Next Article in Journal
Automated High-Speed Approaches for the Extraction of Permanent Magnets from Hard-Disk Drive Components for the Circular Economy
Previous Article in Journal
Unraveling the Magnetic Properties of NiO Nanoparticles: From Synthesis to Nanostructure
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Skyrmion Crystal Induced by Four-Spin Interactions in Itinerant Triangular Magnets

Graduate School of Science, Hokkaido University, Sapporo 060-0810, Japan
Magnetism 2024, 4(3), 281-294; https://doi.org/10.3390/magnetism4030018
Submission received: 16 July 2024 / Revised: 15 August 2024 / Accepted: 27 August 2024 / Published: 6 September 2024

Abstract

:
We investigate the emergence of magnetic skyrmion crystals with swirling topological spin textures in itinerant magnets with an emphasis on momentum-resolved multi-spin interactions. By performing the simulated annealing for the effective spin model with the two-spin and four-spin interactions on a two-dimensional triangular lattice, we show that various types of four-spin interactions become the microscopic origin of the magnetic skyrmion crystal with the skyrmion numbers of one and two. We find that the four-spin interactions between the different wave vectors lead to the skyrmion crystal with the skyrmion number of one, whereas those at the same wave vectors lead to the skyrmion crystals with the skyrmion number of one and two. Our results indicate that the multi-spin interactions arising from the itinerant nature of electrons provide rich topological spin textures in magnetic metals.

1. Introduction

Noncoplnar magnetic states have attracted much interest in condensed matter physics, as they give rise to fascinating physical phenomena induced by an emergent magnetic field through the spin Berry phase mechanism [1,2,3,4,5,6]. One of the examples is the topological Hall effect with neither an external magnetic field nor uniform magnetization [7,8,9,10,11,12,13,14,15,16,17,18,19,20,21]. Microscopically, the uniform alignment of the scalar spin chirality, which is defined by the triple scalar product as S i · ( S j × S k ) ( S i represents the spin at site i), becomes an indicator, since it has the same symmetry as the ferromagnetic spin state. Another example is the nonlinear nonreciprocal transport when noncoplanar magnetic textures break the spatial inversion symmetry in addition to the time-reversal symmetry [22,23]. In this case, the uniform component of the scalar spin chirality is not necessary [24,25]. Moreover, since these phenomena are purely driven by the noncoplanar magnetic orderings, the relativistic spin–orbit coupling plays a less important role, which indicates that candidate materials are not limited to those with heavy elements.
In contrast to rich physical phenomena, it is usually difficult to stabilize the noncoplanar magnetic states compared to the collinear and noncollinear ones. Multiple-Q states, which are expressed as a superposition of multiple spin density waves, naturally give rise to noncoplanar spin textures [26,27,28,29,30]. They include the magnetic skyrmion crystal (SkX) as a superposition of the double-Q (triple-Q) spiral waves on a two-dimensional square (triangular) lattice [31,32,33,34,35,36,37,38,39,40,41,42,43] and magnetic hedgehog crystal as a superposition of the triple-Q or quadruple-Q spiral waves on a three-dimensional cubic lattice [44,45,46,47,48,49,50,51,52,53,54]. Such multiple-Q states have often been found in itinerant magnets, where long-range exchange interactions with the sign change are present. Indeed, a plethora of multiple-Q states with noncoplanar spin textures have been theoretically identified in itinerant triangular-lattice systems [12,55,56,57,58,59,60,61,62,63,64,65] and itinerant tetragonal-lattice systems [66,67,68,69,70,71]. Meanwhile, extracting the essentially important interactions in itinerant magnets is difficult owing to the competition of the interactions including higher-order multi-spin interactions. The approach based on perturbative analyses [72,73,74] and machine learning analyses [75] has mainly been used. However, systematic investigations to clarify the necessary multi-spin interactions to cause the instability toward noncoplanar magnetic states have not been fully performed.
In the present study, we explore a new stabilization mechanism of the SkX by focusing on four-spin interactions that arise from the interplay between charge and spin degrees of freedom in itinerant magnets. We investigate the instability toward the SkX for five-type four-spin interactions. By performing the simulated annealing for the effective spin model incorporating the four-spin interactions on a two-dimensional triangular lattice, we find that four out of five four-spin interactions lead to the SkX as the ground state in an external magnetic field. Among them, the four-spin interactions consisting of the single wave-vector channels give rise to two types of SkXs with a skyrmion number of one and two, while those consisting of multiple wave-vector channels give rise to the SkX with a skyrmion number of one. The present result provides a possibility of stabilizing the SkX by multi-spin interactions in itinerant magnets, which might be relevant to why many of the SkX-hosting materials are metallic.

