Abstract
Although recent experiments and theories have shown a variety of exotic transport properties of nonequilibrium quasiparticles (QPs) in superconductor (SC)-based devices with either Zeeman or exchange spin-splitting, how a QP interplays with magnon spin currents remains elusive. Here, using nonlocal magnon spin-transport devices where a singlet SC (Nb) on top of a ferrimagnetic insulator (Y3Fe5O12) serves as a magnon spin detector, we demonstrate that the conversion efficiency of magnon spin to QP charge via inverse spin-Hall effect (iSHE) in such an exchange-spin-split SC can be greatly enhanced by up to 3 orders of magnitude compared with that in the normal state, particularly when its interface superconducting gap matches the magnon spin accumulation. Through systematic measurements by varying the current density and SC thickness, we identify that superconducting coherence peaks and exchange spin-splitting of the QP density-of-states, yielding a larger spin excitation while retaining a modest QP charge-imbalance relaxation, are responsible for the giant QP iSHE. The latter exchange-field-modified QP relaxation is experimentally proved by spatially resolved measurements with varying the separation of electrical contacts on the spin-split Nb.
Keywords: nonlocal magnon spin transport, exchange-spin-split superconductor, quasiparticle spin-Hall effect, resonant absorption of magnon spin, exchange-field-frozen QP relaxation
Over the past decade, it has been shown that the combination of superconductivity with spintronics leads to a variety of phenomena that do not exist separately.1−8 In particular, recent discovery and progress in the proximity generation and control of spin-polarized triplet Cooper pairs1−3 at carefully engineered superconductor (SC)/ferromagnet (FM) interfaces in equilibrium allow for the development of nondissipative spin-based logic and memory technologies.
Besides triplet Cooper pairs, nonequilibrium quasiparticles (QPs) in a spin-split SC4−6 have also raised considerable interest. This is because their exotic properties resulting from the mutual coupling between different nonequilibrium imbalances of spin, charge, heat, and spin-heat can greatly enhance spintronics functionality.5 For example, the coupling of spin and heat imbalances gives rise to long-range QP spin signals as observed in Al-based nonlocal spin valves9−11 with a Zeeman spin-splitting field. In addition, a temperature gradient between a normal metal (NM) and a spin-split SC separated by a tunnel barrier induces a pure QP spin current12 without an accompanying net charge current, analogous to the spin-dependent Seebeck tunneling.13,14 Substituting the NM by a FM, one can achieve large (spin-dependent) thermoelectric currents15,16 far beyond those commonly found in all-metallic structures.
Magnon spintronics17−19 has been an emerging approach toward computing devices in which magnons, the quanta of spin waves, are used to carry, transport, and process spin information instead of conduction electrons. Especially in the low-damping ferrimagnetic insulator yttrium–iron–garnet (Y3Fe5O12 = YIG),20 a magnon-carried spin current can propagate over extremely long distances (centimeters at best), and it is free from ohmic dissipation due to the absence of electrons in motion.17−19 Despite many recent advances17−19 in this research field, how magnon spin current interacts with and is converted to QP spin and charge currents in a spin-split SC is yet to be investigated.
In this paper, we report three key aspects of the conversion behavior of magnon spin to QP charge via the inverse spin-Hall effect (iSHE) in an exchange-spin-split SC (Nb), directly probed by nonlocal magnon spin-transport18 (Figure 1a). First, the iSHE in the superconducting state of Nb becomes up to 3 orders of magnitude greater than in the normal state. Second, this enhancement appears only in the vicinity of the superconducting transition temperature Tc when the magnon spin current has an energy comparable to the (singlet) superconducting gap 2ΔSC of Nb (Figure 1a). Lastly, its characteristic dependence on a dc current density Jdc and the Nb thickness tNb indicates that a singularity near the gap edge and a spin-splitting field are both essential for the giant transition-state QP iSHE, the latter of which is experimentally confirmed by performing spatial profiling of the transition-state enhancement by varying the separation distance of electrical contacts on the spin-split Nb layer.
Results and Discussion
The nonlocal magnon spin-transport devices (Figure 1b) we study consist of two identical Pt electrodes and a central Nb layer on top of 200-nm-thick YIG films, which are liquid-phase epitaxially grown on a (111)-oriented single-crystalline gadolinium gallium garnet (Gd3Ga5O12, GGG) wafer (see Methods). Control devices, in which a 10-nm-thick Al2O3 spin-blocking layer is inserted between Nb and YIG in an otherwise identical structure, are also prepared for comparison (Figure 1b). Here, we send a dc current Idc through one Pt (using leads 1 and 2 in Figure 1b) while measuring the in-plane (IP) magnetic-field-angle α dependence of the nonlocal open-circuit voltages [VnlPt(α), Vnl(α)] using both the other Pt (leads 7 and 8) and the central Nb (leads 3 and 4). Note that we apply an external in-plane magnetic field μ0Hext of 5 mT, larger than the coercive field of YIG (Figure 1c), to fully align its magnetization MYIG along the field direction. α is defined as the relative angle of μ0Hext (//MYIG) to the long axis of two Pt electrodes, which are collinear.
