This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Inhomogeneous entanglement structure in
monoaxial chiral ferromagnetic quantum spin chain

Kentaro Nishimura1,2,3, Ryosuke Yoshii4,2,3 1Institute of Science and Technology, Niigata University, Niigata 950-2181, Japan
2Research and Education Center for Natural Sciences, Keio University, 4-1-1 Hiyoshi, Yokohama, Kanagawa 223-8521, Japan
3International Institute for Sustainability with Knotted Chiral Meta Matter (SKCM2), Hiroshima University, 1-3-2 Kagamiyama, Higashi-Hiroshima, Hiroshima 739-8511, Japan
4Center for Liberal Arts and Sciences, Sanyo-Onoda City University, 1-1-1 Daigaku-Dori, Sanyo-Onoda, Yamaguchi 756-0884, Japan
Abstract

Chiral magnets, characterized by inhomogeneous magnetic moment arrangements, have attracted significant attention recently due to their topological orders, such as magnetic skyrmion lattices and chiral soliton lattices. In this work, we investigate the entanglement entropy of quantum chiral magnets and demonstrate that it reflects the inhomogeneous nature of the ground state. We perform numerical simulations of a one-dimensional monoaxial chiral ferromagnetic chain with Zeeman term using the density matrix renormalization group method. Our results show that the entanglement entropy exhibits oscillatory behavior, which can be tuned by varying the external magnetic field. Analysis of the local magnetization and spin chirality further confirms that these oscillations correspond to solitonic structures. Moreover, our findings suggest that the entanglement entropy can serve as a probe for detecting the vacuum structure, providing new insights into quantum correlations.

I Introduction

Topological solitons are spatially localized field configurations characterized by nontrivial topological charges. The well-known examples include sine-Gordon solitons, baby Skyrmions, vortices, and instantons. These solitons are ubiquitous objects that arise not only in high-energy physics—such as cosmology, particle physics, and nuclear physics—but also in condensed matter systems, including spin systems and superfluids. It is widely recognized that topological solitons play a crucial role in understanding various physical phenomena (see famous textbooks [1, 2, 3]).

In addition to the importance of the semiclassical solitonic structure, the quantum effects of topological solitons also play an essential role in understanding a wide range of physical phenomena. In high-energy physics, notable examples include the Skyrmion description [4, 5] of nucleons and the confinement mechanism [6], while in spin systems, Ref. [7] provides a relevant example. Nevertheless, the topological solitons are classical solutions to the static equations of motion, and studying their quantum aspects is not straightforward. One of the adopted approaches is the semiclassical method, which incorporates quantum effects by quantizing fluctuations around a classical background [8, 9, 7]. However, the semiclassical approximation breaks down when quantum fluctuations become large, rendering the method inapplicable in such regimes. Therefore, uncovering the quantum counterparts of the topological solitons and understanding their impact on the system through a fully quantum mechanical framework remains a significant and challenging problem.

From this perspective, a particularly suitable model is the monoaxial chiral magnet. Its Hamiltonian consists of the exchange interaction, the Zeeman term, and the Dzyaloshinskii–Moriya interaction (DMI) [10, 11]. One key reason for its relevance is that, in the classical and continuum limit, the model reduces to a (1+1)(1+1)-dimensional sine-Gordon theory with an additional total derivative term [12]. This theory admits soliton solutions, whose energies become negative due to the total derivative contribution. As a result, the ground state forms a periodic array of topological solitons, known as the chiral soliton lattice (CSL) [13], which is experimentally observed in chiral magnets [14].

Interestingly, this is not merely a toy model; such an effective theory can also be found in finite-density chiral perturbation theory under strong magnetic fields [15, 16], as well as dense QCD2\mathrm{QCD}_{2} in the large-NcN_{\mathrm{c}} expansion [17]. More recently, it has been reported that, in the continuum limit, the spin system with DMI can be mapped onto a field theory in a curved background by appropriately tuning the system parameters [18].

A second reason for the suitability of this model lies in recent experimental proposals for realizing S=1/2S=1/2 quantum spin chains with exchange interaction, DMI, and Zeeman terms using quantum simulators based on Rydberg atoms [19]. Therefore, investigating the quantum effects of the topological solitons in monoaxial chiral ferromagnetic spin systems not only offers the potential for experimental justification but also provides the advantage of feedback from experimental data.

The third reason is numerical feasibility: the ground state of this system can be efficiently obtained using the density matrix renormalization group (DMRG) method [20, 21], with the matrix product state (MPS) [22] as a variational ansatz. The other advantage of DMRG is its ability to keep the Hilbert space dimension small, enabling numerical simulations with many sites. Since the study of spatially inhomogeneous structures, such as the topological solitons, requires large system sizes, this model—in which the sine-Gordon solitons naturally appear in the ground state—is particularly well-suited for such investigations.