2. Model and Method

We start from the Kondo lattice model with classical spins S i ( | S i | = 1 ) on a two-dimensional triangular lattice, which is one of the typical models for itinerant magnets. The Kondo lattice model consists of itinerant electrons and localized spins, and it is given by
H KLM = t i , j , σ c i σ c j σ + J K i , σ , σ c i σ σ σ σ · S i c i σ ,
where c i σ ( c i σ ) is a creation (annihilation) operator of an itinerant electron at site i and spin σ . The first term represents the hopping of itinerant electrons on the triangular lattice and the second term represents the exchange coupling between itinerant electron spins and localized spins S i with the coupling constant J K ; σ = ( σ x , σ y , σ z ) is the vector of Pauli matrices. Owing to the centrosymmetric lattice structure, no spin-dependent hopping, which is referred to as the antisymmetric spin–orbit interaction, appears in the Hamiltonian.
When J K is much smaller than the bandwidth of itinerant electrons, the Kondo lattice model reduces to a spin model with effective spin interactions between localized spins after tracing out the itinerant electron degree of freedom [72,73,74]. Up to the four-spin interaction, the effective spin model is generally given by
H = q J q S q · S q + 1 N q 1 , q 2 , q 3 , q 4 K q 1 , q 2 , q 3 , q 4 δ q 1 + q 2 + q 3 + q 4 , l G × ( S q 1 · S q 2 ) ( S q 3 · S q 4 ) + ( S q 1 · S q 4 ) ( S q 2 · S q 3 ) ( S q 1 · S q 3 ) ( S q 2 · S q 4 ) ,
where N is the total number of sites, δ is the Kronecker delta, and G is the reciprocal lattice vector (l is an interger). S q is the Fourier transform of localized spins S i . The first term represents the bilinear interaction with the coupling constant J q proportional to J K 2 , whereas the second term represents the four-spin interaction with the coupling constant K q 1 , q 2 , q 3 , q 4 proportional to J K 4 . The former bilinear interaction is referred to as the Ruderman–Kittel–Kasuya–Yosida (RKKY) interaction [76,77,78]. Since the magnitude of K q 1 , q 2 , q 3 , q 4 is determined by the products of Green’s functions of the itinerant electrons as well as the factor J K 4 in a rigorous way, it can be large depending on the electronic state [72]. It is noted that no Dzyaloshinskii–Moriya interaction [79,80] appears owing to the absence of the antisymmetric spin–orbit interaction in the Kondo lattice model [81]; the isotropic form of the spin interaction is caused by the spin rotational symmetry shown in Equation (1).
In order to investigate the instability toward the SkX in the model in Equation (2), we consider the situation where the interaction at ± Q 1 = ± ( π / 3 , 0 ) corresponds to the dominant interaction by supposing that the nesting of the Fermi surfaces occurs at Q 1 . In other words, we consider the Fermi surfaces, which invokes the instability toward the magnetic orderings with the magnetic modulations by Q 1 . In such a situation, the interactions at ± Q 2 = ± ( π / 6 , 3 π / 6 ) and ± Q 3 = ± ( π / 6 , 3 π / 6 ) also give the dominant contributions owing to the threefold rotational symmetry of the triangular lattice Figure 1. By considering the contributions from the ± Q 1 , ± Q 2 , and ± Q 3 channels, the model in Equation (2) is simplified as follows [74]:
H eff = 2 J ν S Q ν · S Q ν + 2 K 1 N ν ( S Q ν · S Q ν ) 2 + 2 K 2 N ν ( S Q ν · S Q ν ) ( S Q ν · S Q ν ) + K 3 N ν ν ( S Q ν · S Q ν ) ( S Q ν · S Q ν ) + K 4 N ν ν ( S Q ν · S Q ν ) ( S Q ν · S Q ν ) + K 5 N ν ν ( S Q ν · S Q ν ) ( S Q ν · S Q ν ) ,
where ν = 1 –3 is the index for the ordering wave vectors Q ν and J J Q 1 = J Q 2 = J Q 3 . The first term represents the RKKY interaction, while the second to sixth terms represent the four-spin interactions with the coupling constants K 1 K 5 . In the following section, we investigate the effects of the four-spin interactions on the stabilization of the SkX. We set J = 1 as the energy unit of the model. In addition, we also consider the effect of the external magnetic field H in the form of the Zeeman coupling as
H Z = H i S i z .
In the following section, we analyze the total Hamiltonian as
H = H eff + H Z .
It is noted that the ground-state spin configuration is given as the single-Q spiral state irrespective of H when all the four-spin interactions are zero.
To obtain the stable spin configurations of the model in Equation (5), we perform the simulated annealing combined with Monte Carlo simulations based on the standard single-spin-flip Metropolis algorithm. Starting from a high temperature T 0 / J = 1 , we gradually reduce the temperature at the ratio of α = 0.999999 until the final temperature T / J = 0.01 is reached. At each temperature, we perform Monte Carlo sweeps, and we further perform 10 5 10 6 Monte Carlo sweeps at the final temperature for measurements. We also start the simulations from the spin patterns obtained at low temperatures in the vicinity of the phase boundaries.
For the obtained spin configurations, we calculate the spin structure factor given by
S s η η ( q ) = 1 N i , j S i η S j η e i q · ( r i r j ) ,
where η = x , y , z . r i is the position vector at site i, and q is the wave vector. The total spin structure factor is given by S s ( q ) = S s x x ( q ) + S s y y ( q ) + S s z z ( q ) , and its in-plane spin component is given by S s ( q ) = S s x x ( q ) + S s y y ( q ) . The magnetization along the magnetic field direction given by
M = 1 N i S i z ,
is also calculated. Meanwhile, the scalar spin chirality is used to identify whether the spin configurations are topologically nontrivial or not, whose expression is given by
χ sc = 1 N R S j · ( S k × S l ) ,
where R represents the position vector at the centers of triangles; the sites j, k, and l form the triangle at R in the counterclockwise order. The magnetic ordering with nonzero χ sc exhibits the topological Hall effect.

3. Results

In this section, we discuss the stability of the SkX in the presence of the five-type four-spin interactions in Equation (5). In order to understand the role of each four-spin interaction, we investigate its effect one by one below. We discuss the effect of K 1 in Section 3.1, that of K 2 in Section 3.2, that of K 3 in Section 3.3, that of K 4 in Section 3.4, and that of K 5 in Section 3.5. We especially focus on the emergence of the SkX in each case. In the following section, the result for H < 0 is the same as that for H > 0 when the sign of S i z is reversed [82].

3.1. Effect of K 1

First, we discuss the effect of K 1 , which has the form of ( S Q ν · S Q ν ) 2 , by setting K 2 = K 3 = K 4 = K 5 = 0 . This type of interaction corresponds to the biquadratic interaction. Since this interaction plays an important role when the Fermi surface is strongly nested at Q ν [72,73,74], its effect on the stability of the SkX has been investigated in the case of the triangular lattice [83] and square lattice [84]. Thus, we briefly discuss the role of K 1 in the following. Since the SkX does not appear for K 1 < 0 , we focus on the situation for K 1 > 0 .
Figure 2a shows the H dependence of the magnetization M and the scalar spin chirality χ sc . For 0.5 H 1.0 , the magnetic state with nonzero χ sc indicates the emergence of the SkX. We plot the real-space spin and scalar spin chirality configurations in this state in Figure 3a. The skyrmion cores located at S i z = 1 form the triangular lattice, as shown in the left panel of Figure 3a; the vorticity in terms of the x y -spin component around the skyrmion core is given by 1 , which means that the SkX phase exhibits the skyrmion number of + 1 [43]; we denote this as the SkX-I. The nonzero skyrmion number is also found in the distribution of the scalar spin chirality in the right panel of Figure 3a, where the region with the positive scalar spin chirality is much larger than that with the negative scalar spin chirality. It is noted that the SkX with the skyrmion number of + 1 is energetically degenerated with that with the skyrmion number of 1 owing to the spin rotational symmetry in the model. Such a degeneracy can be lifted when considering the symmetric anisotropic exchange interaction that arises from the relativistic spin–orbit coupling [85,86]. Another characteristic feature in the SkX-I is the triple-Q peaks in both x y and z components of the spin structure factor with the equal intensity at Q 1 , Q 2 , and Q 3 , i.e., S s ( Q 1 ) = S s ( Q 2 ) = S s ( Q 3 ) and S s z z ( Q 1 ) = S s z z ( Q 2 ) = S s z z ( Q 3 ) .
When the magnitude of K 1 becomes larger, another SkX phase appears in the low magnetic field region, as shown in the case of K 1 = 0.3 in Figure 3b. This SkX is also characterized by the isotropic triple-Q peaks in the spin structure factor with S s ( Q 1 ) = S s ( Q 2 ) = S s ( Q 3 ) . The scalar spin chirality in this low-field SkX phase is larger than that in the SkX-I. By looking into the real-space spin configuration in this state in the left panel of Figure 3b, one finds that the vorticity around the skyrmion core with positive S i z is given by 2 . Thus, this spin configuration has a skyrmion number of two, which is denoted as the SkX-II [62,87]. The distribution of the scalar spin chirality is shown in the right panel of Figure 3b, where the uniform component exists. Similar SkX with high skyrmion numbers has also been found when the anisotropic exchange interaction [88,89,90], higher-harmonic-wave vector interaction [91], and dynamical electric field [92] are considered, although the present SkX-II does not require such factors. Similar to the SkX-I, the SkX-II has a degeneracy in terms of the sign of the skyrmion number.