As schematically illustrated in Figure 1a, the right Pt acts as a NM injector of magnon spin current across the Pt/YIG interface via either electron-mediated SHE (charge-to-spin conversion)21 or spin Seebeck effect (SSE) (heat-to-spin conversion)22 due to the accompanying Joule heating [ΔT ∝ (Idc)2]. The left Pt serves as an NM detector of the magnon spin current, diffusing through a YIG channel, via electron-mediated iSHE (spin-to-charge conversion), whereas in the same device, the middle Nb functions as an exchange-spin-split SC detector of the diffusive magnon current via QP-mediated iSHE below Tc.8
The total voltage measured across the detector is given by Vnltot = ΔVnl + ΔVnlth + V0, where ΔVnl and ΔVnlth are proportional to the magnon spin current and accumulation created electrically (SHE ∝ Idc)21 and thermally [SSE ∝ (Idc)2],22 respectively. These electrical and thermal magnon currents can be separated straightforwardly by reversing the polarity of Idc, allowing us to determine the magnitude of each component based on their distinctive angular dependences;18 and . V0 is an offset voltage that is independent of the magnon spin-transport.
The typical result of such a measurement using the Pt detector at 300 K is displayed in Figure 1d–i, for the tNb = 15 nm devices with and without the Al2O3 spin-blocking layer. This evidences that both electrically (Figure 1e and h) and thermally (Figure 1f and i) excited magnons transport spin angular momentum over a long distance of 15 μm at room temperature, which is consistent with the original work.18 We note that from reference devices having the Pt injector/detector only, the room-temperature magnon spin-diffusion length lsdm of the YIG is estimated to be around 11(9) μm for the electrically (thermally) driven magnons (Supplementary Section 1). The transporting spin current is absorbed by the middle Nb to a certain extent, given by the difference between the signals with versus without the Al2O3 insertion (see Supplementary Section 2 for the quantitative analysis).
Figure 2a,b,d,e show the temperature T evolution of ΔVnlel(α) and ΔVnl(α) for the tNb = 15 nm devices measured by the Pt detector at a fixed Idc = |0.5| mA (Jdc = |3.3| MA/cm2). As summarized in Figure 2f and g, ⌈ΔVnlel⌉Pt diminishes with decreasing the base temperature Tbase, and it almost vanishes for Tbase ≤ 10 K, whereas ⌈ΔVnl⌉Pt significantly increases at low Tbase. Such distinct Tbase-dependences are in line with previous experiments23,24 and theoretical considerations25,26 that the injection mechanisms for electrical and thermal magnons across the Pt/YIG interface (parametrized by the effective spin conductance and the interface spin Seebeck coefficient, respectively) differ fundamentally. Furthermore, the energy-dependent magnon diffusion and relaxation of the YIG channel may play a role in the transport process.27,28
We below focus on the nonlocal signal from the thermally generated magnons (ΔVnlth) since it remains sufficiently large at low Tbase for allowing a reliable analysis across Tc. In Figure 2f, we first plot the Tbase dependence of ⌈ΔVnl⌉Pt without the Al2O3 layer normalized by that with the Al2O3 layer; ⌈ΔVnlth⌉Pt,no Al2O3/⌈ΔVnl⌉Pt,with Al2O3. This value reflects how much the magnon spin current is absorbed by the Nb layer. Notably, ⌈ΔVnlth⌉Pt,no Al2O3/⌈ΔVnl⌉Pt,with Al2O3 drops abruptly right below Tc (extracted from the Nb resistance RNbversus Tbase plot of Figure 2c), and then it rises progressively as the Nb enters deep into the superconducting state, resulting in a downturn at Tbase/Tc ≈ 0.95 (inset of Figure 2f). Such a nontrivial behavior is compatible with recent theoretical predictions29,30 and experimental reports31,32 on ferromagnetic insulator (FMI)/SC structures, where (spin-singlet) Cooper pairs from the SC cannot leak into the FMI even if the exchange spin-splitting can still penetrate the SC.4 So rather well-developed coherence peaks of the QP density-of-states (DOS) at the FMI/SC interface5 are accessible to the transporting spin current. This gives rise to an anomalous enhancement of spin absorption by the adjacent SC near Tc. Note that in contrast, for metallic/conducting FM/SC proximity-coupled structures,33 2ΔSC is significantly suppressed at the FM/SC interface, and the superconducting coherence peak effect is therefore fading away.29,34−38 A slight rise in ⌈ΔVnlth⌉Pt,no Al2O3/⌈ΔVnl⌉Pt,with Al2O3 far below Tc (inset of Figure 2f) is also explained by the development of a (singlet) superconducting gap and the freeze out of the QP population at a lower Tbase.29,34−38
Next, using the Nb detector in the same device, we confirm the above interpretation and demonstrate that the conversion efficiency of magnon spin to QP charge can be dramatically enhanced in the vicinity of Tc. Figure 3a shows the thermally driven nonlocal signal ⌈ΔVnlth⌉Nb for the tNb = 15 nm devices with and without the Al2O3 (spin-blocking) layer at various Tbase around the superconducting transition of the Nb. In the normal state (Tbase/Tc > 1), a negative ⌈ΔVnl⌉Nb(<0) with several tens of nanovolts is clearly observed. Given ⌈ΔVnlth⌉Pt > 0 (see Figure 2b), this evidences that Nb and Pt have opposite signs in the spin-Hall angle θSH, which is in agreement with recent theoretical and experimental studies.38−40 Intriguingly, upon entering the superconducting state (Tbase/Tc < 1), a significant enhancement of ⌈ΔVnl⌉Nb up to a few microvolts appears immediately below Tc (Tbase/Tc ≈ 0.96), and then it decays toward zero deep in the superconducting state. We note that there exist visible dips in ⌈ΔVnlth⌉Nb at α ≈ 90° and 270° near Tc (Figure 3a), which are also present for the Al2O3-inserted control device (Figure 3b) and thus have nothing to do with the magnon spin-transport and QP iSHE. Similar spin-independent signals have been observed in local measurements on NbN/YIG32 and MoGe/YIG41 bilayers as well and are explained in terms of an Abrikosov-vortex-flow-driven Hall effect under a transverse magnetic field that is close to the upper critical field μ0Hc2 of (type-II) SC.
