.
Chapter 5
5.1 Introduction
There are different types of excitations in solid state resulting in
propagation of mechanical waves (phonons) [1], electromagnetic
waves (photons) [2], free electron gas density oscillations
(plasmons) [3] or collective oscillations of the magnetization,
and spin waves (SWs) (magnons) [4, 5], which can be utilized for
practical applications. The wavelength of SWs is a few ranges of
amplitude smaller than that of electromagnetic waves for the
same frequency. Moreover, they are easy tunable by a magnetic
field [6]. This makes SWs interesting for application, especially
for microwave and electronic devices [7, 8].
Monocrystalline yttrium iron garnet, Y3Fe5O12 (YIG), mainly
in the form of micrometer-thick films, has been the leading
material for SW studies since the 1960s [9–11]. YIG has extremely
low damping, enabling SW propagation over a centimeter range.
Recent fabrication of YIG films with nanometer thickness [12–15]
has opened new frontiers for magnetic studies and made new
applications possible because of both the discovery of new effects,
such as current–spin wave conversion, or the spin-Seebeck
effect [16–20], and possibilities for high-quality patterning in
the submicrometer range, enabling research on SWs of short
wavelengths [21–23].
The rapid progress in studies of SW dynamics observed
recently was possible not only because of development of
fabrication technology but also because of the progress in
measurement techniques. New SW visualization spectroscopic
techniques such as Brillouin light scattering (BLS) allow the
investigation of phenomena related to SWs (such as diffraction,
focusing, reflection or Goos–Hänchen shift), being similar to the
effects known in classical optics [24, 25]. In this chapter, we focus
on the optical effects of SWs in micrometer YIG films, investigated
experimentally with the aid of the BLS technique and simulated
numerically, and in nanometer films, investigated theoretically
Introduction 141
dM a dM(r , t ) , (5.1)
= –| g | m0M × Heff + M(r , t )×
dt MS dt
gm0
f= 2 2
[( Heff,0 + MS lex 2 2
k )( Heff,0 + MS lex k + MS F ( j, J ))]1/2 . (5.3)
2p
1 – e – kt .
P =1–
kt
Figure 5.1 (a) Geometry of the YIG film. (b) Dispersion relation for
SWs propagating in two perpendicular in-plane directions (perpendicular
(DE) and parallel (BVMW) to the direction of the H) for thick and
thin YIG film. The inset at left side shows the relation with enlarged
small k region. (c, d) Surface plots of the SWs’ dispersion in YIG film
of 4.5 μm and 5 nm thickness, respectively. The in-plane magnetic field
was assumed to be 1000 Oe.
Figure 5.2 Sketch of the sample (with definitions of angles j and q),
overlaid by an exemplary BLS intensity 2D map representing SW
interaction with a single antidot.
146 Spin Wave Optics in Patterned Garnet
This is due to the fact that for SWs the direction of the group
velocity defines energy propagation, generally as not parallel to
the wavevector k. Rotation of the antenna and thus the change in
the direction of wavevector k does not influence the vectors of
group velocity, because the caustic beam forms only by linear
regions of the iso-frequency dependence.
The problem of SWs’ self-focusing was studied in some
papers [35, 36]. It was theoretically predicted that for the
observation of self-focusing caustic SWs, it is necessary to have
some kind of SW point source and linear regions of the
iso-frequency curve [35]. On the basis of this idea, so-called
SW active sources of the formation of caustic SWs have been
observed [34, 36] driving SWs a sample from a waveguide region
to the semi-plane one. SW self-focusing based on antidot geometry
as a “passive source” seems to be simpler than one based on
an “active source.”
Figure 5.5 Schematic view of the sample, with gray scale of BLS intensity
I(x, y) of SWs traveling along the line of antidots at the critical angle Fcrit
for total nonreflection, and with definitions of angle F.
Figure 5.6 SW interaction with the square antidots line inclined by the
angle F = 0° and 43° from antenna for (a, c) and (b, d), respectively.
The static magnetic field, H = 1020 Oe, was applied along the y axis.
The gray-scale SW intensity distributions measured by BLS (a, b). SWs
were excited at the sample on the right side and then were propagated
along the x direction. SWs incidence kin and reflection kout vectors
constructions with iso-frequency lines for F = 0° (c) and 43° (d) antidots
line (marked by dashed lines) orientations. Dotted lines are the projection
of the k vectors on the line of antidots.
Spin Waves in Patterned YIG Micrometer Films 151
Figure 5.8 Modeling of SW interaction with the antidots line. (a) The
demagnetizing field in the YIG film (4.5 μm thickness) near the square
antidots row (of 70 μm size). The gray scale shows the demagnetizing
field component along the direction of the H oriented at angle 45° with
respect to the antidots line. (b) Iso-frequency dispersion relation lines
(obtained from Eq. 5.3) at 4.7 GHz in the homogeneous YIG film for
various internal magnetic field magnitudes in the range 950–980 Oe.
The IFDRLs for 980 Oe and 966 Oe are emphasized with the solid and
dash-dotted lines, respectively. The direction of antidots line is marked
as e||. The direction of the refracted SW group velocity is marked with
vg,out.
Figure 5.8b shows IFDRLs plotted over the (kx, ky) plane
obtained from Eq. 5.3 for a frequency of 4.7 GHz with H equal
950, 966, and 980 Oe. The incident SWs are refracted at the
interface between the regions of lower and higher values of
the refractive index (higher and lower H, respectively) before
reaching the edge of the antidots line. The condition kin,|| = kout,||
(index || means the component of the wavevector which is
tangential to the interface) provides the wavevector kout,|| of
the refracted SWs and the direction of the group velocity vg at
the terminal point of kout, which is parallel to the direction of
antidote line, e||. Indeed, this was observed in experiment in
Fig. 5.6b.
Moreover, the direction of energy transfer is not very sensitive
to the assumed decrease in the internal magnetic field, as
indicated by the IFDRLs plotted for various field magnitudes H.
Another interesting conclusion from Fig. 5.8b is that |kout| > |kin|,
i.e., the wavelength of the refracted SWs is significantly shorter
than that of the incident SWs (at 966 Oe lout = 2p/kout = 65 μm).
This decrease is fully confirmed in MS in Fig. 5.9a. In Fig. 5.9,
the results of MS performed by means of Mumax3 [44] are
presented. They show the out-of-plane component of the
magnetization mz or averaged in time square mz (<(mz(t))2>T = 1/f)
proportional to the BLS intensity in a steady state (see Fig. 5.5a or
Fig. 5.5b–d, respectively. Our simulations well describe creation
of strong SWs beam experimentally observed (see Fig. 5.6b).
Furthermore, we simulated SW scattering with long rectangular
hole. The result of these simulations is shown in Fig. 5.9c, and it
clearly demonstrates almost the same features as we found in
Fig. 5.9b for the antiodots line. Thus, we conclude that the
crucial role in the experimentally observed effects is played by
the inhomogeneity of the internal magnetic field, rather than
the edge of the ferromagnetic film itself. Nevertheless, the edge
of the film is the source of the demagnetizing field there.
We have also checked numerically the influence of increasing
SW frequency on the refraction and the beam formation.
Exemplary result of the simulations at frequency 4.98 GHz is
shown in Fig. 5.9d. We clearly see the broadening of the SW
beam with respect to the beam of lower frequencies (compare
with Fig. 5.9c for 4.7 GHz) and the rotation of the beam
propagation direction. This agrees well with the IFDRL analysis
Spin Waves in Patterned YIG Micrometer Films 155
Figure 5.9 Results of MSs showing SWs’ interaction with (a, b) antidots
of side 70 μm line with period a = 100 μm at 4.7 GHz and (c, d), long
rectangular hole (70 μm width) at 4.7 GHz and 4.98 GHz, respectively.
Note that in (a) there is SW amplitude and in (b–d) SW intensity
presented. The SWs are excited by microstrip antenna located on the
right side of the antidots line (rectangular hole) and propagate on the
left side of the antenna. The static magnetic field 980 Oe for (a–d) was
oriented along the y axis.
156 Spin Wave Optics in Patterned Garnet
1940s [46], and then for acoustic waves [47], electrons [48], and
neutron waves [49]. Recently, this phenomenon was investigated
theoretically also for the exchange or dipole-exchange SWs
[42, 50, 51]. The magnetic properties at the film edge were shown
to be crucial for a lateral shift of the SW beam.
Figure 5.11 Schematic plot of the thin YIG film geometry for consideration
of the GHS. The film has thickness t, which is much smaller than the
film’s lateral sizes, Lx and Ly. The (x,y,z) coordinating system defines
the structure with the film edge at y = 0 (hatched area). kin and kout are
wavevectors of incident and reflected SWs beams, respectively. DX is a
total shift of the SW beam reflected at the edge.
arg(R ) 2p tan q in
, (5.4)
DX = = 2
kx p +(k cos q in )2
Figure 5.12 (a) The analytical results of the GHS in the reflection of the
SW beam from the edge of the thin YIG film calculated using Eq. 5.4 in
dependence on the magnetic surface anisotropy constant KS. The gray
square corresponds to the value of GHS for KS = 0. (b) The GHS in
dependence on the angle of incidence obtained from Eq. 5.4 (solid lines)
and MSs (dashed lines with symbols) for KS = 0 for YIG (solid) and Py
(dashed line). (c) Internal magnetic field for YIG (solid line) and Py
(dashed line).
Figure 5.13 (a) The refraction of the wave on the interface between
media with low refractive index (n-th slice, solid black line) and high
refractive index (n + 1 slice, gray dashed line) based on IFDRL analysis.
The conservation of the kx components in the refraction is required by
the translational symmetry along the x axis. (b) Example of two
refractions on the interfaces between n-th and n + 1, and between n + 1
and n + 2 slice corresponding to the different values of Heff,0, as it is
presented in the inset on the right side of (b). The beam shift
DXbending resulting from the bending is marked schematically.
(c) Results of the MSs (gray square points) obtained from the analytical
models presenting dependence of the SW beam shift on the surface
magnetic anisotropy constant in the YIG film in the external magnetic
field of 7 kOe. Solid black line corresponds to the basic analytical model
for GHS (Eq. 5.4) without including SW bending. Dashed gray line
presents results for the analytical model taking into account SW
bending (see details in [42]).
Optics of Spin Waves in Nanometer-Thick YIG Film 161
Figure 5.14 (a) The structure of the magnonic ADL based on the YIG film
of thickness 12 nm. The cylindrical antidots of the diameter 84.6 nm
are arranged in a square lattice (lattice constant 150 nm). (b) Dispersion
relation (band structure) in the first Brillouin zone for the structure
shown in (a) with the external magnetic field applied along the y direction
calculated using plane wave method.
of SWs incoming from HF and entering into ADL for two different
orientations of the magnetic field: perpendicular and parallel
to the HF/ADL interface. For the first case (see Fig. 5.15a),
we found the superprism effect, where a small change in the
incidence angle results in a large change in the refraction angle.
The system accepts the transmission at the direction normal
to the interface because both ADL and HF have magnonic
eigenstates with the tangential component of the wavevector
equal to zero at frequency 7.67 GHz. As a result, the superprism
effect leads to total internal reflection for the incidence angle
exceeding a small critical value. The system in this configuration
works as a low-pass angular filter transmitting through the
HF/ALD interface only those SWs that come in at the incidence
angle lower than some threshold value (we marked this range in
Fig. 5.15 by a bright grey circular sector). The structure considered
can also work in the other operational mode, high-pass angular
filtering. We can switch between these modes by rotating the
external field (or the structure) by 90°. In the mode presented in
Fig. 5.15b, the field is normal to the HF/ALD interface and the
system does not transmit, through the interface, the SWs of
incidence angles lower than a specific critical value.
Figure 5.15 The IFDRLs for ADL (black contours) and homogeneous YIG
film (gray contour) for frequency 7.67 GHz located close to the bottom of
the first magnonic band in ADL in Fig. 5.14. The arrows show the
directions and the magnitudes (in arbitrary units) of the group velocity
(arrows normal to IFDRLs) and phase velocity (arrows attached at the
center of Brillouin zone). The gray (black) arrows refer to the group
and phase velocities in HF (ADL). The dotted squares mark the edges of
the first Brillouin zone. (a) Low-pass angular filtering and (b) high-
pass angular filtering. ADL and HF are made of YIG and have structural
and material parameters as described in Fig. 5.14.
164 Spin Wave Optics in Patterned Garnet
Figure 5.16 All-angle collimation effect. The IFDRLs in ADL and HF for
frequency 9.1 GHz (the bottom of second band in ADL). Due to the flat
IFDRLs in ADL, the divergent beam coming from HF is collimated in
ADL into the beam propagating in the direction normal to the ADL–HF
interface. The meaning of the lines and arrows is the same as in Fig. 5.14.
Summary 165
5.4 Summary
In summary, we presented different optic-like effects of SWs
in patterned thin magnetic film YIG samples. Brillouin light
scattering technique was used for imagining SW interaction with
the single antidot and the line of antidots in YIG micrometer-thick
films. The SW wavelength was driven by the excitation frequency
or the external magnetic field amplitude. Discussion of the
results of SW interaction with single antidot studies was focused
on caustic SW beam diffraction, like SW reflection and interference.
Especially interesting was the SW total nonreflection effect
connected with the creation of a highly focused beam of SWs
propagating along a line of antidots and behind this line. We
showed that a physical phenomenon to obtain highly focused
beams originates from the refraction of SWs in an inhomogeneous
internal magnetic field (connected with the demagnetization
field) near the edge of the antidot line. In the theoretical analysis
of SWs in nanometer-thick magnetic films, we showed that a
Goos–Hänchen shift depends on both the local surface magnetic
anisotropy at the film edge and the variation of the internal
magnetic field. In patterned 2D antidots lattice nanometer
films, in specific ranges of frequency, we reported the effects of
low-pass and high-pass angular filtering and of all-angle
collimation.
The SW total nonreflection effect was used for the creation
of strong SW beams at the 180° switcher. The already available
technology of fabricating low-damping, very thin, nanometer-
thickness films and the various optic-like effects of SWs reported
have a promising potential for desirable miniaturization of SW
devices.
Acknowledgments
This work was supported by the National Science Centre, Poland,
through OPUS grant DEC-2013/09/B/ST3/02669 and Sonata-bis
grant UMO-2012/07/E/ST3/00538. The authors would like to
thank V. D. Bessonov and H. Ulrichs for their contributions to spin
wave studies in YIG micrometer-thick films.
166 Spin Wave Optics in Patterned Garnet
References
30. Gieniusz, R., Ulrichs, H., Bessonov, V. D., Guzowska, U., Stognii, A. I.,
and Maziewski, A. (2013). Single antidot as a passive way to create
caustic spin-wave beams in yttrium iron garnet films, Appl. Phys. Lett.,
102, 102409.
31. Gieniusz, R., Bessonov, V. D., Guzowska, U., Stognii, A. I., and Maziewski,
A. (2014). An antidot array as an edge for total non-reflection of
spin waves in yttrium iron garnet films, Appl. Phys. Lett., 104, 082412.
32. Demokritov, S. O., Hillebrands, B., and Slavin, A. N. (2001). Brillouin
light scattering studies of confined spin waves: linear and nonlinear
confinement, Phys. Rep., 348, 441–489.
33. Demidov, V. E., Demokritov, S. O., Hillebrands, B., Laufenberg, M., and
Freitas, P. P. (2004). Radiation of spin waves by a single micrometer-
sized magnetic element, Appl. Phys. Lett., 85, 2866.
34. Schneider, T., Serga, A. A., Chumak, A. V., Sandweg, C. W., Trudel,
S., Wolff, S., Kostylev, M. P., Tiberkevich, V. S., Slavin, A. N., and
Hillebrands, B. (2010). Nondiffractive subwavelength wave beams
in a medium with externally controlled anisotropy, Phys. Rev. Lett.,
104, 197203.
35. Veerakumar, V., and Camley, R. E. (2006). Magnon focusing in thin
ferromagnetic films, Phys. Rev. B, 74, 214401.
36. Demidov, V. E., Demokritov, S. O., Birt, D., O’Gorman, B., Tsoi, M., and
Li, X. (2009). Radiation of spin waves from the open end of a
microscopic magnetic-film waveguide, Phys. Rev. B, 80, 014429.
37. Mansfeld, S., Topp, J., Martens, K., Toedt, J. N., Hansen, W., Heitmann,
D., and Mendach, S. (2012). Spin wave diffraction and perfect imaging
of a grating, Phys. Rev. Lett., 108, 047204.
38. Gubbiotti, G., Conti, M., Carlotti, G., Candeloro, P., Fabrizio, E.
D., Guslienko, K. Y., Andre, A. Bayer, C., and Slavin, A. N. (2004).
Magnetic field dependence of quantized and localized spin wave
modes in thin rectangular magnetic dots, J. Phys.: Condens. Matter,
16, 7709.
39. Born, M., and Wolf, E. (1998). Principles of Optics, 7th ed. (Cambridge
University Press, Cambridge, UK).
40. Lock, E. H. (2008). The properties of isofrequency dependences
and the laws of geometrical optics, Phys. Usp., 51, 375–393.
41. Joseph, R. J., and Schlomann, E. (1965). Demagnetizing field in
nonellipsoidal bodies, J. Appl. Phys., 36, 1579.
42. Gruszecki, P., Dadoenkova, Yu. S., Dadoenkova, N. N., Lyubchanskii,
I. L., Romero-Vivas, J., Guslienko, K. Y., and Krawczyk, M. (2015).
References 169
6.1 Introduction
Investigations of magnetic micro- and nanostructures as potential
candidates for applications in spintronics and for magnetic logic
devices have become a hot topic recently [1–7]. In particular,
intensive studies of various magnetic structures are performed to
understand properties of perspective materials for above-mentioned
applications. This, in turn, requires investigation of important
physical phenomena related to spin wave dynamics in magnetic
materials and especially in micro- and nanostructured magnetic
( )
m0t ∂2x + ∂2y Y t + ∂2zY t = 0 (6.3)
for the magnetostatic potential inside the ferromagnetic matrix and
inclusions.
As the next step of consideration we suppose that the matrix
surfaces z = 0 and z = d as well as the inclusion end parts z = 0 and
z = d (see Fig. 6.1) are metallized, following the known studies of a
cylindrical ferromagnetic resonator [31]. In this case one can put Yt
(x, y, z) = Yt(x, y) cos(kzz), where kz = np/d (n = 0, 1, 2,…) and
(∂2
x )
+ ∂2y Y t + krt 2Y t = 0. (6.4)
In the 2D Walker equation (Eq. 6.4) the components krt of
the wave vector along the (x, y) plane inside the matrix and the
inclusion are defined by krt = -1 / m0t k z . The desired form Yt(x, y,
z) of the magnetostatic potential satisfies the boundary conditions
bzt ( z = 0, d ) = 0 on the matrix surfaces and the end parts of an inclusion.
The first equation (Eq. 6.2) gives brt ( x , y , z brt (r , )cosk and
t t
hj ( x , y , z ) = hj (r ,j )cos k z z , where r and j are the local cylindrical
coordinates of a cylindrical inclusion (see Fig. 6.3). Solving the
2D Walker equation (Eq. 6.4) with boundary conditions, which
means the continuity of the magnetic induction normal component
and the magnetic field tangential component on each inclusion
lateral surface, and substituting the result into Yt(x, y, z) gives in
magnetostatic approximation the magnetic induction and magnetic
field inside the matrix with inclusions.
176 Spin Waves in Circular and Linear Chains of Discrete Magnetic Elements
Figure 6.3 Sketch showing two ferromagnetic inclusions with their local
coordinate systems.