3.2. Effect of K 2

Next, we consider the effect of K 2 in the form of ( S Q ν · S Q ν ) ( S Q ν · S Q ν ) ; we set K 1 = K 3 = K 4 = K 5 = 0 . In contrast to K 1 , the effect of K 2 has not been investigated so far. In the case of K 2 , the SkX appears only for K 2 < 0 , as shown below.
For K 2 = 0.2 , there is no SkX phase, as shown in Figure 4a; no scalar spin chirality is induced in the whole magnetic field region. Only the topological trivial triple-Q state, whose spin configuration is characterized by a superposition of the dominant in-plane cycloidal spiral wave at Q 1 and subdominant spin density waves at Q 2 and Q 3 appears instead of the single-Q spiral state. With the increase in | K 2 | , both SkX-I and SkX-II appear against H, as shown in Figure 4b. The real-space spin and scalar spin chirality configurations of the SkX-I and SkX-II are shown in Figure 5a and Figure 5b, respectively. For the SkX-I, the spin configuration in Figure 5a is similar to that in Figure 3a, except for the vorticity. Since there is a degeneracy in terms of the vorticity in the presence of K 2 , these states are identical to each other. Meanwhile, the spin configuration of the SkX-II in Figure 5b looks different from that in Figure 3b. Nevertheless, the spin configuration in Figure 5b also exhibits the skyrmion number of two [93]. According to the integer skyrmion number, we refer to this state as the SkX-II. Indeed, the distribution of the scalar spin chirality in Figure 5b is similar to that in Figure 3, which indicates the similarity of the topological properties; a similar triple-Q peak structure is also found in the spin structure factor.
While further increasing | K 2 | , the SkX-II remains stable whereas the SkX-I vanishes, as shown by K 2 = 0.5 in Figure 4c. In contrast to the situation in the presence of K 1 , the data indicate the appearance of another magnetic state with nonzero scalar spin chirality for 0.25 H 1.15 , although this state does not exhibit nonzero skyrmion numbers. The real-space spin and scalar spin chirality configurations are shown in the left and right panels of Figure 5c, respectively. As shown in Figure 5c, the region with the positive scalar spin chirality is almost the same as that with the negative scalar spin chirality, although they are not perfectly canceled out with each other. This is why this state shows a small value of χ sc compared to the SkX-I and SkX-II, as shown in Figure 5c. We refer to this state as the chiral state in order to distinguish it from other topologically trivial states. The spin structure factor in this state shows the dominant double-Q peaks in the x y -spin component at Q 1 and Q 2 , as well as the subdominant double-Q peaks in the z-spin component at 2 Q 1 and 2 Q 2 , where the intensities at Q 1 and Q 2 are slightly different from each other. The SkX-II vanishes for further large K 2 = 0.6 , as shown in Figure 4d; the stability of the chiral state is extended to the zero-field region.

3.3. Effect of K 3

We consider the effect of K 3 in the form of ( S Q ν · S Q ν ) ( S Q ν · S Q ν ) for ν ν by setting K 1 = K 2 = K 4 = K 5 = 0 . When K 3 is negative (positive), this interaction corresponds to the attractive (repulsive) interaction between the different ordering wave vectors. Thus, the multiple-Q instability is expected only for K 3 < 0 ; we confirmed that no SkX appears for K 3 > 0 .
Figure 6 shows the H dependence of M and χ sc for K 3 = 0.1 . Similarly to the case of K 1 = 0.1 in Figure 2a, only the SkX-I appears in the intermediate-field region. The real-space spin and scalar spin chirality configurations shown in Figure 7 are similar to those in Figure 3a. Thus, the K 3 -type four-spin interaction also gives rise to the SkX-I. On the other hand, the SkX-II does not appear when the magnitude of K 3 increases; only the SkX-I appears, at least, up to K 3 = 0.8 .

3.4. Effect of K 4

In the case of K 4 , which is characterized by the different types of four-spin interactions in the form of ( S Q ν · S Q ν ) ( S Q ν · S Q ν ) for ν ν with K 1 = K 2 = K 3 = K 5 = 0 , similar SkX instability occurs. We show the H dependence of M and χ sc at K 4 = 0.2 in Figure 8; only the SkX-I appears against H. The spin and scalar spin chirality configurations are shown in the left and right panels of Figure 9, respectively. Also, in this case, the SkX-II is not stabilized, at least, for K 4 0.8 . Thus, K 4 plays a similar role in stabilizing the SkX-I to K 3 .

3.5. Effect of K 5

Finally, let us discuss the effect of K 5 , which has the form of ( S Q ν · S Q ν ) ( S Q ν · S Q ν ) , where K 1 = K 2 = K 3 = K 4 = 0 . In contrast to the situations with K 1 K 4 , any SkXs do not appear for both positive and negative K 5 ; see the H dependence of χ sc at K 5 = 0.8 in Figure 10. Instead of the SkX, the topologically trivial triple-Q state appears even for a large | K 5 | = 0.8 value, whose spin configuration is shown in Figure 11a; the spin structure factor in this state is characterized by a superposition of the dominant in-plane cycloidal spiral at Q 1 and the subdominant spin density waves at Q 2 and Q 3 , as found in the case of K 2 = 0.2 . Thus, this interaction just leads to the topologically trivial triple-Q structure for K 5 < 0 .
When | K 5 | is larger than J, where the perturbation analysis is no longer valid, a triple-Q magnetic vortex appears in the high-field region. We show the obtained spin configuration at K 5 = 2.5 and H = 1.3 in Figure 11b, which consists of the periodic alignment of vortex and antivortex, as found in frustrated magnets [94,95,96]. In this state, there are isotropic triple-Q peaks at Q 1 , Q 2 , and Q 3 in the x y component of the spin structure factor satisfying S s ( Q 1 ) = S s ( Q 2 ) = S s ( Q 3 ) . These results indicate that K 5 does not play an important role in inducing the SkX.

4. Conclusions

We have investigated the instability toward the SkX driven by four-spin interactions that originate from the interplay between spin and charge degrees of freedom in itinerant magnets. By analyzing the optimal spin configurations for the effective spin model with five-type four-spin interactions on the triangular lattice, we found that four out of five cases lead to the SkX with the skyrmion number of one in the external magnetic field. In addition, we also found that two-type four-spin interactions which correspond to the interactions at the same wave-vector channel give rise to the emergence of the higher-order SkX with the skyrmion number of two even without the external magnetic field. The present results indicate the possibility of further topological spin textures induced by multi-spin interactions in itinerant magnets. Since the SkXs have been found in materials with the interplay between itinerant electrons and localized spins, such as Gd2PdSi3 [18,97,98,99], the materials with similar chemical compositions might be promising for the realization of SkXs based on the four-spin interaction mechanism.

Funding

This research was supported by JSPS KAKENHI Grant Numbers JP21H01037, JP22H00101, JP22H01183, JP23H04869, JP23K03288, JP23K20827, and by JST CREST (JPMJCR23O4).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The original contributions presented in this study are included in this article; further inquiries can be directed to the corresponding author.

Conflicts of Interest

The author declares no conflicts of interest.