To examine the effect of heating power, we measure the Tbase dependence of ⌈ΔVnlth⌉Nb (Figure 3b and c) and the normalized RNb/RT=7K (Figure 3d) at various Idc. As Idc increases, Tc of the Nb detector is systematically reduced and the transition width becomes broad (Figure 3d). We note that the stronger decay of Tc for Idc > 0.5 mA is likely caused by the greater injection/excitation of spin-polarized QPs into the Nb detector (see Supplementary Section 3 for a comparison analysis of Tc data between the Al2O3-absent and Al2O3-present devices). Accordingly, not only a peak of the ⌈ΔVnlth⌉Nb enhancement shifts to a low Tbase, but the enhancement regime widens (Figure 3e). For Idc ≥ |0.7| mA (Jdc ≥ |4.2| MA/cm2), the Nb does not turn fully superconducting down to the lowest Tbase = 2 K studied (inset of Figure 3d). The corresponding ⌈ΔVnl⌉Nb then remains nonzero at 2 K and is larger than the normal state value (Figure 3e). For a quantitative analysis, we plot the normalized voltage ⌈ΔVnlth⌉Nb/⌈ΔVnl⌉T=7KNb as a function of the normalized temperature Tbase/Tc in Figure 3f. We then find that the transition-state enhancement of ⌈ΔVnl⌉Nb/⌈ΔVnlth⌉T=7K can reach up to 3 orders of magnitude at the smallest Idc = |0.1| mA (Jdc = |0.6| MA/cm2). With increasing Idc, its peak amplitude decays rapidly, the full-width-at-half-maximum (fwhm) broadens, and the peak is positioned farther away from Tc (inset of Figure 3f). These results ensure that the depressed superconductivity with increasing the heating power has a negative effect on the transition-state enhancement of the QP iSHE.
We perform similar measurements on an additional set of devices with different tNb (Figure 4a–f), comparable to or smaller than the superconducting coherence length ξSC, and thereby strong tNb-dependent superconducting properties (e.g., QP band structure and DOS). Since thin Nb films usually contain a larger amount of grain boundaries, defects, and disorders from the structural inhomogeneity near the growth interface than thick bulk Nb,42,43 the associated scattering effectively weakens electron–electron and electron–phonon interactions and therefore the smearing-out effect of the QP DOS around the gap edge.44 One would predict a greater enhancement of the QP iSHE if the Nb detector is thicker.
However, experiments give a very different result (Figure 4g–i). As tNb increases, the peak amplitude of ⌈ΔVnlth⌉Nb/⌈ΔVnl⌉T=7KNb rises until reaching 15 nm and then drops strongly for thicker Nb detectors, leading to a maximum at tNb = 15 nm (Figure 4j). The width and position of the transition-state enhancement, on the other hand, behave as expected for highly and quickly developed coherence peaks in the QP DOS of thick Nb when Tc is crossed: a progressive narrowing of fwhm and a peak shift closer to Tc, respectively, with the increase of tNb (inset of Figure 4j). The nontrivial tNb-dependent enhancement (Figure 4j) indicates that there is another key ingredient that controls the enhancement amplitude, that is to say, the exchange spin-splitting field,4−6 which has turned out to considerably modify the QP spin relaxation mechanism via a freezing out of elastic/intravalley spin-flip scattering.4−6 Below, we discuss how this exchange-field-frozen spin-flip scattering4−6 is linked to and modifies the QP charge relaxation.
To theoretically describe our results, we first calculate the excited QP spin current density Js0qp at the YIG/superconducting Nb interface as a function of the normalized temperature T/Tc for different values of the magnon spin accumulation Δμm relative to the zero-T energy gap 2Δ0 (Figure 5a and b). For this calculation, we employ the recent models29,30 that explicitly take the superconducting coherence factor into account (see Supplementary Section 4 for full details). Note that the characteristic energy of incoherent magnons which excite spin-polarized QPs in the Nb detector is set by Δμm, and the Tc (or 2Δ0SC) suppression at a larger Δμm is inferred from our data set (Figures 3 and 4). For a quantitative comparison, Js0 is normalized to its normal state value Js0.
The calculated Js0qp/Js0 increases largely near Tc (0.8Tc – 0.9Tc), and it decreases exponentially when T < 0.8Tc, reflecting the singularity behavior in a nonequilibrium population of spin-polarized QPs.29,30,42 In addition, the peak amplitude of Js0/Js0 is inversely proportional to Δμm/2Δ0SC (inset of Figure 5a and b), explaining qualitatively the heating power dependence of the transition-state enhancement (Figure 3f). Nonetheless, this analysis based on the superconducting coherence factor does not capture the mechanism behind the nontrivial tNb dependence (Figure 4j).