ÂA
imf j
Y inc (r j ,f j ) = 0
jm J m ( kr r j )e . (6.5)
m=-•
Here rj, fj are the local cylindrical coordinates of an observation
point and Jm(u) is the Bessel function, with m being a cylindrical
multipole index. The coefficients Ajm = ejim, where the phase factor
e j = exp(ik 0r ◊ R j ) is dependent on the position of the j th cylinder
center relative to the laboratory system. During the scattering
process scattered Yscat and transmitted Ytrans spin wave fields
appear around inclusions and inside them, respectively. They are
Multiple-Scattering Method 177
ÂB
scat 0 i mf j
Y (r j , f j ) = jm Hm ( kr r j )e , (6.6)
m=-•
•
ÂX
trans 0 imf j
and Y (r j , f j ) = jm J m ( kr r j )e , (6.7)
m=-•
where the angles q1 and q2 are defined in Fig. 6.3. We denote by Rjj¢
= Rj – Rj¢ the vector distance between the j th and the j¢ th inclusion
scat
centers, with Rjj¢ = |Rjj¢|. Bearing in mind that Y21 (r2 ,j2 ) is given
actually by Eq. 6.6, the application of the addition theorem in Eq. 6.9
provides
•
scat
Y21 (r2 ,f2 ) = ÂB
m=-•
0
21m J m ( kr r1 )e
imf1
, (6.10)
•
B21m = ÂB 2l e
i arg R12
(
Hl-m k r0R12 , ) (6.11)
l =-•
and the angle b again defined in Fig. 6.3. arg R12 = p + b denotes
the angle between the vector R12 and the unit vector x . Equation
6.11 can be called the invariant form of the addition theorem for
cylindrical wave functions and enables one to write out the total field
around the second inclusion immediately by simply interchanging
the indices 1 ´ 2 in Eqs. 6.8, 6.10, and 6.11.
178 Spin Waves in Circular and Linear Chains of Discrete Magnetic Elements
( )
' i( l -m )arg R
Gmjj-l = e jj'
Hl-m k r0R jj’ , (6.13)
where arg Rjj¢ denotes the angle between the vector Rjj¢ and the unit
vector x̂ .
The ratio Tm( j ) = Bm
( j)
/ A jm that we meet in the RHS of Eq. 6.12
defines the scattering matrix Tm( j ) of the j th single inclusion, which
is dependent on the inclusion parameters only. In particular, the
scattering matrix of the first single inclusion is obtained from the
algebraic Eq. 6.12 and coincides with the expression found in Ref.
[22].
The derived set of Eq. 6.12 for multiple scattering of FVMSWs
provides a transparent physical interpretation that can be achieved
by an iterative solution of the set of equations. In this case the
incident spin wave is singly scattered by the jth inclusion and then
propagates along the ferromagnetic matrix to the jth inclusion to be
singly-scattered by it, and so on. The spin wave propagation along
the ferromagnetic matrix between the two inclusions is described
via the matrix kernel Eq. 6.12, the first factor of which takes into
account the wave phase shift; the second factor is a component of
the Green function for the 2D Walker equation (Eq. 6.4).
Let us note that Eq. 6.13 defines a difference kernel that makes
it favorable to apply the Bloch substitution for Eq. 6.12 in the form
Multiple-Scattering Method 179
As one can see, the RHS of Eq. 6.14 defined by Eqs. 6.15 and
( j)
6.16 describes the incident spin wave in the first term Bm and
contributions of all other multipoles m ± dm created at scattering.
1
N
U (jk )
B jm = Â
N k =1
B | U ( j ) (m) ,
lk
(6.20)
jj'
B j 'm (6.21)
j π j ' =1
d
( Y1 , Y 2 ) = ÚÚ Ú
dx dy dz Y1* ( x , y , z )Y2 ( x , y , z ).
0
(6.24)
Ú
Y( r ) = Y0( r ) + G0(r , r ¢ )U(r ¢ )Y(r ¢ )dr ¢ , (6.27)
where the inhomogeneous term on the RHS is the magnetic potential
of the incident spin wave
Ú
Y0( r ) = G0(r , r ¢ )j(r ¢ )dr ¢. (6.28)
The T-scattering operator T(r¢, r¢¢) for magnetostatic spin waves
is introduced by writing the solution to the integral Eq. 6.27 in the
form
Radiation Losses 183
Ú
Y( r ) = Y0( r ) + G0(r , r ¢ )T ( r ¢ , r ¢¢ )Y0(r ¢¢ )dr ¢¢ (6.29)
and satisfies the integral Lippmann–Schwinger (LS) equation
Ú
T (r , r ¢ ) = U( r )d (r - r ¢ ) + U( r ) G0(r , r ¢¢ )T ( r ¢¢ , r ¢ )dr ¢¢. (6.30)
The operator T(r¢, r¢¢) depends on both of its arguments and has
non-zero values only inside the inclusions. On the basis of Eqs. 6.27
and 6.29 it is useful to introduce a quantity
Ú
P( r ) = U( r )Y( r ) = T (r , r ¢ )Y(r ¢ )d r ¢ , (6.31)
which has a physical meaning of dynamic magnetization displacement
current excited inside inclusions by spin wave scattering. In terms
of this current the magnetic field potential scattered by inclusions
Ysc(r¢) is written from Eqs. 6.27 and 6.29 as
Ú
Y sc ( r ) = G0(r , r ¢ )P(r ¢ )dr ¢. (6.32)
Because of the scattering potential U hermiticity one can derive
from the LS Eq. 6.30 the optical theorem for the T-scattering operator
in the form
T (r , r ¢ ) - T * ( r ¢ , r ) =
(6.33)
Údr ¢¢ Ú ÈÎG (r ¢¢ , r ¢¢¢) - G (r ¢¢¢ , r ¢¢)˘˚ ¥ T (r ¢¢ , r )T(r ¢¢¢ , r ¢)dr ¢¢¢.
* *
0 0
1 (1) 0
Gn(0)( ) =
H (krn ). (6.36)
4i 0
(1)
On the RHS of this equation Hm (u) denotes the Hankel function
of the first kind and order m = 0 and k 0rn = -1 / m00 (p n / d ) is a
component of the spin wave wave vector along the (x, y) plane inside
the homogeneous matrix for the n th transverse mode. Equation
6.32 for the magnetic field potential Ysc(r) scattered by inclusions
in terms of 2D quantities takes the form
Ú
Y scn ( ) = Gn(0)( - ¢ )Pn ( ¢ )d ¢ , (6.37)
Ú
Pn ( ) = Tn ( , )Y0n ( )d ¢. (6.38)
On the RHS of Eq. 6.38 Tn(, ¢¢) denotes the 2D scattering
operator that is connected to the 3D scattering operator T(r¢, r¢¢)
by transformation similar to one in Eq. 6.35. The quantity Y0n(¢)
appears in expansion of the magnetic potential of the incident spin
wave in Eq. 6.28 along transverse eigenmodes. The application of this
expansion along transverse eigenmodes. The Fourier transformation
Tn(k, ¢) and a denotion of the 2D unit vector s to the optical theorem
Eq. 6.33 give
1 2
Im È Y0* n ( )Pn ( )d ˘ = -
Ú Ú
0
Pn (krn s ) d s. (6.39)
ÎÍ ˚˙ 8p
2p
The obtained relation is a basic optical theorem for magnetostatic
spin waves under consideration. It is useful to rewrite this relation
in the form
Cext = Csc (6.40)
where Cext and Csc are cross sections of extinction and scattering,
respectively, for spin wave scattering by inclusions. These values are
defined by the relations
1
Cext = - Im È Y0* n ( )Pn ( )d ˘ ,
Ú (6.41)
krn ÎÍ
0 ˚˙
1 2
Ú P (k
0
and C sc = 0 n rn s ) d s. (6.42)
8p krn
Radiation Losses 185
Ú Ú
+ d ¢¢ d ¢¢¢ Tn(0)( - R j , ¢¢ - R j ) ¥ Gn(0)( ¢¢ - ¢¢¢ )
(6.46)
N
¥ ÂT (j ¢ )
n ( ¢¢¢ , ¢ ).
j π j' =1
its first argument inside the j th inclusion. According to Eq. 6.38 the
2D current Pn() in the case of N inclusions takes the form of the sum
N
Pn ( ) = ÂP
j =1
( j)
n ( ),
(6.47)
Ú
Pn( j )( ) = Tn( j )( , ¢ )Y0n ( ¢ )d ¢ ,
Ú
j)
Y(scn ( ) = Gn(0)( - ¢ )Pn( j )( ¢ )d ¢.
Now the general optical theorem in Eqs. 6.40–6.42 is taken in
the case of incident plane spin wave scattering by an ensemble of
inclusions in the form similar to a single inclusion case in Eq. 6.44,
with
N
Ts ( s ) = ÂT
j =1
( j)
( s ). (6.49)
ip
- 2 0
Y(j)
scn ( ) ªe 4
0
eikrnrT ( j )(s). (6.52)
Æ• p krn r
With the general expression for scattered magnetostatic
potential, Eq. 6.6, we can show the extinction cross section in one-
multipole approximation using the partial scattered amplitude from
Eq. 6.22:
1
Cext = - 0
Re(Tm Aˆ 1m Bˆ 1m ). (6.53)
krn
Defining the extinction cross section of a single inclusion as
0
C(1)ext = -(4 / krn )ReTm and putting Aˆ 1m ª 1 , we present the ratio
Cext / C(1)ext in the form
Cext ImTm
= ReFN - ImFN , (6.54)
C(1)ext ReTm
where the collective extinction factor FN is defined by
FN = Bˆ 1m . (6.55)
The major property of Eq. 6.50 is that the extinction cross section
of the linear chain in the case of an incident narrow spin wave beam
irradiating only the first inclusion formally coincides with the
extinction cross section in Eq. 6.44 for a single inclusion. Thus only
the irradiated inclusion makes a direct contribution in the collective
extinction cross section despite the fact that the total number of
inclusions in the linear chain that makes the direct summarized
contribution of all other inclusions in spin wave scattering almost
invisible; we call this dark mode.
6.4 Results
As is known, excitation of edge spin waves in different magnetic
structures is connected with reciprocity breaking at the spin
wave scattering by non-uniformity in the antisymmetric magnetic
susceptibility tensor. We give a simple interpretation of this
phenomenon as the effect of spatial phase modulation in the spin
wave scattered field, with the appearance of a helical component in
the scattered field.
188 Spin Waves in Circular and Linear Chains of Discrete Magnetic Elements
Figure 6.4 Helical lines in the distribution of the real part of the magnetic
field sum potentials scattered by two inclusions: the collective wave
parameter kr0R12 = 1.2 for case (a) and kr0R12 = 4.3 for case (b).
Figure 6.5 (a) Dependence of Re(1 / l4( -1) ) on the normalized frequency
W at four almost closely packed cylindrical inclusions with parameter R12 =
6.2 mm (the deeper and narrower peak) and R12 = 6.6 mm (the less deep and
wider peak), and inclusion radius R = 1.8 mm . (b) Dependence of quality
factor Q of peaks on the distance R12.
satisfy conditions q¢¢Æ 0 and q¢Æ 0 in both Eqs. 6.57 and 6.58.
Therefore, Eq. 6.22 leads to the limiting formula
Ê j ˆ Ê 1 ˆ
Bˆ j Æ ( -1) j -1 2Á 1 - ˜ , FN Æ 2Á 1 - . (6.59)
Ë N +1 ¯ Ë N + 1 ˜¯
On the other hand, if the condition q¢¢π 0 or even q¢¢Æ • is
satisfied, then Eq. 6.22 gives
Bˆ j Æ ( -1) j -1 exp[-( j - 1)q ¢¢ ] , FN Æ 1. (6.60)
The limiting formulas of Eq. 6.59 describe the case of the distant
resonant transfer of spin wave excitation along the linear chain of
coupled inclusions at resonant value of the coupling parameter a12
Æ –0.5, which shows a linear dependence of excitation decrease on
the number of particles. The collective extinction factor according to
Eq. 6.59 is equal to approximately FN ª 2 for N ? 1; the fact that FN π
1 shows some indirect effect of the particles influence with numbers
j > 1 on the collective extinction factor via influence on the self-
consistent scattering amplitude B 1 in the first inclusion. Equation
6.59 describes a short transfer of spin wave excitation along the
chain, with exponential decrease of excitation, and the collective
extinction factor has a more physically understandable value FN Æ 1.
Another calculation result is that the case of distant transfer of
the spin wave is obtainable with resonant conditions 2a¢12 < -1 and
q¢¢Æ 0 (Eq. 6.57). Here the distribution of the scattered amplitude
̂
B j over inclusions will not be described by Eq. 6.59 or 6.60 but will
be described by the general expression Eq. 6.22.
In Fig. 6.6 we present dependencies of normalized resonance
frequency Wres = wres/wH on geometrical parameters R and R12 (Fig.
6.2) of the linear chain when resonance conditions in Eqs. 6.58 and
6.59 for the coupling parameter are satisfied. As usually, wH denotes
the ferromagnetic resonance frequency of the ferromagnetic film.
All curves in Fig. 6.6 are obtained by numerical solution of the
dispersion equation (Eq. 6.22) under the additional condition
q¢¢Æ 0 in accordance with the first Eq. 6.58. All calculations were
performed for the following material parameters: external magnetic
field Hext = 5 kOe, saturation magnetization of the film and inclusions
in ferromagnetic materials M0s = 1620 Oe and M1s = 1740 Oe , and
film thickness d = 10 mm.
Results 191
The left boundary linear curve in Fig. 6.6 is related to the case
Eq. 6.59. Parameters presented by this curve can be approximately
described by the relation (Wres – 1)R12/R = const. Other data points
in Fig. 6.6 outside the curve, which represent data for linear decay
of scattering amplitudes, depict the case of distant transfer with
conditions from Eqs. 6.56 and 6.58. This case is represented by
the Fig. 6.7, where scattering amplitudes and collective extinction
factors are described by general equations (Eqs. 6.22 and 6.55).
̂
Figure 6.7 Illustrations of the behavior of scattering amplitudes B j of
inclusions, depending on the inclusion number j for the case N = 24 (a),
N = 23 (b), and collective extinction factor FN (c) of a linear chain under
condition of dark-mode filtering from radiation losses.
6.5 Conclusion
We presented the quantum-mechanical-type T-scattering operator
approach to study FVMSW multiple scattering by a finite ensemble
of cylindrical magnetic inclusions in a ferromagnetic film (matrix)
metallized from both sides. The substantial result consisted of
deriving an optical theorem for the T-scattering operator and as a
consequence deriving a new formula for the collective extinction cross
section of the inclusion ensemble in terms of the incident spin wave
influence on dynamic magnetization displacement currents excited
inside inclusions by spin wave scattering. An analogy to the known
quantum mechanics scattering theory, the Watson composition rule
for the T-scattering operators of particles is formulated in the case
of the magnetic inclusion ensemble. This composition rule was used
to derive the equation set for partial spin wave multiple-scattering
amplitudes. The general results of the T-scattering operator
approach were applied to study the Bloch-like spin wave eigenmodes
propagation along the inclusion circular array and the distant spin
wave excitation transfer along the finite linear array of inclusions.
By studying the Bloch-like spin wave eigenmodes we described
high-quality-factor-specific micro-spin wave resonator consisting
of four magnetic inclusions placed periodically along the circuit. In
connection with distant spin wave excitation transfer along a finite
linear array of inclusions, we concretized our new formula for the
collective extinction cross section of the inclusion ensemble in the
References 193
case of only the first inclusion of the chain being irradiated by the
incident narrow spin wave beam. Such formula showed that directly
only the irradiated inclusion makes a contribution to the collective
extinction cross section despite the total number of inclusions being
big, which makes the direct summing contribution of all another
inclusions in spin wave scattering as unviewed (dark mode). We
found also a resonant mechanism of filtering the dark mode from
radiation losses, which makes the linear chain of magnetic inclusions
to be a micro–spin waveguide, which transfers distantly information
in the form of the above-mentioned dark mode without radiation
losses.
References
1. Stamps, R. L., Breitkreutz, S., Akerman, J., Chumak, A. V., Otani, Y., Bauer,
G. E. W., Thiele, J.-U., Bowen, M., Majetich, S. A., Klui, M., et al. (2014). J.
Phys. D: Appl. Phys., 47, 333001.
2. Blugel, S., Burgler, D., Morgenstern, M., Schneider, C. M., and Waser, R.
(2009). Lecture Notes of the 40th Spring School, Julich, Germany.
3. Zutic, I., and Fuhrer, M. (2005). Nat. Phys., 1, 85.
4. Morris, D., Bromberg, D., Zhu, J., and Pileggi, L. (2012). Int. J. High Speed
Electron. Syst., 21, 1250005.
5. Pulizzi, F. (2012). Nat. Mater., 11, 367.
6. Kruglyak, V. V., Keatley, P. S., Neudert, A., Hicken, R. J., Childress, J. R.,
and Katine, J. A. (2010). Phys. Rev. Lett., 104, 027201.
7. Ding, J., and Adeyeye, A. (2012). Appl. Phys. Lett., 101, 103117.
8. Jorzick, J., Demokritov, S. O., Hillebrands, B., Bailleul, M., Fermon, C.,
Guslienko, K. Y., Slavin, A. N., Berkov, D. V., and Gorn, N. L. (2002). Phys.
Rev. Lett., 88, 047204.
9. Gubbiotti, G., Conti, M., Carlotti, G., Candeloro, P., Fabrizio, E. D.,
Guslienko, K. Y., Andre, A., Bayer, C., and Slavin, A. N. (2004). J. Phys.:
Condens. Matter, 16, 7709.
10. Kruglyak, V. V., Demokritov, S. O., and Grundler, D. (2010). J. Phys. D:
Appl. Phys., 43, 264001.
11. Ciubotaru, F., Chumak, A. V., Obry, B., Serga, A. A., and Hillebrands, B.
(2013). Phys. Rev. B, 88, 134406.
12. Krawczyk, M., and Puszkarski, H. (1998). Acta Phys. Pol. A, 93, 805.
194 Spin Waves in Circular and Linear Chains of Discrete Magnetic Elements
13. Nikitov, S. A., Tailhades, P., and Tsai, C. S. (2001). J. Magn. Magn. Mater.,
236, 320.
14. Gulyaev, Y. V., Nikitov, S. A., et al. (2003). JETP Lett., 77, 567.
15. Vysotsky, S. L., Nikitov, S. A., et al. (2005). ZhETF, 128, 636.
16. Chumak, A. V., Pirro, P., Serga, A. A., Kostylev, M. P., Stamps, R. L.,
Schultheiss, H., Vogt, K., Hermsdoerfer, S. J., Laegel, B., Beck, P. A., et al.
(2009). Appl. Phys. Lett., 95, 262508.
17. Kostylev, M. P., Serga, A. A., Schneider, T., Neumann, T., Leven, B.,
Hillebrands, B., and Stamps, R. L. (2007). Phys. Rev. B, 76, 184419.
18. Demidov, V. E., Hansen, U.-H., and Demokritov, S. O. (2007). Phys. Rev.
Lett., 98, 157203.
19. Chumak, A. V., Serga, A. A., Hillebrands, B., and Kostylev, M. P. (2008).
Appl. Phys. Lett., 93, 022508.
20. Mruczkiewicz, M., Krawczyk, M., Gubbiotti, G., Tacchi, S., Filimonov, Y.
A., Kalyabin, D. V., Lisenkov, I. V., and Nikitov, S. A. (2013). New J. Phys.,
15, 113023.
21. Mruczkiewicz, M., Pavlov, E. S., Vysotsky, S. L., Krawczyk, M., Filimonov,
Y. A., and Nikitov, S. A. (2014). Phys. Rev. B, 90, 174416.
22. Lisenkov, I., Kalyabin, D., and Nikitov, S. (2013). Appl. Phys. Lett., 103,
202402.
23. Shindou, R., Ohe, J.-i., Matsumoto, R., Murakami, S., and Saitoh, E.
(2013). Phys. Rev. B, 87, 174402.
24. Lisenkov, I., Tyberkevych, V., Slavin, A., Bondarenko, P., Ivanov, B. A.,
Bankowski, E., Meitzler, T., and Nikitov, S. (2014). Phys. Rev. B, 90,
104417.
25. Shindou, R., Matsumoto, R., Murakami, S., and Ohe, J.-i. (2013). Phys.
Rev. B, 87, 174427.
26. Sebastian, T., Bracher, T., Pirro, P., Serga, A. A., Hillebrands, B., Kubota,
T., Naganuma, H., Oogane, M., and Ando, Y. (2013). Phys. Rev. Lett., 110,
067201.
27. Burin, A. L. (2006). Phys. Rev. E, 73, 066614.
28. Shindou, R., and Ohe, J.-i. (2014). Phys. Rev. B, 89, 054412.
29. Barabanenkov, Yu., Osokin, S., Kalyabin, D., and Nikitov, S. (2015). Phys.
Rev. B, 91, 214419.
30. Barabanenkov, Yu., Osokin, S., Kalyabin, D., and Nikitov, S. P. (2016).
Phys. Rev. B, 94, 184409.
