Coupling between spin waves (SWs) and other types of waves in nanostructured magnetic media has attracted increased attention in recent years because of the rich physics and the potential to spawn disruptive technologies. Among this family of intriguing phenomena, we recently reported a new one: coupling between SWs and hybridized phonon–plasmon waves, resulting in tripartite coupling of magnons, phonons, and plasmons. Here, this acousto-plasmo-magnonic phenomenon is studied in a two-dimensional periodic array of bilayered Co/Al nanodots on a silicon substrate, where Co is a magnetostrictive constituent responsive to magnetoelastic coupling and Al acts as a source of surface plasmons. Time-resolved magneto-optical-Kerr-effect microscopy revealed parametric amplification and strong coupling between two SW modes mediated by a hybrid phonon–plasmon mode. The strong coupling forms a new quasi-particle: the phonon–plasmon–magnon polariton.

Plasmonics1,2 and magneto-plasmonics3,4 have been busy fields of study, spawning novel technologies such as magneto-optical nano-antennas5,6 and circuits.7 In magneto-plasmonics, one usually examines the influence of magnetic fields on plasmons, while ignoring the reciprocal effect where surface plasmons, potentially hybridized with entities like phonons,8–10 impact the dynamic magnetic properties of magnetic media, specifically SWs. Here, we report a study of this latter effect, which we have recently shown could reveal intriguing features such as the formation of acousto-plasmo-magnonic frequency combs.11 This field is an extension of the well-explored and popular field of magnetoelastic coupling12–23 where pure phonon modes interact with magnon modes (magnon–phonon coupling) in magnetostrictive nanomagnets to produce such entities as magnon–polarons.20,21 The next frontier is obviously tripartite coupling where three different entities—phonons, magnons and plasmons—couple to reveal new physics and portend new applications.

Conventional wisdom will hold that phonon–magnon–plasmon coupling would be an elusive phenomenon since plasma wave frequencies far exceed those of SWs. Hence, any coupling between them will encounter a nearly insurmountable barrier because of the resulting large phase mismatch. Surface plasmons or surface plasmon polaritons in metals, however, can have much lower frequencies than bulk plasmons,24 and, hence, they can couple with acoustic phonons in appropriate media to form hybrid phonon–plasmon modes8–10 whose frequencies are close to those of spin waves in ferromagnetic media. These hybrid modes (phonons + plasmons) can couple strongly with SW modes (magnons) to produce tripartite coupling between phonons, magnons, and plasmons, giving rise to acousto-plasmo-magnonics.

In this paper, we have employed time-resolved magneto-optical Kerr effect (TR-MOKE) microscopy to showcase the intriguing coupling between SWs and hybridized phonon–plasmon waves. Our study employs a two-dimensional periodic array of 100 nm (Co/Al) heterostructured nanodots deposited on an Si substrate. Acousto-plasmo-magnonic modes are excited in this system when it is exposed to ultrashort ( 100 fs) laser pulses. The periodic heating and cooling caused by the laser pulses give rise to acoustic waves (phonons) in the Si substrate owing to periodic stresses caused by the unequal thermal expansion/contraction coefficients of the Si substrate and the nanodots on top. These phonons mix with the surface plasmons or surface plasmon polaritons due to the Al layer to form the hybrid phonon–plasmon modes, which then couple with the SW modes generated in the nanomagnets, with or without an external magnetic field, to form the acousto-plasmo-magnonic modes whose dispersion relations may show no dependence on any magnetic field because they are not magnetic in origin.

The samples are fabricated with e-beam lithography and sequential e-beam evaporation of different metals. An e-beam resist is spun onto a Si substrate and patterned with e-beam in a Raith system to open windows for the nanodots, followed by the sequential evaporation of Ta (10 nm: for adhesion), Al (10 nm), Co (6 nm), and Au (2 nm), followed by lift-off. The Au layer is needed to prevent the oxidation of Co. Both Al and Au are plasmonic metals, but because the Au layer is 5 × thinner than the Al layer and is much more distant from the Si substrate where the phonons reside, its contribution to plasmon–magnon hybrid modes, if any, is much lower than that of Al. We cannot separate out the contributions from Au and Al, but in our context, it is of no consequence. We fabricate two sets of samples: (i) a two-dimensional array of elliptical Co nanomagnets (NMs) on a Si substrate and (ii) an identical array of (Co/Al) bilayered nanodots of the same shape and dimensions as the Co NMs on a similar Si substrate. Only the latter hosts strong enough hybrid plasmon–phonon modes that can couple with magnon modes in the nanomagnets because of the 10 nm thick Al layer close to the Si substrate.

In order to study acousto-plasmo-magnonics and differentiate it from the more familiar acousto-magnonics (i.e., tripartite phonon–plasmon–magnon coupling vs bipartite phonon–magnon coupling), the investigation of ultrafast magnetization dynamics of the samples was conducted by employing a custom-built TR-MOKE microscope25 based on a two-color collinear pump–probe technique under ambient conditions. A schematic representation of this experimental setup is shown in Fig. 1(c).

FIG. 1.

(a) Scanning electron microscope (SEM) images of the Co nanomagnet (NM) arrays on Si substrate. (b) SEM images of the identical (Co/Al) nanodot arrays. (c) Schematic of the TR-MOKE measurement geometry.

FIG. 1.

(a) Scanning electron microscope (SEM) images of the Co nanomagnet (NM) arrays on Si substrate. (b) SEM images of the identical (Co/Al) nanodot arrays. (c) Schematic of the TR-MOKE measurement geometry.