We next consider the QP resistivity ρSCqp (Figure 5c and d) and the volume fraction of QP charge imbalance vQ (Figure 5e and f), which together determine the effective resistivity ρSC (=ρSCqpvQ, inset of Figure 5e and f) of the superconducting Nb.38,45 Here ,38,45 where ly is the spin-active length of the Nb detector, given approximately by the sum of the length of the Pt injector ly and lsdm in our device geometry, and λQ is the QP charge-imbalance relaxation length. ρSC and ρSC∗ are normalized by their normal-state ones ρ0 and ρ0, respectively. We note that if the SC thickness is comparable to or smaller than the QP spin transport length, as relevant to our system,38,46 the QP-mediated iSHE voltage ViSHEqp in the SC can be approximated as , where θSH is the QP spin-Hall angle, which is predicted to slightly increase near Tc(45,47) (see Supplementary Section 4 for details), e is the electron charge, and ℏ is the reduced Planck constant. Consequently, Js0qp and ρSC appear to be governing parameters in ViSHEqp.
The most salient aspect of the calculations is that in the vicinity of Tc, vQ dominates the T-dependent ρSC∗ over ρSC, resulting in ViSHEqp ∝ λQ for given Js0 and ly values. This signifies that the QP charge imbalance relaxation is likely responsible for the nontrivial tNb-dependent transition-state enhancement (Figure 4j) observed in our system.
We thus propose the following mechanism. If QP charge relaxes through the spin-flip scattering 1/τsfqp and the inelastic scattering 1/τin, and 1/τsf > 1/τin, the effective relaxation time τQ∗ for the QP charge imbalance48 is given by where kB is the Boltzmann constant. Based on the exchange-field-frozen spin-flip scattering4−6 and its proximity nature33 in an FMI/SC system, one can reasonably assume τsf ∝ ΔEex ∝ . This leads to and . Qualitatively, we can understand the tNb-dependent transition-state enhancement (Figure 4j) in the following manner. When tNb ≪ ξNb, the superconducting coherence is too weak to inject/excite large QP spin currents across the YIG/Nb interface. In contrast, for tNb > ξNb, the exchange spin-splitting-field cannot propagate over the entire depth of such thick Nb and hence the converted QP charge relaxes faster primarily via the spin-flip scattering process. Overall, these two competing effects control the amplitude of the transition-state enhancement by which one would expect a maximum at the intermediate tNb ≈ ξNb (around 15 nm for Nb thin films). Note also that the enhancement width and position are determined by Js0qp × ρSC, the latter of which decays rapidly to zero below Tc for a strong superconducting Nb.
To check the validity of this proposal, we experimentally investigate how the transition-state enhancement scales with the separation distance ds between Au/Ru electrical contracts on the exchange-spin-split Nb layer (Figure 6a, see Supplementary Section 5). Importantly, while the peak position and width of the transition-state enhancement are almost independent of ds (Figure 6b and c), the peak amplitude increases quasi-exponentially with the increase of ds (inset of Figure 6c), reflecting the characteristics of the QP charge-imbalance relaxation effect (see Supplementary Section 5). From the ds-dependent ⌈ΔVnlth⌉Nb (Figure 6e), we are able to estimate λQ in the vicinity of Tc (Tbase/Tc = 0.94–0.98) for the spin-split Nb to be around 90 μm. This is surprisingly a few orders of magnitude larger than either commonly assumed48 or hitherto reported in Nb films without the presence of spin-splitting fields49 and thereby should indicate the significantly exchange-field-modified QP relaxation in our system.
Finally, we briefly mention other relevant experiments. It has been previously shown that in all-metallic nonlocal spin-Hall devices,8 the giant iSHE (∼2000 times at most) is created by electrical spin injection from Ni8Fe2 through Cu into superconducting NbN far below Tc (Tbase/Tc = 0.3) and attributed to the exponentially increasing QP resistivity at a lower T. By contrast, a recent experiment has reported that for a YIG/NbN vertical junction33 the 2–3 times enhanced iSHE voltage by local SSE is measurable only in a limited T range right below Tc (Tbase/Tc = 0.96). In this work, the superconducting coherence factor is pointed out as a main source for such an enhancement, and a quantitative description of the data is also provided. In metallic/conducting Nb/Ni8Fe2 bilayers,38 a monotonic decay of spin-pumping-induced iSHE appears across Tc, indicating no superconducting coherence effect detectable.