References 195
7.1 Introduction
Magnonic crystals (MCs) offer enhanced control of spin waves
(SWs) via their artificial SW band structures. Periodic magnetic
patterns with lattice constants on the order of the SW wavelength
l allow one to tailor eigenfrequencies w = 2pf and group
velocities dw/dk for a given wavevector k = 2p/l [1–5].
Bicomponent MCs, composed of two different ferromagnetic
materials, have showed backfolding of spin wave (magnon)
dispersion relations, formation of minibands, and opening of
Figure 7.1 Two-dimensional bicomponent magnonic crystal (sketch on
the top right) consisting of two materials [5]. The lattice constant is a.
Measured (open circles) and simulated (dark gray) mode frequencies as a
function of the SW wavevector q in the plane of the MC perpendicular to
magnetic field H0 (see inset). There exist forbidden frequency gaps near
a Brillouin zone boundary (vertical line) due to the periodic modulation
of magnetic properties. On the right side at the bottom two characteristic
mode profiles are depicted for the two low-frequency modes: bright color
indicates large spin-precession amplitude. Adapted from Tacchi et al. [5].
Introduction 199
Figure 7.3 Process using photolithography and etching: (a) YIG (blue)
on the substrate (gray). (b) Photoresist (brown) covering the sample.
(c) After exposure. The exposed area (yellow) becomes soluble in the
developer. (d) After development. The mesa pattern and markers are
transferred to the resist layer. (e) After etching. The ferromagnetic layer has
been etched by argon ion milling. (f) After removal of the resist. Markers
and mesa remain.
Figure 7.4 Cross section of a CPW. The two ground lines with width wg and
the signal line with width ws are separated by a gap of width wgap. t is the
thickness of the dielectric substrate. Field lines of the oscillating magnetic
field surrounding the CPW leads are sketched.
Figure 7.5 The top view of two neighboring CPWs with ground (G) and
signal (S) lines by which a spin wave transmission measurement is
performed. The parameter s denotes the center-to-center separation
between signal lines.
204 Microwave-to-Magnon Transducers for Exchange-Dominated Spin Waves
30p K (d0 )
Zc = ,
eeff K (d0 )
where K is the complete elliptic integral of the first kind and the
effective permittivity eeff is given as
er –1 K (d0 ) K (d1 )
eeff = 1 + ,
2 K (d0 ) K (d1 )
ws
d0 = ; d0 = 1 – d02
ws + wgap
sinh( pws /4t )
d0 = ; d = 1 – d12
sinh{[ p( ws + 2wgap )]/4t } 1
use of a vector network analyzer (VNA). The VNA can apply and
detect microwave signals with different frequencies ranging from
10 MHz to 26.5 GHz. In conventional cavity-based ferromagnetic
resonance (FMR) experiments a small sized magnetic sample is
placed in the standing wave pattern of a microwave cavity of large
volume. Thereby a uniform excitation is possible. For our technique
based on integrated CPWs we intentionally excite the magnetic
sample only locally. Thereby spin waves of non-zero wavevector k
are excited. Depending on their damping (decay) length they can
propagate a certain distance and induce a voltage in a remotely
positioned detector CPW (Fig. 7.5). Using a two-port VNA we can
excite and phase-sensitively detect the propagating spin waves.
Due to the phase-sensitive detection it is possible to analyze the
real and imaginary part of the scattering parameter S12 as discussed
in the following.
Figure 7.6 Sketch of the AESWS setup with a pair of magnet coils
applying a field along the CPW. The VNA is connected via microwave
cables to tips which have a GSG geometry and connect the microwave
source and detector to the CPW. A sample might be mounted in flip-chip
configuration to a preexisting CPW or contain an integrated CPW.
(GSG) geometry which fits to the pad region of the CPW (Fig. 7.5)
and make electrical connections accordingly. The microwave tips
are held by x,y,z translational stages to ensure precise and rigid
positioning. The contacted CPWs plus sample are located between
magnet coils which generate an external magnetic field (Fig. 7.6).
The whole setup is on a vibration-cushioned table in a climate
controlled laboratory.
b1 S11 S 21 a1
b2 S12 S 22 a2
j = k . s.
a21
a11
b=
a21 a12
+
a11 a22
where a11, a12, a21, a22 are extracted from VNA spectra taken in
transmission and reflection configurations. The nonreciprocity
parameter enters when evaluating transmitted amplitudes in
terms of the decay length ld [20]:
s s
– –
a21 = ( b )a11e ld
, a12 = (1 – b)a22e ld .
Figure 7.7 Spectra taken on a (a) CoFeB film and (b) a BGC with nanodisks
in reflection configuration for different angles q. White dotted lines
represent model calculations for the CPW modes kI, kII, and kIII [11].
White squares highlight the resonance frequencies of these three
modes at q = 90°. The BGC contained Py nanodisks on a lattice with
period a = 800 nm. The red dashed lines in (b) represent calculated
angular dependencies for modes with k = kI + G01 (uppermost line) and
k = kI – G01 (second-highest line). Modes attributed to the grating
coupler effect with reciprocal lattice vectors G that are tilted with
respect to G01 are marked with blue arrows. (c) Calculated dispersion
relation f (k) of DE modes at 40 mT in a CoFeB plane film for positive
and negative k (solid lines). The vertical dashed line indicates kI.
The squares highlight frequencies expected for |kI|, |kII|, and |kIII|.
Horizontal arrows and dashed curves indicate the back-folding of spin
wave branches due to the grating coupler. Graphs extracted from
figure “Angular-dependent spin-wave spectroscopy and back-folded
dispersion relation,” rearranged from Yu et al. [11], published under
CC-BY-NC-SA.
Performance of a Spin Wave Grating Coupler 211
The spectra become richer in Fig. 7.9 (bottom row) when the
nanodisk array is present. As discussed, spin waves are detected
up to higher frequencies compared to the bare YIG film.
These high-frequency modes are due to the grating coupler effect.
With angular dependence measurements, we follow how such
modes transform from the DE mode configuration into BVMSW
mode. Some of the modes have “frequency maxima” near 45°
Conclusions and Outlook 215
Acknowledgments
We thank S. Tacchi, G. Duerr, R. Huber, M. Bahr, T. Schwarze, F.
Brandl, M. Madami, S. Neusser, G. Gubbiotti, G. Carlotti, O. d’ Allivy
Kelly, V. Cros, R. Bernard, P. Bortolotti, A. Anane, F. Heimbach,
and Y. Stasinopoulos for their contributions. The experimental
work received funding from the Deutsche Forschungsgemeinschaft
(DFG) via the German excellence cluster Nanosystems Initiative
Munich (NIM), as well as project GR1640/5-1/2 in the framework
of the DFG Priority Program Spincaloric Transport Phenomena
SPP 1538.
We thank the following colleagues for their help in editing
the text: Che Ping, Tobias Stueckler, Sa Tu, Chuanpu Liu, Junfeng Hu,
Jilei Chen, Jianyu Zhang, Xing Chen, and Yinan Wang.
References
10. Khitun, A., Bao, M., and Wang, K. L. (2010). Magnetic cellular nonlinear
network with spin wave bus for image processing. Superlattices
Microstruct., 47, 464–483.
11. Yu, H., et al. (2013). Omnidirectional spin-wave nanograting coupler,
Nat. Commun., 4, 2702.
12. Kelly, O. A., et al. (2014). Inverse spin Hall effect in nanometer-
thick yttrium iron garnet/Pt system, Appl. Phys. Lett., 103, 082408.
13. Yu, H., et al. (2016). Approaching soft X-ray wavelengths in
nanomagnet-based microwave technology, Nat. Commun., 7, 11255.
14. Simons, R. N. (2001). Coplanar Waveguide Circuits, Components,
and Systems (Wiley-Interscience John Wiley & Sons, Inc., New York,
NY, USA).
15. Silva, T. J., Lee, C. S., Crawford, T. M., and Rogers, C. T. (1999).
Inductive measurement of ultrafast magnetization dynamics in
thin-film Permalloy, J. Appl. Phys., 85, 7849–7862.
16. Bailleul, M., Olligs, D., Fermon, C., and Demokritov, S. O. (2001).
Spin waves propagation and confinement in conducting films
at the micrometer scale, Europhys. Lett., 56, 741.
17. Yu, H., et al. (2012). Magnetic thin-film insulator with ultra-low
spin wave damping for coherent nanomagnonics, Sci. Rep., 4, 6848.
18. Yu, H., et al. (2012). High propagating velocity of spin waves and
temperature dependent damping in a CoFeB thin film, Appl. Phys.
Lett., 100, 262412.
19. Sebastian, T., et al. (2012). Low-damping spin-wave propagation
in a micro-structured Co2Mn0.6Fe0.4Si Heusler waveguide, Appl. Phys.
Lett., 100, 112402.
20. Huber, R., Krawczyk, M., Schwarze, T., Yu, H., Duerr, G., Albert, S., and
Grundler, D. (2013). Reciprocal Damon-Eshbach-type spin wave
excitation in a magnonic crystal due to tunable magnetic symmetry,
Appl. Phys. Lett., 102, 012403.
21. Au, Y., et al. (2012). Resonant microwave-to-spin-wave transducer,
Appl. Phys. Lett., 100, 182404.
22. Sandweg, C. W., et al. (2011). Spin pumping by parametrically excited
exchange magnons, Phys. Rev. Lett., 106, 216601.
23. Demidov, V. E., et al. (2011). Excitation of short-wavelength spin
waves in magnonic waveguides, Appl. Phys. Lett., 99, 082507.
24. Madami, M., et al. (2011). Direct observation of a propagating spin
wave induced by spin-transfer torque, Nat. Nanotechnol., 6, 635–638.
218 Microwave-to-Magnon Transducers for Exchange-Dominated Spin Waves
25. Bonetti, S., et al. (2015). Direct observation and imaging of a spin-
wave soliton with p-like symmetry, Nat. Commun., 6, 8889.
26. Vlaminck, V., and Bailleul, M. (2010). Spin-wave transduction at
the submicrometer scale: experiment and modeling, Phys. Rev. B,
81, 014425.
27. Che, P., Bi, L., Yu, H., et al. (2016). Short-wavelength spin waves
in yttrium iron garnet micro-channels on silicon, IEEE. Mag. Lett.,
7, 3508404.
28. Wintz, S., et al. (2016). Magnetic vortex cores as tunable spin-wave
emitters, Nat. Nanotechnol., 11, 948–953.
29. Davies, C. S., and Kruglyak, V. V. (2016). Generation of propagating
spin waves from edges of magnetic nanostructures pumped by
uniform microwave magnetic field, IEEE Trans. Magn., 52, 2300504.
30. Grundler, D. (2016). Spintronics: nanomagnonics around the corner,
Nat. Nanotechnol., 11, 407–408.
Chapter 8
Colorado, USA
[email protected]
8.1 Introduction
Spintronics exists because of an extra degree of freedom provided
by electron spin that can be used for carrying information.
Whereas information can be manipulated and transported using
charge, this comes at a cost due to the concurrent and inherent
generation of Joule heating. An alternative mechanism for
transporting information through the spin variable is available
and, in fact, has been studied for over 80 years. Spin waves and
their particle-like counterpart, magnons, are the low-lying energy
states of spin systems and were first predicted by Bloch [4, 25,
38, 43]. Not only do spin wave excitations exhibit a wide variety
of linear and nonlinear properties (which makes them interesting
for fundamental research), they also exist in the gigahertz to
terahertz region of the frequency spectrum, which is appropriate
for telecommunications and information technologies. New
technologies that allow the fabrication of devices in the nanoscale
together have led to the discovery of phenomena such as spin
pumping [88], spin transfer torque [2, 80], and spin Hall effects
[24, 42]. The field now called magnonics is concerned with
consequences of the fact that the transport and processing of
information can be achieved without physical charge transport.
Challenges addressed in this field pertain to issues related
with spin wave dissipation, device miniaturization [82],
and fabrication of artificial magnonic crystals [48–50]. Most
recently, consequences of the Dzyaloshinskii–Moriya interaction
(DMI) on spin wave properties has been studied extensively,
especially in regards to interface-induced DMI. The DMI arises in
low-symmetry materials with a strong spin-orbit coupling and
is modeled as an antisymmetric form of the exchange interaction.
Dzyaloshinskii first postulated this interaction in order to
explain weak ferromagnetism in antiferromagnets [26]. A few
years later Moriya calculated the second-order energy terms
associated with spin-orbit couplings for the exchange interaction,
thereby establishing a mechanism for the interaction [68, 69].
In noncentrosymmetric magnetic crystals the DMI is
responsible for the spontaneous formation of helicoidal and
skyrmionic structures [6, 7]. From the viewpoint of applications,
the most exciting recent development has been experimental
demonstration that an interface form of the DMI can appear
because of inversion symmetry breaking at the surface between
magnetic films on heavy metal nonmagnetic substrates that
provide the spin-orbit coupling. Experiments have shown that
this induced form of the DMI leads to chiral spin structures in
manganese monolayers on top of tungsten [5] and skyrmion
lattices in iron monolayers on iridium [37]. This interfacial
DMI also exists for sputtered multilayer films such as Pt/Co/Ni
[23], Pt/Co/AlOx [1], Pt/Co/Ir [66], Pt/Co/MgO [9], and
(TaN, Hf, W)/CoFeB/MgO [81], which have strong perpendicular
magnetic anisotropy. This form of DMI has helped explain
Introduction 221
dm dm , (8.1)
= – gm0m × Heff + am ×
dt dt
d dm
= – gm0 (m0 × dHeff + dm × Heff,0 ). (8.3)
dt
where
and
Figure 8.1 DMI-induced drift of a spin wave ripple. (a) Time evolution
of a spin wave ripple at three instants (2, 4, and 8 ns) after the
application of a sinusoidal field pulse at the center of the image, with
D = 1 mJ/m2. The image dimensions are 10 μm × 10 μm. Dx denotes the
displacement of the ripple center. (b) Ripple displacement as a function
of time for three different values of D. Symbols correspond to simulation
data while solid lines are based on Eq. 8.4. Reprinted with
permission from Kim et al. [53]. Copyright 2016 American Physical Society.
226 Spin Waves on Spin Structures
8.2.2 Caustics
The DMI-induced drift in spin waves has interesting consequences
for power flow. With an interfacial DMI and for propagation in the
DE geometry, the dispersion curve is approximately parabolic
but with the minimum shifted away from the origin along the
wavevector axis. Because of this, dw/dk is negative in some regions,
and this indicates the group velocity is opposite to the phase
velocity. However, this simple analysis is not sufficient to capture
all the important features of the anisotropic power flow created
by the DMI.
The study of focusing patterns for bulk [85] and surface
phonons [13] in crystals is well known. The corresponding
investigations in thick-film magnetic systems have begun only
recently with both experimental [19, 22, 78, 79] and theoretical
results [94]. The focusing results have already shown remarkable
behaviors, including focusing effects of energy well below the
expected diffraction limit and an interesting reflection behavior
for energy where the angle of incidence is not equal to the
angle of reflection. In many respects the magnetic system offers
richer phenomena because the external magnetic field offers an
opportunity to tune the dispersion relations and alter the focusing
patterns, something that is not available in phonon focusing.
Let us discuss how this drift leads to focusing effects and
caustics. In general, the far-field radiation pattern of waves excited
by a point source can be predicted from the slowness surface,
that is, a constant frequency curve in k-space. The radiation or
focusing pattern can then be determined from the power flow,
directed along the normal to the slowness surface, and with an
amplitude that is inversely proportional to the square root of the
curvature of the slowness surface [94]. Caustics appear at points
along the slowness surface at which its curvature goes to zero,
resulting in a divergence in the power flow.
To understand how caustics appear for spin waves in ultrathin
films with an interfacial DMI, let us return to the dispersion
relation in Eq. 8.4 from which the slowness surfaces can be
computed. There are three main contributions to the spin wave
energy that are wavevector dependent. First, the exchange term,
wex k2, gives a circular component to the slowness surface
that results in a finite and positive curvature for all propagation
directions in the film plane. As such, radiation of spin wave
Chiral Interactions and Spin Waves 227
Figure 8.2 Spin wave power flow and caustics. (a) Slowness surfaces
for different frequencies determined from Eq. 8.4. vg denotes the group
velocity vector. (b) Predicted focusing patterns based on the slowness
surfaces in (a). (c) Simulated focusing patterns due to a sinusoidal
point source excitation at different frequencies. Each image represents
an area of 20 μm × 20 μm, and the point source is located at the center.
The frequencies are chosen to match the focusing patterns in (b).
Reprinted with permission from Kim et al. [53]. Copyright 2016
American Physical Society.
Chiral Interactions and Spin Waves 229
(a)
hrf
DW
y
z x – 0.01 0 0.01
50
50 GHz
40 bulk
W / 2p (GHz)
30
(c)
20
wall
10 10 GHz
(b)
0
–0.2 0 0.2
kx (nm-1)
50
50 GHz
40 bulk
W / 2p (GHz)
30
(e)
20
wall
10 10 GHz
(d)
0
–0.2 0 0.2
kx (nm-1)
Figure 8.4 Spin wave channeling in domain walls (DWs). (a) Geometry
for channeling along the center of the wall, where a radio-frequency
antenna generating an alternating field, hrf, excites spin waves that
propagate along the x direction. (b, d) Dispersion relation for channeled
Bloch (b) and Néel (d) domain wall spin wave modes in comparison
with bulk spin waves. For the Néel wall case, D = 1.5 mJ/m2. (c, e) Simulation
results of propagating modes for excitation field frequencies in the
bulk (50 GHz) and in the gap (10 GHz) for Bloch (c) and Néel (e)
walls. These driving frequencies are shown as dashed lines in (b, d).
Reprinted with permission from Garcia-Sanchez et al. [28]. Copyright
2015 American Physical Society.
Figure 8.5 Spin wave channeling along curved domain walls. (a) Geometry
of the curved domain wall used in simulation, where a radio-frequency
antenna generating an alternating field, hrf, excites spin waves that
propagate along the wall. (b) Snapshots of channeled spin wave modes
at different wavelengths with the corresponding excitation frequencies.
rd denotes the radius of curvature of the domain wall channel.
Reprinted with permission from Garcia-Sanchez et al. [28]. Copyright 2015
American Physical Society.
w 2A
yc – = l cosh–1 , (8.8)
2 Dl
Figure 8.7 Twisted spin states at boundary edges due to the DMI.
(a) Transverse magnetization component at the boundary edges
(located at y = ±256 nm) of a 512-nm-wide rectangular wire. (b)
Illustration of the magnetization tilts for D > 0, with the shaded
regions representing the tilts shown in panels (a) and (c). The partial
wall (dashed curve) is shown schematically, with yc denoting the wall
center and w the wire width. (c) The perpendicular component at
the boundary edges, where the solid lines correspond to fits to a partial
Néel wall profile. (d) Partial wall center yc as a function of D. Points
are simulation data and the solid line represents Eq. 8.8. Reprinted
with permission from Garcia-Sanchez et al. [29]. Copyright 2014 American
Physical Society.
relation for the edge modes, while the central modes remain
symmetric about kcen = 0 (Fig. 8.8c). For the central modes, the
dispersion relation is well described by exchange-dominated
spin waves, where the theoretical curve using our micromagnetic
parameters agrees well with the simulated curves. For the edge
modes, the shifted dispersion relation for D = 4.5 mJ/m2 is well
described by a reduction in the spin wave gap due to the reduced
anisotropy field at the edge in addition to a linear wavevector
term that describes the nonreciprocity. As Fig. 8.7d shows, the
center of the partial wall is located farther outside the wire for
smaller values of the DMI, which results in a weaker nonreciprocal
channeling effect. This can be seen in the dispersion relation of
the edge modes in Fig. 8.8d, where the shifts become less
pronounced as D decreases. This phenomenon is reminiscent of
edge modes in topological insulators.
Channeling for the wire geometry is robust with regard
to the curvature of the edge in the same way as for curved
domain walls. We now discuss the consequences for finite-size
nanostructures. In a circular dot, for example, it is known that
clockwise (CW)- and counterclockwise (CCW)-propagating
azimuthal spin waves are degenerate in frequency. The inclusion
of the DMI, however, lifts this degeneracy by favoring one
handedness over the other. To appreciate how this might occur,
one can imagine the edge modes in a circular dot constructed
by deforming a rectangular wire bent into a ring-shaped structure.
The lowest-frequency spin waves traveling along the outer
circumference can propagate with only one handedness. Spin
waves traveling along the inner circumference travel with the
opposite handedness at the same frequency.