References

  1. Loss, D.; Goldbart, P.M. Persistent currents from Berry’s phase in mesoscopic systems. Phys. Rev. B 1992, 45, 13544–13561. [Google Scholar] [CrossRef] [PubMed]
  2. Ye, J.; Kim, Y.B.; Millis, A.J.; Shraiman, B.I.; Majumdar, P.; Tešanović, Z. Berry Phase Theory of the Anomalous Hall Effect: Application to Colossal Magnetoresistance Manganites. Phys. Rev. Lett. 1999, 83, 3737–3740. [Google Scholar] [CrossRef]
  3. Haldane, F.D.M. Berry Curvature on the Fermi Surface: Anomalous Hall Effect as a Topological Fermi-Liquid Property. Phys. Rev. Lett. 2004, 93, 206602. [Google Scholar] [CrossRef] [PubMed]
  4. Nagaosa, N.; Sinova, J.; Onoda, S.; MacDonald, A.H.; Ong, N.P. Anomalous Hall effect. Rev. Mod. Phys. 2010, 82, 1539–1592. [Google Scholar] [CrossRef]
  5. Xiao, D.; Chang, M.C.; Niu, Q. Berry phase effects on electronic properties. Rev. Mod. Phys. 2010, 82, 1959–2007. [Google Scholar] [CrossRef]
  6. Zhang, S.S.; Ishizuka, H.; Zhang, H.; Halász, G.B.; Batista, C.D. Real-space Berry curvature of itinerant electron systems with spin-orbit interaction. Phys. Rev. B 2020, 101, 024420. [Google Scholar] [CrossRef]
  7. Ohgushi, K.; Murakami, S.; Nagaosa, N. Spin anisotropy and quantum Hall effect in the kagomé lattice: Chiral spin state based on a ferromagnet. Phys. Rev. B 2000, 62, R6065–R6068. [Google Scholar] [CrossRef]
  8. Taguchi, Y.; Oohara, Y.; Yoshizawa, H.; Nagaosa, N.; Tokura, Y. Spin chirality, Berry phase, and anomalous Hall effect in a frustrated ferromagnet. Science 2001, 291, 2573–2576. [Google Scholar] [CrossRef]
  9. Tatara, G.; Kawamura, H. Chirality-driven anomalous Hall effect in weak coupling regime. J. Phys. Soc. Jpn. 2002, 71, 2613–2616. [Google Scholar] [CrossRef]
  10. Machida, Y.; Nakatsuji, S.; Maeno, Y.; Tayama, T.; Sakakibara, T.; Onoda, S. Unconventional anomalous Hall effect enhanced by a noncoplanar spin texture in the frustrated Kondo lattice Pr2Ir2O7. Phys. Rev. Lett. 2007, 98, 057203. [Google Scholar] [CrossRef]
  11. Shindou, R.; Nagaosa, N. Orbital Ferromagnetism and Anomalous Hall Effect in Antiferromagnets on the Distorted fcc Lattice. Phys. Rev. Lett. 2001, 87, 116801. [Google Scholar] [CrossRef] [PubMed]
  12. Martin, I.; Batista, C.D. Itinerant Electron-Driven Chiral Magnetic Ordering and Spontaneous Quantum Hall Effect in Triangular Lattice Models. Phys. Rev. Lett. 2008, 101, 156402. [Google Scholar] [CrossRef] [PubMed]
  13. Neubauer, A.; Pfleiderer, C.; Binz, B.; Rosch, A.; Ritz, R.; Niklowitz, P.G.; Böni, P. Topological Hall Effect in the A Phase of MnSi. Phys. Rev. Lett. 2009, 102, 186602. [Google Scholar] [CrossRef] [PubMed]
  14. Takatsu, H.; Yonezawa, S.; Fujimoto, S.; Maeno, Y. Unconventional anomalous Hall effect in the metallic triangular-lattice magnet PdCrO2. Phys. Rev. Lett. 2010, 105, 137201. [Google Scholar] [CrossRef] [PubMed]
  15. Ueland, B.; Miclea, C.; Kato, Y.; Ayala-Valenzuela, O.; McDonald, R.; Okazaki, R.; Tobash, P.; Torrez, M.; Ronning, F.; Movshovich, R.; et al. Controllable chirality-induced geometrical Hall effect in a frustrated highly correlated metal. Nat. Commun. 2012, 3, 1067. [Google Scholar] [CrossRef]
  16. Hamamoto, K.; Ezawa, M.; Nagaosa, N. Quantized topological Hall effect in skyrmion crystal. Phys. Rev. B 2015, 92, 115417. [Google Scholar] [CrossRef]
  17. Nakazawa, K.; Bibes, M.; Kohno, H. Topological Hall effect from strong to weak coupling. J. Phys. Soc. Jpn. 2018, 87, 033705. [Google Scholar] [CrossRef]
  18. Kurumaji, T.; Nakajima, T.; Hirschberger, M.; Kikkawa, A.; Yamasaki, Y.; Sagayama, H.; Nakao, H.; Taguchi, Y.; Arima, T.h.; Tokura, Y. Skyrmion lattice with a giant topological Hall effect in a frustrated triangular-lattice magnet. Science 2019, 365, 914–918. [Google Scholar] [CrossRef]
  19. Matsui, A.; Nomoto, T.; Arita, R. Skyrmion-size dependence of the topological Hall effect: A real-space calculation. Phys. Rev. B 2021, 104, 174432. [Google Scholar] [CrossRef]
  20. Pahari, R.; Balasubramanian, B.; Ullah, A.; Manchanda, P.; Komuro, H.; Streubel, R.; Klewe, C.; Valloppilly, S.R.; Shafer, P.; Dev, P.; et al. Peripheral chiral spin textures and topological Hall effect in CoSi nanomagnets. Phys. Rev. Mater. 2021, 5, 124418. [Google Scholar] [CrossRef]
  21. Ullah, A.; Balasubramanian, B.; Tiwari, B.; Giri, B.; Sellmyer, D.J.; Skomski, R.; Xu, X. Topological spin textures and topological Hall effect in centrosymmetric magnetic nanoparticles. Phys. Rev. B 2023, 108, 184432. [Google Scholar] [CrossRef]
  22. Hayami, S.; Yanagi, Y.; Kusunose, H. Spontaneous antisymmetric spin splitting in noncollinear antiferromagnets without spin-orbit coupling. Phys. Rev. B 2020, 101, 220403(R). [Google Scholar] [CrossRef]
  23. Eto, R.; Pohle, R.; Mochizuki, M. Low-Energy Excitations of Skyrmion Crystals in a Centrosymmetric Kondo-Lattice Magnet: Decoupled Spin-Charge Excitations and Nonreciprocity. Phys. Rev. Lett. 2022, 129, 017201. [Google Scholar] [CrossRef] [PubMed]
  24. Cheon, S.; Lee, H.W.; Cheong, S.W. Nonreciprocal spin waves in a chiral antiferromagnet without the Dzyaloshinskii-Moriya interaction. Phys. Rev. B 2018, 98, 184405. [Google Scholar] [CrossRef]
  25. Hayami, S.; Yatsushiro, M. Nonlinear nonreciprocal transport in antiferromagnets free from spin-orbit coupling. Phys. Rev. B 2022, 106, 014420. [Google Scholar] [CrossRef]
  26. Bak, P.; Lebech, B. Triple-q Modulated Magnetic Structure and Critical Behavior of Neodymium. Phys. Rev. Lett. 1978, 40, 800–803. [Google Scholar] [CrossRef]