Indeed, the chiral ferromagnetic quantum spin chains and related models have attracted significant attention in the research community. Notably, several studies employing fully quantum mechanical methods—such as exact diagonalization, DMRG, and analytical calculations—have reported fascinating phenomena, including the quantum many-body scars [19] and spin parity effects [23, 24].

Since we are interested in the quantum effects on the topological solitons, we focus on the entanglement structure inside the system. Entanglement is a fundamental feature of quantum mechanics that has no counterpart in classical physics. Initially introduced to highlight the incompleteness of quantum theory—due to its nonlocal nature [25] and apparent incompatibility with local realism—it is now widely regarded as a hallmark of quantum behavior [26]. Subsequently, this intrinsic nonlocality has been harnessed for applications in secure communication and quantum computation [27, 28].

More recently, entanglement entropy has emerged as a key tool for quantifying entanglement between two subsystems (see Refs. [29, 30] and references therein). Its introduction has significantly advanced the study of quantum correlations across diverse systems and has found applications in various research domains (See, for instance, Refs. [31, 32, 33]).

Entanglement entropy in a local subsystem plays a crucial role in understanding thermalization in isolated quantum systems [34, 35]. One of the most widely accepted mechanisms for quantum thermalization, consistent with unitary time evolution, is the eigenstate thermalization hypothesis (ETH). According to ETH, while the total system evolves unitarily, the expectation value of a local observable is determined by the properties of the reduced density matrix, as the complementary degrees of freedom are traced out during measurement. A key characteristic of this reduced density matrix is entanglement entropy. Recently, the cousin of the entanglement entropy, the so-called Rényi entanglement entropy, has been successfully observed in Refs. [36, 37], and the thermalization of the subsystem is exhibited.

Since the entanglement structure encodes information about the vacuum state of a system, its behavior has been extensively studied in various settings [38, 39]. However, its properties in the presence of a nontrivial background remain largely unexplored. In this work, we investigate a model with an inhomogeneous background and demonstrate that the entanglement entropy reflects this inhomogeneity. Our key findings are twofold. First, the entropy’s oscillatory pattern corresponds to soliton locations, verified by magnetization and chirality profiles. Second, these oscillations vanish in ferromagnetic regimes, suggesting entanglement entropy acts as a sensitive probe of vacuum inhomogeneity. This aligns with its role in perturbed field theories [40] and offers a new lens for understanding quantum solitonic systems. We outline the model and methods, present results, and discuss implications for quantum matter.

II Model and Methods

We consider a spin-1/2 chain along the zz-direction with NN sites, governed by the following Hamiltonian:

H^=\displaystyle\hat{H}= n=1N1(Jδij+Dϵzij)S^niS^n+1j+Bn=1NS^nx,\displaystyle-\sum_{n=1}^{N-1}\left(J\delta_{ij}+D\epsilon_{zij}\right)\hat{S}_{n}^{i}\hat{S}_{n+1}^{j}+B\sum_{n=1}^{N}\hat{S}_{n}^{x}\,, (1)

where SniS_{n}^{i} is the spin operator’s ii-component (i=x,y,zi=x,y,z), δij\delta_{ij} and ϵzij\epsilon_{zij} are, respectively, the Kronecker delta and the Levi-Civita symbol, JJ (set to 11 as the energy scale) governs exchange interaction, D/J=2D/J=2 is the DMI strength, and BB is the magnetic field along xx. Here we use the Einstein summation rule. This model captures helical and soliton-like orders via the competition between these interactions (see Fig. 1).

Refer to caption
Figure 1: Schematic of the model. Sites align along the zz-direction, with the magnetic field applied along xx. The system is divided into subsystem AA (sites 1 to ii) and its complement A¯\bar{A}.

This section presents numerical results obtained using DMRG with the ITensor Library [41]. The simulations were performed with a maximum bond dimension of 4040 and a truncation error cutoff of 10810^{-8} in the final sweeps. A few hundred sweeps were typically required for convergence. We evaluate three key observables:

  • Entanglement entropy: SA=TrρAlogρAS_{A}=-\text{Tr}\rho_{A}\log\rho_{A}, where ρA=TrA¯ρ\rho_{A}=\text{Tr}_{\bar{A}}\rho is the reduced density matrix for subsystem AA (sites 1 to ii).

  • Magnetization: Mni=S^niM_{n}^{i}=\langle\hat{S}_{n}^{i}\rangle (total magnetization is defined as Mi=n=1NMniM^{i}=\sum_{n=1}^{N}M^{i}_{n}), revealing spin textures.

