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

Isolated Skyrmions in the CP2CP^{2} nonlinear σ\sigma-model with a Dzyaloshinskii-Moriya type interaction

Yutaka Akagi Department of Physics, Graduate School of Science, The University of Tokyo, Bunkyo, Tokyo 113-0033, Japan Yuki Amari BLTP, JINR, Dubna 141980, Moscow Region, Russia Department of Mathematical Physics, Toyama Prefectural University, Kurokawa 5180, Imizu, Toyama, 939-0398, Japan Nobuyuki Sawado Department of Physics, Tokyo University of Science, Noda, Chiba 278-8510, Japan Yakov Shnir BLTP, JINR, Dubna 141980, Moscow Region, Russia
Abstract

We study two dimensional soliton solutions in the CP2CP^{2} nonlinear σ\sigma-model with a Dzyaloshinskii-Moriya type interaction. First, we derive such a model as a continuous limit of the SU(3)SU(3) tilted ferromagnetic Heisenberg model on a square lattice. Then, introducing an additional potential term to the derived Hamiltonian, we obtain exact soliton solutions for particular sets of parameters of the model. The vacuum of the exact solution can be interpreted as a spin nematic state. For a wider range of coupling constants, we construct numerical solutions, which possess the same type of asymptotic decay as the exact analytical solution, both decaying into a spin nematic state.

1 Introduction

In the 1960s, Skyrme introduced a (3+1)-dimensional O(4)O(4) nonlinear (NL) σ\sigma-model [1, 2] which is now well-known as a prototype of a classical field theory that supports topological solitons (See Ref. [3], for example). Historically, the Skyrme model has been proposed as a low-energy effective theory of atomic nuclei. In this framework, the topological charge of the field configuration is identified with the baryon number.

The Skyrme model, apart from being considered a good candidate for the low-energy QCD effective theory, has attracted much attention in various applications, ranging from string theory and cosmology to condensed matter physics. One of the most interesting developments here is related to a planar reduction of the NLσ\sigma-model, the so-called baby Skyrme model [4, 5, 6]. This (2+1)-dimensional simplified theory resembles the basic properties of the original Skyrme model in many aspects.

The baby Skyrme model finds a number of physical realizations in different branches of modern physics. Originally, it was proposed as a modification of the Heisenberg model [7, 4, 5]. Then, it was pointed out that Skyrmion configurations naturally arise in condensed matter systems with intrinsic and induced chirality [8, 9, 10, 11, 12]. These baby Skyrmions, often referred to as magnetic Skyrmions, were experimentally observed in non-centrosymmetric or chiral magnets [13, 14, 15]. This discovery triggered extensive research on Skyrmions in magnetic materials. This direction is a rapidly growing area both theoretically and experimentally [16].

A typical stabilizing mechanism of magnetic skyrmions is the existence of Dzyaloshinskii-Moriya (DM) interaction [17, 18], which stems from the spin-orbit coupling. In fact, the magnetic Skyrmions in chiral magnets can be well described by the continuum effective Hamiltonian

H=d2x[J2(𝒎)2+κ𝒎(×𝒎)Bm3+A{|𝒎|2+(m3)2}],H=\int{\mathrm{d}}^{2}x\left[\frac{J}{2}\left(\nabla\bm{m}\right)^{2}+\kappa\leavevmode\nobreak\ \bm{m}\cdot\left(\nabla\times\bm{m}\right)-Bm^{3}+A\left\{|\bm{m}|^{2}+\left(m^{3}\right)^{2}\right\}\right], (1.1)

where 𝒎(𝒓)=(m1,m2,m3)\bm{m}(\bm{r})=\left(m^{1},m^{2},m^{3}\right) is a three component unit magnetization vector which corresponds to the spin expectation value at position 𝒓\bm{r}. The first term in Eq. (1.1) is the continuum limit of the Heisenberg exchange interaction, i.e., the kinetic term of the O(3)O(3) NLσ\sigma-model, which is often referred to as the Dirichlet term. The second term there is the DM interaction term, the third one is the Zeeman coupling with an external magnetic field BB, and the last, symmetry breaking term A{|𝒎|2+(m3)2}A\left\{|\bm{m}|^{2}+\left(m^{3}\right)^{2}\right\} represents the uniaxial anisotropy.

