The quantum skyrmion Hall effect in electron systems
Abstract
The flow of electric current through a two-dimensional material in a magnetic field gives rise to the family of Hall effects. The quantum versions of these effects accommodate robust electronic edge channels and fractional charges. Recently, the Hall effect of skyrmions, classical magnetic quasiparticles with a quantized topological charge, has been theoretically and experimentally reported, igniting ideas on a quantum version of this effect. To this end, we perform dynamical mean field theory calculations on localized electrons coupled to itinerant electrons in the presence of spin-orbit interaction and a magnetic field. Our calculations reveal localized nano quantum skyrmions that start moving transversally when a charge current in the itinerant electrons is applied. The results show the time-transient build-up of the quantum skyrmion Hall effect, accompanied by an Edelstein effect and a magnetoelectric effect that rotate the spins. This work motivates studies about the steady state of the quantum skyrmion Hall effect, looking for eventual quantum skyrmion edge channels and their transport properties.
I Introduction
From fundamental physical processes to application-oriented information storage and processing, the stability of a physical effect is paramount. Some physical effects, especially quantum Hall effects, accommodate observables that are topologically protected, i.e., they are robust against a continuous deformation of selected parameters. Recently, experimental and theoretical studies have found topologically protected classical magnetic structures in thin films or effectively two-dimensional systems, which have been coined magnetic skyrmions Bogdanov (1995); Bogdanov and Hubert (1999); Mühlbauer et al. (2009); Yu et al. (2010); Heinze et al. (2011), connected to earlier ideas in particle physics Skyrme (1962). The stability of these objects and the possibility of creating them by electrical currents or time-controlled magnetic boundary conditions Everschor-Sitte et al. (2017); Stier et al. (2017); Schäffer et al. (2020); Siegl et al. (2022a) promote the idea of using them in spintronics and as information carriers Parkin et al. (2008); Tomasello et al. (2014); Fert et al. (2017). Furthermore, in sight of the ongoing miniaturization of magnetic skyrmions, they have also been proposed as ingredients in quantum computing Psaroudaki and Panagopoulos (2021).
Classical magnetic skyrmions experience an additional drag transversal to the direction of an applied electric current, which leads to an accumulation of skyrmions at one side Thiele (1973); Everschor et al. (2011, 2012); Iwasaki et al. (2013a, b); Sampaio et al. (2013); Moon et al. (2022); Jiang et al. (2016); Litzius et al. (2016). The angle between the direction of motion and the direction of the current is called the Magnus angle of the skyrmion Hall effect. The question arises if there is a quantum version of the skyrmion Hall effect and, if so, which characteristics of the electronic Hall effects transfer, including hypothetical skyrmion edge channels with their quantized conductance. A previous study treating the quantum skyrmion as a product state and calculating an effective action for the quantum skyrmion demonstrated the existence of the Magnus force Takashima et al. (2016). Yet, the challenge in describing general quantum skyrmions comes with the large Hilbert spaces that need to be considered in two-dimensional spin systems carrying quantum skyrmions of a size of minimally spins Sotnikov et al. (2021); Lohani et al. (2019); Siegl et al. (2022b), which demand special theoretical techniques like density matrix renormalization group Haller et al. (2022) or artificial neural networks Yoshi et al. (2023) to investigate them numerically.
Another method to analyze quantum skyrmions is the use of localized spins from interacting electrons, like electrons, to represent the skyrmions in a correlated electronic system Kobayashi and Hayami (2022). Representing the skyrmions as electronic degrees of freedom has the advantage that we can treat considerably large quantum spin systems with established advanced numerical techniques for correlated electronic systems and that we can apply an electric current within the model without further assumptions. Interestingly, skyrmions in -electron systems have been experimentally detected in EuAl4, in which the skyrmions have been treated classically Takagi et al. (2022). Yet, the quantum nature of skyrmions in strongly correlated electronic systems is not well studied. Ultimately, a quantum skyrmion Hall effect could have direct practical applications extenting the manifold of suggested technical applications of magnetic skyrmions Back et al. (2020) to the quantum world. Moreover, fundamental questions about the topological nature of quantum skyrmions, which, strictly speaking, gets lost because of quantum spin slip processes Kim and Tserkovnyak (2016); Posske and Thorwart (2019); Siegl et al. (2022b); Vijayan et al. (2023), could be answered when quantum skyrmions are connected to quantum Hall effects and their unambiguous topological origin.
In this paper, we numerically study a square lattice of localized electrons that are coupled to two-dimensional itinerant conduction () electrons in the presence of spin-orbit coupling and a small magnetic field perpendicular to the two-dimensional plane. Using dynamical mean-field theory, we reliably identify regions in parameter space that host quantum nano skyrmions. We subsequently study the effect of a current in the itinerant electrons on the quantum skyrmion in linear response theory and find a strong initial drag into the direction perpendicular to the current, marking the onset of the quantum skyrmion Hall effect. The shift of the skyrmion is accompanied by an Edelstein effect and a magnetoelectric effect,Edelstein (1990); Culcer and Winkler (2007); Chernyshov et al. (2009); Manchon and Zhang (2008); Garate and MacDonald (2009); Peters and Yanase (2018); Fiebig (2005) which leads to a rotation of the localized -electron spins. Our study stimulates further investigation of the quantum skyrmion Hall effect, especially its steady state, and possibly quantized skyrmion edge channels.