  • Spin chirality: χni=ϵijkS^njS^n+1k\chi^{i}_{n}=\langle\epsilon^{ijk}\hat{S}_{n}^{j}\hat{S}_{n+1}^{k}\rangle, quantifying spin twisting and soliton presence.

Here A^\langle\hat{A}\rangle stands for the expectation value of the operator A^\hat{A} with respect to the ground state.

We briefly explain the third quantity. In this study, the topological solitons in the chiral magnets—the structures where the spins are locally twisted due to the DM interaction—play a crucial role. To quantify the degree of twisting between neighboring spins, the spin chirality is an appropriate measure (see Refs. [42, 23] for its role in quantum spin systems). In the classical limit, the spin operator S^ni\hat{S}^{i}_{n} can be identified with a classical vector SniS^{i}_{n} of fixed length |𝑺n|=1/2|\bm{S}_{n}|=1/2. Assuming the spins lie in the xyxy-plane, the zz-component of the chirality [𝑺n×𝑺n+1]z[\bm{S}_{n}\times\bm{S}_{n+1}]_{z} becomes 14sin(Δφn)\frac{1}{4}\sin(\Delta\varphi_{n}), where Δφn\Delta\varphi_{n} is the angle between neighboring spins Δφn=φn+1φn\Delta\varphi_{n}=\varphi_{n+1}-\varphi_{n}. It reaches its maximum at Δφn=π/2\Delta\varphi_{n}=\pi/2, where the spins are maximally twisted (see appendix A for details).

III Results and Discussion

Refer to caption
Figure 2: The spatial dependencies of the magnitude and the phase, namely (Mnx)2+(Mny)2\sqrt{(M^{x}_{n})^{2}+(M^{y}_{n})^{2}} and atan2(Mny,Mnx)\mathrm{atan}2(M^{y}_{n},M^{x}_{n}), are plotted in the panels for B=0.0B=0.0 (a), B=0.933B=0.933 (b), B=0.935B=0.935 (c), and B=0.941B=0.941 (d).

We simulate the model across various magnetic field strengths BB, revealing rich entanglement and topological behavior. In the classical limit, the ground state at small BB is known to be the CSL (see [12] and references therein). To assess whether this configuration persists in the quantum case, we compute the phase atan2(Mny,Mnx)\mathrm{atan}2(M^{y}_{n},M^{x}_{n}) 111 The function atan2(y,x)\operatorname{atan2}(y,x) returns a unique angle θ\theta in the range π<θπ-\pi<\theta\leq\pi satisfying x=rcosθx=r\cos\theta and y=rsinθy=r\sin\theta where r>0r>0. and the magnitude (Mnx)2+(Mny)2\sqrt{(M^{x}_{n})^{2}+(M^{y}_{n})^{2}} of the spin expectation values in the xyxy-plane. Figure 2 summarizes the results.

The phase is nearly proportional to the site index nn (with π\pi and π-\pi identified), suggesting a constant pitch angle similar to the classical helical structure [12]. However, the magnitude remains zero across all nn, indicating the spin degeneracy—a clear deviation from classically expected behavior; (Mnx)2+(Mny)21/2\sqrt{(M^{x}_{n})^{2}+(M^{y}_{n})^{2}}\approx 1/2 when Mnz0M^{z}_{n}\approx 0 (in appendix B, we numerically show that Mnz0M^{z}_{n}\approx 0 at all site nn).

For B=0.933B=0.933, 0.9350.935, and 0.9410.941, the magnitude shows spatial modulation with 3, 2, and 1 local minima, respectively. While the phase suggests spins align along the x-x axis, this could misleadingly imply a ferromagnetic state driven by the magnetic field. Although the phase is a useful diagnostic in classical systems, it is much less informative in the quantum regime.

Refer to caption
Figure 3: Spin chirality at site nn (a) and the accumulated spin chirality from the left boundary (b) for B=0,0.933,0.935,0.941B=0,0.933,0.935,0.941.

We examine the numerical results for the spin chirality to understand the local minima in (Mnx)2+(Mny)2\sqrt{(M^{x}_{n})^{2}+(M^{y}_{n})^{2}}. Panel (a) in Fig. 3 shows χnz\chi_{n}^{z} (also see the numerical calculation of χnx,y\chi^{x,y}_{n} in appendix A). For B=0B=0, it remains nearly constant except near the boundaries, indicating a quantum spin configuration similar to the classical helical structure.

For B=0.933B=0.933, 0.9350.935, and 0.9410.941, χnz\chi_{n}^{z} exhibits local maxima precisely where the spin magnitude shows local minima, indicating strong twisting between neighboring spins. This behavior is reminiscent of the classical CSL state, with the peaks in χnz\chi_{n}^{z} corresponding to the quantum analogs of the classical solitons.