Close modal

Figures 1(a) and 1(b) show the scanning electron microscope (SEM) images of systems (i) and (ii), which are without and with the Al layer, which is the major source of surface plasmons and surface plasmon polaritons. The Co NMs are shaped like elliptical disks with major axis dimension 105 nm, minor axis dimension 100 nm, and thickness 6 nm. Along the major axes (x axis), the edge-to-edge separation between the NMs is 45 nm, and in the direction of the minor axes, it is 50 nm. The lateral dimensions of the nanomagnets and the edge-to-edge separation between them exhibit a maximum deviation of ± 5 % and ± 8 %, respectively.

The fundamental laser beam (wavelength = 800 nm, fluence = 2 mJ cm 2, pulse width = 80 fs) from a mode-locked Ti-sapphire laser (Tsunami, Spectra-Physics) is used to probe the polar Kerr rotation (hereafter referred to as the probe beam) and its frequency-doubled counterpart (wavelength = 400 nm, fluence = 16 mJ cm 2, pulse width = 100 fs), hereafter referred to as the pump beam, is used to excite the magnetization dynamics of the samples. The latter has high enough intensity that it can penetrate through the Au and Co layers to reach the Al layer and excite plasma oscillations there (surface plasmons and surface plasmon polaritons). The Kerr signal is measured using an optical bridge detector (OBD) as a function of time delay between the pump and probe beams. Achieving the spatial overlap of these two beams is critical. The slightly defocused pump beam (of diameter 1 μm) at the focal plane of the tightly focused probe beam (of diameter 800 nm) is made collinear and overlapping on the sample plane using a single microscope objective of numerical aperture 0.65. The OBD simultaneously measures both the reflectivity and Kerr rotation and separates these intertwined signals with the help of two lock-in amplifiers in a phase-sensitive manner to attain high sensitivity. To maintain temporal synchrony, the pump beam is subjected to periodic modulation at a frequency of about 2 kHz by a mechanical chopper. This modulated pump beam serves as a reference frequency and is conveyed to the lock-in amplifiers, anchoring the phase relationships within the system. Furthermore, the probe beam is systematically positioned at the precise center of the pump beam using an x y z piezoelectric scanning controller, guided by a feedback loop and a white-light illumination system, enhancing the fidelity of the experiment. A static magnetic field with varying magnitude is applied with a small tilt (of 10 ° 15 °) from the sample plane to introduce the necessary out-of-plane demagnetizing field along the direction of the pump beam, which helps to induce precessional magnetization dynamics (SWs) in the sample during pumping. The experimental time window of 2 ns was found to be sufficient to resolve the SW peaks with a temporal resolution of 10 ps from the fast Fourier transform (FFT) of the bi-exponential background-subtracted time-resolved traces.

We compared the experimental results of the samples without Al (no surface plasmons or surface plasmon polaritons present) with theoretical results obtained from micromagnetic simulations using Object Oriented Micromagnetic Framework (OOMMF) software26 , employing a 4 × 4 array of nanomagnets whose specifications are based on what we observed in the SEM images. The discretized array featured rectangular parallelepiped-shaped cells ( 3 × 3 × 6 nm 3) with a two-dimensional periodic boundary condition. The unit cell length was kept below the exchange length of Co ( 4.93 nm), to allow the incorporation of exchange interactions. Simulation parameters used included saturation magnetization, M s = 1400 emu cm 3, anisotropy constant, K = 0, γ = 17.6 MHz Oe 1, the exchange stiffness constant, A e x = 3.0 × 10 6 erg cm 1,27 and Gilbert damping coefficient, α = 0.008.28 

To simulate magnetization dynamics, we first obtained the static magnetic state after system relaxation to equilibrium at a specific bias-field. Optical excitation was emulated with a square pulsed magnetic field (10 ps rise and fall time, 200 ps width, 20 Oe peak amplitude) perpendicular to the sample plane. The FFT of the simulated time-resolved out-of-plane magnetization ( m z) revealed the SW spectra, consistently showing three dominant peaks (M1, M2, and M3) across various bias magnetic fields. These theoretical results showed excellent agreement with the experimental observations.

We were not able to simulate the magnetization dynamics in the samples containing Al since no software is available to us that can incorporate the effect of surface plasmons or surface plasmon polaritons in magneto-dynamics. Therefore, we are unable to compare theory with experiments in the Al-containing samples. Our experiments, however, showed Stark differences between the plasmonic (Al-present) and non-plasmonic (Al-absent) samples, which can only accrue from coupling with surface plasmons and/or surface plasmon polaritons.

In Figs. 2(a) and 2(b), we show the background-subtracted time-resolved data for reflectivity of the non-plasmonic (without aluminum) and plasmonic (with aluminum) samples as a function of delay between the pump and the probe, obtained at 16 mJ cm 2 pump fluence. We will discuss the non-plasmonic results first.

FIG. 2.

Background-subtracted time-resolved data for the reflectivity of (a) the Co NM arrays and (b) the (Co/Al) NM arrays, on the Si substrate as a function of delay between the pump and the probe, obtained at 16 mJ cm 2 pump fluence. (c) and (d) FFT of the respective oscillations. The two detected frequencies in the non-plasmonic sample correspond to surface acoustic wave modes excited in the Si substrate by the laser heating/cooling, whose wavelengths are commensurate with distinct periodicities in real space and reciprocal space of the magnonic crystal made of the NM arrays [inset of (c)]. These are bending wave modes (pure phonons) excited by the pump beam periodically heating and cooling the substrate. The spectrum of the Co/Al plasmonic sample shows that the primary peak has blueshifted and its full width at half maximum has increased considerably, compared to that of the non-plasmonic sample. There are also three satellite peaks. These are hybrid phonon–plasmon modes; the phonons are excited by the pump beam and they mix with plasmons emitted by the Al layer to form the hybrid phonon–plasmon wave. (e) and (f) Surface plot of the reflectivity spectra of the Co NM arrays on the Si substrate and identical (Co/Al) NM arrays, respectively.