Figure 8.9 illustrates the spin wave eigenmode spectra for
a circular dot 100 nm in diameter and a square dot 100 nm in
width. A key feature is the frequency splitting of certain modes
as the strength of the DMI is increased. The frequency of other
modes, on the other hand, are only slightly affected by the DMI.
For a similar dot size, the magnitude of the splitting appears to be
larger for the circular dots, which suggests that the azimuthal
component of the eigenmodes plays an important role. For the
circular dots, the frequency splitting with increasing DMI is
associated with lifting in the degeneracy of eigenmodes with a
strong azimuthal character, such as Modes 2 and 3 in Fig. 8.9c.
240 Spin Waves on Spin Structures
are also seen in the square dots, but the distinction between
radial and azimuthal modes is not as sharp. One difference can
be seen in Mode 4 in Fig. 8.9d, which represents a mixed radial-
azimuthal excitation for which splitting due to the DMI results
in an asymmetric profile at higher frequencies.
8.4 Outlook
We close with a few comments regarding the important topics of
charge and spin transport. It may be possible to utilize and control
Outlook 251
Acknowledgments
The authors acknowledge fruitful discussions with J.-P. Adam,
M. Bailleul, V. Cros, T. Devolder, Y. Henry, S. Rohart, J. Sampaio, R.
Soucaille, A. Thiaville, and A. Vansteenkiste. This work was partially
supported by the Agence Nationale de la Recherche (France)
under grant numbers ANR-11-BS10-003 (NanoSWITI), ANR-14-
CE26-0012 (Ultrasky), and ANR-16-CE24-0027 (Swangate); the
Engineering and Physical Sciences Research Council (UK); the
University of Glasgow; and the National Council of Science and
Technology of Mexico (CONACyT). The work of Y. L. and R. L. S.
was supported by the China Scholarship Council and EPSRC
(grant number EP/L002922/1). P. B. acknowledges support
from a Grant-in-Aid for Scientific Research on Innovative Areas
(Grant No.26103006) from the Ministry of Education, Culture,
Sports, Science and Technology (MEXT), Japan. We would like
to acknowledge that some of the results of micromagnetics
simulations presented here made use of the Emerald High
Performance Computing facility made available by the Centre
for Innovation. The Centre for Innovation is formed by the
universities of Oxford, Southampton, Bristol, and University
College London in partnership with the STFC Rutherford-Appleton
Laboratory.
References 253
References
1. Belmeguenai, M., Adam, J.-P., Roussigné, Y., Eimer, S., Devolder, T.,
Kim, J.-V., Cherif, S. M., Stashkevich, A., and Thiaville, A. (2015).
Interfacial Dzyaloshinskii-Moriya interaction in perpendicularly
magnetized Pt/Co/AlOx ultrathin films measured by Brillouin light
spectroscopy, Phys. Rev. B, 91, 180405.
2. Berger, L. (1996). Emission of spin waves by a magnetic multilayer
traversed by a current, Phys. Rev. B, 54, 9353–9358.
3. Bhat, V. S., Sklenar, J., Farmer, B., Woods, J., Hastings, J. T., Lee, S.
J., Ketterson, J. B., and De Long, L. E. (2013). Controlled magnetic
reversal in Permalloy films patterned into artificial quasicrystals,
Phys. Rev. Lett., 111, 077201.
4. Bloch, F. (1930). Zur Theorie des Ferromagnetismus, Z. Phys. A, 61,
206–219.
5. Bode, M., Heide, M., von Bergmann, K., Ferriani, P., Heinze, S., Bihlmayer,
G., Kubetzka, A., Pietzsch, O., Blügel, S., and Wiesendanger, R. (2007).
Chiral magnetic order at surfaces driven by inversion asymmetry,
Nature, 447, 190–193.
6. Bogdanov, A., and Hubert, A. (1994). Thermodynamically stable
magnetic vortex states in magnetic crystals, J. Magn. Magn. Mater.,
138, 255–269.
7. Bogdanov, A. N., and Rößler, U. K. (2001). Chiral symmetry breaking
in magnetic thin films and multilayers, Phys. Rev. Lett., 87, 037203.
8. Borys, P., Garcia-Sanchez, F., Kim, J.-V., and Stamps, R. L. (2016).
Spin-wave eigenmodes of Dzyaloshinskii domain walls, Adv. Electron.
Mater., 2, 1500202.
9. Boulle, O., et al. (2016). Room-temperature chiral magnetic
skyrmions in ultrathin magnetic nanostructures, Nat. Nanotechnol.,
11, 449–454.
10. Braun, H.-B. (2012). Topological effects in nanomagnetism: from
superparamagnetism to chiral quantum solitons, Adv. Phys., 61,
1–116.
11. Budrikis, Z., Politi, P., and Stamps, R. L. (2012). A network model for
field and quenched disorder effects in artificial spin ice, New J. Phys.,
14, 045008.
12. Buess, M., Knowles, T. P. J., Höllinger, R., Haug, T., Krey, U., Weiss, D.,
Pescia, D., Scheinfein, M. R., and Back, C. H. (2005). Excitations with
negative dispersion in a spin vortex, Phys. Rev. B, 71, 104415.
254 Spin Waves on Spin Structures
27. Farhan, A., Kleibert, A., Derlet, P. M., Anghinolfi, L., Balan, A., Chopdekar,
R. V., Wyss, M., Gliga, S., Nolting, F., and Heyderman, L. J. (2014).
Thermally induced magnetic relaxation in building blocks of artificial
kagome spin ice, Phys. Rev. B, 89, 214405.
28. Garcia-Sanchez, F., Borys, P., Soucaille, R., Adam, J.-P., Stamps, R. L.,
and Kim, J.-V. (2015). Narrow magnonic waveguides based on domain
walls, Phys. Rev. Lett., 114, 247206.
29. Garcia-Sanchez, F., Borys, P., Vansteenkiste, A., Kim, J.-V., and Stamps,
R. L. (2014). Nonreciprocal spin-wave channeling along textures
driven by the Dzyaloshinskii-Moriya interaction, Phys. Rev. B, 89,
224408.
30. Gilbert, I., Chern, G.-W., Zhang, S., O’Brien, L., Fore, B., Nisoli, C.,
and Schiffer, P. (2014). Emergent ice rule and magnetic charge
screening from vertex frustration in artificial spin ice, Nat. Phys., 10,
670–675.
31. Gilbert, I., Lao, Y., Carrasquillo, I., O’Brien, L., Watts, J. D., Manno,
M., Leighton, C., Scholl, A., Nisoli, C., and Schiffer, P. (2015).
Emergent reduced dimensionality by vertex frustration in artificial
spin ice, Nat. Phys., 12, 162–165.
32. Gubbiotti, G., Albini, L., Carlotti, G., De-Crescenzi, M., Di Fabrizio,
E., Gerardino, A., Donzelli, O., Nizzoli, F., Koo, H., and Gomez, R. D.
(2000). Finite size effects in patterned magnetic Permalloy films, J.
Appl. Phys., 87, 5633.
33. Gubbiotti, G., Malago, P., Fin, S., Tacchi, S., Giovannini, L., Bisero,
D., Madami, M., Carlotti, G., Ding, J., Adeyeye, A. O., and Zivieri, R.
(2014). Magnetic normal modes of bicomponent Permalloy/cobalt
structures in the parallel and antiparallel ground state, Phys. Rev. B,
90, 024419.
34. Gubbiotti, G., Tacchi, S., Carlotti, G., Vavassori, P., Singh, N., Goolaup,
S., Adeyeye, A. O., Stashkevich, A., and Kostylev, M. (2005).
Magnetostatic interaction in arrays of nanometric Permalloy wires:
a magneto-optic Kerr effect and a Brillouin light scattering study,
Phys. Rev. B, 72, 224413.
35. Hamrle, J., Gaier, O., Min, S.-G., Hillebrands, B., Sakuraba, Y., and
Ando, Y. (2009). Determination of exchange constants of Heusler
compounds by Brillouin light scattering spectroscopy: application
to Co2MnSi, J. Phys. D: Appl. Phys., 42, 084005.
36. Heide, M., Bihlmayer, G., and Blügel, S. (2008). Dzyaloshinskii-
Moriya interaction accounting for the orientation of magnetic
domains in ultrathin films: Fe/W(110), Phys. Rev. B, 78, 140403.
256 Spin Waves on Spin Structures
37. Heinze, S., von Bergmann, K., Menzel, M., Brede, J., Kubetzka, A.,
Wiesendanger, R., Bihlmayer, G., and Blügel, S. (2011). Spontaneous
atomic-scale magnetic skyrmion lattice in two dimensions, Nat. Phys.,
7, 713–718.
38. Herring, C., and Kittel, C. (1951). On the theory of spin waves in
ferromagnetic media, Phys. Rev., 81, 869–880.
39. Heyderman, L. J., Mengotti, E., Zanin, D., Chopdekar, R. V., Braun,
H.-B., Hügli, R. V., and Duff, G. (2013). Method of controlling the
states and vortex chirality in hexagonal ring structures comprising
nanoscale magnetic elements, US Patent Specification, 8450047 B2.
40. Heyderman, L. J., and Stamps, R. L. (2013). Artificial ferroic systems:
novel functionality from structure interactions and dynamics,
J. Phys.: Condens. Matter, 25, 363201.
41. Hillebrands, B., Mathieu, C., Bauer, M., Demokritov, S. O., Bartenlian,
B., Chappert, C., Decanini, D., Rousseaux, F., and Carcenac, F. (1997).
Brillouin light scattering investigations of structured Permalloy
films, J. Appl. Phys., 81, 4993–4995.
42. Hirsch, J. E. (1999). Spin Hall effect, Phys. Rev. Lett., 83, 1834–1837.
43. Holstein, T., and Primakoff, H. (1940). Field dependence of the
intrinsic domain magnetization of a ferromagnet, Phys. Rev., 58,
1098–1113.
44. Hügli, R. V., Duff, G., O’Conchuir, B., Mengotti, E., Fraile Rodríguez,
A., Nolting, F., Heyderman, L. J., and Braun, H.-B. (2012). Artificial
kagome spin ice: dimensional reduction, avalanche control and
emergent magnetic monopoles, Philos. Trans. R. Soc. London, Ser. A,
370, 5767–5782.
45. Iacocca, E., Gliga, S., Stamps, R. L., and Heinonen, O. (2016).
Reconfigurable wave band structure of an artificial square ice, Phys.
Rev. B, 93, 134420.
46. Imre, A., Csaba, G., Ji, L., Orlov, A., Bernstein, G. H., and Porod, W. (2006).
Majority logic gate for magnetic quantum-dot cellular automata,
Science, 311, 205–208.
47. John, S. (1987). Strong localization of photons in certain disordered
dielectric superlattices, Phys. Rev. Lett., 58, 2486–2489.
48. Khanikaev, A., Mousavi, S. H., Tse, W.-K., Kargarian, M., MacDonald,
A., and Shvets, G. (2013). Photonic topological insulators, Nat. Mater.,
12, 233–239.
49. Khitun, A. (2013). Magnonic holographic devices for special type
data processing, J. Appl. Phys., 113, 164503.
References 257
50. Khitun, A., Bao, M., and Wang, K. L. (2010). Magnonic logic circuits,
J. Phys. D: Appl. Phys., 43, 264005.
51. Kikuchi, T., Koretsune, T., Arita, R., and Tatara, G. (2016). Dzyaloshinskii-
Moriya interaction as a consequence of a Doppler shift due to
spin-orbit-induced intrinsic spin current, Phys. Rev. Lett., 116,
247201.
52. Kim, J.-V., and Stamps, R. L. (2005). Hysteresis from antiferromagnet
domain-wall processes in exchange-biased systems: magnetic defects
and thermal effects, Phys. Rev. B, 71, 094405.
53. Kim, J.-V., Stamps, R. L., and Camley, R. E. (2016). Spin wave power
flow and caustics in ultrathin ferromagnets with the Dzyaloshinskii-
Moriya interaction, Phys. Rev. Lett., 117, 197204.
54. Kostylev, M. P. (2014). Interface boundary conditions for dynamic
magnetization and spin wave dynamics in a ferromagnetic layer with
the interface Dzyaloshinskii-Moriya interaction, J. Appl. Phys., 115,
233902.
55. Lan, J., Yu, W., Wu, R., and Xiao, J. (2015). Spin-wave diode, Phys. Rev. X,
5, 041049.
56. Lenk, B., Ulrichs, H., Garbs, F., and Münzenberg, M. (2011). The
building blocks of magnonics, Phys. Rep., 507, 107–136.
57. Li, Y., Gubbiotti, G., Casoli, F., Gonçalves, F. J. T., Morley, S. A.,
Rosamond, M. C., Linfield, E. H., Marrows, C. H., McVitie, S., and Stamps,
R. L. (2016). Brillouin light scattering study of magnetic-element
normal modes in a square artificial spin ice geometry, J. Phys. D: Appl.
Phys., 50, 015003.
58. Li, J., Ke, X., Zhang, S., Garand, D., Nisoli, C., Lammert, P., Crespi, V. H.,
and Schiffer, P. (2010). Comparing artificial frustrated magnets by
tuning the symmetry of nanoscale Permalloy arrays, Phys. Rev. B, 81,
092406.
59. Macià, F., Kent, A. D., and Hoppensteadt, F. C. (2011). Spin-wave
interference patterns created by spin-torque nano-oscillators for
memory and computation, Nanotechnology, 22, 095301.
60. Mathieu, C., et al. (1997). Anisotropic magnetic coupling of
Permalloy micron dots forming a square lattice, Appl. Phys. Lett.,
70, 2912–2914.
61. McMichael, R. D., and Stiles, M. D. (2005). Magnetic normal modes
of nanoelements, J. Appl. Phys., 97, 10J901.
62. Mengotti, E., Heyderman, L. J., Rodriguez, A. F., Nolting, F., Hügli, R.
V., and Braun, H.-B. (2011). Real-space observation of emergent
258 Spin Waves on Spin Structures
88. Tserkovnyak, Y., Brataas, A., and Bauer, G. E. W. (2002). Enhanced gilbert
damping in thin ferromagnetic films, Phys. Rev. Lett., 88, 117601.
89. Uchida, K., Xiao, J., Adachi, H., Ohe, J., Takahashi, S., Ieda, J., Ota, T.,
Kajiwara, Y., Umezawa, H., Kawai, H., Bauer, G. E. W., Maekawa, S., and
Saitoh, E. (2010). Spin Seebeck insulator, Nat. Mater., 9, 894–897.
90. Udvardi, L., and Szunyogh, L. (2009). Chiral asymmetry of the spin-
wave spectra in ultrathin magnetic films, Phys. Rev. Lett., 102,
207204.
91. Ulrichs, H., Demidov, V. E., Demokritov, S. O., and Urazhdin, S. (2012).
Spin-torque nano-emitters for magnonic applications, Appl. Phys.
Lett., 100, 162406.
92. Vassilios, K., Arnalds, U. B., Adam, H. C., Papaioannou, E. T., Karimipour,
M., Korelis, P., Taroni, A., Holdsworth, P. C. W., Bramwell, S. T., and
Hjörvarsson, B. (2012). Melting artificial spin ice, New J. Phys., 14,
035009.
93. Vansteenkiste, A., Leliaert, J., Dvornik, M., Helsen, M., Garcia-Sanchez,
F., and Van Waeyenberge, B. (2014). The design and verification of
MuMax3, AIP Adv., 4, 107133.
94. Veerakumar, V., and Camley, R. E. (2006). Magnon focusing in thin
ferromagnetic films, Phys. Rev. B, 74, 214401.
95. Wagner, K., Kakay, A., Schultheiss, K., Henschke, A., Sebastian, T., and
Schultheiss, H. (2016). Magnetic domain walls as reconfigurable
spin-wave nanochannels, Nat. Nanotechnol., 11, 432–436.
96. Wang, R. F., et al. (2006). Artificial ‘spin ice’ in a geometrically frustrated
lattice of nanoscale ferromagnetic islands, Nature, 439, 303–306.
97. Wang, Z. K., Zhang, V. L., Lim, H. S., Ng, S. C., Kuok, M. H., Jain, S., and
Adeyeye, A. O. (2009). Observation of frequency band gaps in a one-
dimensional nanostructured magnonic crystal, Appl. Phys. Lett., 94,
083112.
98. Winter, J. M. (1961). Bloch wall excitation. Application to nuclear
resonance in a Bloch wall, Phys. Rev., 124, 452–459.
99. Yablonovitch, E. (1987). Inhibited spontaneous emission in solid-
state physics and electronics, Phys. Rev. Lett., 58, 2059–2062.
100. Zakeri, Kh., Zhang, Y., Prokop, J., Chuang, T.-H., Sakr, N., Tang, W. X.,
and Kirschner, J. (2010). Asymmetric spin-wave dispersion on
Fe(110): direct evidence of the Dzyaloshinskii-Moriya interaction,
Phys. Rev. Lett., 104, 137203.
Chapter 9
9.1 Introduction
The main idea behind magnonics is to explore the foundations of
a novel type of information processing technology, in which logic
operations are performed by the superposition of waves rather
than by the motion of electric charges [3–8]. But what are the
reasons behind searching for alternatives for state of the art CMOS
technologies?
Since the information transport by magnons does not rely
on charge transport it does not suffer from Ohmic losses and the
associated Joule heating, one of the drawbacks and limitations of
conventional electronics and their scalability. Moreover, the wave-
based phenomena such as interference, diffraction or refraction
allow for novel types of logic architectures and a broad frequency
Introduction 263
Figure 9.2 (a) Intensity profile of two counter propagating magnon beams
in a Permalloy stripe with a width of 4 µm revealing a clear stationary
interference pattern. (b) Plot of the magnon frequency as a function of
the wavevector extracted from the periodicity of the interference pattern.
Images adopted from [15].
Figure 9.3 (a) Magnon transport in 2.5 µm wide waveguides. One straight and
the other with a skew section, where the waveguide is shifted transversally
by 1 µm over a distance of 3 µm. A magnetic field of 50.7 mT was applied
perpendicularly to the waveguide and the excitation frequency was 7 GHz.
(b) Magnons propagating towards a Y junction, excited by microwaves in
the lower part (not shown). Magnon intensities were measured with µBLS.
Images adapted from Clausen et al. [32] and Vogt et al. [33].
Figure 9.5 (a) Magnon intensities measured in the vicinity of the microwave
antenna used for magnon excitation as a function of the excitation frequency
and the externally applied magnetic field or (b) a direct current flowing in
the Py(30)Au(50) magnon waveguide. The black solid line indicates the
lower limit of the magnon band ( k^ = 0 ), the white dashed line marks the
magnetic field and frequency combination, where the resulting magnon
wavevector coincides with the minimum of the microwave antenna’s
excitation efficiency.
Figure 9.6 (a) Empty symbols show the BLS intensity measured as a function
of an externally applied magnetic field for excitation frequencies ranging
from 2.1 to 3.9 GHz. Filled symbols show the results of the BLS measurement
with dc current pulses through the Au/permalloy hybrid waveguide and
without applying an external magnetic field. The measurement positions
of panels A, B, and C are indicated in (b); the results were obtained 1 µm,
8 µm, and 11 µm away from the antenna. (c) Two-dimensional magnon
intensity distribution with an externally applied magnetic field of 12.3 mT
at an excitation frequency of 2.1 GHz and (d–g) applied dc pulses with an
amplitude of 66.7 mA for excitation frequencies ranging from 2.1 GHz to
3.9 GHz. The intensity scale is logarithmic and color-coded. Images adopted
from [36].
Steering and Multiplexing Magnons by Current-Induced, Local Magnetic Fields 275
Figure 9.8 Magnon intensities measured with µBLS in the left (red circles) and
right (blue squares) part of the magnon waveguide in 4.5 μm distance to the
excitation antenna (effective propagation distance of the magnons). Magnons
are excited at frequencies ranging from 2 to 4 GHz and an electric current of
100 mA was flowing in (a) the left and (b) the right arm of the Y structure.
Images adapted from Vogt et al. [33].
of 100 mA flowing in the left (Fig. 9.8a) and right (Fig. 9.8b) part of
the Y-junction magnon intensities were measured with µBLS in both
the top-right arm (red squares) and the top left arm (blue circles)
for frequencies ranging from 2 GHz to 4 GHz. There is clear evidence
for an overall asymmetry in the magnon propagation between
the left and right arm of the Y junction and maximum magnon
intensity is observed always in the part of the structure in which the
magnetization is aligned by the current-induced magnetic field.