The remainder of this paper is structured as follows: In Sec. II, we introduce the model and the method. In Sec. III, we analyze the stability and structure of the quantum skyrmions for different model parameters. This is followed by Sec. IV, where we demonstrate the onset of the skyrmion Hall effect using linear response theory. Finally, in Sec. V, we discuss our results and conclude the paper.
II Model and Method
Motivated by the discovery of magnetic skyrmions in Eu-compounds Takagi et al. (2022), including partially filled electrons, we focus here on magnetically ordered ground states and low-energy metastable states in -electron systems on a square lattice with a lattice constant of on the order of half a nanometer. In particular, we study the ground states of a noncentrosymmetric -electron system described by a periodic Anderson model Yanase and Sigrist (2008); Peters and Yanase (2018); Michishita and Peters (2019). It is important to note that we explicitly start with an electronic Hamiltonian instead of a quantum spin model. Thus, charge fluctuations and other effective interactions besides the effective Heisenberg and Dzyaloshinskii–Moriya (DM) interaction generally affect the ground state. Furthermore, due to the hybridization between the itinerant conduction () electrons and the electrons, the magnetic moments generated by the electrons are intrinsically coupled to the electrons. Such a coupling, which is necessary to observe skyrmion Hall and skyrmion drag effects, does hence not need to be inserted manually but is naturally included.
Our model Hamiltonian can be split into a single-particle part, , and an interaction part, . The single-particle Hamiltonian is
(1) | |||||
where is the spinor containing the momentum space annihilation operators of the itinerant electrons and electrons, respectively, corresponding to the real-space operators and at site of a square lattice with spin . The matrices and denote the Pauli matrices on the spin and sublattice space, respectively, where are the bare Pauli matrices. The particle number operators are and . The strength of the nearest neighbor hopping of the electrons on the square lattice is denoted by . Throughout this paper, we use as the unit of energy. and are local energies of the and electrons, respectively. is a local hybridization between the and electrons as commonly used in the periodic Anderson model. corresponds to a small magnetic field applied in the direction. Finally, we include a spin-orbit coupling between the and electrons with hopping amplitude . This spin-orbit coupling corresponds to a Rashba-type spin-orbit interaction as present in noncentrosymmetric -electron systems Yanase and Sigrist (2008). The interaction part of the Hamiltonian is
(2) |
corresponding to a density-density interaction between electrons on the same lattice site. The full Hamiltonian is
(3) |
The calculations are performed on a finite lattice with periodic boundary conditions.
To find the ground state of this quantum model, we use the real-space dynamical mean-field theory (RDMFT)Georges et al. (1996); Potthoff and Nolting (1999); Vahedi et al. (2021); Peters and Kawakami (2015, 2014). RDMFT maps each atom of a unit cell (finite lattice) on its own quantum impurity model by calculating the local Green’s function
(4) |
where is the single-particle matrix of the Fourier transform of in Eq. (1) into real space, i.e., . Here, and are super indices including the lattice positions, the - sublattice, and the spin. Furthermore, is a matrix including the local self-energies of each lattice site in the finite lattice, where, by the defining approximation of RDMFT, vanishes when the spatial components of and differ. Writing the local Green’s function as
(5) |
we can map each lattice site on a quantum impurity model, where is the local hybridization of the impurity model. This hybridization function describes the environment for one lattice site created by the rest of the lattice. Here, the self-energy differs for each lattice site, and hence this hybridization function is different for each lattice site. Summarizing the numerical procedure, the local hybridization functions define quantum impurity models, which are solved to obtain the local self-energy for each lattice site. These updated self-energies are then used in Eq. (4), which defines a self-consistency problem. To calculate the self-energy of each lattice site, we use the numerical renormalization group (NRG) Wilson (1975); Bulla et al. (2008); Peters et al. (2006), which can calculate accurate Green’s functions and self-energies at low temperatures.
The magnetic properties of the periodic Anderson model without Rashba spin-orbit interaction are well understood within the DMFT approximation Georges et al. (1996). At half-filling, , on a square lattice, the periodic Anderson model orders antiferromagnetically for weak hybridization strengths Peters and Kawakami (2015). For large hybridization strengths, the periodic Anderson model at half-filling becomes a Kondo insulator. On the other hand, when the number of electrons is small and the electrons are nearly half-filled, the system orders ferromagnetically Peters et al. (2012). This paper aims to study the existence and properties of magnetic skyrmions in a ferromagnetic periodic Anderson model, including Rashba spin-orbit interaction. We thus look for parameters where the electrons are nearly half-filled, and the -electron filling is about .