Conclusions
The key findings of our study that help understand these puzzling results are as follows. The spin-to-charge conversion mediated by QPs is substantially enhanced in the normal-to-superconducting transition regime, where the interface superconducting gap matches the magnon spin accumulation. The conversion efficiency and characteristics depend crucially on the driving/heating power and the SC thickness, which is understood based on the two competing effects: the superconducting coherence29,30,42 and the exchange-field-modified QP relaxation.4−6,48 The validity of these competing mechanisms is experimentally confirmed by spatially resolved measurements with varying the separation of electrical contacts on the spin-split Nb layer. A quantitative reproduction of the result remains an open question for a theory. The coupling between different nonequilibrium imbalances (magnon, spin, charge, heat, magnon-heat, and spin-heat)4,12 with exchange spin-splitting and the nonlinear kinetic equations4 in the superconducting state should be taken into account rigorously. Moreover, how the magnetic-field-induced screening supercurrents in a spin-split SC contribute to the QP spin-to-charge conversion when coupled with these nonequilibrium modes50 remains to be addressed. We speculate that the giant transition-state QP SHE is generic in any FMI/SC system, and its efficiency gets even larger especially with two-dimensional (2D) SCs51 where the exchange spin-splitting can readily proximity-penetrate the entire depth of the 2D SCs. We also anticipate that such a giant spin-to-charge conversion phenomenon (involving nonequilibrium QPs) can be used as an extremely sensitive probe of spin currents in emergent quantum materials.52
Methods
Device Fabrication
We fabricated the magnon spin-transport devices (Figure 1b) based on 200-nm-thick single-crystalline YIG films (from Matesy GmbH) by repeating a sequence of optical lithography, deposition, and lift-off steps. Note that these YIG films exhibited a very low Gilbert damping of 0.6 × 10–4 at room temperature, determined via ferromagnetic resonance line width measurements (by Matesy GmbH, https://www.matesy.de/en/products/materials/yig-single-crystal). We first defined the central Nb detector with a lateral dimension of 9 × 90 μm2, which was grown by accelerated Ar-ion beam sputtering at a working pressure of 1.5 × 10–4 mbar. For the control device, a 10-nm-thick Al2O3 spin-blocking layer was in situ deposited prior to the Nb deposition. We then defined a pair of Pt electrodes of 1.5 × 50 μm2, which were deposited by dc magnetron plasma sputtering at an Ar pressure of 4 × 10–3 mbar. These Pt electrodes are separated by a center-to-center distance dPt–Pt of 15 μm, which is comparable to the typical lsdm of single-crystalline YIG films18 and also to the estimated values from our Pt-only reference devices with different dPt–Pt (Supplementary Section 1). The Nb thickness ranges from 10 to 35 nm, whereas the Pt thickness is fixed at 10 nm. Finally, we defined Au(80 nm)/Ru(2 nm) electrical leads and bonding pads, which were deposited by the Ar-ion beam sputtering. Before depositing the Au/Ru layers, the Nb and Pt surfaces were gently Ar-ion beam etched for transparent electrical contacts between them.
Nonlocal Measurement
We measured the nonlocal magnon spin-transport (Figure 1a and b) in a quantum design physical property measurement system at a temperature varying between 2 and 300 K. A dc current Idc in the range of 0.1 to 1 mA was applied to the first Pt using a Keithley 6221 current source, and the nonlocal voltages [VnlPt(α), Vnl(α)] across the second Pt and the central Nb are simultaneously recorded as a function of in-plane magnetic-field-angle α by a Keithley 2182A nanovoltmeter. α is defined as the relative angle of μ0Hext (//MYIG) to the long axis of two Pt electrodes that are collinear.
Acknowledgments
This work was supported by the Alexander von Humboldt Foundation.
Supporting Information Available
The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acsnano.0c07187.
Estimation of the magnon spin-diffusion length of YIG, quantification of spin currents leaking into the central Nb at room temperature, first-order estimate of the YIG-induced internal field at the Nb/YIG interface, theoretical description of the conversion efficiency of magnon spin to QP charge in the superconducting Nb, spatially resolved measurements by varying the separation of electrical contacts on the spin-split Nb layer (PDF)
Author Contributions
K.-R.J. conceived and designed the experiments. The magnon spin-transport devices were fabricated by K.-R.J. with help from J.-C.J., X.Z., and A.M. The nonlocal transport measurements were carried out by K.-R.J. with the help of J.Y. and J.-C.J. K.-R.J. performed the data analysis and model calculation. S.P.P.P. supervised the project. All authors discussed the results and commented on the manuscript, which was written by K.-R.J.
The authors declare no competing financial interest.