Ultimate evidence for the functionality of the magnon multiplexer
is obtained from the measured two-dimensional magnon intensity
distributions in Fig. 9.9. Excitation with a frequency of 2.75 GHz,
which showed maximum magnon transmission in Fig. 9.8 for an
electric current of 100 mA, demonstrates switching of magnon beams
in the direction of the current flow. Figure 9.9c shows that magnon
propagation is entirely blocked behind the Y junction if a uniform
external magnetic field is applied, as discussed in the previous
section. Magnons are known to adiabatically adjust their wavevector
in regions of varying internal magnetic fields to compensate for the
resulting shifts of the dispersion [38–43]. But, apparently, magnons
do not propagate in stripe-shaped magnon waveguides if the relative
orientation between the magnetization direction and the magnon
wavevector changes.
along the wall [50]. By targeting this class of magnons in Néel walls,
we focus on the potential of using domain walls as nanometer-scaled
channels to open new perspectives for energy-efficient control of
magnon propagation in two dimensions.
The top part of Fig. 9.10a shows a schematic of a 180° Néel wall.
The magnetic moments rotate within the sample plane, giving rise
to magnetic volume charges, with opposite sign on the two sides of
the domain-wall center. The red curve in Fig. 9.10a displays these
volume charges as a function of position perpendicular to the domain
wall calculated by means of micromagnetic simulations. They create
a strong dipolar magnetic field (blue curve in Fig. 9.10a) oriented
antiparallel to the magnetization in the center of the domain wall,
resulting in a locally decreased effective magnetic field. Since the
magnon energy depends on this internal magnetic field, the domain
wall effectively forms a potential well for magnons comparable to
magnons localized near edges of magnetic elements [38, 51]. The
width of this potential well is given by the domain wall width and
strongly depends on various parameters of the magnetic material.
The Néel wall studied here has a width of about 40 nm. However,
note that it can be tuned towards even smaller sizes if other
materials are used. In the following, we demonstrate experimentally
and numerically that such a potential well hosts magnons that are
strongly localized to the domain-wall width but still travel freely
along the wall.
Figure 9.10b presents a scanning electron microscopy (SEM)
image of the structure used in the experiment. A 40 nm thick
permalloy film is patterned into a magnon waveguide that is 5 μm
wide at one end and gradually broadens until it reaches a constant
width of 10 μm. This variable width stabilizes the 180° Néel wall
under the microwave antenna. A domain wall parallel to the long axis
of the waveguide is initiated by applying a sinusoidal, exponentially
decaying magnetic field parallel to its short axis, as depicted in the
bottom right inset in Fig. 9.10b. Magneto-optical Kerr microscopy
was used to confirm the magnetic remanence state of the magnon
waveguide. The overlay in Fig. 9.10c shows a Kerr micrograph,
with the black–white color code representing the magnetization
component along the y direction. A Landau-like domain pattern is
formed with a 180° Néel wall in the center separating two domains
with opposite magnetization. Additional micromagnetic simulations
282 Steering Magnons by Noncollinear Spin Textures
color representing the field component along the short axis of the
rectangle. The data show strong effective fields across the domain
wall oriented antiparallel to the magnetization direction.
An out-of-plane field pulse with a Gaussian spatial profile was
used to locally excite spin dynamics at the domain-wall position
at a distance of 1.6 µm from the bottom edge (green dot in Fig.
9.12c). The subsequent analysis yields the magnon spectrum and
dispersion relation, which is discussed in the next paragraph.
To illustrate the mode profiles of the magnons, the response to
a continuous microwave excitation at four different frequencies
was simulated. In Fig. 9.12c, the normalized z component of the
magnetization is plotted for a given time once the system reached
a steady state. For the lower frequencies of 0.52, 1.28 and 2.16 GHz,
the magnons are strongly localized inside the domain wall. However,
a general trend is that the strong localization within the wall softens
with increasing frequency. Particularly for modes with higher
frequencies, for example 5.68 GHz, the radiation from the domain
wall indicates the onset of the magnon band in the domains and,
hence, loss of the channeling effect. Although the strong localization
for lower frequencies inside the domain wall is in good agreement
with the experiment, the simulated domain spectra appear at higher
frequencies than in the experiment. In general, magnons shift to
higher frequencies when reducing the domain size, which therefore
allows the guiding of magnons inside the wall, even for higher
frequencies and smaller wavelengths.
In addition to the spatial and spectral characteristics, the
micromagnetic simulations also shed light on the dispersion
relation. The wavelength for a given frequency—illustrated by the
green bars in Fig. 9.12c—can be determined by a Fourier analysis
of the dynamic magnetization along the domain wall. Figure 9.12d
shows the resulting positive dispersion that enables the transport
of information via magnons propagating within the domain wall.
Even though the magnons are confined transverse to the wall on a
length scale given by the domain wall width, the dispersion is mainly
dominated by dipolar energy. For the first three modes shown in
Fig. 9.12c, the dynamic part of the dipolar energy originating from
the magnons is by a factor of three to five larger than the dynamic
286 Steering Magnons by Noncollinear Spin Textures
Figure 9.13 Magnon intensities measured across the waveguide for different
external magnetic fields applied parallel to the long axis of the magnon
waveguide for 0.52 GHz. The measurements show that the domain wall
can be easily shifted with small applied fields over micrometer distances in
both directions, therefore allowing for fine control of the magnon channel
position. Inset: displacement of the magnon propagation path as a function
of applied magnetic field, yielding a proportionality constant of 5.57 µm per
mT. Image adapted from Wagner et al. [50].
Acknowledgments
The samples studied in the original publications were prepared at
the Center for Nanoscale Materials (CNM) and the Materials Science
Division (MSD) at the Argonne National Laboratory as well as at the
Nanofabrication Facilities (NanoFaRo) at the Institute for Ion Beam
Physics and Materials Research at the Helmholtz-Center Dresden-
Rossendorf (HZDR). Without these facilities and the commitment
of the people behind them we would not be able to conduct our
experimental research. Therefore, special thanks go to L. E. Ocola
and R. Divan (CNM), J. E. Pearson (MSD), and T. Schönherr, C. Neisser,
B. Scheumann, and A. Erbe (HZDR). Furthermore, we are grateful to
A. Henschke and T. Hula for their support in designing, building, and
maintaining the experimental setups.
The work discussed in this chapter was inspired by ideas,
contributions, and discussions from F. Y. Fradin, S. D. Bader, A.
Hoffmann, B. Hillebrands, T. Sebastian, J. Lindner, and J. Fassbender.
K.S. acknowledges funding from the Helmholtz
Postdoc Programme. Financial support by the Deutsche
Forschungsgemeinschaft is gratefully acknowledged within
programme SCHU2922/1-1.
References
8. Khitun, A., Bao, M., and Wang, K. L. (2010). Magnonic logic circuits, J.
Phys. D: Appl. Phys., 43, 264005.
9. Schneider, T., Serga, A., Leven, B., Hillebrands, B., Stamps, R. L., and
Kostylev, M. (2008). Realization of spin-wave logic gates, Appl. Phys.
Lett., 92, 022505.
10. Chumak, A. V., Serga, A. A., and Hillebrands, B. (2014). Magnon
transistor for all-magnon data processing, Nat. Commun., 5, 4700.
11. Damon, R. W., and Eshbach, J. R. (1961). Magnetostatic modes of a
ferromagnet slab, J. Phys. Chem. Solids, 19, 308.
12. Hurben, M. J., and Patton, C. E. (1995). Theory of magnetostatic waves
for inplane magnetized isotropic films, J. Magn. Magn. Mater., 139, 263.
13. Kalinikos, B., and Slavin, A. (1986). Theory of dipole-exchange spin
wave spectrum for ferromagnetic films with mixed exchange boundary
conditions, J. Phys.: Condens. Matter, 19, 7013.
14. Schultheiss, H., Schäfer, S., Candeloro, P., Leven, B., Hillebrands, B., and
Slavin, A. N. (2008). Observation of coherence and partial decoherence
of quantized spin waves in nanoscaled magnetic ring structures, Phys.
Rev. Lett., 100, 047204.
15. Vogt, K., Schultheiss, H., Hermsdoerfer, S., Pirro, P., Serga, A., and
Hillebrands, B. (2009). All-optical detection of phase fronts of
propagating spin waves in a Ni81Fe19 microstripe, Appl. Phys. Lett.,
95, 182508.
16. Demidov, V. E., Demokritov, S. O., Rott, K., Krzysteczko, P., and Reiss,
G. (2007). Self-focusing of spin waves in Permalloy microstripes, Appl.
Phys. Lett., 91, 252504.
17. Demidov, V. E., Demokritov, S. O., Rott, K., Krzysteczko, P., and Reiss, G.
(2008). Mode interference and periodic self-focusing of spin waves in
Permalloy microstripes, Phys. Rev. B, 77, 064406.
18. Pirro, P., Braecher, T., Vogt, K., Obry, B., Schultheiss, H., Leven, B., and
Hillebrands, B. (2011). Interference of coherent spin waves in micron-
sized ferromagnetic waveguides, Phys. Status Solidi B, 248, 2404.
19. Schultheiss, H., Pearson, J. E., Bader, S., and Hoffmann, A. (2012).
Thermoelectric detection of spin waves, Phys. Rev. Lett., 109, 237204.
20. Hertel, R., Wulfhekel, W., and Kirschner, J. (2004). Domain-wall induced
phase shifts in spin waves, Phys. Rev. Lett., 93, 257202.
21. Bayer, C., Schultheiss, H., Hillebrands, B., and Stamps, R. L. (2005).
Phase shift of spin waves traveling through a 180 degrees Bloch-
domain wall, IEEE Trans. Magn., 41, 3094.
References 291
22. Mühlbauer, S., Binz, B., Jonietz, F., Pfleiderer, C., Rosch, A., Neubauer,
A., Georgii, R., and Böni, P. (2009). Skyrmion lattice in chiral magnet,
Science, 323, 915.
23. Jiang, W., Upadhyaya, P., Zhang, W., Yu, G., Jungfleisch, M. B., Fradin, F.
Y., Pearson, J. E., Tserkovnyak, Y., Wang, K. L., Heinonen, O., te Velthuis,
S. G. E., and Hoffmann, A. (2015). Blowing magnetic skyrmion bubbles,
Science, 349, 283.
24. Büttner, F., Moutafis, C., Schneider, M., Krüger, B., Günther, C. M.,
Geilhufe, J., Korff Schmising, C. V., Mohanty, J., Pfau, B., Schaffert, S.,
Bisig, A., Foerster, M., Schulz, T., Vaz, C. A. F., Franken, J. H., Swagten,
H. J. M., Kläui, M., and Eisebitt, S. (2015). Dynamics and inertia of
skyrmionic spin structures, Nat. Phys., 11, 225.
25. Demokritov, S. O., Hillebrands, B., and Slavin, A. (2001). Brillouin
light scattering studies of confined spin waves: linear and nonlinear
confinement, Phys. Rep., 348, 441.
26. Kostylev, M. P., Gubbiotti, G., Hu, J. G., Carlotti, G., Ono, T., and Stamps, R.
L. (2007). Dipole-exchange propagating spin-wave modes in metallic
ferromagnetic stripes, Phys. Rev. B, 76, 054422.
27. Guslienko, K. Y., Demokritov, S. O., Hillebrands, B., and Slavin, A. (2002).
Effective dipolar boundary conditions for dynamic magnetization in
thin magnetic stripes, Phys. Rev. B, 66, 132402.
28. Sebastian, T., Ohdaira, Y., Kubota, T., Pirro, P., Brächer, T., Vogt, K., Serga,
A. A., Naganuma, H., Oogane, M., Ando, Y., Hillebrands, B., and Kubota,
T. (2012). Low-damping spin-wave propagation in a micro-structured
Co2Mn0.6Fe0.4Si Heusler waveguide, Appl. Phys. Lett., 100, 112402.
29. Sebastian, T., Brächer, T., Pirro, P., Serga, A. A., Hillebrands, B., Kubota,
T., Naganuma, H., Oogane, M., and Ando, Y. (2013). Nonlinear emission
of spin-wave caustics from an edge mode of a microstructured
Co2Mn0.6Fe0.4Si waveguide, Phys. Rev. Lett., 110, 067201.
30. Demokritov, S. O., and Demidov, V. E. (2008). Micro-Brillouin light
scattering spectroscopy of magnetic nanostructures, IEEE Trans.
Magn., 44, 6.
31. Sebastian, T., Schultheiss, K., Obry, B., Hillebrands, B., and Schultheiss,
H. (2015). Micro-focused Brillouin light scattering: imaging spin waves
at the nanoscale, Front. Phys., 3, 35.
32. Clausen, P., Vogt, K., Schultheiss, H., Schäfer, S., Obry, B., Wolf, G., Pirro,
P., Leven, B., and Hillebrands, B. (2011). Mode conversion by symmetry
breaking of propagating spin waves, Appl. Phys. Lett., 99, 162505.
292 Steering Magnons by Noncollinear Spin Textures
33. Vogt, K., Fradin, F. Y., Pearson, J. E., Sebastian, T., Bader, S. D., Hillebrands,
B., Hoffmann, A., and Schultheiss, H. (2014). Realization of a spin-wave
multiplexer, Nat. Commun., 5, 3727.
34. Davies, C. S., Francis, A., Sadovnikov, A. V., Chertopalov, S. V., Bryan,
M. T., Grishin, S. V., Allwood, D. A., Sharaevskii, Y. P., Nikitov, S. A., and
Kruglyak, V. V. (2015). Towards graded-index magnonics: steering
spin waves in magnonic networks, Phys. Rev. B, 92, 020408.
35. Yu, H., Duerr, G., Huber, R., Bahr, M., Schwarze, T., Brandl, F., and
Grundler, D. (2013). Omnidirectional spin-wave nanograting coupler,
Nat. Commun., 4, 2702.
36. Vogt, K., Schultheiss, H., Jain, S., Pearson, J. E., Hoffmann, A., Bader, S.
D., and Hillebrands, B. (2012). Spin waves turning a corner, Appl. Phys.
Lett., 101, 042410.
37. Sekiguchi, K., Yamada, K., Seo, S. M., Lee, K. J., Chiba, D., Kobayashi, K.,
and Ono, T. (2012). Time-domain measurement of current-induced
spin wave dynamics, Phys. Rev. Lett., 108, 017203.
38. Jorzick, J., Demokritov, S. O., Hillebrands, B., Bailleul, M., Fermon, C.,
Guslienko, K. Y., Slavin, A. N., Berkov, D. V., and Gorn, N. L. (2002). Spin
wave wells in nonellipsoidal micrometer size magnetic elements, Phys.
Rev. Lett., 88, 047204.
39. Demidov, V. E., Kostylev, M. P., Rott, K., Münchenberger, J., Reiss, G., and
Demokritov, S. O. (2011). Excitation of short-wavelength spin waves in
magnonic waveguides, Appl. Phys. Lett., 99, 082507.
40. Obry, B., Vasyuchka, V. I., Chumak, A. V., Serga, A. A., and Hillebrands,
B. (2012). Spin-wave propagation and transformation in a thermal
gradient, Appl. Phys. Lett., 101, 192406.
41. Stigloher, J., Decker, M., Körner, H. S., Tanabe, K., Moriyama, T., Taniguchi,
T., Hata, H., Madami, M., Gubbiotti, G., Kobayashi, K., Ono, T., and Back,
C. H. (2016). Snell’s law for spin waves, Phys. Rev. Lett., 117, 037204.
42. Toedt, J. N., Mansfeld, S., Mellem, D., Hansen, W., Heitmann, D., and
Mendach, S. (2016). Interface modes at step edges of media with
anisotropic dispersion, Phys. Rev. B, 93, 184416.
43. Toedt, J.-N., Mundkowski, M., Heitmann, D., Mendach, S., and Hansen,
W. (2016). Design and construction of a spin-wave lens, Sci. Rep., 6,
33169.
44. Schoen, M. A. W., Thonig, D., Schneider, M. L., Silva, T. J., Nembach, H.
T., Eriksson, O., Karis, O., and Shaw, J. M. (2016). Ultra-low magnetic
damping of a metallic ferromagnet, Nat. Phys., 12, 839.
45. Bali, R., Wintz, S., Meutzner, F., Hübner, R., Boucher, R., Ünal, A. A.,
Valencia, S., Neudert, A., Potzger, K., Bauch, J., Kronast, F., Facsko, S.,
References 293
58. Kozhevnikov, A., Gertz, F., Dudko, G., Filimonov, Y., and Khitun, A. (2015).
Pattern recognition with magnonic holographic memory device, Appl.
Phys. Lett., 106, 142409.
59. Vogel, A., Wintz, S., Kimling, J., Bolte, M., Strache, T., Fritzsche, M., Im, M.
Y., Fischer, P., Meier, G., and Fassbender, J. (2010). Domain-wall pinning
and depinning at soft spots in magnetic nanowires, IEEE Trans. Magn.,
46, 1708.
60. Fassbender, J., Strache, T., Liedke, M. O., Marko, D., Wintz, S., Lenz, K.,
Keller, A., Facsko, S., Mönch, I., and McCord, J. (2009). Introducing
artificial length scales to tailor magnetic properties, New J. Phys., 11,
125002.
61. Garcia-Sanchez, F., Borys, P., Soucaille, R., Adam, J. P., Stamps, R. L.,
and Kim, J.-V. (2015). Narrow magnonic waveguides based on domain
walls, Phys. Rev. Lett., 114, 247206.
62. Borys, P., Garcia-Sanchez, F., Kim, J.-V., and Stamps, R. L. (2016). Spin‐
wave eigenmodes of Dzyaloshinskii domain walls, Adv. Electron. Mater.,
2, 1500202.
63. Xing, X., Pong, P. W. T., Akerman, J., and Zhou, Y. (2017). Paving spin-
wave fibers in magnonic nanocircuits using spin-orbit torque, Phys.
Rev. Appl., 7, 054016.
64. Jiang, W., Upadhyaya, P., Zhang, W., Yu, G., Jungfleisch, M. B., Fradin, F.
Y., Pearson, J. E., Tserkovnyak, Y., Wang, K. L., Heinonen, O., te Velthuis,
S. G. E., and Hoffmann, A. (2015). Blowing magnetic skyrmion bubbles,
Science, 349, 6245.
65. Zakeri, K., Zhang, Y., Prokop, J., Chuang, T. H., Sakr, N., Tang, W. X., and
Kirschner, J. (2010). Asymmetric spin-wave dispersion on Fe(110):
direct evidence of the Dzyaloshinskii-Moriya interaction, Phys. Rev.
Lett., 104, 137203.
66. Nembach, H. T., Shaw, J. M., Weiler, M., Jué, E., and Silva, T. J. (2015).
Linear relation between Heisenberg exchange and interfacial
Dzyaloshinskii-Moriya interaction in metal films, Nat. Phys., 11, 825.
67. Wintz, S., Tiberkevich, V., weigand, M., Raabe, J., Lindner, J., Erbe,
A., Slavin, A. N., and Fassbender, J. (2016). Magnetic vortex cores as
tunable spin-wave emitters, Nat. Nanotechnol., 11, 948.
68. Otálora, J. A., Yan, M., Schultheiss, H., Hertel, R., and Kákay, A. (2016).
Curvature-induced asymmetric spin-wave dispersion, Phys. Rev. Lett.,
117, 227203.
69. Schütte, C., and Garst, M. (2014). Magnon-skyrmion scattering in chiral
magnets, Phys. Rev. B, 90, 094423.
Chapter 10
10.1 Introduction
Since the discovery of spin transfer torque (STT) in the late
1990s, a number of current-induced effects have been explored
experimentally, interpreted theoretically, and, for some of them,
even put in application. Let us mention current-induced domain
wall motion in nanostrips and current-induced magnetic switching
or current-induced magnetic oscillations in magnetoresistive
nanopillars or nanocontacts. All these effects involve the flow of
spin-polarized electrons (the so-called spin current) in single-
or multilayer ferromagnets having a non-uniform magnetic
configuration. A spin wave, whose instantaneous magnetization
distribution provides such a non-uniform magnetic configuration,
is also subjected to spin transfer. In this case, the STT translates
into a simple shift of the spin wave frequency. This shift can be
identified as a Doppler effect. Indeed, everything happens as if the
whole spin system were drifting at an effective velocity u = Q/Ms,
where Q is the spin current expressed in Bohr magneton per unit
time and MS is the saturation magnetization. This so-called current-
induced spin wave Doppler shift (CISWDS) had been predicted
theoretically in 1966. It has been observed experimentally in
2008. Since that time, the CISWDS has proven to be a relevant
tool for measuring spin-polarized current in different materials
and for different experimental conditions. In 2012, another effect,
namely the current-induced modification of spin wave relaxation,
has been observed that is related to a non-adiabatic correction
of the standard (adiabatic) STT.