An exhaustive search of the parameter space of the periodic Anderson model for stable quantum skyrmions in the ground state is numerically unfeasible. In advance to our fully quantum-mechanical calculations, we therefore first identify candidate parameter regions where the ground state or low-energy metastable states accommodate magnetic skyrmions. We do so by mapping Eq. (1) to a classical Heisenberg spin model with nearest-neighbor coupling by integrating out the electrons using second-order perturbation theory, which obtains the RKKY spin-spin interactions Ruderman and Kittel (1954); Kasuya (1956); Yosida (1957). We then use classical Monte Carlo methods to find the ground states of these spin models Neuhaus-Steinmetz et al. (2022). In particular, we have varied in this procedure the local hybridization , the spin-orbit coupling , and the -electron level position, . We subsequently transfer parameter configurations where we find a classical skyrmion to the quantum model and conduct RDMFT calculations to obtain the system’s ground state. Here, the presence of magnetic skyrmions in the corresponding classical model generally is a good indicator for quantum skyrmions in the quantum model. Setting and , corresponding to half-filling of the electrons, we find quantum skyrmions in a ferromagnetic background for , , and a finite spin-orbit coupling in combination with a magnetic field, in agreement with previous results on classical and quantum magnetic skyrmions Sotnikov et al. (2021); Lohani et al. (2019); Siegl et al. (2022b); Haller et al. (2022). In the RDMFT calculations, we vary the strength of the spin-orbit interaction, , and the strength of the magnetic field, , in the region according to the results of the classical calculations.
III Structure and stability of magnetic skyrmions in the periodic Anderson model
To unambiguously identify a magnetic skyrmion, we break the spin translation symmetry of the model in the first DMFT iteration. By this, we select a specific state of the translationally invariant space of ground states. We use two different strategies in our calculations. We either start with a ferromagnetic solution where all electrons point downwards and flip a single electron upwards. Alternatively, we directly start with a magnetic skyrmion solution obtained for a different set of parameters. Then, by iterating the DMFT self-consistency equation, we find possible, stable magnetic skyrmion solutions when the algorithm converges. In the skyrmion phase, both initial states lead to identical DMFT solutions. We show the convergence of the self-energies for a typical skyrmion solution in Appendix B.
To verify the existence of a magnetic skyrmion, we calculate the spin expectation values of the and electrons for each lattice site,
(6) | |||||
(7) |
where corresponds to the coordinates of a lattice site, and is the vector containing the spin space Pauli matrices. Using these spin expectation values, we calculate the local lattice skyrmion density for the and electrons based on the solid angle spanned by three vectors as
(8) | ||||
where either stands for or electrons, , , and are nearest-neighbor lattice sites spanning an elemental triangle in the densest triangular tessellation of the lattice. The sum of this skyrmion density over all triangles spanning the square lattice yields the skyrmion number
(9) |
Unlike in a classical calculation, the spin expectation values in a quantum model do not need to be in magnitude. In fact, these expectation values are usually smaller due to quantum fluctuations, . We thus calculate two types of skyrmion densities: One is the quantum skyrmion density/number using unnormalized spin expectation values. The second type is a classical skyrmion density, where we normalize all spin expectation values to before using them in Eq. (8). The skyrmion number is an integer when using normalized spin expectation values. When using unnormalized spin expectation values, the skyrmion number is not quantized, and instead, its magnitude is an indicator of the skyrmion stability Siegl et al. (2022b), similar to the scalar chirality defined in Ref. Sotnikov et al. (2021).

A representative magnetic skyrmion solution is shown in Fig. 1 calculated for a spin-orbit coupling and a magnetic field . We note that within the accuracy of our calculations, we cannot find discernible energy differences between the ferromagnetic configuration and the magnetic skyrmion. The described skyrmions can, therefore, be metastable excitations on a ferromagnetic background with almost vanishing excitation energy or present in the ground state itself. Such an almost degenerate situation is favorable for applications in racetrack devices. If skyrmions were energetically strongly favorable, a skyrmion lattice would form instead of individual skyrmions. Figure 1 shows the spin texture of the and electrons underlaid with the local skyrmion density for normalized spin expectation values as 2D color plot, see Eq. (8). Due to the local hybridization, , which leads to an effective antiferromagnetic interaction between the and electrons, the spins of the and electrons mostly point in opposite directions, with deviations in the skyrmion’s center, where the Rashba interaction and the itinerant character of the electrons play a stronger role. The combined state corresponds to a bound skyrmion-antiskyrmion pair where the electrons form a magnetic skyrmion with skyrmion number and the electrons form a magnetic antiskyrmion with skyrmion number . However, in this system, the spin expectation values of the and electrons are of very distinct origins and magnitudes. Because the electrons are strongly interacting, they form localized magnetic moments, and their spin expectation values in this calculation are approximate . They are not perfectly polarized due to the entanglement between the and the electrons. On the other hand, the electrons are noninteracting, and their spin expectation values vary around . The electrons’ polarization is a direct cause of the hybridization with the magnetized electrons and, thus, a secondary effect. In this situation, the skyrmion of the electrons and the antiskyrmion of the electrons do not annihilate. This is revealed by the finite total skyrmion number calculated with unnormalized spin expectation values. This is indeed different from classical skyrmions, where the magnitude of the spin vectors is normalized. The reduced polarization decreases the topological protection of the magnetic skyrmion. The smaller the spin expectation value, the easier the spin can be flipped, and the magnetic skyrmion is destroyed Siegl et al. (2022b). On the other hand, this facilitates manipulating them as necessary for technical applications.