Supplementary Material
References
- Linder J.; Robinson J. W. A. Superconducting Spintronics. Nat. Phys. 2015, 11, 307–315. 10.1038/nphys3242. [DOI] [Google Scholar]
- Eschrig M. Spin-Polarized Supercurrents for Spintronics: A Review of Current Progress. Rep. Prog. Phys. 2015, 78, 104501. 10.1088/0034-4885/78/10/104501. [DOI] [PubMed] [Google Scholar]
- Birge N. O. Spin-Triplet Supercurrents in Josephson Junctions Containing Strong Ferromagnetic Materials. Philos. Trans. R. Soc., A 2018, 376, 20150150. 10.1098/rsta.2015.0150. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Beckmann D. Spin Manipulation in Nanoscale Superconductors. J. Phys.: Condens. Matter 2016, 28, 163001. 10.1088/0953-8984/28/16/163001. [DOI] [PubMed] [Google Scholar]
- Bergeret F. S.; Silaev M.; Virtanen P.; Heikkilä T. T. Colloquium: Nonequilibrium Effects in Superconductors with a Spin-Splitting Field. Rev. Mod. Phys. 2018, 90, 041001 10.1103/RevModPhys.90.041001. [DOI] [Google Scholar]
- Quay C. H. L.; Aprili M. Out-of-Equilibrium Spin Transport in Mesoscopic Superconductors. Philos. Trans. R. Soc., A 2018, 376, 20150342. 10.1098/rsta.2015.0342. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Yang H.; Yang S.-H.; Takahashi S.; Maekawa S.; Parkin S. S. P. Extremely Long Quasiparticle Spin Lifetimes in Superconducting Aluminium Using MgO Tunnel Spin Injectors. Nat. Mater. 2010, 9, 586–593. 10.1038/nmat2781. [DOI] [PubMed] [Google Scholar]
- Wakamura T.; Akaike H.; Omori Y.; Niimi Y.; Takahashi S.; Fujimaki A.; Maekawa A.; Otani Y. Quasiparticle-Mediated Spin Hall Effect in a Superconductor. Nat. Mater. 2015, 14, 675–678. 10.1038/nmat4276. [DOI] [PubMed] [Google Scholar]
- Hübler F.; Lemyre J. C.; Beckmann D.; Löhneysen H. v. Charge Imbalance in Superconductors in the Low-Temperature Limit. Phys. Rev. B: Condens. Matter Mater. Phys. 2010, 81, 184524. 10.1103/PhysRevB.81.184524. [DOI] [Google Scholar]
- Hübler F.; Wolf M. J.; Beckmann D.; Löhneysen H. v. Long-Range Spin-Polarized Quasiparticle Transport in Mesoscopic Al Superconductors with a Zeeman Splitting. Phys. Rev. Lett. 2012, 109, 207001. 10.1103/PhysRevLett.109.207001. [DOI] [PubMed] [Google Scholar]
- Quay C. H. L.; Chevallier D.; Bena C.; Aprili M. Spin Imbalance and Spin-Charge Separation in a Mesoscopic Superconductor. Nat. Phys. 2013, 9, 84–88. 10.1038/nphys2518. [DOI] [Google Scholar]
- Silaev M.; Virtanen P.; Bergeret F. S.; Heikkilä T. T. Long-Range Spin Accumulation from Heat Injection in Mesoscopic Superconductors with Zeeman Splitting. Phys. Rev. Lett. 2015, 114, 167002. 10.1103/PhysRevLett.114.167002. [DOI] [PubMed] [Google Scholar]
- Walter M.; Walowski J.; Zbarsky V.; Münzenberg M.; Schäfers M.; Ebke D.; Reiss G.; Thomas A.; Peretzki P.; Seibt M.; Moodera J. S.; Czerner M.; Bachmann M.; Heiliger C. Seebeck Effect in Magnetic Tunnel Junctions. Nat. Mater. 2011, 10, 742–746. 10.1038/nmat3076. [DOI] [PubMed] [Google Scholar]
- Le Breton J. C.; Sharma S.; Saito H.; Yuasa S.; Jansen R. Thermal Spin Current from a Ferromagnet to Silicon by Seebeck Spin Tunneling. Nature 2011, 475 (2011), 82–85. 10.1038/nature10224. [DOI] [PubMed] [Google Scholar]
- Machon P.; Eschrig M.; Belzig W. Nonlocal Thermoelectric Effects and Nonlocal Onsager relations in a Three-Terminal Proximity-Coupled Superconductor-Ferromagnet Device. Phys. Rev. Lett. 2013, 110, 047002 10.1103/PhysRevLett.110.047002. [DOI] [PubMed] [Google Scholar]
- Kolenda S.; Wolf M. J.; Beckmann D. Observation of Thermoelectric Currents in High-Field Superconductor-Ferromagnet Tunnel Junctions. Phys. Rev. Lett. 2016, 116, 097001 10.1103/PhysRevLett.116.097001. [DOI] [PubMed] [Google Scholar]
- Chumak A. V.; Vasyuchka V. I.; Serga A. A.; Hillebrands B. Magnon Spintronics. Nat. Phys. 2015, 11, 453–461. 10.1038/nphys3347. [DOI] [Google Scholar]
- Cornelissen L. J.; Liu J.; Duine R. A.; Ben Youssef J.; van Wees B. J. Long-Distance Transport of Magnon Spin Information in a Magnetic Insulator at Room Temperature. Nat. Phys. 2015, 11, 1022–1026. 10.1038/nphys3465. [DOI] [Google Scholar]