In this chapter, we review the work done on current-induced
modification of spin wave dynamics. For this purpose, we first
discuss the general idea of a Doppler effect for spin waves and
derive the expression of the frequency shift from the standard
equations of the STT (Section 10.2). Then, we present the three
experimental approaches that have been used to observe the
effect (Section 10.3) and discuss how such measurements can be
used to extract spin-dependent resistivities, which are essential
ingredients in the description of spin-polarized diffusive electrical
A Doppler Shift for Spin Waves 297
Figure 10.1 (a) Sketches illustrating the Doppler effect for water waves
and spin waves in a drifting medium. (Top) Water surface waves
generated by the oscillation of a wake maker. (Middle) Same in the
presence of a water current. (Bottom) Spin waves generated locally
in a magnetic medium subjected to a global drift. (b) Calculated spin
wave dispersion in the absence of drift (full line) and for large drift
velocities (dashed lines). The long dashed line shows a hypothetical
situation in which the electric current is strong enough to induce
instability of the uniform ground state. After McMichael and Stiles [1]
and Fernandez-Rossier et al. [4].
dM M dM M
= – 0 gM × Heff + a × –(u . )M + b [(u . )M)],
dt Ms dt Ms
(10.1)
– g B JP
u= , (10.2)
2 Ms | e |
iw bu . k
i( w – u . k )m = – 0 g(m × Heq + Meq × h )+ a – M × m ,
MS w eq
(10.3)
where Heq and h are the equilibrium and dynamic parts of the
effective field. Interestingly, one can note that the here-above
equation is similar to the spin wave equation in the absence of
the STT if one proceeds to the following substitutions [8, 9]:
bu . k
a a– (non-adiabatic STT). (10.5)
w
DVG = u. (10.6)
306 Current-Induced Spin Wave Doppler Shift
scanning the laser spot over the permalloy stripe, an image of the
excited spin wave mode is thus acquired (Figs. 10.5 and 10.6a–c).
Alternatively, the position of the laser spot can be kept at a
given position, while the magnitude of the static magnetic field
is swept across the different resonance conditions, allowing
therefore to probe locally spin wave resonance spectra (not
shown).
The analysis of a spin wave line scan (see, for example, inset
of Fig. 10.5) leads to a direct access to its spatial characteristics:
the real and imaginary parts of the k-vector. A full characterization
of spin wave modes is performed, including spin wave mode
imaging, which enables the evaluation of important additional
quantities such as group velocity (Fig. 10.6f). Similarly to
the inductive studies, the permalloy stripe is subjected to an
electric current (±1.2 mA, equivalent to ±0.935 × 1011 A/m²). A
TR-MOKE study will thereby aim at addressing the
current-induced changes in the real and imaginary parts of the
Parametrizing the Two-Current Model 309
k-vector [25]. Note that the case of the real part of the k-vector
is the exact realization of McMichael and Stiles’s picture of
the spin wave Doppler effect (see Fig. 10.1a) [1]. The current-
induced shift in the resonant magnetic field was also measured;
however, unlike PSWS, TR-MOKE is not suitable to measure
frequency shifts accurately. Indeed, because of the stroboscopic
approach, the excitation frequency has to be a multiple of the
laser repetition rate, that is, 80 MHz. The frequency resolution
is therefore limited to this value. The current-induced shift of the
real part of the spin wave vector can be related to the spin
drift velocity (u) as follows:
VG Dk
u= Re , (10.7)
2 k
k Dk
Im
w k (10.10)
b= + a,
k Dk
Re
gu0H k
Dk
Im
–1 k Im( k )t
b – + a, (10.11)
k Dk 1 – Re( k )t
VG Re Re
gì 0H k
The so-called MSSWs, that is, spin waves with a wave vector
perpendicular to the equilibrium magnetization, both vectors
lying in the film plane, have specific non-reciprocal properties,
in that the amplitudes and sometimes also the frequencies of two
counterpropagating spin waves are different. This is illustrated
in Fig. 10.10c, which shows the two counterpropagating PSWS
signals measured on a 10 nm permalloy film at a wave vector
of +/–7.8 rad/µm. One recognizes clearly a large difference
of amplitude of the two waveforms, the k < 0 one being about
four times more intense than the k > 0 one. One recognizes also
a clear frequency shift, the k < 0 waveform lying 34 MHz higher
in frequency than the k > 0 one.
8 g K Sbot – K Stop k .
fNR (10.12)
p3 MS L2 p2
1+ 2
t
K bot – K Stop and
8 gHere 8kg K bot –are
K Stop the ksurface magnetic anisotropy
.
fNR 3 S fNR 32 2S .
constants
p MS for theL p top
p M and
S
bottomL2 pfilm
2
surfaces, respectively,
1
and L is the exchange+ 2
1 +
length.t Note that the interfacial
2
t
Dzyaloshinskii–Moriya interaction, of importance in ferromagnetic
metal/heavy metal bilayers, also results in an MSSW frequency
non-reciprocity (see chapter 8 by Stamps et al. for a discussion
of the specificities of spin waves in the presence of this interaction).
Note also that both amplitude and frequency non-reciprocities
reverse when the direction of the equilibrium magnetization
is reversed. Consequently, these two effects can be ruled out
by comparing (+k, +H) and (–k, –H) spectra, as done by Zhu,
Thomas, Sugimoto, and coworkers [13–15, 18].
Other Types of Spin Wave Frequency Shifts 321
4 2 Jt P –P
dW = Q01 00 11 . (10.13)
2
p Ms W– W
References
17. Haidar, M., Bailleul, M., Kostylev, M., and Lao, Y. (2014).
Nonreciprocal Oersted field contribution to the current-induced
frequency shift of magnetostatic surface waves, Phys. Rev. B, 89,
094426.
18. Sugimoto, S., Rosamond, M., Linfield, E. H., and Marrows, C. H.
(2016). Observation of spin-wave Doppler shift in Co90Fe10/Ru
microstrips for evaluating spin polarization, Appl. Phys. Lett., 109,
122405.
19. Covington, M., Crawford, T. M., and Parker, G. J. (2002). Time-resolved
measurement of propagating spin waves in ferromagnetic thin
films, Phys. Rev. Lett., 89, 237202.
20. Sekiguchi, K., et al. (2012). Time-domain measurement of current-
induced spin wave dynamics, Phys. Rev. Lett., 108, 017203.
21. Hubert, A., and Schäfer, R. (1998). Magnetic Domains: The Analysis
of Magnetic Microstructures (Springer-Verlag, Berlin, Heidelberg).
22. Hiebert, W. K., Stankiewicz, A., and Freeman, M. R. (1997). Direct
observation od magnetic relaxation in a small Permalloy disk by
time-resolved scanning kerr microscopy, Phys. Rev. Lett., 79, 1134.
23. Liu, Z., Giesen, F., Zhu, X., Sydora, R. D., and Freeman, M. R. (2007).
Spin wave dynamics and the determination of intrinsic damping
in locally excited Permalloy thin films, Phys. Rev. Lett., 98, 087201.
24. Perzlmaier, K., Woltersdorf, G., and Back, C. H. (2008). Observation
of the propagation and interference of spin waves in ferromagnetic
thin films, Phys. Rev. B, 77, 054425.
25. Chauleau, J.-Y., et al. (2014). Self-consistent determiantion of the
key spin-transfer torque parameters from spin-wave Doppler
experiments, Phys. Rev. B, 89, R020403.
26. Fert, A., and Campbell, I. A. (1976). Electrical resistivity of
ferromagnetic nickel and iron based alloys, J. Phys. F: Metal Phys., 5,
849.
27. Bass, J., and Pratt Jr., W. P. (1999). Current-perpendicular (CPP)
magnetoresistance in magnetic metallic multilayers, J. Magn. Magn.
Mater., 200, 274.
28. Mertig, I. (1999). Transport properties of dilute alloys, Rep. Prog. Phys.,
62, 237.
29. Starikov, A. A., Kelly, P. J., Brataas, A., Tserkovnyak, Y., and Bauer,
G. E. W. (2010). Unified first-principles study of Gilbert damping,
spin-flip diffusion, and resistivity in transition metal alloys, Phys.
Rev. Lett., 105, 236601.
328 Current-Induced Spin Wave Doppler Shift
11.1 Introduction
The unique features of spin waves, such as the possibility to achieve
submicrometer wavelength at microwave frequencies and electronic
controllability by static magnetic fields make these waves uniquely
suited for implementation of novel integrated electronic devices
characterized by high speed, low power consumption, and extended
functionalities. The utilization of spin waves for integrated electronic
applications is addressed within the emerging field of magnonics
[1–8]. Although the application of spin waves for microwave signal
processing has been intensively explored for many decades (see,
e.g., [9, 10]), recent advances in spintronics and nanomagnetism,
as well as the development of novel techniques for nanofabrication
and measurements of high-frequency magnetization dynamics,
Figure 11.3 (a) Measured map of the intensity of spin waves emitted by the
STNO. The positions of the STNO and the CoFe nanowire are schematically
shown. Spatial decay is numerically compensated. (b) Dependence of the
transverse width of the spin-wave intensity distribution on the propagation
coordinate. Symbols are experimental data, horizontal line is the mean
value. (c) Dependence of the integral spin-wave intensity on the propagation
coordinate. Symbols are experimental data, line is the exponential fit. ©
(2015) IEEE. Reprinted, with permission, from [6].
Figure 11.6 (a) Dependences of the BLS intensity on the propagation coordinate
recorded for I = 0 and 10 mA by moving the laser spot along the waveguide
axis. Excitation frequency is 6.4 GHz. Symbols are experimental data, lines are
exponential fits (note the logarithmic scale of the vertical axis). (b) The ratio
between the spin-wave propagation lengths measured at I = 10 mA and I = 0
vs. the excitation frequency. f0 marks the frequency of the quasi-uniform
ferromagnetic resonance. © (2015) IEEE. Reprinted, with permission, from
[6].
Figure 11.7 (a) Normalized spatial intensity map of the propagating spin
wave excited by the antenna. The map was recorded for I = 2.55 mA. The
mapping was performed by rastering the probing spot over the area 1.6
by 10 µm, which is larger than the waveguide width of 1 µm. Dashed lines
show the edges of the waveguide. Inset shows the transverse profile of the
spin-wave intensity. (b) Dependences of the spin-wave intensity on the
propagation coordinate for different currents, as labeled, in the log-linear
scale. Lines show the exponential fit of the experimental data. Reprinted
from [21], with the permission of AIP Publishing.
Figure 11.8 Current dependences of the propagation length and of the decay
constant. Vertical dashed line marks the critical current IC, at which the
damping is completely compensated by the spin current. Solid line is the
linear fit of the experimental data at I < IC. The data were obtained at H0 =
1000 Oe. Reprinted from [21], with the permission of AIP Publishing.
100
10
1.0 1.5 2.0 2.5
I, mA
Figure 11.9 Current dependence of the intensity of the spin wave at the output
of a 10 µm long transmission line calculated based on the experimentally
measured propagation length. The intensity is normalized by the value at I
= 0. Reprinted from [21], with the permission of AIP Publishing.
Figure 11.11 (a) Solid lines: dispersion spectra of spin waves in an extended
Py film with the thickness of 5 and 25 nm, as labeled. Symbols: dispersion
spectrum of a spin-wave mode in a 500 nm-wide and 20 nm-thick stripe
waveguide manufactured on top of 5 nm-thick Py film. Calculations were
performed for H0 = 1000 Oe. Dashed horizontal line marks the FMR
frequency. (b) Current dependences of the auto-oscillation frequency
(point-down triangles) and the intensity (point-up triangles) of the NLSI
nano-oscillator. H0 = 1000 Oe. Reprinted from [76], with the permission
from Elsevier.
350 Excitation and Amplification of Propagating Spin Waves by Spin Currents
The maps shown in Fig. 11.14 demonstrate that the pulse of the
current-induced magnetization precession of the NLSI oscillator
efficiently couples to spin waves in the waveguide, producing a
propagating spin-wave packet. At t = 1.6 ns, a dynamic signal emerges
in the waveguide near its edge facing the NLSI oscillator. At t = 2.4
ns, the increased-intensity region spreads away from the oscillator,
indicating the propagation of the leading front of the spin-wave
354 Excitation and Amplification of Propagating Spin Waves by Spin Currents
Figure 11.15 (a) Temporal profile of the wave packet at x = 0, fitted by the
Gaussian function. (b) Propagation-coordinate dependence of the temporal
width of the spin-wave packet. The data were obtained at H0 = 1000 Oe. The
width of the driving-current pulse is 3 ns and its amplitude is 7 mA. The
spin-wave frequency is 8 GHz. Reprinted from [8], with the permission of
AIP Publishing.
11.6 Conclusions
In conclusion, we would like to emphasize the large importance
of the advancements in current-induced excitation and control of
spin waves for the development of the research field of magnonics,
which until recently has evolved independently from the field of
spin-torque phenomena. Although the limitations imposed by
the geometry of traditional spin-torque devices have discouraged
researchers from using them in magnonic circuits, we believe that
recent advances in studies of pure spin currents will dramatically
accelerate the integration of spin-torque and magnonic devices. An
additional encouraging benefit of pure spin currents for magnonics
is the possibility to excite and control spin waves in magnetic
insulators such as yttrium iron garnet, which is presently viewed as
the most suitable material for future nanomagnonic circuits.
Acknowledgments
We would like to acknowledge A. Anane, J. Ben Youssef, V. Bessonov,
P. Bortolotti, M. Collet, V. Cros, B. Divinskiy, M. Evelt, K. Garcia-
Hernandez, T. Kendziorczyk, O. Klein, T. Kuhn, J. Leuthold, R. Liu,
G. de Loubens, M. Munoz, V. V. Naletov, J. L. Prieto, G. Reiss, A. B.
356 Excitation and Amplification of Propagating Spin Waves by Spin Currents
References
14. Hirsch, J. E. (1999). Spin Hall effect, Phys. Rev. Lett., 83, 1834–1837.
15. Demidov, V. E., Urazhdin, S., and Demokritov, S. O. (2010). Direct
observation and mapping of spin waves emitted by spin-torque nano-
oscillators, Nat. Mater., 9, 984–988.
16. Madami, M., Bonetti, S., Consolo, G., Tacchi, S., Carlotti, G., Gubbiotti,
G., Mancoff, F. B., Yar, M. A., and Åkerman, J. (2011). Direct observation
of a propagating spin wave induced by spin-transfer torque, Nat.
Nanotechnol., 6, 635–638.
17. Wang, Z., Sun, Y., Wu, M., Tiberkevich, V., and Slavin, A. (2011). Control
of spin waves in a thin film ferromagnetic insulator through interfacial
spin scattering, Phys. Rev. Lett., 107, 146602.
18. Padron-Hernandez, E., Azevedo, A., and Rezende, S. M. (2011).
Amplification of spin waves in yttrium iron garnet films through the
spin Hall effect, Appl. Phys. Lett., 99, 192511.
19. Demidov, V. E., Urazhdin, S., Rinkevich, A. B., Reiss, G., and Demokritov,
S. O. (2014). Spin Hall controlled magnonic microwaveguides, Appl.
Phys. Lett., 104, 152402.
20. An, K., Birt, D. R., Pai, C.-F., Olsson, K., Ralph, D. C., Buhrman, R. A., and Li,
X. (2014). Control of propagating spin waves via spin transfer torque
in a metallic bilayer waveguide, Phys. Rev. B, 89, 140405(R).
21. Evelt, M., Demidov, V. E., Bessonov, V., Demokritov, S. O., Prieto, J. L.,
Muñoz, M., Ben Youssef, J., Naletov, V. V., de Loubens, G., Klein, O., Collet,
M., Garcia-Hernandez, K., Bortolotti, P., Cros, V., and Anane, A. (2016).
High-efficiency control of spin-wave propagation in ultra-thin yttrium
iron garnet by the spin-orbit torque, Appl. Phys. Lett., 108, 172406.
22. Demidov, V. E., Urazhdin, S., Liu, R., Divinskiy, B., Telegin, A., and
Demokritov, S. O. (2016). Excitation of coherent propagating spin
waves by pure spin currents, Nat. Commun., 7, 10446.
23. Ralph, D. C., and Stiles, M. D. (2008). Spin transfer torques, J. Magn.
Magn. Mater., 320, 1190–1216.
24. Brataas, A., Kent, A. D., and Ohno, H. (2012). Current-induced torques
in magnetic materials, Nat. Mater., 11, 372–381.
25. Locatelli, N., Cros, V., and Grollier, J. (2104). Spin-torque building
blocks, Nat. Mater., 13, 11–20.
26. Chen, T., Dumas, R. K., Eklund, A., Muduli, P. K., Houshang, A., Awad, A.
A., Dürrenfeld, P., Malm, B. G., Rusu, A., and Åkerman, J. (2016). Spin-
torque and spin-Hall nano-oscillators, Proc. IEEE, 104, 1919–1945.
358 Excitation and Amplification of Propagating Spin Waves by Spin Currents
27. Kajiwara, Y., Harii, K., Takahashi, S., Ohe, J., Uchida, K., Mizuguchi, M.,
Umezawa, H., Kawai, H., Ando, K., Takanashi, K., Maekawa, S., and
Saitoh, E. (2010). Transmission of electrical signals by spin-wave
interconversion in a magnetic insulator, Nature, 464, 262–266.
28. Cherepanov, V., Kolokolov, I., and L’vov, V. (1993). The saga of YIG:
spectra, thermodynamics, interaction and relaxation of magnons in a
complex magnet, Phys. Rep., 229, 81–144.
29. Jedema, F. J., Filip, A. T., and van Wees, B. J. (2001). Electrical spin
injection and accumulation at room temperature in an all-metal
mesoscopic spin valve, Nature, 410, 345–348.
30. Otani, Y., and Kimura, T. (2011). Manipulation of spin currents in
metallic systems, Philos. Trans. R. Soc. London, Ser. A, 369, 3136–3149.
31. Xue, L., Wang, C., Cui, Y.-T., Liu, L., Swander, A., Sun, J. Z., Buhrman, R.
A., and Ralph, D. C. (2012). Resonance measurement of nonlocal spin
torque in a three-terminal magnetic device, Phys. Rev. Lett., 108,
147201.
32. Demokritov, S. O., Hillebrands, B., and Slavin, A. N. (2001). Brillouin
light scattering studies of confined spin waves: linear and nonlinear
confinement, Phys. Rep., 348, 441–489.
33. Demidov, V. E., Demokritov, S. O., Hillebrands, B., Laufenberg, M., and
Freitas, P. P. (2004). Radiation of spin waves by a single micrometer-
sized magnetic element, Appl. Phys. Lett., 85, 2866–2868.
34. Demokritov, S. O., and Demidov, V. E. (2008). Micro-Brillouin light
scattering spectroscopy of magnetic nanostructures, IEEE Trans.
Magn., 44, 6–12.
35. Demidov, V. E., Kostylev, M. P., Rott, K., P. Krzysteczko, Reiss, G., and
Demokritov, S. O. (2009). Excitation of microwaveguide modes by a
stripe antenna, Appl. Phys. Lett., 95, 112509.
36. Demidov, V. E., Kostylev, M. P., Rott, K., Münchenberger, J., Reiss, G., and
Demokritov, S. O. (2011). Excitation of short-wavelength spin waves in
magnonic waveguides, Appl. Phys. Lett., 99, 082507.
37. Gubbiotti, G., Carlotti, G., Madami, M., Tacchi, S., Vavassori, P., and
Socino, G. (2009). Setup of a new Brillouin light scattering apparatus
with submicrometric lateral resolution and its application to the study
of spin modes in nanomagnets, J. Appl. Phys., 105, 07D521.
38. Demidov, V. E., Urazhdin, S., Ulrichs, H., Tiberkevich, V., Slavin, A.,
Baither, D., Schmitz, G., and Demokritov, S. O. (2012). Magnetic nano-
oscillator driven by pure spin current, Nat. Mater., 11, 1028–1031.