It is remarkable that in the limiting case A=κ2/2J,B=0A=\kappa^{2}/2J,B=0, the Hamiltonian (1.1) can be written as the static version of the SU(2)SU(2) gauged O(3)O(3) NLσ\sigma-model [19, 20]

H=J2d2x(k𝒎+𝑨k×𝒎)2,k=1,2\displaystyle H=\frac{J}{2}\int{\mathrm{d}}^{2}x\left({\partial}_{k}\bm{m}+\bm{A}_{k}\times\bm{m}\right)^{2},\qquad k=1,2 (1.2)

with a background gauge field 𝑨1=(κ/J, 0, 0),𝑨2=(0,κ/J, 0)\bm{A}_{1}=(-\kappa/J,\leavevmode\nobreak\ 0,\leavevmode\nobreak\ 0),\leavevmode\nobreak\ \bm{A}_{2}=(0,-\kappa/J,\leavevmode\nobreak\ 0). Though the DM term is usually introduced phenomenologically, a mathematical derivation of the Hamiltonian (1.2) with arbitrary 𝑨k\bm{A}_{k} has been developed recently [19], i.e.; it has been shown that the Hamiltonian can be derived mathematically in a continuum limit of the tilted (quantum) Heisenberg model

=Jij(𝒲iSia𝒲i1)(𝒲jSja𝒲j1),{\cal H}=-J\sum_{\langle ij\rangle}\left({\cal W}_{i}S^{a}_{i}{\cal W}^{-1}_{i}\right)\left({\cal W}_{j}S^{a}_{j}{\cal W}^{-1}_{j}\right)\leavevmode\nobreak\ , (1.3)

where the sum ij\langle ij\rangle is taken over the nearest-neighbor sites, SiaS^{a}_{i} denotes the aa-th component of spin operators at site ii and 𝒲iSU(2){\cal W}_{i}\in SU(2). It was reported that the tilting Heisenberg model can be derived from a Hubbard model at half-filling in the presence of spin-orbit coupling [21]. Therefore, the background field 𝑨k\bm{A}_{k} can still be interpreted as an effect of the spin-orbit coupling.

There are two advantages of utilizing the expression (1.2) for the theoretical study of baby Skyrmions in the presence of the so-called Lifshitz invariant, an interaction term which is linear in a derivative of an order parameter [22, 23], like the DM term. The first advantage of the form Eq. (1.2) is that one can study a NLσ\sigma-model with various form of Lifshitz invariants which are mathematically derived by choice of the background field 𝑨k\bm{A}_{k}, although Lifshitz invariants have, in general, a phenomenological origin corresponding to the crystallographic handedness of a given sample. The second advantage of the model (1.2) is that it allows us to employ several analytical techniques developed for the gauged NLσ\sigma-model. It has been recently reported in Ref. [20] that the Hamiltonian (1.2) with a specific choice of the potential term exactly satisfies the Bogomol’nyi bound, and the corresponding Bogomol’nyi-Prasad-Sommerfield (BPS) equations have exact closed-form solutions [20, 24, 25].

Geometrically, the planar Skyrmions are very nicely described in terms of the CP1CP^{1} complex field on the compactified domain space S2S^{2} [6]. Further, there are various generalizations of this model; for example, two-dimensional CP2CP^{2} Skyrmions have been studied in the pure CP2CP^{2} NLσ\sigma-model [26, 27, 28] and in the Faddeev-Skyrme type model [29, 30].

Remarkably, the two dimensional CP2CP^{2} NLσ\sigma-model can be obtained as a continuum limit of the SU(3)SU(3) ferromagnetic (FM) Heisenberg model [31, 32] on a square lattice defined by the Hamiltonian

=J2ijTimTjm,{\cal H}=-\frac{J}{2}\sum_{\langle ij\rangle}T^{m}_{i}T^{m}_{j}, (1.4)

where JJ is a positive constant, and TimT^{m}_{i} (m=1,,8m=1,...,8) stand for the SU(3)SU(3) spin operators of the fundamental representation at site ii satisfying the commutation relation

[Til,Tim]=iflmnTin.\left[T^{l}_{i},T^{m}_{i}\right]=if_{lmn}T^{n}_{i}. (1.5)

Here, the structure constants are given by flmn=i2Tr(λl[λm,λn])f_{lmn}=-\frac{i}{2}\mathrm{Tr}\left(\lambda_{l}\left[\lambda_{m},\lambda_{n}\right]\right), where λm\lambda_{m} are the usual Gell-Mann matrices.