Next, we analyze the stability of the magnetic skyrmion for different magnetic field strengths, as shown in Fig. 2. As stated above, we generally apply a small magnetic field which helps to stabilize the magnetic skyrmion against spin spiral solutions Sotnikov et al. (2021); Haller et al. (2022). In Fig. 2(a), we show the skyrmion number using normalized and unnormalized spins, respectively. We observe that, while the skyrmion number (normalized) is constantly one for , the skyrmion number using unnormalized spin expectation values is and gradually drops for an increased magnetic field until is reached. The difference between these numbers demonstrates the relevance of quantum effects to the system at hand. We furthermore show the average spin expectation values of the and electrons, indicating that the electrons are considerably more stronger polarized than the electrons. Furthermore, for magnetic fields stronger than , we only find ferromagnetic solutions. Figures 2(b) and (c) give representative spin textures of the electrons for the corresponding parameter regimes, i.e., small and large magnetic fields.

We next analyze the structure of the skyrmion depending on the strength of the Rashba spin-orbit coupling . We show the skyrmion number using normalized spin expectation values, the skyrmion number using unnormalized spin expectation values, and the average spin expectation values ( and ) in Fig. 3(a). Increasing the Rashba interaction, the -electron spin expectation value is slightly suppressed, while the -electron spin expectation value slightly increases. This increase in the electron magnetization can be explained by the stronger coupling between the and electrons. While we need to create a finite DM interaction that stabilizes the magnetic skyrmion, we see that for increasing the skyrmion gets destabilized and for , magnetic skyrmions become unstable indicated by the vanishing skyrmion number. To analyze this transition further, we calculate the average size of the skyrmion. First, the center of the skyrmion created by the electrons is
(10) |
where , , and are the coordinates of the lattice sites spanning the elemental triangle as explained below Eq. (8). The extension of the skyrmion in the and the -direction is then given as
(11) | |||||
(12) |
where () is the () component of the position and the center of the skyrmion is . In Fig. 3(b), we show the extension of the skyrmion in the and -direction depending on the Rashba spin-orbit interaction. We see that while the average extension of the skyrmion in the direction remains unchanged when increasing , the magnetic skyrmion is strongly elongated in the direction. At , the magnetic skyrmion changes into a spiral phase, again consistent with findings for quantum skyrmions on nonelectronic spin lattices Sotnikov et al. (2021); Haller et al. (2022). Representative spin textures of the electrons are shown in Fig. 3(c)-(e), depicting a skyrmion for small (c), an elongated skyrmion close to the phase transition (d), and a spiral wave for large (e). For larger values of the spin-orbit coupling, we do not find stable skyrmion solutions.
Finally, we note that we have confirmed the stability of the quantum skyrmion phase for smaller lattice sizes, such as 7x7. Magnetic skyrmions remain stable as long as their elongation is smaller than the lattice width. Furthermore, we do not find an even/odd effect in the lattice width, which can be understood by the fact that all spins are ferromagnetically aligned far away from the magnetic skyrmion, irrespective of changes of the lattice sizes once it exceeds the size of the quantum skyrmion.
IV Charge-driven quantum skyrmions — the onset of the quantum skyrmion Hall effect
Finally, we study the response of the identified stable skyrmion textures to an applied charge current in the itinerant electrons. To do this, we calculate the change in the spin expectation values of all lattice sites in linear response theory. We focus on describing the time-transient behavior of the system. A description of the nonequilibrium steady state poses considerable numerical challenges, as discussed in the concluding remarks.
In linear response, the change in an expectation value of operator resulting from a perturbation is given by
(13) | |||||
(14) |
where is the Heaviside step function. Because we are interested in the linear response of the spin expectation values of the electrons to a charge current in the itinerant electrons, we use
(15) | |||||
(16) |
where corresponds exactly to the local spin of an electron, and is the charge current operator in the electrons.




For these operators, Eq. (14) corresponds to a nonlocal two-particle Green’s function. Using the DMFT approximation, where vertex corrections in nonlocal Green’s functions vanish, we write these two-particle Green’s functions as the convolution of two single-particle Green’s functions. Then, we calculate the time evolution of the spin expectation values of all spins. Because the self-energy depends on the lattice site, also the time evolution of the spin expectation value depends on the lattice site. This is shown in Fig. 4, where we show the change of the , , and component of the spin expectation values , , and along the -direction of the lattice across the center of the skyrmion solution for (shown in Fig. 1). Specifically, the spin expectation values are shown for lattice sites , where is the center of the skyrmion, see Eq. (10). and show a strong dependence on the position close to the center of the skyrmion. changes even its sign when changing the position from the left of the center to the right of the center. On the other hand, is nearly independent of the lattice site. Also, while becomes small for spins far away from the skyrmion center, and are nonzero. Thus, even in the ferromagnetic region away from the skyrmion center, and change. This rotation of the spin in the ferromagnetic state when a charge current is applied is explained by the Edelstein and the magnetoelectric effect Peters and Yanase (2018); in a system where the Fermi surface is split due to the Rashba spin-orbit coupling, a charge current results in an accumulation of spin. This can be seen here as a rotation of the spin expectation values in the and the direction, even far away from the magnetic skyrmion. Furthermore, we emphasize that the linear-response results only remain valid within sufficiently small times . In Fig. 4, we see that the initial linear trend in is reduced, and, as an expected artifact from linear response theory, all spin expectation values start oscillating after a certain individual time. The change of all spin expectation values for and are visualized in Fig. 5 as arrows. Each arrow corresponds to the direction of the local change in the spin expectation values . The actual length of each change is and for and ,respectively. We see that in the ferromagnetic region, all spins are rotated in the same direction. Only close to the magnetic skyrmion, the change in the spin expectation values significantly depends on the lattice site.