FIG. 2.

Background-subtracted time-resolved data for the reflectivity of (a) the Co NM arrays and (b) the (Co/Al) NM arrays, on the Si substrate as a function of delay between the pump and the probe, obtained at 16 mJ cm 2 pump fluence. (c) and (d) FFT of the respective oscillations. The two detected frequencies in the non-plasmonic sample correspond to surface acoustic wave modes excited in the Si substrate by the laser heating/cooling, whose wavelengths are commensurate with distinct periodicities in real space and reciprocal space of the magnonic crystal made of the NM arrays [inset of (c)]. These are bending wave modes (pure phonons) excited by the pump beam periodically heating and cooling the substrate. The spectrum of the Co/Al plasmonic sample shows that the primary peak has blueshifted and its full width at half maximum has increased considerably, compared to that of the non-plasmonic sample. There are also three satellite peaks. These are hybrid phonon–plasmon modes; the phonons are excited by the pump beam and they mix with plasmons emitted by the Al layer to form the hybrid phonon–plasmon wave. (e) and (f) Surface plot of the reflectivity spectra of the Co NM arrays on the Si substrate and identical (Co/Al) NM arrays, respectively.

Close modal

1. Non-plasmonic sample

The FFT of the reflectivity oscillations in the non-plasmonic sample [Fig. 2(c)] reveals the dominance of two peaks at 1.5 GHz ( f 1) and 2.7 GHz ( f 2) in the spectrum. Because the maximum delay between the pump and probe in our setup is slightly more than 3 ns, we can observe only 5 periods of the oscillation in the reflectivity but that is enough to allow resolution of the two peaks in the spectrum. These modes (peaks) are absent in the FFT of time-resolved reflectivity oscillations from the bare Si substrate (without NM), which attests to the fact that the NM array plays a role in determining these peaks.

The origin of these peaks or modes in the non-plasmonic sample can be explained by invoking bending waves that can be excited in the Si substrate when heated by the pump laser pulse. The periodic strain produced by the unequal expansion/contraction of the NMs and the Si substrate spawns the bending waves (BWs). For simplicity, we will assume a vertical rod-like geometry for the NMs in our analysis. When the wavelength of the BW is large compared with the thickness of the rod and the oscillations are in a direction perpendicular to the axis of the rod, the frequency ( ω) and the wavevector ( k) obey the dispersion relation,29 
(1)
where E, ρ, and σ are the Young modulus, density and the Poisson ratio of Co, respectively, and h is the thickness of the Co nanodot. The material parameters for Co are E = 200 GPa, ρ = 8900 kg  m 3, and σ = 0.32.30 The dominant peak frequency ν 1 ( = 1.5 GHz ) observed in Fig. 2 obeys Equation (1), when the wave vector corresponds to the X point of the 2D Brillouin zone (BZ) associated with the NM array, | k 1 | = 2 π / l ( l is the NM array periodicity). The secondary peak frequency ν 2 ( = 2.7 GHz ) corresponds to the wave vector along the diagonal of the two-dimensional lattice (M point of the BZ), i.e., | k 2 | = 2 2 π / l. The calculated velocity of propagation of the BWs for this periodicity of the system is of the order of 1 km s 1 , which is very reasonable. Note that this is the velocity of BWs in Si, which is not the same as the surface acoustic wave (SAW) velocity in Si. We, therefore, believe that the oscillations in the reflectivity of the non-plasmonic samples are due to BWs generated by the laser heating/cooling and the oscillation period or frequency is determined by the NM array parameters. We emphasize, however, that this is a qualitative picture and not expected to be quantitatively accurate. We also recognize that Eq. (1) may not be exactly applicable to our system, which is why we offer this only as a plausible scenario and do not insist on its rigor.

2. Plasmonic sample

For the plasmonic sample, the FFT of the oscillations reveals the dominance of four (instead of two) peaks at 2 ( f 1 ), 5.5 ( f 2 ), 7.1 ( f 3 ), and 8.6 GHz ( f 4 ), where the intensities of the last three are much lower than that of the first but still well above our noise floor. Thus, the reflectivity signals of the plasmonic and non-plasmonic samples are both qualitatively and quantitatively different.

Before we proceed to delve into the origin of the reflectivity modes in the plasmonic sample, we point out that the observed frequencies f n ( n = 1 , 2) (non-plasmonic sample) and f n ( n = 1 , 2 , 3 , 4) (plasmonic sample) are insensitive to the presence or absence of any magnetic field at any given fluence and show exceptional stability over the broad range of magnetic field, confirming the non-magnetic origin of these oscillations.

There is a compelling reason to believe that the four modes ( f n ) in the plasmonic sample arise from mixing of the BWs caused by laser heating/cooling (phonons) with plasma waves (surface plasmons or surface plasmon polaritons) caused by the excitation of plasma oscillations in the Al layer by the laser beam. This is the hybridization of phonon–plasmon modes. Non-linear mixing will cause additional frequencies to appear, which is why we observe two additional modes (four instead of two) when the Al layer is present. There is also a blueshift, i.e., the mode frequencies are higher in the Al-containing sample. That is expected if energy is coupled from the plasma waves into the BWs to increase their frequencies. The phonons in the BW are absorbing plasmons to increase their energy and, hence, the frequency.