References 359
39. Jersch, J., Demidov, V. E., Fuchs, H., Rott, K., Krzysteczko, P.,
Münchenberger, J., Reiss, G., and Demokritov, S. O. (2010). Mapping of
localized spin-wave excitations by near-field Brillouin light scattering,
Appl. Phys. Lett., 97, 152502.
40. Demidov, V. E., Urazhdin, S., and Demokritov, S. O. (2009). Control of
spin-wave phase and wavelength by electric current on the microscopic
scale, Appl. Phys. Lett., 95, 262509.
41. Slavin, A., and Tiberkevich, V. (2005). Spin wave mode excited by
spin-polarized current in a magnetic nanocontact is a standing self-
localized wave bullet, Phys. Rev. Lett., 95, 237201.
42. Berkov, D. V., Boone, C. T., and Krivorotov, I. N. (2011). Micromagnetic
simulations of magnetization dynamics in a nanowire induced by a
spin-polarized current injected via a point contact, Phys. Rev. B, 83,
054420.
43. Consolo, G., Lopez-Diaz, L., Azzerboni, B., Krivorotov, I., Tiberkevich,
V., and Slavin, A. (2013). Excitation of spin waves by a current-driven
magnetic nanocontact in a perpendicularly magnetized waveguide,
Phys. Rev. B, 88, 014417.
44. Demidov, V. E., Urazhdin, S., Tiberkevich, V., Slavin, A., and Demokritov,
S. O. (2011). Control of spin-wave emission from spin-torque nano-
oscillators by microwave pumping, Phys. Rev. B, 83, 060406(R).
45. Ulrichs, H., Demidov, V. E., Demokritov, S. O., and Urazhdin, S. (2012).
Spin-torque nano-emitters for magnonic applications, Appl. Phys. Lett.,
100, 162406.
46. Ando, K., Takahashi, S., Harii, K., Sasage, K., Ieda, J., Maekawa, S., and
Saitoh, E. (2008). Electric manipulation of spin relaxation using the
spin Hall effect, Phys. Rev. Lett., 101, 036601.
47. Liu, L., Moriyama, T., Ralph, D. C., and Buhrman, R. A. (2011). Spin-
torque ferromagnetic resonance induced by the spin Hall effect, Phys.
Rev. Lett., 106, 036601.
48. Demidov, V. E., Urazhdin, S., Edwards, E. R. J., and Demokritov, S. O.
(2011). Wide-range control of ferromagnetic resonance by spin Hall
effect, Appl. Phys. Lett., 99, 172501.
49. Hamadeh, A., d’Allivy Kelly, O., Hahn, C., Meley, H., Bernard, R.,
Molpeceres, A. H., Naletov, V. V., Viret, M., Anane, A., Cros, V., Demokritov,
S. O., Prieto, J. L., Muñoz, M., de Loubens, G., and Klein, O. (2014). Full
control of the spin-wave damping in a magnetic insulator using spin-
orbit torque, Phys. Rev. Lett., 113, 197203.
360 Excitation and Amplification of Propagating Spin Waves by Spin Currents
50. Demidov, V. E., Demokritov, S. O., Rott, K., Krzysteczko, P., and Reiss, G.
(2008). Mode interference and periodic self-focusing of spin waves in
Permalloy microstripes, Phys. Rev. B, 77, 064406.
51. Demidov, V. E., Demokritov, S. O., Rott, K., Krzysteczko, P., and Reiss, G.
(2008). Nano-optics with spin waves at microwave frequencies, Appl.
Phys. Lett., 92, 232503.
52. Sun, Y., Song, Y.-Y., Chang, H., Kabatek, M., Jantz, M., Schneider, W., Wu,
M., Schultheiss, H., and Hoffmann, A. (2012). Growth and ferromagnetic
resonance properties of nanometer-thick yttrium iron garnet films,
Appl. Phys. Lett., 101, 152405.
53. d’Allivy Kelly, O., Anane, A., Bernard, R., Ben Youssef, J., Hahn, C.,
Molpeceres, A. H., Carrétéro, C., Jacquet, E., Deranlot, C., Bortolotti, P.,
Lebourgeois, R., Mage, J.-C., de Loubens, G., Klein, O., Cros, V., and Fert,
A. (2013). Inverse spin Hall effect in nanometer-thick yttrium iron
garnet/Pt system, Appl. Phys. Lett., 103, 082408.
54. Yu, H., d’Allivy Kelly, O., Cros, V., Bernard, R., Bortolotti, P., Anane,
A., Brandl, F., Huber, R., Stasinopoulos, I., and Grundler, D. (2014).
Magnetic thin-film insulator with ultra-low spin wave damping for
coherent nanomagnonics, Sci. Rep., 4, 6848.
55. Hauser, C., Richter, T., Homonnay, N., Eisenschmidt, C., Qaid, M., Deniz,
H., Hesse, D., Sawicki, M., Ebbinghaus, S. G., and Schmidt, G. (2016).
Yttrium iron garnet thin films with very low damping obtained by
recrystallization of amorphous material, Sci. Rep., 6, 20827.
56. Collet, M., de Milly, X., d’Allivy Kelly, O., Naletov, V. V., Bernard, R.,
Bortolotti, P., Ben Youssef, J., Demidov, V. E., Demokritov, S. O., Prieto, J.
L., Munoz, M., Cros, V., Anane, A., de Loubens, G., and Klein, O. (2016).
Generation of coherent spin-wave modes in yttrium iron garnet
microdiscs by spin–orbit torque, Nat. Commun., 7, 10377.
57. Demidov, V. E., Evelt, M., Bessonov, V., Demokritov, S. O., Prieto, J. L.,
Muñoz, M., Ben Youssef, J., Naletov, V. V., de Loubens, G., Klein, O., Collet,
M., Bortolotti, P., Cros, V., and Anane, A. (2016). Direct observation of
dynamic modes excited in a magnetic insulator by pure spin current,
Sci. Rep., 6, 32781.
58. Gurevich, A. G., and Melkov, G. A. (1996). Magnetization Oscillations
and Waves (CRC, New York, USA).
59. Demidov, V. E., Urazhdin, S., Edwards, E. R. J., Stiles, M. D., McMichael,
R. D., and Demokritov, S. O. (2011). Control of magnetic fluctuations by
spin current, Phys. Rev. Lett., 107, 107204.
References 361
60. Liu, L., Pai, C.-F., Ralph, D. C., and Buhrman, R. A. (2012). Magnetic
oscillations driven by the spin Hall effect in 3-terminal magnetic
tunnel junction devices, Phys. Rev. Lett., 109, 186602.
61. Liu, R. H., Lim, W. L., and Urazhdin, S. (2013). Spectral characteristics
of the microwave emission by the spin Hall nano-oscillator, Phys. Rev.
Lett., 110, 147601.
62. Demidov, V. E., Ulrichs, H., Gurevich, S. V., Demokritov, S. O., Tiberkevich,
V. S., Slavin, A. N., Zholud, A., and Urazhdin, S. (2014). Synchronization
of spin Hall nano-oscillators to external microwave signals, Nat.
Commun., 5, 3179.
63. Demidov, V. E., Urazhdin, S., Zholud, A., Sadovnikov, A. V., and
Demokritov, S. O. (2014). Nanoconstriction-based spin-Hall nano-
oscillator, Appl. Phys. Lett., 105, 172410.
64. Duan, Z., Smith, A., Yang, L., Youngblood, B., Lindner, J., Demidov, V. E.,
Demokritov, S. O., and Krivorotov, I. N. (2014). Nanowire spin torque
oscillator driven by spin orbit torques, Nat. Commun., 5, 5616.
65. Liu, R. H., Lim, W. L., and Urazhdin, S. (2015). Dynamical skyrmion
state in a spin current nano-oscillator with perpendicular magnetic
anisotropy, Phys. Rev. Lett., 114, 137201.
66. Awad, A., Dürrenfeld, P., Houshang, A., Dvornik, M., Iacocca, E., Dumas,
R., and Åkerman, J. (2017). Long-range mutual synchronization of spin
Hall nano-oscillators, Nat. Phys., 13, 292–299.
67. Demidov, V. E., Urazhdin, S., Zholud, A., Sadovnikov, A. V., Slavin, A. N.,
and Demokritov, S. O. (2015). Spin-current nano-oscillator based on
nonlocal spin injection, Sci. Rep., 5, 8578.
68. Demidov, V. E., Urazhdin, S., Divinskiy, B., Rinkevich, A. B., and
Demokritov, S. O. (2015). Spectral linewidth of spin-current nano-
oscillators driven by nonlocal spin injection, Appl. Phys. Lett., 107,
202402.
69. Urazhdin, S., Demidov, V. E., Cao, R., Divinskiy, B., Tyberkevych, V.,
Slavin, A., Rinkevich, A. B., and Demokritov, S. O. (2016). Mutual
synchronization of nano-oscillators driven by pure spin current, Appl.
Phys. Lett., 109, 162402.
70. Demidov, V. E., Urazhdin, S., Zholud, A., Sadovnikov, A. V., and
Demokritov, S. O. (2015). Dipolar field-induced spin-wave waveguides
for spin-torque magnonics, Appl. Phys. Lett., 106, 022403.
71. Satoh, T., Terui, Y., Moriya, R., Ivanov, B. A., Ando, K., Saitoh, E., Shimura,
T., and Kuroda, K. (2012). Directional control of spin-wave emission by
spatially shaped light, Nat. Photonics, 6, 662–666.
362 Excitation and Amplification of Propagating Spin Waves by Spin Currents
72. Au, Y., Dvornik, M., Davison, T., Ahmad, E., Keatley, P. S., Vansteenkiste,
A., Van Waeyenberge, B., and Kruglyak, V. V. (2013). Direct excitation of
propagating spin waves by focused ultrashort optical pulses, Phys. Rev.
Lett., 110, 097201.
73. Iihama, S., Sasaki, Y., Sugihara, A., Kamimaki, A., Ando, Y., and Mizukami,
S. (2016). Quantification of a propagating spin-wave packet created by
an ultrashort laser pulse in a thin film of a magnetic metal, Phys. Rev. B,
94, 020401(R).
74. Kalinikos, B. A., Kovshikov, N. G., and Slavin, A. (1988). Envelope
solitons and modulational instability of dipole-exchange spin waves in
yttrium-iron garnet films, Sov. Phys. JETP, 67, 303–312.
75. Kovshikov, N. G., Kalinikos, B. A., Patton, C. E., Wright, E. S., and Nash,
J. M. (1996). Formation, propagation, reflection, and collision of
microwave envelope solitons in yttrium iron garnet films, Phys. Rev. B,
54, 15210.
76. Demidov, V. E., Urazhdin, S., de Loubens, G., Klein, O., Cros, V., Anane,
A., and Demokritov, S. O. (2017). Magnetization oscillations and waves
driven by pure spin currents, Phys. Rep., 673, 1–31.
Chapter 12
12.1 Introduction
Spin torque oscillators (STOs) comprise a diverse class of
nanomagnetic devices that exhibit ultrawide operating frequencies
and modulation rates. Furthermore, their manufacturing processes
are compatible with radio frequency (RF) complementary metal-
oxide semiconductor (CMOS) fabrication standards, which makes
them particularly well-suited for easy integration into existing
and future technologies. STOs combine several spintronic and
nanomagnetic phenomena for their operation, such as giant
magnetoresistance (GMR), tunneling magnetoresistance (TMR),
spin transfer torque (STT), and, depending on device architecture
and the constituent magnetic materials employed, a plethora of
possible fundamentally intriguing magnetodynamic modes.
Fig. 12.2 is strictly only valid in the far field, that is, at a
considerable distance from the NC. It is important to instead
think in terms of a spatially varying FMR frequency, calculated
on a cell-by-cell basis, that takes into account all relevant
magnetic fields. Such an FMR frequency spatial map near an NC
is shown in Fig. 12.3a. For this particular geometry, the in-plane
component of the external field points to the right and the flow
of electrons is into the plane of the page, establishing the sense
of the indicated Oersted field shown in Fig. 12.3b. Therefore,
regions near the top (bottom) of the NC experience a smaller
(larger) effective field and the FMR frequency varies, as indicated
in Fig. 12.3a. Clearly, the current-induced Oersted field has
dramatic consequences on the local FMR frequency [63], which
shows variations on the order of 6 GHz. The regions with a
higher local FMR frequency will tend to block the propagation of
spin waves in that direction, downward in this case. The spatial
variation of the power of the simulated spin wave propagation
is shown in Fig. 12.3b for an applied field angle of qex = 70°.
A clear spin wave beam is observed, with the majority of the spin
wave energy travelling upward toward the regions with locally
smaller FMR frequencies.
Figure 12.3 The calculated FMR frequency landscape for a single NC with
a diameter of 90 nm. The inhomogeneous FMR frequency landscape
promotes the formation of spin wave beams, as shown in the micromagnetic
simulations (b).
372 Propagating Spin Waves in Nanocontact Spin Torque Oscillators
Figure 12.5 SEM images (a, e), calculated FMR frequency landscapes
(b, f), simulated spin wave intensity maps (c, g), and experimentally
measured frequency spectra (d, h) of two NC-STOs spatially arranged so
the in-plane component of the external field is either parallel or
orthogonal to the line joining the NCs, that is, either a horizontal (left
panels) or a vertical (right panels) array geometry. Due to the asymmetric
spin wave beam propagation, synchronization is much more preferred
in the vertical array geometry and can be observed for NC separations
up to 1300 nm. Adapted by permission from Macmillan Publishers Ltd:
Nature Nanotechnology [53].
376 Propagating Spin Waves in Nanocontact Spin Torque Oscillators
Figure 12.6 (a, top) Experimentally measured frequency spectra of 5
NC-STOs in a vertical array geometry showing robust synchronization
over the entire range of bias currents. When the in-plane component
of the applied field is rotated by 30°, synchronization is broken (a, bottom).
By comparing the integrated power and linewidth (b) of the fully and
partially synchronized states shown in (a) it can be concluded that there
exists pairwise synchronization of four of the oscillators and a single
unlocked oscillator in (a, bottom). Reprinted by permission from Macmillan
Publishers Ltd: Nature Nanotechnology [53].
Acknowledgments
This work was supported, in part, by the Swedish Research
Council (VR), the Swedish Foundation for Strategic Research
(SSF), and the Knut and Alice Wallenberg Foundation. It was also
partially supported by the European Research Council (ERC)
grant no. 307144 “MUSTANG” and the European Commission
FP7-ICT-2011 contract no. 317950 “MOSAIC.” We would also
like to thank M. Madami for the m-BLS measurements presented
in this chapter.
References
1. Slonczewski, J. (1996). Current-driven excitation of magnetic
multilayers, J. Magn. Magn. Mater., 159(1–2), L1–L7.
2. Berger, L. (1996). Emission of spin waves by a magnetic multilayer
traversed by a current, Phys. Rev. B, 54(13), 9353–9358.
3. Sankey, J. C., Cui, Y.-T., Sun, J. Z., Slonczewski, J. C., Buhrman, R. A., and
Ralph, D. C. (2007). Measurement of the spin-transfer-torque vector
in magnetic tunnel junctions, Nat. Phys., 4(1), 67–71.
4. Muduli, P. K., Heinonen, O. G., and Åkerman, J. (2011). Bias dependence
of perpendicular spin torque and of free- and fixed-layer eigenmodes
in MgO-based nanopillars, Phys. Rev. B, 83(18), 184410.
5. Zhou, Y., and Åkerman, J. (2009). Perpendicular spin torque promotes
synchronization of magnetic tunnel junction based spin torque
oscillators, Appl. Phys. Lett., 94(11), 112503.
6. Ralph, D., and Stiles, M. (2008). Spin transfer torques, J. Magn. Magn.
Mater., 320(7), 1190–1216.
7. Hoffmann, A. (2013). Spin Hall effects in metals, IEEE Trans. Magn.,
49(10), 5172–5193.
8. Sinova, J., Valenzuela, S. O., Wunderlich, J., Back, C. H., and Jungwirth, T.
(2015). Spin Hall effects, Rev. Mod. Phys., 87(4), 1213–1260.
9. Demidov, V. E., et al. (2012). Magnetic nano-oscillator driven by
pure spin current, Nat. Mater., 11(12), 1028–1031.
10. Demidov, V. E., Urazhdin, S., Zholud, A., Sadovnikov, A. V., and
Demokritov, S. O. (2014). Nanoconstriction-based spin-Hall nano-
oscillator, Appl. Phys. Lett., 105(17), 172410.
11. Awad, A. A., et al. (2017). Long-range mutual synchronization of
spin Hall nano-oscillators, Nat. Phys., 13, 292–299.
380 Propagating Spin Waves in Nanocontact Spin Torque Oscillators
12. Dürrenfeld, P., Awad, A. A., Houshang, A., Dumas, R. K., and Åkerman, J.
(2017). A 20 nm spin Hall nano-oscillator, Nanoscale, 1–12.
13. Mazraati, H., et al. (2016). Low operational current spin Hall
nano-oscillators based on NiFe/W bilayers, Appl. Phys. Lett., 109(24),
242402.
14. Collet, M., et al. (2016). Generation of coherent spin-wave modes
in yttrium iron garnet microdiscs by spin–orbit torque, Nat. Commun.,
7, 10377.
15. Haidar, M., et al. (2016). Controlling Gilbert damping in a YIG film
using nonlocal spin currents, Phys. Rev. B, 94(18), 180409.
16. Neusser, S., and Grundler, D. (2009). Magnonics: spin waves on the
nanoscale, Adv. Mater., 21, 2927–2932.
17. Kruglyak, V. V., Demokritov, S. O., and Grundler, D. (2010). Magnonics,
J. Phys. D: Appl. Phys., 43(26), 260301.
18. Serga, A. A., Chumak, A. V., and Hillebrands, B. (2010). YIG magnonics,
J. Phys. D: Appl. Phys., 43(26), 264002.
19. Lenk, B., Ulrichs, H., Garbs, F., and Münzenberg, M. (2011).
The building blocks of magnonics, Phys. Rep., 507(4–5), 107–136.
20. Dumas, R. K., and Åkerman, J. (2014). Spintronics: channelling spin
waves, Nat. Nanotechnol., 9(7), 503–504.
21. Chen, T., et al. (2016). Spin-torque and spin-Hall nano-oscillators,
Proc. IEEE, 104(10), 1919–1945.
22. Tsoi, M., et al. (1998). Excitation of a magnetic multilayer by an
electric current, Phys. Rev. Lett., 80(19), 4281–4284.
23. Myers, E., Ralph, D., Katine, J., Louie, R., and Buhrman, R. (1999).
Current-induced switching of domains in magnetic multilayer
devices, Science, 285, 867–870.
24. Rippard, W. H., Pufall, M. R., and Silva, T. J. (2003). Quantitative
studies of spin-momentum-transfer-induced excitations in Co/Cu
multilayer films using point-contact spectroscopy, Appl. Phys. Lett.,
82(8), 1260.
25. Pufall, M. R., Rippard, W. H., and Silva, T. J. (2003). Materials
dependence of the spin-momentum transfer efficiency and critical
current in ferromagnetic metal/Cu multilayers, Appl. Phys. Lett.,
83(2), 323.
26. Rippard, W., Pufall, M., Kaka, S., Russek, S., and Silva, T. (2004).
Direct-current induced dynamics in Co90Fe10/Ni80Fe20 point
contacts, Phys. Rev. Lett., 92(2), 27201.
References 381
55. Bonetti, S., Puliafito, V., Consolo, G., Tiberkevich, V. S., Slavin, A. N.,
and Åkerman, J. (2012). Power and linewidth of propagating and
localized modes in nanocontact spin-torque oscillators, Phys. Rev. B,
85(17), 174427.
56. Slavin, A., and Tiberkevich, V. (2005). Spin wave mode excited by
spin-polarized current in a magnetic nanocontact is a standing self-
localized wave bullet, Phys. Rev. Lett., 95(23), 237201.
57. Consolo, G., Lopez-Diaz, L., Torres, L., and Azzerboni, B. (2007).
Magnetization dynamics in nanocontact current controlled
oscillators, Phys. Rev. B, 75(21), 214428.
58. Slonczewski, J. (1999). Excitation of spin waves by an electric current,
J. Magn. Magn. Mater., 195(2), 261–268.
59. Madami, M., et al. (2011). Direct observation of a propagating
spin wave induced by spin-transfer torque, Nat. Nanotechnol., 6(10),
635–638.