Panel (b) shows the accumulated spin rotation in the xyxy-plane from site 1 to nn. The rotation increases linearly for B=0B=0, consistent with a uniform pitch. In contrast, for the above BB values, sharp changes occur at the soliton positions. The number of solitons—i.e., the number of local maxima in χnz\chi_{n}^{z}—determines the total rotation up to n=50n=50. For clarity in subsequent discussions, we define the soliton number QQ as the number of local maxima in SAS_{A}.

Refer to caption
Figure 4: Plots of the entanglement entropy SAS_{A} and zz-component of the spin chirality χnz\chi^{z}_{n} for N=50N=50 with various magnetic field strengths. The horizontal axis represents the subsystem size for SAS_{A} and the position for χnz\chi^{z}_{n}. Panel (a) shows a state with constant winding, while panels (b), (c), and (d) correspond to soliton numbers QQ 33, 22, and 11, respectively.

The entanglement entropy SAS_{A} (Fig. 4) is our central result. For N=50N=50, SAS_{A} versus subsystem size are plotted (Fig. 4(a-d)). The peaks (dips) in SAS_{A} align with soliton maxima (minima in the absolute values (Mnx)2+(Mny)2\sqrt{(M^{x}_{n})^{2}+(M^{y}_{n})^{2}}), deviating from the smooth curve seen in homogeneous solutions. At high BB, where the spins align in a ferromagnetic manner (See Appendix B for details), SAS_{A} drops to 0, indicating that the system is described by the local product states. We note that this oscillation is tunable: increasing BB reduces the soliton number QQ, smoothing SAS_{A}. For the case of B=0B=0, the solitons have huge overlapping and thus the peak-dip structure is smeared out. This causes a spatially homogeneous configuration, and thus we have a smooth curve similar to the curve typically appearing in the homogeneous case (Fig. 4 (a)) [44].

Due to the background inhomogeneity and the transitions between states with different soliton numbers as the magnetic field varies, we find highly nontrivial BB-dependence of both MiM^{i} and SAS_{A}. In Fig. 5 (a), sudden changes are found in the total magnetization which reflect changes in soliton number; however, in the weak-field regime, the soliton overlap suppresses this singular behavior (See related studies for discussions of the jump behavior in both classical [45] and quantum systems [19, 23, 24]).

Figure 5 (b) shows SAS_{A} when the system is bipartitioned at the center. If a soliton is initially located at the center of the system, a change in the soliton number displaces it from the center. Since SAS_{A} is enhanced where the solitons are present and suppressed in ferromagnetic regions, SAS_{A} exhibits a nontrivial BB-dependence, characterized by alternating large and small values. Panels (c) and (d) of Figure 5, which present the enlarged views of the region exhibiting sudden change at large BB, reveal clear plateau structures. We emphasize that each plateau is characterized by different soliton numbers QQ (see Appendix C for details).

Refer to caption
Figure 5: BB-dependence of the magnetization MiM^{i} and entanglement entropy SAS_{A}. Panel (a) shows MxM^{x}, MyM^{y}, and MzM^{z} as functions of BB. In panel (b), SAS_{A} is plotted with the subsystem size set to N/2N/2. Panels (c) and (d) are magnified views of the panels (a) and (b), respectively, in the range 0.93B0.950.93\leq B\leq 0.95. The corresponding topological number QQ are indicated.

We also present NN-dependence of MxM^{x} and SAS_{A} (Fig. 6). As NN increases, the soliton emerges into the system since the twisted spin configurations are energetically more favorable than the ferromagnetic alignment. As a result, the NN-dependence of MxM^{x} and SAS_{A} becomes highly nontrivial, especially near the transition points.

As shown in Fig. 6, SAS_{A} exhibits the discontinuities that signal the transitions, while the variation in MxM^{x} is more gradual near these points. This suggests that SAS_{A} is more sensitive to the transitions than the total magnetization.

Refer to caption
Figure 6: Total size dependence of the magnetization MiM^{i} and entanglement entropy SAS_{A}. In the panel (a), we plot MxM^{x}, MyM^{y}, and MzM^{z} as functions of NN. For the plot of SAS_{A} in (b), we set the size of the subsystem to be N/2N/2.

These findings link entanglement to topological inhomogeneity. Classically, solitons arise from DM-induced twisting [46]; quantum mechanically, they imprint on SAS_{A}, suggesting it encodes vacuum structure. This mirrors the sensitivity of the entanglement in perturbed free fields [40], but extends to interacting spin systems. The soliton-entropy correspondence holds across BB, with chirality providing a bridge to classical analogs [42]. Unlike classical phase analysis, which misleads in quantum regimes , SAS_{A} and χn\chi_{n} reliably detect topological features. Our results imply entanglement entropy can probe chiral magnet ground states, potentially aiding quantum simulation platforms. Future work could explore dynamics or higher dimensions.