It may be argued that the plasmonic nanodots consist of Ta/Al/Co/Au layers, while the non-plasmonic nanodots consist of Ta/Co/Au layers so that there is a mass difference between the two, which could make a difference between the modes observed in the two types of samples. Further consideration rules out this possibility. These modes accrue from the unequal thermal expansion/contraction coefficients of the substrate and the layer that is in contact with the substrate, which is Ta in both cases. Hence, the compositional or mass difference between the two types of nanodots should not have a significant impact. Furthermore, the plasmonic sample, being heavier, should experience a redshift in the frequency of the modes, but we see a blueshift. Finally, the mass difference is not likely to spawn additional modes. We think it is unlikely that the compositional or mass difference plays a significant role.

We also notice that the peak linewidths are larger in the plasmonic sample spectra, indicating that the hybrid phonon–plasmon modes are more lossy than pure phonon (BW) modes. Plasmons couple very strongly with electromagnetic waves,31 and, hence, the hybrid phonon–plasmon wave may radiate some photons (electromagnetic waves) causing more rapid loss of energy. This is a speculation, and we intend to verify this in future by measuring any possible electromagnetic wave emission from these samples. To ensure that the features we have reported are not mere artifacts of structural, geometrical, or other trivial differences between the Al-containing and Al-lacking samples, we have moved the laser spots of the pump and probe beams in the TR-MOKE setup to different regions of the wafer. The laser spot size is 1 μm, the nanodot diameter is 100 nm, and the edge-to-edge separation is 50 nm. Hence, the spot covers an array of 6 × 6 nanomagnets, which constitutes a measured “sample.” By simply moving the spot elsewhere on the wafer, we measure a different “sample.” We have done that and observe similar differences between the Al-containing and Al-lacking samples, no matter where we focus the laser spot, which confirms that what we report is not an artifact. This study is reported in the accompanying supplementary material.

1. Non-plasmonic sample

Figure 3 shows the experimentally measured background-subtracted Kerr oscillation data obtained from the Co NMs on Si (no Al present) at different bias magnetic fields. The observed peaks in the FFT spectra exhibit a systematic upward shift in frequency as the bias magnetic field increases, confirming their magnetic origin. Despite a modest signal-to-noise ratio in the data, three distinct peaks denoted as “1”, “2”, and “3” have been clearly discerned in the FFT spectra. We could fit the magnetic field dependence of mode “2” with the Kittel formula:32 
(2)
where the gyromagnetic ratio γ = g μ B / , H is the bias magnetic field, g is the Landé g factor ( g = 2), μ B is the Bohr magneton, and is the reduced Planck constant. The theoretical fits are depicted in Fig. 3(d), demonstrating that mode “2” conforms well to a Kittel-like mode. The effective magnetization ( M e f f) required for the fit was 1250 emu cm 3, a value slightly lower than the intrinsic saturation magnetization of Cobalt (Co) at 1400 emu cm 3. This discrepancy is attributed to a somewhat non-uniform magnetization profile within each nanomagnet.
FIG. 3.

(a) Background-subtracted time-resolved Kerr oscillations at a magnetic field of 1 kOe from Co NM arrays on the Si substrate (no Al present) at the pump fluence of 16  mJ cm 2, (b) FFT of the corresponding oscillations. (c) FFT of the simulated temporal evolution of the m z component (out-of-plane component of magnetization) at a magnetic field of 1 kOe using OOMMF, showing good agreement with the experimentally measured spectra in (b). (d) Bias magnetic field dependence of the observed three dominant frequencies in the Kerr oscillations. The circles represent the experimental data, and the solid line is Kittel fit for mode “2.” The dashed lines are the micromagnetic simulation results.

FIG. 3.

(a) Background-subtracted time-resolved Kerr oscillations at a magnetic field of 1 kOe from Co NM arrays on the Si substrate (no Al present) at the pump fluence of 16  mJ cm 2, (b) FFT of the corresponding oscillations. (c) FFT of the simulated temporal evolution of the m z component (out-of-plane component of magnetization) at a magnetic field of 1 kOe using OOMMF, showing good agreement with the experimentally measured spectra in (b). (d) Bias magnetic field dependence of the observed three dominant frequencies in the Kerr oscillations. The circles represent the experimental data, and the solid line is Kittel fit for mode “2.” The dashed lines are the micromagnetic simulation results.

Close modal

In contrast, modes “1” and “3” cannot be fitted with either the Kittel formula or the perpendicular standing SW (PSSW) mode relation. Mode “1” is likely an edge mode, while mode ”3” is identified as a hybrid magneto-dynamical mode, consistent with previous reports in these systems.30 The corresponding SW mode analysis can be found in the supplementary material.

2. Plasmonic sample

In stark contrast to the non-plasmonic sample with no Al present, the sample with Al (where plasmonic coupling is present) exhibits much richer spectra in a magnetic field as shown in Fig. 4(a). The corresponding FFT, showing five modes in the Kerr oscillations (M1, M2, M3, M4, *), is shown in Fig. 4(b). Note from Fig. 4(a) that the oscillations were measured up to a time delay (between the pump and probe beam) of 1.8 ns, which allows for a frequency resolution of 0.55 GHz in Fig. 4(b). This is at least four times smaller than the smallest separation between adjacent modes, which is why we can confidently claim that five modes are present in these data and that the mode frequencies have an error bar of no more than 0.55  GHz. Despite the modest signal-to-noise ratio available in these experiments where the probe beam can probe ( 6 × 6 ) NMs, we still have enough measurement accuracy to justify the claims made in this paper.