  27. Shapiro, S.M.; Gurewitz, E.; Parks, R.D.; Kupferberg, L.C. Multiple-q Magnetic Structure in CeAl2. Phys. Rev. Lett. 1979, 43, 1748–1751. [Google Scholar] [CrossRef]
  28. Bak, P.; Jensen, M.H. Theory of helical magnetic structures and phase transitions in MnSi and FeGe. J. Phys. C Solid State Phys. 1980, 13, L881. [Google Scholar] [CrossRef]
  29. Forgan, E.M.; Gibbons, E.P.; McEwen, K.A.; Fort, D. Observation of a Quadruple-q Magnetic Structure in Neodymium. Phys. Rev. Lett. 1989, 62, 470–473. [Google Scholar] [CrossRef]
  30. Hayami, S.; Motome, Y. Topological spin crystals by itinerant frustration. J. Phys. Condens. Matter 2021, 33, 443001. [Google Scholar] [CrossRef]
  31. Bogdanov, A.N.; Yablonskii, D.A. Thermodynamically stable “vortices” in magnetically ordered crystals: The mixed state of magnets. Sov. Phys. JETP 1989, 68, 101. [Google Scholar]
  32. Bogdanov, A.; Hubert, A. Thermodynamically stable magnetic vortex states in magnetic crystals. J. Magn. Magn. Mater. 1994, 138, 255–269. [Google Scholar] [CrossRef]
  33. Rößler, U.K.; Bogdanov, A.N.; Pfleiderer, C. Spontaneous skyrmion ground states in magnetic metals. Nature 2006, 442, 797–801. [Google Scholar] [CrossRef] [PubMed]
  34. Mühlbauer, S.; Binz, B.; Jonietz, F.; Pfleiderer, C.; Rosch, A.; Neubauer, A.; Georgii, R.; Böni, P. Skyrmion lattice in a chiral magnet. Science 2009, 323, 915–919. [Google Scholar] [CrossRef] [PubMed]
  35. Yi, S.D.; Onoda, S.; Nagaosa, N.; Han, J.H. Skyrmions and anomalous Hall effect in a Dzyaloshinskii-Moriya spiral magnet. Phys. Rev. B 2009, 80, 054416. [Google Scholar] [CrossRef]
  36. Yu, X.Z.; Onose, Y.; Kanazawa, N.; Park, J.H.; Han, J.H.; Matsui, Y.; Nagaosa, N.; Tokura, Y. Real-space observation of a two-dimensional skyrmion crystal. Nature 2010, 465, 901–904. [Google Scholar] [CrossRef]
  37. Butenko, A.B.; Leonov, A.A.; Rößler, U.K.; Bogdanov, A.N. Stabilization of skyrmion textures by uniaxial distortions in noncentrosymmetric cubic helimagnets. Phys. Rev. B 2010, 82, 052403. [Google Scholar] [CrossRef]
  38. Münzer, W.; Neubauer, A.; Adams, T.; Mühlbauer, S.; Franz, C.; Jonietz, F.; Georgii, R.; Böni, P.; Pedersen, B.; Schmidt, M.; et al. Skyrmion lattice in the doped semiconductor Fe1−xCoxSi. Phys. Rev. B 2010, 81, 041203. [Google Scholar] [CrossRef]
  39. Adams, T.; Mühlbauer, S.; Neubauer, A.; Münzer, W.; Jonietz, F.; Georgii, R.; Pedersen, B.; Böni, P.; Rosch, A.; Pfleiderer, C. Skyrmion lattice domains in Fe1−xCoxSi. J. Phys. Conf. Ser. 2010, 200, 032001. [Google Scholar] [CrossRef]
  40. Yu, X.Z.; Kanazawa, N.; Onose, Y.; Kimoto, K.; Zhang, W.; Ishiwata, S.; Matsui, Y.; Tokura, Y. Near room-temperature formation of a skyrmion crystal in thin-films of the helimagnet FeGe. Nat. Mater. 2011, 10, 106–109. [Google Scholar] [CrossRef]
  41. Heinze, S.; von Bergmann, K.; Menzel, M.; Brede, J.; Kubetzka, A.; Wiesendanger, R.; Bihlmayer, G.; Blügel, S. Spontaneous atomic-scale magnetic skyrmion lattice in two dimensions. Nat. Phys. 2011, 7, 713–718. [Google Scholar] [CrossRef]
  42. Adams, T.; Mühlbauer, S.; Pfleiderer, C.; Jonietz, F.; Bauer, A.; Neubauer, A.; Georgii, R.; Böni, P.; Keiderling, U.; Everschor, K.; et al. Long-Range Crystalline Nature of the Skyrmion Lattice in MnSi. Phys. Rev. Lett. 2011, 107, 217206. [Google Scholar] [CrossRef] [PubMed]
  43. Nagaosa, N.; Tokura, Y. Topological properties and dynamics of magnetic skyrmions. Nat. Nanotechnol. 2013, 8, 899–911. [Google Scholar] [CrossRef] [PubMed]
  44. Binz, B.; Vishwanath, A. Theory of helical spin crystals: Phases, textures, and properties. Phys. Rev. B 2006, 74, 214408. [Google Scholar] [CrossRef]
  45. Park, J.H.; Han, J.H. Zero-temperature phases for chiral magnets in three dimensions. Phys. Rev. B 2011, 83, 184406. [Google Scholar] [CrossRef]
  46. Tanigaki, T.; Shibata, K.; Kanazawa, N.; Yu, X.; Onose, Y.; Park, H.S.; Shindo, D.; Tokura, Y. Real-space observation of short-period cubic lattice of skyrmions in MnGe. Nano Lett. 2015, 15, 5438–5442. [Google Scholar] [CrossRef]
  47. Yang, S.G.; Liu, Y.H.; Han, J.H. Formation of a topological monopole lattice and its dynamics in three-dimensional chiral magnets. Phys. Rev. B 2016, 94, 054420. [Google Scholar] [CrossRef]
  48. Kanazawa, N.; Seki, S.; Tokura, Y. Noncentrosymmetric magnets hosting magnetic skyrmions. Adv. Mater. 2017, 29, 1603227. [Google Scholar] [CrossRef]
  49. Fujishiro, Y.; Kanazawa, N.; Nakajima, T.; Yu, X.Z.; Ohishi, K.; Kawamura, Y.; Kakurai, K.; Arima, T.; Mitamura, H.; Miyake, A.; et al. Topological transitions among skyrmion-and hedgehog-lattice states in cubic chiral magnets. Nat. Commun. 2019, 10, 1059. [Google Scholar] [CrossRef]
  50. Okumura, S.; Hayami, S.; Kato, Y.; Motome, Y. Magnetic hedgehog lattices in noncentrosymmetric metals. Phys. Rev. B 2020, 101, 144416. [Google Scholar] [CrossRef]
  51. Kanazawa, N.; Kitaori, A.; White, J.S.; Ukleev, V.; Rønnow, H.M.; Tsukazaki, A.; Ichikawa, M.; Kawasaki, M.; Tokura, Y. Direct Observation of the Statics and Dynamics of Emergent Magnetic Monopoles in a Chiral Magnet. Phys. Rev. Lett. 2020, 125, 137202. [Google Scholar] [CrossRef] [PubMed]
  52. Aoyama, K.; Kawamura, H. Hedgehog-lattice spin texture in classical Heisenberg antiferromagnets on the breathing pyrochlore lattice. Phys. Rev. B 2021, 103, 014406. [Google Scholar] [CrossRef]
  53. Aoyama, K.; Kawamura, H. Hedgehog lattice and field-induced chirality in breathing-pyrochlore Heisenberg antiferromagnets. Phys. Rev. B 2022, 106, 064412. [Google Scholar] [CrossRef]
  54. Eto, R.; Mochizuki, M. Theory of Collective Excitations in the Quadruple-Q Magnetic Hedgehog Lattices. Phys. Rev. Lett. 2024, 132, 226705. [Google Scholar] [CrossRef] [PubMed]