IV Conclusion

In this study, we explored the entanglement entropy of a monoaxial chiral ferromagnetic quantum spin chain, revealing its sensitivity to the system’s topological structure. Using the DMRG method, we analyzed a spin-1/2 model with exchange interaction, DMI, and Zeeman effect, uncovering spontaneous oscillatory patterns in the entanglement entropy. These oscillations, tunable by the external magnetic field, correspond closely to the positions of topological solitons, as confirmed by local magnetization and spin chirality profiles. At low magnetic fields, the entanglement entropy reflects the inhomogeneous background of the chiral soliton lattice, while in high-field ferromagnetic regime, it diminishes to zero, signaling a transition to a trivial, unentangled state. The size dependence of the entanglement entropy also shows a nontrivial behavior due to the inhomogeneity. Although the maximum value seems to be bounded by 1\sim 1, whether it obeys the area law or not is an open question in this case [39]. Our findings establish entanglement entropy as a powerful probe for detecting quantum vacuum inhomogeneity and topological features in chiral magnets, bridging classical soliton physics and quantum correlations.

From the perspective of entanglement entropy, our results suggest that the domain structure of the entanglement is expected to be spontaneously generated in higher dimensions with infinite volume where the Hamiltonian has translational invariance, since our model is known to have the CSL ground state even in such cases. The island structures of entanglement are expected to be ubiquitously found in various systems and play crucial roles for understanding the clustering mechanism of quantum information. This work not only deepens the understanding of quantum effects in topological systems but also suggests practical applications for characterizing ground states and manipulating entanglement in quantum simulation platforms, opening avenues for future investigations into dynamics and higher-dimensional systems.

Acknowledgments

We thank Yusuke Kato for drawing our attention to Refs. [19, 23]. This work is supported by the JSPS Grant-in-Aid for Scientific Research (KAKENHI Grant No. 19K14616, 20H01838, and 25K07156) (RY) and the WPI program “Sustainability with Knotted Chiral Meta Matter (SKCM2)” at Hiroshima University (KN and RY).

Appendix A Classical and continuum limit

Refer to caption
Figure 7: Schematic picture of the spin vectors

In the main text, we focused exclusively on the zz-component of the spin chirality. In this appendix, we clarify the meaning of the xx- and yy-components and confirm that they are not relevant in the context of our study. A schematic illustration of the spin vector in polar coordinates is shown in Fig. 7. We assume φn=π/2\varphi_{n}=\pi/2 for all nn to simplify the discussion. In this case, the spins lie in the yzyz-plane, and only the xx-component of the spin chirality becomes nonzero, given by χnx=S2sin(θnθn+1)\chi^{x}_{n}=S^{2}\sin(\theta_{n}-\theta_{n+1}). Thus, χnx\chi^{x}_{n} quantifies the rotation of spin vectors within the yzyz-plane. Analogously, χny\chi^{y}_{n} reflects the rotation in the zxzx-plane.

Figure 8 shows the numerical results for χnx\chi^{x}_{n} and χny\chi^{y}_{n}, indicating that both components remain approximately zero and show no dependence on nn, indicating that the spin expectation values do not rotate in the yzyz- or zxzx-planes. This justifies our choice to focus only on χnz\chi^{z}_{n} in the main text.

Refer to caption
Figure 8: Plots of χnx\chi^{x}_{n} (a) and χny\chi^{y}_{n} (a) at site nn and the accumulated spin chiralities =1nχx\sum_{\ell=1}^{n}\chi^{x}_{\ell} (c) and =1nχy\sum_{\ell=1}^{n}\chi^{y}_{\ell} (d).

We now discuss the continuum and classical limits of χnz\chi^{z}_{n}. In the classical limit, the spin operator S^nμ\hat{S}^{\mu}_{n} is replaced by a classical spin vector (Snx,Sny,Snz)=S𝒏n(S^{x}_{n},S^{y}_{n},S^{z}_{n})=S\bm{n}_{n} with S=1/2S=1/2, where the unit vector 𝒏n\bm{n}_{n} is parametrized in polar coordinates as 𝒏n=(sinθncosφn,sinθnsinφn,cosθn)𝖳\bm{n}_{n}=(\sin\theta_{n}\cos\varphi_{n},\,\sin\theta_{n}\sin\varphi_{n},\,\cos\theta_{n})^{\mathsf{T}}.