Finally, we take the time evolution of each spin on the lattice and calculate the skyrmion density and the time-dependent size and position of the skyrmion according to Eqs. (10-11). By Eq. (14), we find that the center of the skyrmion moves almost perpendicularly to the applied current, as shown in Fig. 6(a). While the current is applied in the direction, the skyrmion moves in the positive direction. Thus, our results demonstrate the onset of a quantum skyrmion Hall effect with a Magnus angle close to degrees. Notably, the size of the skyrmion effectively remains constant during the motion, shown in Fig. 6(b).
V Discussion
In conclusion, we show that noncentrosymmetric -electron systems with spin-orbit coupling in the presence of a small external magnetic field can host nano quantum skyrmions in the ground state, and we demonstrate the onset of the quantum skyrmion Hall effect upon applying a charge current, which is accompanied by an Edelstein and magnetoelectric effect.
The reason for the stability of the quantum skyrmion is an effective DM interaction generated by the spin-orbit interaction and a local density-density interaction. Despite the itinerant electrons being magnetized like an antiskyrmion, the quantum skyrmions of the electrons remain stable and dominate the physical behavior of the system because of its considerably stronger polarization due to strong correlations. Concerning the quantum skyrmion Hall effect, we observe a Magnus angle close to . This is consistent with the behavior of classical skyrmions, where the Magnus angle increases when the size of the skyrmions is smaller or when dissipative effects are small Litzius et al. (2016); Jiang et al. (2016). Both is the case for the observed nano quantum skyrmions. Furthermore, no quantum skyrmion pinning is visible in our study. We note that our method can only describe the onset of the skyrmion motion. In particular, in a full nonequilibrium calculation, time-dependent spin expectation values would lead to time-dependent self-energies. The system would adapt to the changed spin expectation values and backaction effects would alter our conclusions when the linear-response regime is left. For example, linear response theory can permanently decrease the polarization locally, ultimately resulting in a site with vanishing spin polarization. However, this situation is energetically unfavorable due to the strong density-density interaction. Thus, in a full nonequilibrium calculation, self-energies will change in a way that an atom with vanishing spin polarization is prevented, rendering the quantum magnetic skyrmion stable and letting it continue its motion perpendicular to an applied current. Yet, a full nonequilibrium calculation, as well as a steady-state analysis, goes beyond the scope of the current paper and is left for future work.
We note that other forms of spin-orbit interaction also lead to stable nano quantum skyrmions in the -electron system at hand. We show the results for a different form of the spin-orbit interaction, where the momenta couple to the same spin direction, in Appendix A. Also in these systems, the spin-orbit interaction results in a spin accumulation when a current is applied, which leads to a site-dependent change of the spin expectation values, and to a skyrmion Hall effect. These results emphasize that the existence of magnetic skyrmions in strongly correlated -electron systems with spin-orbit coupling and the skyrmion Hall effect is a general effect.
Acknowledgements.
All authors acknowledge funding by the Kyoto University - Hamburg University (KU-UHH) international partnership funding program for 2021 and 2022. R.P. is supported by JSPS KAKENHI No. JP18K03511 and JP23K03300. Parts of the numerical simulations in this work have been done using the facilities of the Supercomputer Center at the Institute for Solid State Physics, the University of Tokyo. J. N.-S. acknowledges support by the Cluster of Excellence “CUI: Advanced Imaging of Matter” of the Deutsche Forschungsgemeinschaft (DFG) – EXC 2056 – project ID 390715994 and the Universität Hamburg’s Next Generation Partnership funded under the Excellence Strategy of the Federal Government and the Länder. T. P. acknowledges funding by the DFG (project no. 420120155) and the European Union (ERC, QUANTWIST, project number 101039098). Views and opinions expressed are however those of the authors only and do not necessarily reflect those of the European Union or the European Research Council. Neither the European Union nor the granting authority can be held responsible for them.Appendix A Different form of spin-orbit interaction


To demonstrate that our results are robust for different types of spin-orbit coupling, we repeat our analysis using a spin-orbit interaction of the form
(17) | |||||
The rest of the Hamiltonian, including the two-particle interaction, is unchanged compared to the main text. In Fig. 7, we show two RDMFT solutions, including magnetic skyrmions, for , where we again use a small magnetic field, , to stabilize the magnetic skyrmion Sotnikov et al. (2021); Lohani et al. (2019); Siegl et al. (2022b); Haller et al. (2022). The change in the sign of the spin-orbit interaction leads to a change in the rotation direction of the spin texture.