- Kajiwara Y.; Harii K.; Takahashi S.; Ohe J.; Uchida K.; Mizuguchi M.; Umezawa H.; Kawai H.; Ando K.; Takanashi K.; Maekawa S.; Saitoh E. Transmission of Electrical Signals by Spin-Wave Interconversion in a Magnetic Insulator. Nature 2010, 464, 262–266. 10.1038/nature08876. [DOI] [PubMed] [Google Scholar]
- Yang F.; Hammel R. C. FMR-Driven Spin Pumping in Y3Fe5O12-Based Structures. J. Phys. D: Appl. Phys. 2018, 51, 253001. 10.1088/1361-6463/aac249. [DOI] [Google Scholar]
- Sinova J.; Valenzuela S. O.; Wunderlich J.; Back C. H.; Jungwirth T. Spin Hall Effects. Rev. Mod. Phys. 2015, 87, 1213–1259. 10.1103/RevModPhys.87.1213. [DOI] [Google Scholar]
- Ganzhorn K.; Wimmer T.; Cramer J.; Schlitz R.; Geprägs S.; Jakob G.; Gross R.; Huebl H.; Kläui M.; Goennenwein S. T. B. Temperature Dependence of the Non-Local Spin Seebeck Effect in YIG/Pt Nanostructures. AIP Adv. 2017, 7, 085102 10.1063/1.4986848. [DOI] [Google Scholar]
- Goennenwein S. T. B.; Schlitz R.; Pernpeintner M.; Ganzhorn K.; Althammer M.; Gross R.; Huebl H. Non-Local Magnetoresistance in YIG/Pt Nanostructures. Appl. Phys. Lett. 2015, 107, 172405. 10.1063/1.4935074. [DOI] [Google Scholar]
- Cornelissen L. J.; Shan J.; van Wees B. J. Temperature Dependence of the Magnon Spin Diffusion Length and Magnon Spin Conductivity in the Magnetic Insulator Yttrium Iron Garnet. Phys. Rev. B: Condens. Matter Mater. Phys. 2016, 94 (R), 180402. 10.1103/PhysRevB.94.180402. [DOI] [Google Scholar]
- Cornelissen L. J.; Peters K. J. H.; Bauer G. E. W.; Duine R. A.; van Wees B. J. Magnon Spin Transport Driven by the Magnon Chemical Potential in a Magnetic Insulator. Phys. Rev. B: Condens. Matter Mater. Phys. 2016, 94, 014412 10.1103/PhysRevB.94.014412. [DOI] [Google Scholar]
- Bender S. A.; Duine R. A.; Tserkovnyak Y. Electronic Pumping of Quasi equilibrium Bose–Einstein-Condensed Magnons. Phys. Rev. Lett. 2012, 108, 246601. 10.1103/PhysRevLett.108.246601. [DOI] [PubMed] [Google Scholar]
- Shan J.; Cornelissen L. J.; Vlietstra N.; Ben Youssef J.; Kuschel T.; Duine R. A.; van Wees B. J. Influence of Yttrium Iron Garnet Thickness and Heater Opacity on the Nonlocal Transport of Electrically and Thermally Excited Magnons. Phys. Rev. B: Condens. Matter Mater. Phys. 2016, 94, 174437. 10.1103/PhysRevB.94.174437. [DOI] [Google Scholar]
- Giles B. L.; Yang Z.; Jamison J. S.; Gomez-Perez J. M.; Vélez S.; Hueso L. E.; Casanova F.; Myers R. C. Thermally Driven Long-Range Magnon Spin Currents in Yttrium Iron Garnet Due to Intrinsic Spin Seebeck Effect. Phys. Rev. B: Condens. Matter Mater. Phys. 2017, 96 (R), 180412. 10.1103/PhysRevB.96.180412. [DOI] [Google Scholar]
- Inoue M.; Ichioka M.; Adachi H. Spin Pumping into Superconductors: A New Probe of Spin Dynamics in a Superconducting Thin Film. Phys. Rev. B: Condens. Matter Mater. Phys. 2017, 96, 024414 10.1103/PhysRevB.96.024414. [DOI] [Google Scholar]
- Kato T.; Ohnuma Y.; Matsuo M.; Rech J.; Jonckheere T.; Martin T. Microscopic Theory of Spin Transport at the Interface between a Superconductor and a Ferromagnetic Insulator. Phys. Rev. B: Condens. Matter Mater. Phys. 2019, 99, 144411. 10.1103/PhysRevB.99.144411. [DOI] [Google Scholar]
- Yao Y.; Song Q.; Takamura Y.; Cascales J. P.; Yuan W.; Ma Y.; Yun Y.; Xie X.; Moodera J. S.; Han W. Probe of Spin Dynamics in Superconducting NbN Thin Films via Spin Pumping. Phys. Rev. B: Condens. Matter Mater. Phys. 2018, 97, 224414. 10.1103/PhysRevB.97.224414. [DOI] [Google Scholar]
- Umeda M.; Shiomi Y.; Kikkawa T.; Niizeki T.; Lustikova J.; Takahashi S.; Saitoh E. Spin-Current Coherence Peak in Superconductor/Magnet Junctions. Appl. Phys. Lett. 2018, 112, 232601. 10.1063/1.5027456. [DOI] [Google Scholar]
- Buzdin A. I. Proximity Effects in Superconductor-Ferromagnet Heterostructures. Rev. Mod. Phys. 2005, 77, 935–976. 10.1103/RevModPhys.77.935. [DOI] [Google Scholar]
- Wakamura T.; Hasegawa N.; Ohnishi K.; Niimi Y.; Otani Y. Spin Injection into a Superconductor with Strong Spin-Orbit Coupling. Phys. Rev. Lett. 2014, 112, 036602 10.1103/PhysRevLett.112.036602. [DOI] [PubMed] [Google Scholar]