We adopt the spin chirality to characterize solitonic structures rather than employing the soliton number operator introduced in Ref. [23]. While this operator is conserved when J=0J=0 and is helpful in analytically constructing eigenstates, it is not conserved in our case with J0J\neq 0. Therefore, we focus on the spin chirality, which is related to the topological charge of the classical sine-Gordon solitons, as discussed below. Taking the continuum limit (with lattice spacing a0a\to 0 and z=naz=na), the spin chirality becomes

χz(z)=S2sin2θ(z)dφ(z)dz.\displaystyle\chi^{z}(z)=S^{2}\sin^{2}\theta(z)\frac{\text{d}\varphi(z)}{\text{d}z}\,. (2)

In particular, when the spins lie in the xyxy-plane (θ=π/2\theta=\pi/2), χz(z)\chi^{z}(z) simply represents the spatial variation of the phase φ(z)\varphi(z). Thus, large values of χz(z)\chi^{z}(z) correspond to regions where the spins are strongly twisted—i.e., the locations of the sine-Gordon solitons.

Finally, we emphasize that the quantity

Ntopo:=12πdzdφ(z)dz\displaystyle N_{\mathrm{topo}}:=\frac{1}{2\pi}\int\text{d}z\frac{\text{d}\varphi(z)}{\text{d}z}\in\mathbb{Z} (3)

is the topological charge, provided that φ(z)\varphi(z) takes the same value (mod 2π2\pi) at both boundaries. Naturally, the quantized topological number discussed above is well-defined only in the classical and continuum limits. Nonetheless, a key advantage of the spin chirality is that it offers an intuitive understanding of such structures within these limiting regimes.

Appendix B Numerical results of local magnetization

Refer to caption
Figure 9: The spatial dependencies of the magnetization MniM^{i}_{n} are shown in the panels for B=0.0B=0.0 (a), B=0.933B=0.933 (b), B=0.935B=0.935 (c), and B=0.941B=0.941 (d).

In this appendix, we present numerical results for the spatial dependence of the magnetization MniM^{i}_{n}. Panel (b) of Fig. 9 shows that for B=0B=0, MniM^{i}_{n} vanishes and is independent of nn, indicating a degeneracy in the spin expectation values. In the classical counterpart, the spin configuration forms a helimagnetic structure, where the spins rotate around the zz-axis with a constant pitch angle.

As BB increases, the spins tend to align along the x-x direction due to the Zeeman term, leading to an increase in |Mnx||M^{x}_{n}|. Panels (b), (c), and (d) of Fig. 9 confirm that |Mnx||M^{x}_{n}| increases and displays spatial oscillations, while MnyM^{y}_{n} oscillates around zero. The number of maxima in MnxM^{x}_{n} is 3, 2, and 1 for B=0.933B=0.933, 0.9350.935, and 0.9410.941, respectively. For sufficiently large BB, except near the boundaries, the spins align along the x-x direction, in agreement with classical expectations.

Appendix C Plateau structure in BB-dependence of MxM^{x} and SAS_{A}

Refer to caption
Figure 10: Plots of the entanglement entropy SAS_{A} and zz-component of the spin chirality χnz\chi^{z}_{n} for B=0.932B=0.932 (a), B=0.9341B=0.9341 (b), B=0.937B=0.937 (c), B=0.9396B=0.9396 (d), B=0.9413B=0.9413 (e), and B=0.94185B=0.94185 (f).

In the main text, we presented χnz\chi^{z}_{n} and SAS_{A} for a representative point within each soliton-number plateau labeled by QQ. To further confirm these results, we now show the BB-dependence of χnz\chi^{z}_{n} and SAS_{A} for additional values of BB, as shown in panels (a), (c), and (e) of Figure 10. For magnetic fields corresponding to the soliton number QQ, χnz\chi^{z}_{n} exhibits QQ local maxima. However, there is some subtlety in defining the soliton number QQ.

To demonstrate this, we examine χnz\chi^{z}_{n} and SAS_{A} at intermediate values of BB where the system undergoes a transition between different soliton numbers. For instance, panel (b) of Figure 10 corresponds to the transition from Q=3Q=3 to Q=2Q=2. In this regime, χnz\chi^{z}_{n} exhibits strong oscillations as a function of nn, while SAS_{A} takes smaller values than in panel (a) and displays fine fluctuations near the center of the system. These fine oscillations make it difficult to count the number of maxima unambiguously.

Similar behavior is also observed in panels (d) and (f), corresponding to the transitions from Q=2Q=2 to Q=1Q=1 and from Q=1Q=1 to Q=0Q=0, respectively. In fact, the large oscillations of χnz\chi^{z}_{n} can be attributed to the fact that χnz\chi^{z}_{n} takes negative values in panel (f). Such negative values indicate that the spin twist is being unwound. As a result, the ground state tends toward the ferromagnetic configuration induced by the magnetic field, and the entanglement entropy SAS_{A} is significantly reduced compared to that in panel (e).