  55. Akagi, Y.; Motome, Y. Spin Chirality Ordering and Anomalous Hall Effect in the Ferromagnetic Kondo Lattice Model on a Triangular Lattice. J. Phys. Soc. Jpn. 2010, 79, 083711. [Google Scholar] [CrossRef]
  56. Barros, K.; Kato, Y. Efficient Langevin simulation of coupled classical fields and fermions. Phys. Rev. B 2013, 88, 235101. [Google Scholar] [CrossRef]
  57. Venderbos, J.W.F.; Kourtis, S.; van den Brink, J.; Daghofer, M. Fractional Quantum-Hall Liquid Spontaneously Generated by Strongly Correlated t2g Electrons. Phys. Rev. Lett. 2012, 108, 126405. [Google Scholar] [CrossRef]
  58. Jiang, K.; Zhang, Y.; Zhou, S.; Wang, Z. Chiral Spin Density Wave Order on the Frustrated Honeycomb and Bilayer Triangle Lattice Hubbard Model at Half-Filling. Phys. Rev. Lett. 2015, 114, 216402. [Google Scholar] [CrossRef]
  59. Venderbos, J.W.F. Multi-Q hexagonal spin density waves and dynamically generated spin-orbit coupling: Time-reversal invariant analog of the chiral spin density wave. Phys. Rev. B 2016, 93, 115108. [Google Scholar] [CrossRef]
  60. Barros, K.; Venderbos, J.W.F.; Chern, G.W.; Batista, C.D. Exotic magnetic orderings in the kagome Kondo-lattice model. Phys. Rev. B 2014, 90, 245119. [Google Scholar] [CrossRef]
  61. Ghosh, S.; O’Brien, P.; Henley, C.L.; Lawler, M.J. Phase diagram of the Kondo lattice model on the kagome lattice. Phys. Rev. B 2016, 93, 024401. [Google Scholar] [CrossRef]
  62. Ozawa, R.; Hayami, S.; Motome, Y. Zero-Field Skyrmions with a High Topological Number in Itinerant Magnets. Phys. Rev. Lett. 2017, 118, 147205. [Google Scholar] [CrossRef] [PubMed]
  63. Takagi, R.; White, J.; Hayami, S.; Arita, R.; Honecker, D.; Rønnow, H.; Tokura, Y.; Seki, S. Multiple-q noncollinear magnetism in an itinerant hexagonal magnet. Sci. Adv. 2018, 4, eaau3402. [Google Scholar] [CrossRef]
  64. Wang, Z.; Su, Y.; Lin, S.Z.; Batista, C.D. Skyrmion Crystal from RKKY Interaction Mediated by 2D Electron Gas. Phys. Rev. Lett. 2020, 124, 207201. [Google Scholar] [CrossRef] [PubMed]
  65. Saito, H.; Kon, F.; Hidaka, H.; Amitsuka, H.; Kwanghee, C.; Hagihala, M.; Kamiyama, T.; Itoh, S.; Nakajima, T. In-plane anisotropy of single-q and multiple-q ordered phases in the antiferromagnetic metal CeRh2Si2. Phys. Rev. B 2023, 108, 094440. [Google Scholar] [CrossRef]
  66. Venderbos, J.W.F.; Daghofer, M.; van den Brink, J.; Kumar, S. Switchable Quantum Anomalous Hall State in a Strongly Frustrated Lattice Magnet. Phys. Rev. Lett. 2012, 109, 166405. [Google Scholar] [CrossRef] [PubMed]
  67. Solenov, D.; Mozyrsky, D.; Martin, I. Chirality Waves in Two-Dimensional Magnets. Phys. Rev. Lett. 2012, 108, 096403. [Google Scholar] [CrossRef]
  68. Shahzad, M.; Sengupta, P. Noncollinear magnetic ordering in a frustrated magnet: Metallic regime and the role of frustration. Phys. Rev. B 2017, 96, 224402. [Google Scholar] [CrossRef]
  69. Shahzad, M.; Sengupta, P. Phase diagram of the Shastry-Sutherland Kondo lattice model with classical localized spins: A variational calculation study. J. Phys. Condens. Matter 2017, 29, 305802. [Google Scholar] [CrossRef]
  70. Marcus, G.G.; Kim, D.J.; Tutmaher, J.A.; Rodriguez-Rivera, J.A.; Birk, J.O.; Niedermeyer, C.; Lee, H.; Fisk, Z.; Broholm, C.L. Multi-q Mesoscale Magnetism in CeAuSb2. Phys. Rev. Lett. 2018, 120, 097201. [Google Scholar] [CrossRef]
  71. Seo, S.; Wang, X.; Thomas, S.M.; Rahn, M.C.; Carmo, D.; Ronning, F.; Bauer, E.D.; dos Reis, R.D.; Janoschek, M.; Thompson, J.D.; et al. Nematic State in CeAuSb2. Phys. Rev. X 2020, 10, 011035. [Google Scholar] [CrossRef]
  72. Akagi, Y.; Udagawa, M.; Motome, Y. Hidden Multiple-Spin Interactions as an Origin of Spin Scalar Chiral Order in Frustrated Kondo Lattice Models. Phys. Rev. Lett. 2012, 108, 096401. [Google Scholar] [CrossRef] [PubMed]
  73. Hayami, S.; Motome, Y. Multiple-Q instability by (d-2)-dimensional connections of Fermi surfaces. Phys. Rev. B 2014, 90, 060402(R). [Google Scholar] [CrossRef]
  74. Hayami, S.; Ozawa, R.; Motome, Y. Effective bilinear-biquadratic model for noncoplanar ordering in itinerant magnets. Phys. Rev. B 2017, 95, 224424. [Google Scholar] [CrossRef]
  75. Sharma, V.; Wang, Z.; Batista, C.D. Machine learning assisted derivation of minimal low-energy models for metallic magnets. npj Comput. Mater. 2023, 9, 192. [Google Scholar] [CrossRef]
  76. Ruderman, M.A.; Kittel, C. Indirect Exchange Coupling of Nuclear Magnetic Moments by Conduction Electrons. Phys. Rev. 1954, 96, 99–102. [Google Scholar] [CrossRef]
  77. Kasuya, T. A Theory of Metallic Ferro- and Antiferromagnetism on Zener’s Model. Prog. Theor. Phys. 1956, 16, 45–57. [Google Scholar] [CrossRef]
  78. Yosida, K. Magnetic Properties of Cu-Mn Alloys. Phys. Rev. 1957, 106, 893–898. [Google Scholar] [CrossRef]
  79. Dzyaloshinsky, I. A thermodynamic theory of “weak” ferromagnetism of antiferromagnetics. J. Phys. Chem. Solids 1958, 4, 241–255. [Google Scholar] [CrossRef]
  80. Moriya, T. Anisotropic superexchange interaction and weak ferromagnetism. Phys. Rev. 1960, 120, 91. [Google Scholar] [CrossRef]
  81. Hayami, S.; Motome, Y. Néel- and Bloch-Type Magnetic Vortices in Rashba Metals. Phys. Rev. Lett. 2018, 121, 137202. [Google Scholar] [CrossRef] [PubMed]
  82. Ullah, A.; Li, X.; Jin, Y.; Pahari, R.; Yue, L.; Xu, X.; Balasubramanian, B.; Sellmyer, D.J.; Skomski, R. Topological phase transitions and Berry-phase hysteresis in exchange-coupled nanomagnets. Phys. Rev. B 2022, 106, 134430. [Google Scholar] [CrossRef]