Furthermore, we apply a charge current in the direction for both solutions and find that the center of the quantum magnetic skyrmion dominantly moves into the positive -direction. This is explained as follows: The change in the sign of the spin-orbit interaction leads not only to a reversal of the spin rotation inside the skyrmion but also changes the sign of the Edelstein and magnetoelectric effect. Thus, spins in these two examples are rotated in opposite directions when current is applied. As a result, both magnetic skyrmions move into the same, the positive direction.
Appendix B Convergence of the real-space DMFT


In this appendix, we demonstrate the convergence of the real-space DMFT in the magnetic skyrmion phase for . In Figs. 8 and 9, we show representative self-energies of the electrons for two different lattice sites and different DMFT iterations. Panel (a) shows the off-diagonal self-energy, , and panel (b) shows the diagonal self-energies, and . On average, we need - DMFT iterations (depending on the parameters) to obtain a converged magnetic skyrmion solution. In both figures, we see that while the self-energy of the 10th iteration qualitatively shows the same behavior as iterations 26 and 27, quantitatively, it still differs from the converged self-energy. On the other hand, the self-energies of the 26th and 27th iterations lie on top of each other.
References
- Bogdanov (1995) A Bogdanov, “New localized solutions of the nonlinear field-equations,” JETP Letters 62, 247–251 (1995).
- Bogdanov and Hubert (1999) A Bogdanov and A Hubert, “The stability of vortex-like structures in uniaxial ferromagnets,” Journal of Magnetism and Magnetic Materials 195, 182–192 (1999).
- Mühlbauer et al. (2009) S. Mühlbauer, B. Binz, F. Jonietz, C. Pfleiderer, A. Rosch, A. Neubauer, R. Georgii, and P. Böni, “Skyrmion lattice in a chiral magnet,” Science 323, 915–919 (2009).
- Yu et al. (2010) X. Z. Yu, Y. Onose, N. Kanazawa, J. H. Park, J. H. Han, Y. Matsui, N. Nagaosa, and Y. Tokura, “Real-space observation of a two-dimensional skyrmion crystal,” Nature 465, 901–904 (2010).
- Heinze et al. (2011) Stefan Heinze, Kirsten Von Bergmann, Matthias Menzel, Jens Brede, André Kubetzka, Roland Wiesendanger, Gustav Bihlmayer, and Stefan Blügel, “Spontaneous atomic-scale magnetic skyrmion lattice in two dimensions,” Nature Physics 7, 713–718 (2011).
- Skyrme (1962) T. H.R. Skyrme, “A unified field theory of mesons and baryons,” Nuclear Physics 31, 556–569 (1962).
- Everschor-Sitte et al. (2017) Karin Everschor-Sitte, Matthias Sitte, Thierry Valet, Artem Abanov, and Jairo Sinova, “Skyrmion production on demand by homogeneous dc currents,” New Journal of Physics 19, 92001 (2017).
- Stier et al. (2017) Martin Stier, Wolfgang Häusler, Thore Posske, Gregor Gurski, and Michael Thorwart, “Skyrmion-anti-skyrmion pair creation by in-plane currents,” Physical Review Letters 118, 267203 (2017).
- Schäffer et al. (2020) Alexander F. Schäffer, Pia Siegl, Martin Stier, Thore Posske, Jamal Berakdar, Michael Thorwart, Roland Wiesendanger, and Elena Y. Vedmedenko, “Rotating edge-field driven processing of chiral spin textures in racetrack devices,” Scientific Reports 10, 20400 (2020).
- Siegl et al. (2022a) Pia Siegl, Martin Stier, Alexander F Schäffer, Elena Y Vedmedenko, Thore Posske, Roland Wiesendanger, and Michael Thorwart, “Creating arbitrary sequences of mobile magnetic skyrmions and antiskyrmions,” Physical Review B 106, 14421 (2022a).
- Parkin et al. (2008) Stuart S P Parkin, Masamitsu Hayashi, and Luc Thomas, “Magnetic domain-wall racetrack memory,” Science 320, 190–194 (2008).
- Tomasello et al. (2014) R Tomasello, E Martinez, R Zivieri, L Torres, M Carpentieri, and G Finocchio, “A strategy for the design of skyrmion racetrack memories,” Sci. Rep. 4, 6784 (2014).
- Fert et al. (2017) Albert Fert, Nicolas Reyren, and Vincent Cros, “Magnetic skyrmions: advances in physics and potential applications,” Nature Reviews Materials 2, 17031 (2017).
- Psaroudaki and Panagopoulos (2021) Christina Psaroudaki and Christos Panagopoulos, “Skyrmion qubits: A new class of quantum logic elements based on nanoscale magnetization,” Physical Review Letters 127, 067201 (2021).
- Thiele (1973) A. A. Thiele, “Steady-state motion of magnetic domains,” Physical Review Letters 30, 230–233 (1973).
- Everschor et al. (2011) Karin Everschor, Markus Garst, R. A. Duine, and Achim Rosch, “Current-induced rotational torques in the skyrmion lattice phase of chiral magnets,” Phys. Rev. B 84, 064401 (2011).
- Everschor et al. (2012) Karin Everschor, Markus Garst, Benedikt Binz, Florian Jonietz, Sebastian Mühlbauer, Christian Pfleiderer, and Achim Rosch, “Rotating skyrmion lattices by spin torques and field or temperature gradients,” Phys. Rev. B 86, 054432 (2012).