FIG. 4.

(a) Bias magnetic field dependence of the background-subtracted time-resolved Kerr oscillations from (Co/Al) NM arrays at the pump fluence of 16 mJ cm 2. (b) The FFT of the corresponding Kerr oscillations showing a field-independent mode at 30.5 GHz, which is independent of the bias magnetic field.

FIG. 4.

(a) Bias magnetic field dependence of the background-subtracted time-resolved Kerr oscillations from (Co/Al) NM arrays at the pump fluence of 16 mJ cm 2. (b) The FFT of the corresponding Kerr oscillations showing a field-independent mode at 30.5 GHz, which is independent of the bias magnetic field.

Close modal

One easily discernible high-frequency magnetic field-independent mode at 30.5 GHz is always present, in addition to other field-dependent modes. The former has more power than all other modes. This field-independent mode is only present when the Al layer is there, confirming its plasmonic origin. Figure 4(b) illustrates this stable mode across four different bias magnetic fields (it is the mode marked with an asterisk). Further insight into this mode will require framing a rigorous theory of plasmon–phonon coupling in this complex system, which is outside the scope of this work.

Figure 5(a) reveals the presence of field-dependent modes in the plasmonic Al-containing sample in addition to the high-frequency field-independent mode. These are magnonic modes. The dispersions of modes M4 and M3 are fitted by the blue and red lines with arrowheads. These lines are merely guide to the eye. A distinct frequency gap of magnitude 2 g opens up at a magnetic field of 820 Oe. This gap attests to the coupling between M4 and M3 (both magnonic modes), which is obviously mediated by the hybrid phonon–plasmon mode. This is magnon–magnon coupling mediated by the hybrid mode.

FIG. 5.

(a) Experimental bias magnetic field dependence of Kerr oscillation modes in (Co/Al) NM arrays (plasmonic samples). The lines with arrow heads are guide to the eye for the modes M4 and M3. A gap of frequency 2 g opens up between these two dispersion curves at a magnetic field of 820 Oe, showing the existence of coupling between them at that magnetic field. The inset shows the FFT power spectrum of the observed field-independent and field-dependent modes at the magnetic field of H = 820 Oe (black solid line). Here, 2 k 1 and 2 k 2 are the full width at half maximum (FWHM) of the Lorentzian fits to the spectra of M4 and M3. (b) Average peak-to-peak Kerr rotation amplitude (averaged over all modes) in the plasmonic sample (with Al) and in the non-plasmonic sample (without Al), as a function of bias magnetic field. (c) Amplitude of power in the FFT spectra obtained from Kerr rotation of plasmonic and non-plasmonic samples at magnetic field H = 1 kOe.

FIG. 5.

(a) Experimental bias magnetic field dependence of Kerr oscillation modes in (Co/Al) NM arrays (plasmonic samples). The lines with arrow heads are guide to the eye for the modes M4 and M3. A gap of frequency 2 g opens up between these two dispersion curves at a magnetic field of 820 Oe, showing the existence of coupling between them at that magnetic field. The inset shows the FFT power spectrum of the observed field-independent and field-dependent modes at the magnetic field of H = 820 Oe (black solid line). Here, 2 k 1 and 2 k 2 are the full width at half maximum (FWHM) of the Lorentzian fits to the spectra of M4 and M3. (b) Average peak-to-peak Kerr rotation amplitude (averaged over all modes) in the plasmonic sample (with Al) and in the non-plasmonic sample (without Al), as a function of bias magnetic field. (c) Amplitude of power in the FFT spectra obtained from Kerr rotation of plasmonic and non-plasmonic samples at magnetic field H = 1 kOe.

Close modal

The gap 2 g and the loss rates (half width at half maximum) of modes M4 and M3, denoted as k 1 and k 2, are depicted in the inset in Fig. 5(a). To quantify these parameters, we applied a Lorentzian fit to the peaks corresponding to M4 and M3. The extracted values are, g = 0.78 GHz, k 1 = 0.36 GHz, and k 2 = 0.35 GHz. The cooperativity factor for coupling is defined as C = g 2 / ( k 1 k 2 ), which, in our case, has a value of 4.8. In accordance with established criteria, strong coupling is delineated by C > 1 and g > k 1 , k 2. Because we satisfy both conditions, we have strong coupling33 between modes M4 and M3. The coupling agent is obviously the hybrid phonon–plasmon mode, which we observed in the reflectivity spectrum of Fig. 2(d) and it is coupling two magnon (SW) modes M4 and M3. This is tripartite coupling involving phonons, plasmons and magnons, and the coupling is strong.

We understand that M3 and M4 curves are fitted to dispersed data and that the dispersion in the data is considerable. Nonetheless, there is a clear gap opening up in Fig. 5(a) and even if we are inaccurate in estimating its magnitude, we do not overestimate it by as much as a factor of 4. Hence, irrespective of our inaccuracy, the value of the cooperativity factor surely exceeds unity, which attests to “strong” coupling.