60. Gerhart, G., Bankowski, E., Melkov, G., Tiberkevich, V., and Slavin, A.
(2007). Angular dependence of the microwave-generation threshold
in a nanoscale spin-torque oscillator, Phys. Rev. B, 76(2), 24437.
61. Consolo, G., et al. (2008). Micromagnetic study of the above-
threshold generation regime in a spin-torque oscillator based on a
magnetic nanocontact magnetized at an arbitrary angle, Phys. Rev. B,
78(1), 1–7.
62. Muduli, P. K., Heinonen, O. G., and Åkerman, J. (2012). Decoherence
and mode hopping in a magnetic tunnel junction based spin torque
oscillator, Phys. Rev. Lett., 108(20), 207203.
63. Hoefer, M., Silva, T., and Stiles, M. (2008). Model for a collimated
spin-wave beam generated by a single-layer spin torque nanocontact,
Phys. Rev. B, 77, 144401.
64. Demidov, V. E., Urazhdin, S., and Demokritov, S. O. (2010). Direct
observation and mapping of spin waves emitted by spin-torque
nano-oscillators, Nat. Mater., 9(12), 984–988.
65. Kaka, S., Pufall, M. R., Rippard, W. H., Silva, T. J., Russek, S. E., and Katine,
J. (2005). Mutual phase-locking of microwave spin torque nano-
oscillators, Nature, 437(7057), 389–392.
66. Mancoff, F. B., Rizzo, N. D., Engel, B. N., and Tehrani, S. (2005).
Phase-locking in double-point-contact spin-transfer devices, Nature,
437(7057), 393–395.
67. Pufall, M., Rippard, W., Russek, S., Kaka, S., and Katine, J. (2006).
Electrical measurement of spin-wave interactions of proximate
spin transfer nanooscillators, Phys. Rev. Lett., 97(8), 87206.
384 Propagating Spin Waves in Nanocontact Spin Torque Oscillators
68. Grollier, J., Querlioz, D., and Stiles, M. D. (2016). Spintronic nanodevices
for bioinspired computing, Proc. IEEE, 104(10), 2024–2039.
69. Nakada, K., and Miura, K. (2016). Pulse-coupled spin torque nano
oscillators with dynamic synapses for neuromorphic computing,
in IEEE 16th International Conference on Nanotechnology (IEEE-
NANO), pp. 397–400.
70. Lequeux, S., et al. (2016). A magnetic synapse: multilevel spin-torque
memristor with perpendicular anisotropy, Sci. Rep., 6, 31510.
Chapter 13
Earth Sciences, University of Messina, V.le F. d’Alcontres 31, Messina I-98166, Italy
dDepartment of Physics, Oakland University, 2200 N. Squirrel Road, Rochester,
MI 48309, USA
13.1 Introduction
The research field of magnonics has attracted growing attention
due to the potential applications of propagating spin waves (SW)
(or magnons) in the next generation of signal-processing devices
as information carriers replacing the electrons that are used
for this purpose in traditional CMOS devices [8, 26, 28]. SWs as
carriers of information have several important advantages: (i) small
wavelengths, , down to tens of nanometers [1], (ii) the possibility
to vary SW dispersion both by the patterning of a ferromagnetic
film [11, 41] and by dynamic manipulation of the external bias
magnetic field [7, 26], (iii) the possibility to achieve nonreciprocal
SW propagation [16, 24, 54, 63], and (iv) the possibility to control
SWs by various nonlinear and parametric processes [16, 31]. All
these features allow one to design nanoscale devices for both digital
and analog data processing [12, 21, 46].
One of the important drawbacks of the existing SW technology is
that the excitation and control of SW are performed using external
magnetic fields, which are, usually, created by electric currents in
adjacent conducting lines. The generation of microwave currents
Introduction 387
 (c m e ( )
È i kx -w k t )
˘
M (r , t ) = M s Íe z + k k + c.c. ˙ , (13.1)
ÍÎ k ˙˚
one can obtain the following equation for the SW amplitude ck under
the action of a parallel parametric pumping [34]:
dck Lg
dt
+ iw k ck + G k ck = Âi l
k¢ x
Vkk ¢ bk + k ¢ e
- iw pt *
ck ' . (13.2)
Excitation of Spin Waves 391
Here bk is the Fourier image of the spatial profile of the effective
pumping magnetic field bp(x), which, in our case, is equal to the
voltage-induced anisotropy field bp(x) = ΔBan(x) = 2βE(x)/(Msh), Lg
and lx are the lengths of the gate electrode and the magnetic nanowire
in the x direction, respectively, and the superscript * denotes the
complex conjugation. For a spatially uniform gate of the length Lg the
Fourier profile bk = sinc[kLg /2] , and the parametric pumping most
efficiently couples to the SWs having opposite wavevectors k and –k,
which is a consequence of the momentum conservation law.
Note that, in general, the SW spectrum of a nanowire consists of
infinitely many SW modes, which differ by the mode profile along
the nanowire width (i.e., mk = mk,n(y)). However, since the pumping
is uniform along the nanowire width, it does not couple the SW
modes having different width profiles, and the dynamics of each SW
mode is described by the same Eq. 13.2) (with replacement ωk Æ
ωk,n, etc.). Below we shall restrict our attention to the case of the
lowest SW mode n = 0 of the nanowire, which has a uniform profile
along the nanowire width. Notes on the excitation of higher-order
SW modes are given in Section 13.2.1.3.
The parameter Vkk’ describes the efficiency of the parametric
coupling of SWs with the pumping. In this case, when the condition
of exact parametric resonance ωk = ωp/2 is satisfied for a certain SW
wavevector k, the coupling parameter is equal to
m*k ◊ m*k
Vk( - k ) = Vkk = -g , (13.3)
2 A k
where γ is the gyromagnetic ratio of the ferromagnetic material,
angular brackets mean averaging over the nanowire width, and
A k is the norm of the SW mode (see Appendix A, which provides
details on the calculation of the SW spectrum and the mode vector
structure mk). Noting that for a perpendicular static magnetization
the SW mode structure can be represented as mk = |mk,x|ex + i|mk,y|ey,
one can easily find that the parametric interaction efficiency
is proportional to the difference of the squares of the dynamic
magnetization components, Vkk ~ (|mk,x|2 – |mk,y|2). In other words,
|Vkk| depends on the ellipticity of magnetization precession. So, for
circular precession, when |mk,x| = |mk,y|, the parametric coupling is
equal to zero, and this coupling increases with the increase in the
precession ellipticity.
392 Parametric Excitation and Amplification of Spin Waves
(m )
2
Vkk = g *
k ,z /2 A k . (13.7)
the material parameters are the same as in Figs. 13.2 and 13.4). Thus,
one can conclude that in the case of the in-plane static magnetization
the parametric excitation of SWs by microwave VCMA is more
effective than in the case of the out-of-plane magnetization. This
is related to the different mechanisms of coupling of the pumping
with SWs—via precession ellipticity in the case of the out-of-plane
magnetization, and via perpendicular dynamic magnetization
component in the case of the in-plane magnetization.
in the subthreshold regime bp < bp,th; otherwise the excited SWs will
significantly distort the incoming signal SW. As one can see from Fig.
13.8, when the ratio ΓLg/v is large (wide pumping gate), the region
of proper amplification • > K > 1 is rather narrow, and it might
be difficult to operate in this regime given the uncertainties of the
experimental parameters of nanoscale VCMA gates. Thus, the use of
narrow VCMA gates ΓLg/v < 1 is preferable in practical applications.
If the incident SW does not exactly satisfy the parametric resonance
condition, the amplification rate decreases. This frequency
mismatch Δω = ωk – ωp/2 can be accounted for by the replacement
Γk Æ Γk –i Δω, in Eq. 13.10 [20].
can see from Fig. 13.9, the stabilization properties are better (i.e., a
smaller derivative, daout/dain, can be achieved) for larger amplitudes
bp of the parametric pumping.
wk = (w H + w M ÎÈlex
2 2
)(
k + Fkxx ˚˘ w H + w M ÎÈlex
2 2 zz ˘
k + Fkzz - Nan ) *
˚ - wM D k
.
(13.11)
Here k = kx, ωH = γBint, Bint is the static internal field in the
nanowire, tensors Fk and Nan are defined in the Appendix A, and D* =
2Ddi/(μ0Ms2h).
It is clear from Eq. 13.11 that the SW spectrum of a ferromagnetic
nanowire in the presence of IDMI is nonreciprocal, ωk ≠ ω–k. If the
IDMI is sufficiently strong, then the minimum of the SW spectrum
is located at a non-zero wavevector kmin ≠ 0 and is monotonic for
the positive SW wavevectors k, while being nonmonotonic for the
negative values of k (or vice versa). The sufficient condition for the
appearance of the spectrum minimum at kmin ≠ 0 is D* > h(ωH + ωM)/
(4ω0), which is derived in the approximation that the waveguide
width in substantially larger that the magnetic film thickness wy
>> h. An example of a SW spectrum having minimum at kmin ≠ 0 is
shown in Fig. 13.10. This spectrum was calculated for a permalloy
nanowire deposited on a platinum substrate. The parameters for
414 Parametric Excitation and Amplification of Spin Waves
13.5 Summary
In this chapter we described several phenomena, which can be used
for the excitation and processing of SWs in ultrathin ferromagnetic
nanowires using microwave electric field via the VCMA effect.
First, it has been shown that microwave pumping using the
VCMA effect can excite propagating SWs. In the absence of an
external magnetic field for both in-plane and out-of-plane ground
states only the parametric excitation is possible, when SWs are
excited at half the driving frequency. The in-plane magnetized
geometry is more effective for parametric excitation, since in this
geometry the excitation efficiency is proportional to the out-of-plane
magnetization component, rather than the SW precession ellipticity,
which determines the threshold in the case of the out-of-plane
magnetization. Consequently, in the in-plane magnetized case it is
possible to excite SW in a significantly wider frequency range with
the magnitudes of the threshold electric fields below 1 V/nm (for a
typical Fe/MgO structure). We have also considered a nonlinear stage
of parametric excitation and have shown that excited SW amplitudes
are determined by two mechanisms: (i) 4-magnon “pair” interaction
and (ii) amplitude dependence of SW group velocity, resulting from
non-zero nonlinear frequency shift. The last mechanism becomes
important only at the nanoscale.
Second, it has been shown that the application of a parametric
pumping at a certain location along the path of a propagating SW
leads to the compensation of the propagation losses and, if the
pumping is sufficiently large, to the amplification of a propagating
SW. If the amplitude of the incident SW becomes sufficiently large,
the nonlinear SW interaction leads to a decrease in the amplification
rate. It is also shown that in a certain range of the SW amplitudes
the parametric amplifier can be used for the stabilization of SW
Calculation of Spin Wave Dispersion and Vector Structure 419
amplitude. The output SWs can have the same mean value, but
significantly smaller amplitude spread compared to the input SWs.
Finally, it has been shown how the IDMI-induced nonreciprocity
can improve the characteristics of a VCMA SW parametric amplifier.
Under certain conditions, the signal and idler SWs in a parametric
amplifier are co-propagating, which removes the disruptive influence
of the idler SW on the operation of all the preceding SW processing
gates. At the optimal conditions, the idler SW has a nonpropagating
evanescent character, which minimizes both the energy losses of the
idler SW and its parasitic influence on other gates in a SW processing
device. Also, it has been found that the parametric excitation of co-
propagating SWs is impossible in the case of spatially extended
(“adiabatic”) pumping, which prevents spurious noise generation in
the parametric amplifier.
All the described features make the VCMA-based processing
of SWs in ferromagnetic nanowires (especially in the ones with
IDMI) attractive for applications in the SW-based microwave signal
processing devices at the nanoscale.
Appendix A. C
alculation of Spin Wave
Dispersion and Vector Structure
The vector structure of SWs in ferromagnetic nanowires and SW
dispersion relation can be calculated using the general formalism
of linear collective SW excitation [53]. The SW frequency ωkx and
vector structure mkx (which describes ellipticity) can be determined
from the following equation:
- kx mkx = ¥ Wkx ◊ mkx , (A.1)
where μ is the unit vector in the direction of static magnetization
and
( )
2
Wkx = g B + w M lex k x2 + k 2y - w M Nan + w M Fkx (A.2)
is the Hamiltonian tensor. Here B = ◊ (Be - w M F0 + w M Nan ) is
the static internal magnetic field, Be is the external field, ky is the
effective wavenumber describing the SW profile across the nanowire
width, ωM = γμ0Ms. The uniaxial anisotropy tensor is defined as
Nan = (e z ¢ ƒ e z ¢ )K an /( m0 M s ) , where Kan is the anisotropy constant,
420 Parametric Excitation and Amplification of Spin Waves
(A.3)
where σky = ∫g(y)exp[–ikyy]dy is the Fourier transform of the
normalized SW profile g(y) across the nanowire width, f(kh) = 1– (1–
exp[–kh])/(kh) and k2 = kx2+ky2 (note that in the main text we used
brief notation k = kx, which is not applicable here). For the lowest SW
mode with uniform profile g(y) = 1 the function σky = sinc[kywy/2]. In
the approximation of free boundary conditions normalized profiles
of higher order modes are given by g(y) = 21/2cos[πny/wy], n = 2,
4, 6 … for even modes and by g(y) = 21/2sin[πny/wy], n = 1, 3, 5 …
for odd modes. In more complex cases, one should use theoretical
formalism from [19] for calculation of the mode profiles. The norm
of a SW mode, used in Eqs. 13.3 and 13.7, is equal to Ãk = imk*·μ×mk.
References
1. Balashov, T., Buczek, P., Sandratskii, L., Ernst, A., and Wulfhekel, W.
(2014). Magnon dispersion in thin magnetic films, J. Phys.: Condens.
Matter, 26, 394007.
2. Bauer, U., Emori, S., and Beach, G. S. D. (2012). Voltage-gated modulation
of domain wall creep dynamics in an ultrathin metallic ferromagnet,
Appl. Phys. Lett., 101, 172403.
3. Bloembergen, N. (1965). Nonlinear Optics (Addison-Wesley Publishing
Co).
4. Bode, M., Heide, M., von Bergmann, K., Ferriani, P., Heinze, S., Bihlmayer,
G., Kubetzka, A., Pietzsch, O., Blügel, S., and Wiesendanger, R. (2007).
Chiral magnetic order at surfaces driven by inversion asymmetry,
Nature, 447, 190.
5. Brächer, T., Heussner, F., Pirro, P., Meyer, T., Fischer, T., Geilen, M., Heinz,
B., Lägel, B., Serga, A. A., and Hillebrands, B. (2016). Phase-to-intensity
References 421
32. Maruyama, T., Shiota, Y., Nozaki, T., Ohta, K., Toda, N., Mizuguchi, M.,
Tulapurkar, A. A., Shinjo, T., Shiraishi, M., Mizukami, S., Ando, Y., and
Suzuki, Y. (2009). Large voltage-induced magnetic anisotropy change
in a few atomic layers of iron, Nat. Nanotechnol., 4, 158.
33. Matsukura, F., Tokura, Y., and Ohno, H. (2015). Control of magnetism by
electric fields, Nat. Nanotechnol., 10, 209–220.
34. Melkov, G. A., Serga, A. A., Tiberkevich, V. S., Kobljanskij, Y. V., and Slavin,
A. N. (2001). Nonadiabatic interaction of a propagating wave packet
with localized parametric pumping, Phys. Rev. E, 63, 066607.
35. Melkov, G. A., Vasyuchka, V. I., Lazovskiy, V. V., Tiberkevich, V. S., and
Slavin, A. N. (2006). Wave front reversal with frequency conversion in
a nonreciprocal medium, Appl. Phys. Lett., 89, 252510.
36. Mills, D. L., and Dzyaloshinskii, I. E. (2008). Influence of electric
fields on spin waves in simple ferromagnets: role of the flexoelectric
interaction, Phys. Rev. B, 78, 184422.
37. Moon, J.-H., Seo, S.-M., Lee, K.-J., Kim, K.-W., Ryu, J., Lee, H.-W., McMichael,
R. D., and Stiles, M. D. (2013). Spin-wave propagation in the presence of
interfacial Dzyaloshinskii-Moriya interaction, Phys. Rev. B, 88, 184404.
38. Moriya, T. (1960). Anisotropic superexchange interaction and weak
ferromagnetism, Phys. Rev., 120, 91–99.
39. Nakamura, K., Shimabukuro, R., Fujiwara, Y., Akiyama, T., Ito, T., and
Freeman, A. J. (2009). Giant modification of the magnetocrystalline
anisotropy in transition-metal monolayers by an external electric field,
Phys. Rev. Lett., 102, 187201.
40. Nawaoka, K., Shiota, Y., Miwa, S., Tomita, H., Tamura, E., Mizuochi, N.,
Shinjo, T., and Suzuki, Y. (2015). Voltage modulation of propagating
spin waves in Fe, J. Appl. Phys., 117, 17A905.
41. Neusser, S., and Grundler, D. (2009). Magnonics: spin waves on the
nanoscale, Adv. Mater., 21, 2927–2932.
42. Niranjan, M. K., Duan, C.-G., Jaswal, S. S., and Tsymbal, E. Y. (2010).
Electric field effect on magnetization at the Fe/MgO(001) interface,
Appl. Phys. Lett., 96, 222504.
43. Nozaki, T., Shiota, Y., Miwa, S., Murakami, S., Bonell, F., Ishibashi, S.,
Kubota, H., Yakushiji, K., Saruya, T., Fukushima, A., Yuasa, S., Shinjo, T.,
and Suzuki, Y. (2012). Electric-field-induced ferromagnetic resonance
excitation in an ultrathin ferromagnetic metal layer, Nat. Phys., 8, 491.
44. Nozaki, T., Kozioł-Rachwał, A., Skowroński, W., Zayets, V., Shiota, Y.,
Tamaru, S., Kubota, H., Fukushima, A., Yuasa, S., and Suzuki, Y. (2016).
Large voltage-induced changes in the perpendicular magnetic
424 Parametric Excitation and Amplification of Spin Waves
57. Verba, R., Carpentieri, M., Finocchio, G., Tiberkevich, V., and Slavin,
A. (2016). Excitation of propagating spin waves in ferromagnetic
nanowires by microwave voltage-controlled magnetic anisotropy, Sci.
Rep., 6, 25018.
58. Verba, R., Carpentieri, M., Finocchio, G., Tiberkevich, V., and Slavin,
A. (2017). Excitation of spin waves in an in-plane-magnetized
ferromagnetic nanowire using voltage-controlled magnetic anisotropy,
Phys. Rev. Appl., 7, 064023.
59. Wang, W.-G., Li, M., Hageman, S., and Chien, C. L. (2012). Electric-field-
assisted switching in magnetic tunnel junctions, Nat. Mater., 11, 64.
60. Weisheit, M., Fähler, S., Marty, A., Souche, Y., Poinsignon, C., and Givord,
D. (2007). Electric field-induced modification of magnetism in thin-
film ferromagnets, Science, 315, 349–351.
61. Yang, H. X., Chshiev, M., Dieny, B., Lee, J. H., Manchon, A., and Shin, K. H.
(2011). First-principles investigation of the very large perpendicular
magnetic anisotropy at Fe|MgO and Co|MgO interfaces, Phys. Rev. B,
84, 054401.
62. Zakharov, V., L’vov, V., and Starobinets, S. (1975). Spin-wave turbulence
beyond the parametric excitation threshold, Sov. Phys. Usp., 18, 896.
63. Zhang, V. L., Di, K., Lim, H. S., Ng, S. C., Kuok, M. H., Yu, J., Yoon, J., Qiu,
X., and Yang, H. (2015). In-plane angular dependence of the spin-
wave nonreciprocity of an ultrathin film with Dzyaloshinskii-Moriya
interaction, Appl. Phys. Lett., 107, 022402.
64. Zhang, X., Zhou, Y., Ezawa, M., Zhao, G. P., and Zhao, W. (2015). Magnetic
skyrmion transistor: skyrmion motion in a voltage-gated nanotrack,
Sci. Rep., 5, 11369.
65. Zhu, J., Katine, J. A., Rowlands, G. E., Chen, Y.-J., Duan, Z., Alzate, J. G.,
Upadhyaya, P., Langer, J., Amiri, P. H., Wang, K. L., and Krivorotov, I. N.
(2012). Voltage-induced ferromagnetic resonance in magnetic tunnel
junctions, Phys. Rev. Lett., 108, 197203.
Index