  83. Hayami, S. Multiple-Q magnetism by anisotropic bilinear-biquadratic interactions in momentum space. J. Magn. Magn. Mater. 2020, 513, 167181. [Google Scholar] [CrossRef]
  84. Hayami, S.; Kato, Y. Widely-sweeping magnetic field–temperature phase diagrams for skyrmion-hosting centrosymmetric tetragonal magnets. J. Magn. Magn. Mater. 2023, 571, 170547. [Google Scholar] [CrossRef]
  85. Hayami, S.; Yambe, R. Degeneracy Lifting of Néel, Bloch, and Anti-Skyrmion Crystals in Centrosymmetric Tetragonal Systems. J. Phys. Soc. Jpn. 2020, 89, 103702. [Google Scholar] [CrossRef]
  86. Hayami, S.; Motome, Y. Noncoplanar multiple-Q spin textures by itinerant frustration: Effects of single-ion anisotropy and bond-dependent anisotropy. Phys. Rev. B 2021, 103, 054422. [Google Scholar] [CrossRef]
  87. Hayami, S.; Okubo, T.; Motome, Y. Phase shift in skyrmion crystals. Nat. Commun. 2021, 12, 6927. [Google Scholar] [CrossRef] [PubMed]
  88. Amoroso, D.; Barone, P.; Picozzi, S. Spontaneous skyrmionic lattice from anisotropic symmetric exchange in a Ni-halide monolayer. Nat. Commun. 2020, 11, 5784. [Google Scholar] [CrossRef]
  89. Yambe, R.; Hayami, S. Skyrmion crystals in centrosymmetric itinerant magnets without horizontal mirror plane. Sci. Rep. 2021, 11, 11184. [Google Scholar] [CrossRef]
  90. Amoroso, D.; Barone, P.; Picozzi, S. Interplay between Single-Ion and Two-Ion Anisotropies in Frustrated 2D Semiconductors and Tuning of Magnetic Structures Topology. Nanomaterials 2021, 11, 1873. [Google Scholar] [CrossRef]
  91. Hayami, S. Multiple skyrmion crystal phases by itinerant frustration in centrosymmetric tetragonal magnets. J. Phys. Soc. Jpn. 2022, 91, 023705. [Google Scholar] [CrossRef]
  92. Yambe, R.; Hayami, S. Dynamical generation of skyrmion and bimeron crystals by a circularly polarized electric field in frustrated magnets. Phys. Rev. B 2024, 110, 014428. [Google Scholar] [CrossRef]
  93. Hayami, S.; Motome, Y. Effect of magnetic anisotropy on skyrmions with a high topological number in itinerant magnets. Phys. Rev. B 2019, 99, 094420. [Google Scholar] [CrossRef]
  94. Kamiya, Y.; Batista, C.D. Magnetic Vortex Crystals in Frustrated Mott Insulator. Phys. Rev. X 2014, 4, 011023. [Google Scholar] [CrossRef]
  95. Wang, Z.; Kamiya, Y.; Nevidomskyy, A.H.; Batista, C.D. Three-Dimensional Crystallization of Vortex Strings in Frustrated Quantum Magnets. Phys. Rev. Lett. 2015, 115, 107201. [Google Scholar] [CrossRef]
  96. Hayami, S.; Lin, S.Z.; Kamiya, Y.; Batista, C.D. Vortices, skyrmions, and chirality waves in frustrated Mott insulators with a quenched periodic array of impurities. Phys. Rev. B 2016, 94, 174420. [Google Scholar] [CrossRef]
  97. Saha, S.R.; Sugawara, H.; Matsuda, T.D.; Sato, H.; Mallik, R.; Sampathkumaran, E.V. Magnetic anisotropy, first-order-like metamagnetic transitions, and large negative magnetoresistance in single-crystal Gd2PdSi3. Phys. Rev. B 1999, 60, 12162–12165. [Google Scholar] [CrossRef]
  98. Kumar, R.; Iyer, K.K.; Paulose, P.L.; Sampathkumaran, E.V. Magnetic and transport anomalies in R2RhSi3 (R = Gd, Tb, and Dy) resembling those of the exotic magnetic material Gd2PdSi3. Phys. Rev. B 2020, 101, 144440. [Google Scholar] [CrossRef]
  99. Spachmann, S.; Elghandour, A.; Frontzek, M.; Löser, W.; Klingeler, R. Magnetoelastic coupling and phases in the skyrmion lattice magnet Gd2PdSi3 discovered by high-resolution dilatometry. Phys. Rev. B 2021, 103, 184424. [Google Scholar] [CrossRef]
Figure 1. Schematic picture of the ordering wave vectors Q 1 , Q 2 , and Q 3 in the first Brillouin zone.
Figure 1. Schematic picture of the ordering wave vectors Q 1 , Q 2 , and Q 3 in the first Brillouin zone.
Magnetism 04 00018 g001
Figure 2. H dependence of the magnetization M (black squares) and scalar spin chirality χ sc (red circles) at (a) K 1 = 0.1 and (b) K 1 = 0.3 . The backgrounds in blue and green represent the regions where the SkX with the skyrmion number of one (SkX-I) and the SkX with the skyrmion number of two (SkX-II) are stabilized, respectively.
Figure 2. H dependence of the magnetization M (black squares) and scalar spin chirality χ sc (red circles) at (a) K 1 = 0.1 and (b) K 1 = 0.3 . The backgrounds in blue and green represent the regions where the SkX with the skyrmion number of one (SkX-I) and the SkX with the skyrmion number of two (SkX-II) are stabilized, respectively.
Magnetism 04 00018 g002
Figure 3. (Left panel) Real-space spin configurations of (a) the SkX-I at K 1 = 0.1 and H = 0.8 and (b) the SkX-II at K 1 = 0.3 and H = 0.1 . The directions and the magnitudes of the arrows represent the x y - and z-spin moments, respectively. The contour shows the z-spin component. (Right panel) Real-space scalar spin chirality configurations of (a) the SkX-I and (b) the SkX-II.
Figure 3. (Left panel) Real-space spin configurations of (a) the SkX-I at K 1 = 0.1 and H = 0.8 and (b) the SkX-II at K 1 = 0.3 and H = 0.1 . The directions and the magnitudes of the arrows represent the x y - and z-spin moments, respectively. The contour shows the z-spin component. (Right panel) Real-space scalar spin chirality configurations of (a) the SkX-I and (b) the SkX-II.
Magnetism 04 00018 g003
Figure 4. H dependence of the magnetization M (black squares) and scalar spin chirality χ sc (red circles) at (a) K 2 = 0.2 , (b) K 2 = 0.3 , (c) K 2 = 0.5 , and (d) K 2 = 0.6 . The backgrounds in blue and green represent the regions where the SkX with the skyrmion number of one (SkX-I) and the SkX with the skyrmion number of two (SkX-II) are stabilized, respectively. Owing to the degeneracy in terms of the sign of the scalar spin chirality, we show the case for the data where the magnetic states with a positive scalar spin chirality are obtained.