- Iwasaki et al. (2013a) Junichi Iwasaki, Masahito Mochizuki, and Naoto Nagaosa, “Universal current-velocity relation of skyrmion motion in chiral magnets,” Nature Communications 4, 1463 (2013a).
- Iwasaki et al. (2013b) Junichi Iwasaki, Masahito Mochizuki, and Naoto Nagaosa, “Current-induced skyrmion dynamics in constricted geometries,” Nature Nanotechnology 8, 742–747 (2013b).
- Sampaio et al. (2013) J. Sampaio, V. Cros, S. Rohart, A. Thiaville, and A. Fert, “Nucleation, stability and current-induced motion of isolated magnetic skyrmions in nanostructures,” Nature Nanotechnology 8, 839–844 (2013).
- Moon et al. (2022) Kyoung-Woong Moon, Jungbum Yoon, Changsoo Kim, Jae-Hun Sim, Se Kwon Kim, Soong-Geun Je, and Chanyong Hwang, “An alternative understanding of the skyrmion hall effect based on one-dimensional domain wall motion,” Applied Physics Express 15, 123001 (2022).
- Jiang et al. (2016) Wanjun Jiang, Xichao Zhang, Guoqiang Yu, Wei Zhang, Xiao Wang, M. Benjamin Jungfleisch, John E. Pearson, Xuemei Cheng, Olle Heinonen, Kang L. Wang, Yan Zhou, Axel Hoffmann, and Suzanne G.E. Te Velthuis, “Direct observation of the skyrmion hall effect,” Nature Physics 13, 162–169 (2016).
- Litzius et al. (2016) Kai Litzius, Ivan Lemesh, Benjamin Krüger, Pedram Bassirian, Lucas Caretta, Kornel Richter, Felix Büttner, Koji Sato, Oleg A. Tretiakov, Johannes Förster, Robert M. Reeve, Markus Weigand, Iuliia Bykova, Hermann Stoll, Gisela Schütz, Geoffrey S.D. Beach, and Mathias Klaüi, “Skyrmion hall effect revealed by direct time-resolved x-ray microscopy,” Nature Physics 13, 170–175 (2016).
- Takashima et al. (2016) Rina Takashima, Hiroaki Ishizuka, and Leon Balents, “Quantum skyrmions in two-dimensional chiral magnets,” Physical Review B 94, 134415 (2016).
- Sotnikov et al. (2021) O. M. Sotnikov, V. V. Mazurenko, J. Colbois, F. Mila, M. I. Katsnelson, and E. A. Stepanov, “Probing the topology of the quantum analog of a classical skyrmion,” Physical Review B 103, L060404 (2021).
- Lohani et al. (2019) Vivek Lohani, Ciarán Hickey, Jan Masell, and Achim Rosch, “Quantum skyrmions in frustrated ferromagnets,” Physical Review X 91, 1–14 (2019).
- Siegl et al. (2022b) Pia Siegl, Elena Y. Vedmedenko, Martin Stier, Michael Thorwart, and Thore Posske, “Controlled creation of quantum skyrmions,” Physical Review Research 4, 023111 (2022b).
- Haller et al. (2022) Andreas Haller, Solofo Groenendijk, Alireza Habibi, Andreas Michels, and Thomas L. Schmidt, “Quantum skyrmion lattices in heisenberg ferromagnets,” Physical Review Research 4, 043113 (2022).
- Yoshi et al. (2023) Ashish Yoshi, Robert Peters, and Thore Posske, “Describing quantum skyrmions with neural network quantum states,” ArXiv (2023), in preparation.
- Kobayashi and Hayami (2022) Kaito Kobayashi and Satoru Hayami, “Skyrmion and vortex crystals in the hubbard model,” Physical Review B 106, L140406 (2022).
- Takagi et al. (2022) Rina Takagi, Naofumi Matsuyama, Victor Ukleev, Le Yu, Jonathan S. White, Sonia Francoual, José R. L. Mardegan, Satoru Hayami, Hiraku Saito, Koji Kaneko, Kazuki Ohishi, Yoshichika Ōnuki, Taka-hisa Arima, Yoshinori Tokura, Taro Nakajima, and Shinichiro Seki, “Square and rhombic lattices of magnetic skyrmions in a centrosymmetric binary compound,” Nature Communications 13, 1472 (2022).
- Back et al. (2020) C Back, V Cros, H Ebert, K Everschor-Sitte, A Fert, M Garst, Tianping Ma, S Mankovsky, T L Monchesky, M Mostovoy, N Nagaosa, S S P Parkin, C Pfleiderer, N Reyren, A Rosch, Y Taguchi, Y Tokura, K von Bergmann, and Jiadong Zang, “The 2020 skyrmionics roadmap,” Journal of Physics D: Applied Physics 53, 363001 (2020).
- Kim and Tserkovnyak (2016) Se Kwon Kim and Yaroslav Tserkovnyak, “Topological effects on quantum phase slips in superfluid spin transport,” Physical Review Letters 116, 127201 (2016).