The SU(3)SU(3) FM Heisenberg model may play an important role in diverse physical systems ranging from string theory [33] to condensed matter, or quantum optical three-level systems [34]. It can be derived from a spin-1 bilinear-biquadratic model with a specific choice of coupling constants, so-called FM SU(3)SU(3) point, see, e.g., Ref. [35]. The SU(3)SU(3) spin operators can be defined in terms of the SU(2)SU(2) spin operators SaS^{a} (a=1,2,3a=1,2,3) as

(T7T5T2)=(S1S2S3),(T3T8T1T4T6)=((S1)2(S2)213[𝑺𝑺3(S3)2]S1S2+S2S1S3S1+S1S3S2S3+S3S2).\left(\begin{array}[]{c}T^{7}\\ T^{5}\\ T^{2}\end{array}\right)=\left(\begin{array}[]{c}S^{1}\\ -S^{2}\\ S^{3}\end{array}\right),\qquad\left(\begin{array}[]{c}T^{3}\\ T^{8}\\ T^{1}\\ T^{4}\\ T^{6}\end{array}\right)=-\left(\begin{array}[]{c}\left(S^{1}\right)^{2}-\left(S^{2}\right)^{2}\\ \frac{1}{\sqrt{3}}\left[\bm{S}\cdot\bm{S}-3\left(S^{3}\right)^{2}\right]\\ S^{1}S^{2}+S^{2}S^{1}\\ S^{3}S^{1}+S^{1}S^{3}\\ S^{2}S^{3}+S^{3}S^{2}\end{array}\right). (1.6)

Using the SU(2)SU(2) commutation relation [Sia,Sib]=iεabcSic\left[S^{a}_{i},S^{b}_{i}\right]=i\varepsilon_{abc}S^{c}_{i} where εabc\varepsilon_{abc} denotes the anti-symmetric tensor, one can check that the operators (1.6) satisfy the SU(3)SU(3) commutation relation (1.5).

In the present paper, we study baby Skyrmion solutions of an extended CP2CP^{2} NLσ\sigma-model composed of the CP2CP^{2} Dirichlet term, a DM type interaction term, i.e., the Lifshitz invariant, and a potential term. The Lifshitz invariant, instead of being introduced ad hoc in the continuum Hamiltonian, can be derived in a mathematically well-defined way via consideration of a continuum limit of the SU(3)SU(3) tilted Heisenberg model. Below we will implement this approach in our derivation of the Lifshitz invariant. In the extended CP2CP^{2} NLσ\sigma-model, we derive exact soliton solutions for specific combinations of coupling constants called the BPS point and solvable line. For a broader range of coupling constants, we construct solitons by solving the Euler-Lagrange equation numerically.

The organization of this paper is the following: In the next section, we derive an SU(3)SU(3) gauged CP2CP^{2} NLσ\sigma-model from the SU(3)SU(3) tilted Heisenberg model. Similar to the SU(2)SU(2) case described as Eq. (1.2), the term linear in a background field can be viewed as a Lifshitz invariant term. In Sec. 3, we study exact Skyrmionic solutions of the SU(3)SU(3) gauged CP2CP^{2} NLσ\sigma-model in the presence of a potential term for the BPS point and solvable line using the BPS arguments. The numerical construction of baby Skyrmion solutions off the solvable line is given in Sec. 4. Our conclusions are given in Sec. 5.

2 Gauged 𝐂𝐏𝟐\bm{CP^{2}} NL𝛔\bm{\sigma}-model from a spin system

To find Lifshitz invariant terms relevant for the CP2CP^{2} NLσ\sigma-model, we begin to derive an SU(3)SU(3) gauged CP2CP^{2} NLσ\sigma-model, a generalization of Eq. (1.2), from a spin system on a square lattice. By analogy with Eq. (1.2), the Lifshitz invariant, in that case, can be introduced as a term linear in a non-dynamical background gauge potential of the gauged CP2CP^{2} model.

Following the procedure to obtain a gauged NLσ\sigma-model from a spin system, as discussed in Ref. [19], we consider a generalization of the SU(3)SU(3) Heisenberg model defined by the Hamiltonian

=J2ijTim(U^ij)mnTjn,{\cal H}=-\frac{J}{2}\sum_{\langle ij\rangle}T^{m}_{i}(\hat{U}_{ij})_{mn}T^{n}_{j}, (2.1)

where U^ij\hat{U}_{ij} is a background field which can be recognized as a Wilson line operator along with the link from the point ii to the point jj, which is an element of the SU(3)SU(3) group in the adjoint representation. As in the SU(2)SU(2) case [19], the field U^ij\hat{U}_{ij} may describe effects originated from spin (nematic)-orbital coupling, complicated crystalline structure, and so on. This Hamiltonian can be viewed as the exchange interaction term for the tilted operator T~im=𝒲iTim𝒲i1\tilde{T}^{m}_{i}={\cal W}_{i}T^{m}_{i}{\cal W}^{-1}_{i} where 𝒲iSU(3){\cal W}_{i}\in SU(3), because one can write 𝒲jTjm𝒲j1=(Rj)mnTjn{\cal W}_{j}T^{m}_{j}{\cal W}^{-1}_{j}=(R_{j})_{mn}T_{j}^{n} where RjR_{j} is an element of SU(3)SU(3) in the adjoint representation. Clearly, U^ij=RiTRj\hat{U}_{ij}=R_{i}^{\rm T}R_{j}, where T\rm T stands for the transposition.