Figure 4. H dependence of the magnetization M (black squares) and scalar spin chirality χ sc (red circles) at (a) K 2 = 0.2 , (b) K 2 = 0.3 , (c) K 2 = 0.5 , and (d) K 2 = 0.6 . The backgrounds in blue and green represent the regions where the SkX with the skyrmion number of one (SkX-I) and the SkX with the skyrmion number of two (SkX-II) are stabilized, respectively. Owing to the degeneracy in terms of the sign of the scalar spin chirality, we show the case for the data where the magnetic states with a positive scalar spin chirality are obtained.
Magnetism 04 00018 g004
Figure 5. (Left panel) Real-space spin configurations of (a) the SkX-I at K 2 = 0.3 and H = 0.8 , (b) the SkX-II at K 2 = 0.3 and H = 0.1 , and (c) the chiral state at K 2 = 0.5 and H = 0.4 . The directions and the magnitudes of the arrows represent the x y - and z-spin moments, respectively. The contour shows the z-spin component. (Right panel) Real-space scalar spin chirality configurations of (a) the SkX-I, (b) the SkX-II, and (c) the chiral state, where all the spin configurations exhibit a uniform scalar spin chirality.
Figure 5. (Left panel) Real-space spin configurations of (a) the SkX-I at K 2 = 0.3 and H = 0.8 , (b) the SkX-II at K 2 = 0.3 and H = 0.1 , and (c) the chiral state at K 2 = 0.5 and H = 0.4 . The directions and the magnitudes of the arrows represent the x y - and z-spin moments, respectively. The contour shows the z-spin component. (Right panel) Real-space scalar spin chirality configurations of (a) the SkX-I, (b) the SkX-II, and (c) the chiral state, where all the spin configurations exhibit a uniform scalar spin chirality.
Magnetism 04 00018 g005
Figure 6. H dependence of the magnetization M (black squares) and scalar spin chirality χ sc at K 3 = 0.1 (red circles). The backgrounds in blue represent the regions where the SkX with the skyrmion number of one (SkX-I) is stabilized.
Figure 6. H dependence of the magnetization M (black squares) and scalar spin chirality χ sc at K 3 = 0.1 (red circles). The backgrounds in blue represent the regions where the SkX with the skyrmion number of one (SkX-I) is stabilized.
Magnetism 04 00018 g006
Figure 7. (Left panel) Real-space spin configurations of the SkX-I at K 3 = 0.1 and H = 0.8 . The directions and the magnitudes of the arrows represent the x y - and z-spin moments, respectively. The contour shows the z-spin component. (Right panel) Real-space scalar spin chirality configurations of the SkX-I.
Figure 7. (Left panel) Real-space spin configurations of the SkX-I at K 3 = 0.1 and H = 0.8 . The directions and the magnitudes of the arrows represent the x y - and z-spin moments, respectively. The contour shows the z-spin component. (Right panel) Real-space scalar spin chirality configurations of the SkX-I.
Magnetism 04 00018 g007
Figure 8. H dependence of the magnetization M (black squares) and scalar spin chirality χ sc (red circles) at K 4 = 0.2 . The backgrounds in blue represent the regions where the SkX with the skyrmion number of one (SkX-I) is stabilized.
Figure 8. H dependence of the magnetization M (black squares) and scalar spin chirality χ sc (red circles) at K 4 = 0.2 . The backgrounds in blue represent the regions where the SkX with the skyrmion number of one (SkX-I) is stabilized.
Magnetism 04 00018 g008
Figure 9. (Left panel) Real-space spin configurations of the SkX-I at K 4 = 0.2 and H = 0.8 . The directions and the magnitudes of the arrows represent the x y - and z-spin moments, respectively. The contour shows the z-spin component. (Right panel) Real-space scalar spin chirality configurations of the SkX-I.
Figure 9. (Left panel) Real-space spin configurations of the SkX-I at K 4 = 0.2 and H = 0.8 . The directions and the magnitudes of the arrows represent the x y - and z-spin moments, respectively. The contour shows the z-spin component. (Right panel) Real-space scalar spin chirality configurations of the SkX-I.
Magnetism 04 00018 g009
Figure 10. H dependence of the magnetization M (black squares) and scalar spin chirality χ sc (red circles) at K 5 = 0.8 .
Figure 10. H dependence of the magnetization M (black squares) and scalar spin chirality χ sc (red circles) at K 5 = 0.8 .
Magnetism 04 00018 g010
Figure 11. Real-space spin configurations of the magnetic states at (a) K 5 = 0.8 and H = 0.2 and (b) K 5 = 2.5 and H = 1.3 . The directions and the magnitudes of the arrows represent the x y - and z-spin moments, respectively. The contour shows the z-spin component.
Figure 11. Real-space spin configurations of the magnetic states at (a) K 5 = 0.8 and H = 0.2 and (b) K 5 = 2.5 and H = 1.3 . The directions and the magnitudes of the arrows represent the x y - and z-spin moments, respectively. The contour shows the z-spin component.
Magnetism 04 00018 g011
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Hayami, S. Skyrmion Crystal Induced by Four-Spin Interactions in Itinerant Triangular Magnets. Magnetism 2024, 4, 281-294. https://doi.org/10.3390/magnetism4030018

AMA Style

Hayami S. Skyrmion Crystal Induced by Four-Spin Interactions in Itinerant Triangular Magnets. Magnetism. 2024; 4(3):281-294. https://doi.org/10.3390/magnetism4030018

Chicago/Turabian Style

Hayami, Satoru. 2024. "Skyrmion Crystal Induced by Four-Spin Interactions in Itinerant Triangular Magnets" Magnetism 4, no. 3: 281-294. https://doi.org/10.3390/magnetism4030018

Article Metrics

Back to TopTop