In a regime of strong coupling, individual modes transform, giving rise to a novel quasi-particle, akin to the magnon–polaron observed in SW coupling with pure acoustic waves.20,21 Our system exhibits robust tripartite coupling involving phonons, plasmons, and magnons, suggesting the emergence of a distinct phonon–plasmon–magnon polariton—a new quasi-particle. This phenomenon, reminiscent of strongly coupled magnon–plasmon polaritons reported in graphene-2D ferromagnet heterostructures,34 underscores the significance of plasmons in our system.

Notably, the absence of coupling in the sample without aluminum (no plasmons) is evidenced by the lack of anti-crossing (gap opening) in the dispersion curves for field-dependent modes. This unequivocally demonstrates that plasmons introduce qualitative distinctions, as also seen in the reflectivity spectra, not merely quantitative ones, and coupling between SW modes occurs in their presence because the hybrid phonon–plasmon wave mediates the coupling.

Our experiments have unveiled two other intriguing findings: First, the magneto-optical Kerr effect (MOKE) signal, specifically the average peak-to-peak Kerr rotation amplitude across all modes at an arbitrarily picked bias magnetic field of 1 kOe, exhibits a notable 30% increase in the plasmonic sample with aluminum (Al) compared to the non-plasmonic sample without Al. This enhancement persists across a broad range of magnetic fields. Second, the amplitude of power in the FFT spectra corresponding to the anti-crossed acousto-plasmo-magnon SW modes in the Al-containing sample is approximately four times greater than that in the field-dependent modes of the Al-lacking sample when averaged across all modes. Finally, the intensity of the field-independent mode is ten times higher than that of the field-dependent modes, as illustrated in Fig. 5(c).

The significant increase in the MOKE signal, specifically a 30% rise in the average peak-to-peak Kerr rotation amplitude across all modes in the sample with aluminum (Al) compared to the Al-lacking sample, suggests the presence of parametric amplification.35 In this context, energy is transferred from the hybridized phonon–plasmon mode to the intrinsic SW modes in the Al-containing sample, thereby amplifying the amplitude and power of the resultant acousto-plasmo-SW mode. This observation serves as confirmation of the robust tripartite plasmon–phonon–magnon coupling, since such power transfer can only happen if there is strong coupling. Parametric amplification requires the presence of a non-linearity. Note from Fig. 5(c) that the presence of the Al has not only increased the power amplitudes of the modes but also shifted the frequencies. Only a non-linearity can change the frequencies and, hence, we must conclude that a non-linearity is present. Non-linear SW coupling in magnonic crystals (which is what we have) is well known.36 Hence, we expect the nonlinearity to be present. Chasing down the exact nature of the non-linearity, however, is outside the scope of this work. One other point needs some discussion. In a lossless system, ideal parametric amplification would require that the idler frequency ω 1, the signal frequency ω 2, and the pump frequency ω 3 obey the relation ω 3 = ω 1 + ω 2, as mandated by energy conservation. We certainly do not have a lossless system, and, hence, we do not expect this relation to hold. We merely call the observed effect “parametric amplification” because it mimics parametric amplification.

In conclusion, we report (i) tripartite coupling between phonons, plasmons, and magnons in periodic arrays of bilayered nanodots of a magnetostrictive material and a plasmonic material, spawning a high-frequency magnetic-field-independent acousto-plasmo SW mode; (ii) parametric amplification of SW modes by hybrid phonon–plasmon modes, and (iii) strong coupling between two SW modes mediated by the hybrid phonon–plasmon wave, leading to the formation of the magnon–plasmon–phonon polariton. These findings not only contribute to our understanding of fundamental interactions in magnonic and plasmonic systems but also hold significant implications for practical applications. The observed parametric amplification can be leveraged for SW device applications. Furthermore, these results open avenues for controlling spin dynamics through the interplay of plasmon–phonon–photon dynamics. This approach has the potential to mitigate intrinsic limitations associated with ohmic and other losses, offering exciting prospects for plasmon-driven coherent spintronics or magneto-optical activities.

See the supplementary material for the (I) position dependence of the reflectivity spectra for plasmonic and non-plasmonic samples and (II) spin-wave mode analysis for co nanomagnets (non-plasmonic samples).

A.B. gratefully acknowledges the Department of Science and Technology, Govt. of India (Grant No. DST/NM/TUE/QM-3/2019-1C-SNB) for financial assistance. A.B. and S.B. acknowledge the support from the Indo-US Science and Technology Fund Center grant “Center for Nanomagnetics for Energy-Efficient Computing, Communications, and Data Storage” (No. IUSSTF/JC-030/2018). The work of R.F. and S.B. was supported by the US National Science Foundation under Grant No. ECCS-2235789. S.P. and P.K.P. acknowledge the Council of Scientific and Industrial Research (CSIR), Govt. of India, for the respective senior research fellowships.

The authors have no conflicts to disclose.

S. P. and P. K. P. performed the TR-MOKE measurements. S.P. analyzed the data and performed the micromagnetic simulations. R. F. fabricated the samples. S. B. and A. B. conceived the idea and supervised the project. All authors contributed to writing the paper.

S. Pal: Data curation (lead); Formal analysis (lead); Investigation (lead); Methodology (lead); Software (lead); Writing – original draft (lead). P. K. Pal: Data curation (equal); Formal analysis (supporting); Investigation (equal); Methodology (equal). R. Fabiha: Methodology (equal). S. Bandyopadhyay: Conceptualization (equal); Funding acquisition (equal); Project administration (equal); Supervision (equal); Validation (equal); Writing – review & editing (equal). A. Barman: Conceptualization (equal); Funding acquisition (equal); Project administration (equal); Supervision (equal); Validation (equal); Writing – review & editing (equal).