References

  • Vilenkin and Shellard [2000] A. Vilenkin and E. P. S. Shellard, Cosmic Strings and Other Topological Defects (Cambridge University Press, 2000).
  • Manton and Sutcliffe [2004] N. S. Manton and P. Sutcliffe, Topological solitons, Cambridge Monographs on Mathematical Physics (Cambridge University Press, 2004).
  • Weinberg [2012] E. J. Weinberg, Classical solutions in quantum field theory: Solitons and Instantons in High Energy Physics, Cambridge Monographs on Mathematical Physics (Cambridge University Press, 2012).
  • Witten [1983] E. Witten, Current Algebra, Baryons, and Quark Confinement, Nucl. Phys. B 223, 433 (1983).
  • Adkins et al. [1983] G. S. Adkins, C. R. Nappi, and E. Witten, Static Properties of Nucleons in the Skyrme Model, Nucl. Phys. B 228, 552 (1983).
  • Polyakov [1975] A. M. Polyakov, Compact Gauge Fields and the Infrared Catastrophe, Phys. Lett. B 59, 82 (1975).
  • Braun and Loss [1996a] H.-B. Braun and D. Loss, Berry’s phase and quantum dynamics of ferromagnetic solitons, Phys. Rev. B 53, 3237 (1996a).
  • Dashen et al. [1974a] R. F. Dashen, B. Hasslacher, and A. Neveu, Nonperturbative methods and extended hadron models in field theory 1. semiclassical functional methods, Phys. Rev. D 10, 4114 (1974a).
  • Dashen et al. [1974b] R. F. Dashen, B. Hasslacher, and A. Neveu, Nonperturbative methods and extended hadron models in field theory 2. two-dimensional models and extended hadrons, Phys. Rev. D 10, 4130 (1974b).
  • Dzyaloshinsky [1958] I. Dzyaloshinsky, A thermodynamic theory of “weak” ferromagnetism of antiferromagnetics, Journal of Physics and Chemistry of Solids 4, 241 (1958).
  • Moriya [1960] T. Moriya, Anisotropic superexchange interaction and weak ferromagnetism, Phys. Rev. 120, 91 (1960).
  • Kishine and Ovchinnikov [2015] J.-i. Kishine and A. Ovchinnikov, Theory of monoaxial chiral helimagnet, Solid state physics 66, 1 (2015).
  • Dzyaloshinskii [1965] I. Dzyaloshinskii, Theory of helicoidal structures in antiferromagnets. 3, SOVIET PHYSICS JETP-USSR 20, 665 (1965).
  • Togawa et al. [2012] Y. Togawa, T. Koyama, K. Takayanagi, S. Mori, Y. Kousaka, J. Akimitsu, S. Nishihara, K. Inoue, A. S. Ovchinnikov, and J. Kishine, Chiral magnetic soliton lattice on a chiral helimagnet, Phys. Rev. Lett. 108, 107202 (2012).
  • Son and Stephanov [2008] D. T. Son and M. A. Stephanov, Axial anomaly and magnetism of nuclear and quark matter, Phys. Rev. D 77, 014021 (2008)arXiv:0710.1084 [hep-ph] .
  • Brauner and Yamamoto [2017] T. Brauner and N. Yamamoto, Chiral soliton lattice and charged pion condensation in strong magnetic fields, JHEP 04, 132arXiv:1609.05213 [hep-ph] .
  • Kojo [2012] T. Kojo, A (1+1) dimensional example of Quarkyonic matter, Nucl. Phys. A 877, 70 (2012)arXiv:1106.2187 [hep-ph] .
  • Kinoshita et al. [2024] S. Kinoshita, K. Murata, D. Yamamoto, and R. Yoshii, Spin systems as quantum simulators of quantum field theories in curved spacetimes (2024), arXiv:2410.07587 [hep-th] .
  • Kunimi et al. [2024] M. Kunimi, T. Tomita, H. Katsura, and Y. Kato, Proposal for simulating quantum spin models with the dzyaloshinskii-moriya interaction using rydberg atoms and the construction of asymptotic quantum many-body scar states, Phys. Rev. A 110, 043312 (2024).
  • White [1992] S. R. White, Density matrix formulation for quantum renormalization groups, Phys. Rev. Lett. 69, 2863 (1992).
  • Schollwöck [2011] U. Schollwöck, The density-matrix renormalization group in the age of matrix product states, Annals of Physics 326, 96 (2011), january 2011 Special Issue.
  • Perez-Garcia et al. [2007] D. Perez-Garcia, F. Verstraete, M. M. Wolf, and J. I. Cirac, Matrix product state representations, Quant. Inf. Comput. 7, 401 (2007)arXiv:quant-ph/0608197 .