- Ohnishi K.; Ono Y.; Nomura T.; Kimura T. Significant Change of Spin Transport Property in Cu/Nb Bilayer Due to Superconducting Transition. Sci. Rep. 2015, 4, 6260. 10.1038/srep06260. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Morten J. P.; Brataas A.; Bauer G. E. W.; Belzig W.; Tserkovnyak Y. Proximity-Effect–Assisted Decay of Spin Currents in Superconductors. Eur. Phys. Lett. 2008, 84, 57008. 10.1209/0295-5075/84/57008. [DOI] [Google Scholar]
- Bell C.; Milikisyants S.; Huber M.; Aarts J. Spin Dynamics in a Superconductor-Ferromagnet Proximity System. Phys. Rev. Lett. 2008, 100, 047002 10.1103/PhysRevLett.100.047002. [DOI] [PubMed] [Google Scholar]
- Jeon K.-R.; Ciccarelli C.; Kurebayashi H.; Wunderlich J.; Cohen L. F.; Komori S.; Robinson J. W. A.; Blamire M. G. Spin-Pumping-Induced Inverse Spin-Hall Effect in Nb/Ni80Fe20 Bilayers and Its Strong Decay Across the Superconducting Transition Temperature. Phys. Rev. Appl. 2018, 10, 014029 10.1103/PhysRevApplied.10.014029. [DOI] [Google Scholar]
- Tanaka T.; Kontani H.; Naito M.; Naito T.; Hirashima D. S.; Yamada K.; Inoue J. Intrinsic Spin Hall Effect and Orbital Hall Effect in 4D and 5D Transition Metals. Phys. Rev. B: Condens. Matter Mater. Phys. 2008, 77, 165117. 10.1103/PhysRevB.77.165117. [DOI] [Google Scholar]
- Morota M.; Niimi Y.; Ohnishi K.; Wei D. H.; Tanaka T.; Kontani H.; Kimura T.; Otani Y. Indication of Intrinsic Spin Hall Effect in 4d and 5d Transition Metals. Phys. Rev. B: Condens. Matter Mater. Phys. 2011, 83, 174405. 10.1103/PhysRevB.83.174405. [DOI] [Google Scholar]
- Lustikova J.; Shiomi Y.; Yokoi N.; Kabeya N.; Kimura N.; Ienaga K.; Kaneko S.; Okuma S.; Takahashi S.; Saitoh E. Vortex Rectenna Powered by Environmental Fluctuations. Nat. Commun. 2018, 9, 4922. 10.1038/s41467-018-07352-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Minhaj M. S. M.; Meepagala S.; Chen J. T.; Wenger L. E. Thickness Dependence on the Superconducting Properties of Thin Nb Films. Phys. Rev. B: Condens. Matter Mater. Phys. 1994, 49, 15235. 10.1103/PhysRevB.49.15235. [DOI] [PubMed] [Google Scholar]
- Gubin A. I.; Il’in K. S.; Vitusevich S. A. Dependence of Magnetic Penetration Depth on the Thickness of Superconducting Nb Thin Flms. Phys. Rev. B: Condens. Matter Mater. Phys. 2005, 72, 064503 10.1103/PhysRevB.72.064503. [DOI] [Google Scholar]
- Crow J. E.; Strongin M.; Thomson R. S.; Kammerer O. F. The Superconducting Transition Temperatures of Disordered Nb, W, and Mo Films. Phys. Lett. A 1969, 30, 161. 10.1016/0375-9601(69)90910-4. [DOI] [Google Scholar]
- Takahashi S.; Maekawa S. Hall Effect Induced by a Spin-Polarized Current in Superconductors. Phys. Rev. Lett. 2002, 88, 116601. 10.1103/PhysRevLett.88.116601. [DOI] [PubMed] [Google Scholar]
- Gu J. Y.; Caballero J. A.; Slater R. D.; Loloee R.; Pratt W. P. Direct Measurement of Quasiparticle Evanescent Waves in a Dirty Superconductor. Phys. Rev. B: Condens. Matter Mater. Phys. 2002, 66, 140507. 10.1103/PhysRevB.66.140507. [DOI] [Google Scholar]
- Takahashi S.; Maekawa S. Spin Hall Effect in Superconductors. Jpn. J. Appl. Phys. 2012, 51, 010110 10.7567/JJAP.51.010110. [DOI] [Google Scholar]
- Stuivinga M.; Ham C. L. G.; Klapwijk T. M.; Mooij J. E. Phase-Slip Centers in Superconducting Aluminum Strip. J. Low Temp. Phys. 1983, 53, 633–671. 10.1007/BF00683998. [DOI] [Google Scholar]
- Johnson M. Spin Coupled Resistance Observed in Ferromagnet-Superconductor-Ferromagnet Trilayers. Appl. Phys. Lett. 1994, 65, 1460–1462. 10.1063/1.112015. [DOI] [Google Scholar]
- Aikebaier F.; Silaev M. A.; Heikkilä T. T. Supercurrent-Induced Charge-Spin Conversion in Spin-Split Superconductors. Phys. Rev. B: Condens. Matter Mater. Phys. 2018, 98, 024516 10.1103/PhysRevB.98.024516. [DOI] [Google Scholar]
- Žutić I.; Matos-Abiague A.; Scharf B.; Dery H.; Belashchenko K. Proximitized Materials. Mater. Today 2019, 22, 85–107. 10.1016/j.mattod.2018.05.003. [DOI] [Google Scholar]
- Han W.; Maekawa S.; Xie X.-C. Spin Current as a Probe of Quantum Materials. Nat. Mater. 2020, 19, 139–152. 10.1038/s41563-019-0456-7. [DOI] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.