The data that support the findings of the study are available within the article.

1.
L.
Novotny
and
B.
Hecht
,
Principles of Nano-optics
(
Cambridge University Press
,
2006
).
2.
S. A.
Maier
,
M. L.
Brongersma
,
P. G.
Kik
,
S.
Meltzer
,
A. A. G.
Requicha
, and
H. A.
Atwater
, “
Plasmonics—A route to nanoscale optical devices
,”
Adv. Mater.
13
,
1501
1505
(
2001
).
3.
G.
Armelles
,
A.
Cebollada
,
A.
García-Martín
, and
M. U.
González
, “
Magnetoplasmonics: Combining magnetic and plasmonic functionalities
,”
Adv. Opt. Mater.
1
,
10
35
(
2013
).
4.
V. V.
Temnov
, “
Ultrafast acousto-magneto-plasmonics
,”
Nat. Photonics
6
,
728
736
(
2012
).
5.
P.
Chen
,
F.
Monticone
,
C.
Argyropoulos
, and
A.
Alù
,
Modern Plasmonics
(
Elsevier
,
2014
).
6.
V.
Giannini
,
A. I.
Fernández-Domínguez
,
S. C.
Heck
, and
S. A.
Maier
, “
Plasmonic nanoantennas: Fundamentals and their use in controlling the radiative properties of nanoemitters
,”
Chem. Rev.
111
,
3888
3912
(
2011
).
7.
A.
Bile
,
R.
Pepino
, and
E.
Fazio
, “
Study of magnetic switch for surface plasmon-polariton circuits
,”
AIP Adv.
11
,
045222
(
2021
).
8.
A.
Ahmed
,
R.
Gelfand
,
S. D.
Storm
,
A.
Lee
,
A.
Klinkova
,
J. R.
Guest
, and
M.
Pelton
, “
Low-frequency oscillations in optical measurements of metal-nanoparticle vibrations
,”
Nano Lett.
22
,
5365
5371
(
2022
).
9.
H. M.
Dong
,
Z. H.
Tao
,
Y. F.
Duan
,
F.
Huang
, and
C. X.
Zhao
, “
Coupled plasmon–phonon modes in monolayer MoS 2
,”
J. Phys.: Condens. Matter
32
,
125703
(
2019
).
10.
E. H.
Hwang
,
R.
Sensarma
, and
S.
Das Sarma
, “
Plasmon-phonon coupling in graphene
,”
Phys. Rev. B
82
,
195406
(
2010
).
11.
S.
Pal
,
P. K.
Pal
,
R.
Fabiha
,
S.
Bandyopadhyay
, and
A.
Barman
, “
Acousto-plasmo-magnonics: Coupling spin waves with hybridized phonon-plasmon waves in a 2d artificial magnonic crystal deposited on a plasmonic material
,”
Adv. Funct. Mater.
33
,
2304127
(
2023
).
12.
A.
Alekhin
,
A. M.
Lomonosov
,
N.
Leo
,
M.
Ludwig
,
V. S.
Vlasov
,
L.
Kotov
,
A.
Leitenstorfer
,
P.
Gaal
,
P.
Vavassori
, and
V.
Temnov
, “
Quantitative ultrafast magnetoacoustics at magnetic metasurfaces
,”
Nano Lett.
23
,
9295
9302
(
2023
).
13.
J.
Janušonis
,
C. L.
Chang
,
T.
Jansma
,
A.
Gatilova
,
V. S.
Vlasov
,
A. M.
Lomonosov
,
V. V.
Temnov
, and
R. I.
Tobey
, “
Ultrafast magnetoelastic probing of surface acoustic transients
,”
Phys. Rev. B
94
,
024415
(
2016
).
14.
A.
De
,
J. L.
Drobitch
,
S.
Majumder
,
S.
Barman
,
S.
Bandyopadhyay
, and
A.
Barman
, “
Resonant amplification of intrinsic magnon modes and generation of new extrinsic modes in a two-dimensional array of interacting multiferroic nanomagnets by surface acoustic waves
,”
Nanoscale
13
,
10016
10023
(
2021
).
15.
R.
Fabiha
,
J.
Lundquist
,
S.
Majumder
,
E.
Topsakal
,
A.
Barman
, and
S.
Bandyopadhyay
, “
Spin wave electromagnetic nano-antenna enabled by tripartite phonon-magnon-photon coupling
,”
Adv. Sci.
9
,
2104644
(
2022
).
16.
J. L.
Drobitch
,
A.
De
,
K.
Dutta
,
P. K.
Pal
,
A.
Adhikari
,
A.
Barman
, and
S.
Bandyopadhyay
, “
Extreme subwavelength magnetoelastic electromagnetic antenna implemented with multiferroic nanomagnets
,”
Adv. Mater. Technol.
5
,
2000316
(
2020
).
17.
Y.
Yahagi
,
B.
Harteneck
,
S.
Cabrini
, and
H.
Schmidt
, “
Controlling nanomagnet magnetization dynamics via magnetoelastic coupling
,”
Phys. Rev. B
90
,
140405
(
2014
).
18.
Y.
Yahagi
,
C.
Berk
,
B.
Hebler
,
S.
Dhuey
,
S.
Cabrini
,
M.
Albrecht
, and
H.
Schmidt
, “
Optical measurement of damping in nanomagnet arrays using magnetoelastically driven resonances
,”
J. Phys. D: Appl. Phys.
50
,
17LT01
(
2017
).
19.
N. K. P.