  • Kodama et al. [2023] S. Kodama, A. Tanaka, and Y. Kato, Spin parity effects in a monoaxial chiral ferromagnetic chain, Phys. Rev. B 107, 024403 (2023).
  • Sumita et al. [2024] S. Sumita, A. Tanaka, and Y. Kato, Anisotropy-induced spin parity effects, Phys. Rev. B 110, L100403 (2024).
  • Einstein et al. [1935] A. Einstein, B. Podolsky, and N. Rosen, Can quantum-mechanical description of physical reality be considered complete?, Phys. Rev. 47, 777 (1935).
  • Bell [1964] J. S. Bell, On the einstein podolsky rosen paradox, Physics Physique Fizika 1, 195 (1964).
  • Bennett et al. [1993] C. H. Bennett, G. Brassard, C. Crépeau, R. Jozsa, A. Peres, and W. K. Wootters, Teleporting an unknown quantum state via dual classical and einstein-podolsky-rosen channels, Phys. Rev. Lett. 70, 1895 (1993).
  • Nielsen and Chuang [2000] M. A. Nielsen and I. L. Chuang, Quantum Computation and Quantum Interaction (Cambridge University Press, Cambridge, UK, 2000).
  • Amico et al. [2008] L. Amico, R. Fazio, A. Osterloh, and V. Vedral, Entanglement in many-body systems, Rev. Mod. Phys. 80, 517 (2008).
  • Horodecki et al. [2009] R. Horodecki, P. Horodecki, M. Horodecki, and K. Horodecki, Quantum entanglement, Rev. Mod. Phys. 81, 865 (2009).
  • Laflorencie [2016] N. Laflorencie, Quantum entanglement in condensed matter systems, Physics Reports 646, 1 (2016), quantum entanglement in condensed matter systems.
  • Nishioka [2018] T. Nishioka, Entanglement entropy: Holography and renormalization group, Rev. Mod. Phys. 90, 035007 (2018).
  • Vidal et al. [2003] G. Vidal, J. I. Latorre, E. Rico, and A. Kitaev, Entanglement in quantum critical phenomena, Phys. Rev. Lett. 90, 227902 (2003).
  • Deutsch [1991] J. M. Deutsch, Quantum statistical mechanics in a closed system, Phys. Rev. A 43, 2046 (1991).
  • Srednicki [1994] M. Srednicki, Chaos and quantum thermalization, Phys. Rev. E 50, 888 (1994).
  • Islam et al. [2015] R. Islam, R. Ma, P. M. Preiss, M. Eric Tai, A. Lukin, M. Rispoli, and M. Greiner, Measuring entanglement entropy in a quantum many-body system, Nature 528, 77 (2015).
  • Kaufman et al. [2016] A. Kaufman, E. Tai, A. Lukin, M. Rispoli, R. Schittko, P. Preiss, and M. Greiner, Quantum thermalization through entanglement in an isolated many-body system, Science 353, 794 (2016).
  • Calabrese and Cardy [2004] P. Calabrese and J. L. Cardy, Entanglement entropy and quantum field theory, J. Stat. Mech. 0406, P06002 (2004)arXiv:hep-th/0405152 .
  • Eisert et al. [2010] J. Eisert, M. Cramer, and M. B. Plenio, Colloquium: Area laws for the entanglement entropy, Rev. Mod. Phys. 82, 277 (2010).
  • Casini and Huerta [2009] H. Casini and M. Huerta, Entanglement entropy in free quantum field theory, J. Phys. A 42, 504007 (2009)arXiv:0905.2562 [hep-th] .
  • Fishman et al. [2022] M. Fishman, S. R. White, and E. M. Stoudenmire, The itensor software library for tensor network calculations, SciPost Phys. Codeb. 2022, 4 (2022)arXiv:2007.14822 [cs.MS] .
  • Braun and Loss [1996b] H.-B. Braun and D. Loss, Chiral quantum spin solitons, Journal of applied physics 79, 6107 (1996b).
  • Note [1] The function atan2(y,x)\operatorname{atan2}(y,x) returns a unique angle θ\theta in the range π<θπ-\pi<\theta\leq\pi satisfying x=rcosθx=r\cos\theta and y=rsinθy=r\sin\theta where r>0r>0.
  • Bianchi and Donà [2019] E. Bianchi and P. Donà, Typical entanglement entropy in the presence of a center: Page curve and its variance, Phys. Rev. D 100, 105010 (2019).
  • Kishine et al. [2014] J.-i. Kishine, I. G. Bostrem, A. S. Ovchinnikov, and V. E. Sinitsyn, Topological magnetization jumps in a confined chiral soliton lattice, Phys. Rev. B 89, 014419 (2014).
  • Roessler et al. [2006] U. K. Roessler, A. Bogdanov, and C. Pfleiderer, Spontaneous skyrmion ground states in magnetic metals, Nature 442, 797 (2006).