- Posske and Thorwart (2019) Thore Posske and Michael Thorwart, “Winding up quantum spin helices: How avoided level crossings exile classical topological protection,” Physical Review Letters 122, 097204 (2019).
- Vijayan et al. (2023) Vipin Vijayan, L. Chotorlishvili, A. Ernst, S. S. P. Parkin, M. I. Katsnelson, and S. K. Mishra, “Topological dynamical quantum phase transition in a quantum skyrmion phase,” Physical Review B 107, L100419 (2023).
- Edelstein (1990) V.M. Edelstein, “Spin polarization of conduction electrons induced by electric current in two-dimensional asymmetric electron systems,” Solid State Communications 73, 233–235 (1990).
- Culcer and Winkler (2007) Dimitrie Culcer and R. Winkler, “Generation of spin currents and spin densities in systems with reduced symmetry,” Physical Review Letters 99, 226601 (2007).
- Chernyshov et al. (2009) Alexandr Chernyshov, Mason Overby, Xinyu Liu, Jacek K. Furdyna, Yuli Lyanda-Geller, and Leonid P. Rokhinson, “Evidence for reversible control of magnetization in a ferromagnetic material by means of spin–orbit magnetic field,” Nature Physics 5, 656–659 (2009).
- Manchon and Zhang (2008) A. Manchon and S. Zhang, “Theory of nonequilibrium intrinsic spin torque in a single nanomagnet,” Physical Review B 78, 212405 (2008).
- Garate and MacDonald (2009) Ion Garate and A. H. MacDonald, “Influence of a transport current on magnetic anisotropy in gyrotropic ferromagnets,” Physical Review B 80, 134403 (2009).
- Peters and Yanase (2018) Robert Peters and Youichi Yanase, “Strong enhancement of the edelstein effect in -electron systems,” Physical Review B 97, 115128 (2018).
- Fiebig (2005) Manfred Fiebig, “Revival of the magnetoelectric effect,” Journal of Physics D: Applied Physics 38, R123 (2005).
- Yanase and Sigrist (2008) Youichi Yanase and Manfred Sigrist, “Superconductivity and magnetism in non-centrosymmetric system: Application to cept3si,” Journal of the Physical Society of Japan 77, 124711 (2008).
- Michishita and Peters (2019) Yoshihiro Michishita and Robert Peters, “Impact of the rashba spin-orbit coupling on -electron materials,” Physical Review B 99, 155141 (2019).
- Georges et al. (1996) Antoine Georges, Gabriel Kotliar, Werner Krauth, and Marcelo J. Rozenberg, “Dynamical mean-field theory of strongly correlated fermion systems and the limit of infinite dimensions,” Rev. Mod. Phys. 68, 13–125 (1996).
- Potthoff and Nolting (1999) M. Potthoff and W. Nolting, “Surface metal-insulator transition in the hubbard model,” Physical Review B 59, 2549–2555 (1999).
- Vahedi et al. (2021) Javad Vahedi, Robert Peters, Ahmed Missaoui, Andreas Honecker, and Guy Trambly de Laissardière, “Magnetism of magic-angle twisted bilayer graphene,” SciPost Phys. 11, 083 (2021).
- Peters and Kawakami (2015) Robert Peters and Norio Kawakami, “Large and small fermi-surface spin density waves in the kondo lattice model,” Physical Review B 92, 075103 (2015).
- Peters and Kawakami (2014) Robert Peters and Norio Kawakami, “Spin density waves in the hubbard model: A dmft approach,” Physical Review B 89, 155134 (2014).
- Wilson (1975) Kenneth G. Wilson, “The renormalization group: Critical phenomena and the kondo problem,” Rev. Mod. Phys. 47, 773–840 (1975).
- Bulla et al. (2008) Ralf Bulla, Theo A. Costi, and Thomas Pruschke, “Numerical renormalization group method for quantum impurity systems,” Rev. Mod. Phys. 80, 395–450 (2008).
- Peters et al. (2006) Robert Peters, Thomas Pruschke, and Frithjof B. Anders, “Numerical renormalization group approach to green’s functions for quantum impurity models,” Physical Review B 74, 245114 (2006).
- Peters et al. (2012) Robert Peters, Norio Kawakami, and Thomas Pruschke, “Spin-selective kondo insulator: Cooperation of ferromagnetism and the kondo effect,” Physical Review Letters 108, 086402 (2012).
- Ruderman and Kittel (1954) M. A. Ruderman and C. Kittel, “Indirect exchange coupling of nuclear magnetic moments by conduction electrons,” Physical Review 96, 99 (1954).
- Kasuya (1956) Tadao Kasuya, “A theory of metallic ferro- and antiferromagnetism on zener’s model,” Progress of Theoretical Physics 16, 45–57 (1956).
- Yosida (1957) Kei Yosida, “Magnetic properties of cu-mn alloys,” Physical Review 106, 893 (1957).
- Neuhaus-Steinmetz et al. (2022) Jannis Neuhaus-Steinmetz, Elena Y. Vedmedenko, Thore Posske, and Roland Wiesendanger, “Complex magnetic ground states and topological electronic phases of atomic spin chains on superconductors,” Phys. Rev. B 105, 165415 (2022).