Babu
,
A.
Trzaskowska
,
P.
Graczyk
,
G.
Centała
,
S.
Mieszczak
,
H.
Głowiński
,
M.
Zdunek
,
S.
Mielcarek
, and
J. W.
Kłos
, “
The interaction between surface acoustic waves and spin waves: The role of anisotropy and spatial profiles of the modes
,”
Nano Lett.
21
,
946
951
(
2021
).
20.
C.
Berk
,
M.
Jaris
,
W.
Yang
,
S.
Dhuey
,
S.
Cabrini
, and
H.
Schmidt
, “
Strongly coupled magnon–phonon dynamics in a single nanomagnet
,”
Nat. Commun.
10
,
2652
(
2019
).
21.
N.
Vidal-Silva
,
E.
Aguilera
,
A.
Roldán-Molina
,
R. A.
Duine
, and
A. S.
Nunez
, “
Magnon polarons induced by a magnetic field gradient
,”
Phys. Rev. B
102
,
104411
(
2020
).
22.
F.
Godejohann
,
A. V.
Scherbakov
,
S. M.
Kukhtaruk
,
A. N.
Poddubny
,
D. D.
Yaremkevich
,
M.
Wang
,
A.
Nadzeyka
,
D. R.
Yakovlev
,
A. W.
Rushforth
,
A. V.
Akimov
, and
M.
Bayer
, “
Magnon polaron formed by selectively coupled coherent magnon and phonon modes of a surface patterned ferromagnet
,”
Phys. Rev. B
102
,
144438
(
2020
).
23.
T.
Hioki
,
Y.
Hashimoto
, and
E.
Saitoh
, “
Coherent oscillation between phonons and magnons
,”
Commun. Phys.
5
,
115
(
2022
).
24.
F.
Zhou
,
F.
Liu
,
L.
Xiao
,
K.
Cui
,
X.
Feng
,
W.
Zhang
, and
Y.
Huang
, “
Extending the frequency range of surface plasmon polariton mode with metamaterial
,”
Nano-Micro Lett.
9
,
9
(
2017
).
25.
A.
Barman
and
J.
Sinha
,
Spin Dynamics and Damping in Ferromagnetic Thin Films and Nanostructures
(
Springer
,
Berlin
,
2018
).
26.
M.
Donahue
and
D.
Porter
, “OOMMF user’s guide, version 1.0,” Interagency Report NISTIR 6376, National Institute of Standards and Technology, Gaithersburg, MD (1999).
27.
C.
Eyrich
,
W.
Huttema
,
M.
Arora
,
E.
Montoya
,
F.
Rashidi
,
C.
Burrowes
,
B.
Kardasz
,
E.
Girt
,
B.
Heinrich
,
O. N.
Mryasov
,
M.
From
, and
O.
Karis
, “
Exchange stiffness in thin film Co alloys
,”
J. Appl. Phys.
111
,
07C919
(
2012
).
28.
E.
Barati
,
M.
Cinal
,
D. M.
Edwards
, and
A.
Umerski
, “
Gilbert damping in magnetic layered systems
,”
Phys. Rev. B
90
,
014420
(
2014
).
29.
L.
Landau
and
E.
Lifshitz
,
Theory of Elasticity
(
Pergamon Press
,
1959
). Vol.
7
.
30.
S.
Mondal
,
M. A.
Abeed
,
K.
Dutta
,
A.
De
,
S.
Sahoo
,
A.
Barman
, and
S.
Bandyopadhyay
, “
Hybrid magnetodynamical modes in a single magnetostrictive nanomagnet on a piezoelectric substrate arising from magnetoelastic modulation of precessional dynamics
,”
ACS Appl. Mater. Interfaces
10
,
43970
43977
(
2018
).
31.
K.
Nishikawa
, “
Parametric excitation of coupled waves. ii. Parametric plasmon-photon interaction
,”
J. Phys. Soc. Jpn.
24
,
1152
1158
(
1968
).
32.
C.
Kittel
, “
On the gyromagnetic ratio and spectroscopic splitting factor of ferromagnetic substances
,”
Phys. Rev.
76
,
743
748
(
1949
).
33.
X.
Zhang
,
C.-L.
Zou
,
L.
Jiang
, and
H. X.
Tang
, “
Strongly coupled magnons and cavity microwave photons
,”
Phys. Rev. Lett.
113
,
156401
(
2014
).
34.
A. T.
Costa
,
M. I.
Vasilevskiy
,
J.
Fernández-Rossier
, and
N. M. R.
Peres
, “
Strongly coupled magnon–plasmon polaritons in graphene-two-dimensional ferromagnet heterostructures
,”
Nano Lett.
23
,
4510
4515
(
2023
).
35.
T.
Kaihara
,
T.
Ando
,
H.
Shimizu
,
V.
Zayets
,
H.
Saito
,
K.
Ando
, and
S.
Yuasa
, “
Enhancement of magneto-optical Kerr effect by surface plasmons in trilayer structure consisting of double-layer dielectrics and ferromagnetic metal
,”
Opt. Express
23
,
11537
11555
(
2015
).
36.
A. V.
Sadovnikov
,
E. N.
Beginin
,
M. A.
Morozova
,
Y. P.
Sharaevskii
,
S. V.
Grishin
,
S. E.
Sheshukova
, and
S. A.
Nikitov
, “
Nonlinear spin wave coupling in adjacent magnonic crystals
,”
Appl. Phys. Lett.
109
,
042407
(
2016
).