Let us now find the classical counterpart of the quantum Hamiltonian (2.1). It can be defined as an expectation value of Eq. (2.1) in a state possessing over-completeness, through a path integral representation of the partition function. In order to construct such a state for the spin-1 system, it is convenient to introduce the Cartesian basis

|x1=i2(|+1|1),|x2=12(|+1+|1),|x3=i|0,\left|{x^{1}}\right>=\frac{i}{\sqrt{2}}\left(\left|{+1}\right>-\left|{-1}\right>\right),\qquad\left|{x^{2}}\right>=\frac{1}{\sqrt{2}}\left(\left|{+1}\right>+\left|{-1}\right>\right),\qquad\left|{x^{3}}\right>=-i\left|{0}\right>, (2.2)

where |m=|S=1,m\left|{m}\right>=\left|{S=1,m}\right> (m=0m=0, ±1\pm 1). In terms of the Cartesian basis, an arbitrary spin-1 state at a site jj can be expressed as a linear combination |Zj=Za(𝒓j)|xaj\left|{Z}\right>_{j}=Z^{a}(\bm{r}_{j})\left|{x^{a}}\right>_{j} where 𝒓j{\bm{r}}_{j} stands for the position of the site jj, and 𝒁=(Z1,Z2,Z3)T\bm{Z}=\left(Z^{1},Z^{2},Z^{3}\right)^{\rm T} is a complex vector of unit length [36, 31]. Since the state |Zj\left|{Z}\right>_{j} satisfies an over-completeness relation, one can obtain the classical Hamiltonian using the state

|Z=j|Zj=jZa(𝒓j)|xaj.\left|{Z}\right>=\otimes_{j}\left|{Z}\right>_{j}=\otimes_{j}Z^{a}(\bm{r}_{j})\left|{x^{a}}\right>_{j}\leavevmode\nobreak\ . (2.3)

Since 𝒁\bm{Z} is normalized and has the gauge degrees of freedom corresponding to the overall phase factor multiplication, it takes values in S5/S1CP2S^{5}/S^{1}\approx CP^{2}. In terms of the basis (2.2), the SU(3)SU(3) spin operators can be defined as

Tm=(λm)ab|xaxb|m=1,2,,8,T^{m}=\left(\lambda_{m}\right)_{ab}\left|{x^{a}}\right>\left<{x^{b}}\right|\qquad m=1,2,\cdots,8, (2.4)

where λm\lambda_{m} is the mm-th component of the Gell-Mann matrices. One can check that they satisfy the SU(3)SU(3) commutation relation (1.5). The expectation values of the SU(3)SU(3) operators in the state (2.3) are given by

Tjmnm(𝒓j)=(λm)abZ¯a(𝒓j)Zb(𝒓j),\langle{T^{m}_{j}}\rangle\equiv n^{m}(\bm{r}_{j})=\left(\lambda_{m}\right)_{ab}\bar{Z}^{a}(\bm{r}_{j})Z^{b}(\bm{r}_{j}), (2.5)

where Z¯a\bar{Z}^{a} denotes the complex conjugation of ZaZ^{a}. In the context of QCD, the field nmn^{m} is usually termed a color (direction) field [37]. The color field satisfies the constraints

nmnm=43,nm=32dmpqnpnq,n^{m}n^{m}=\frac{4}{3},\qquad n^{m}=\frac{3}{2}d_{mpq}n^{p}n^{q}\leavevmode\nobreak\ , (2.6)

where dmpq=14Tr(λm{λp,λq})d_{mpq}=\frac{1}{4}\mathrm{Tr}\left(\lambda_{m}\left\{\lambda_{p},\lambda_{q}\right\}\right). Consequently, the number of degrees of freedom of the color field reduces to four. Note that, combining the constraints (2.6), one can get the Casimir identity dmpqnmnpnq=8/9d_{mpq}n^{m}n^{p}n^{q}=8/9.

In terms of the color field, the classical Hamiltonian is given by

H\displaystyle H Z||Z=J2ijnl(𝒓i)(U^ij)lmnm(𝒓j).\displaystyle\equiv\left<{Z}\right|{\cal H}\left|{Z}\right>=-\frac{J}{2}\sum_{\langle ij\rangle}n^{l}(\bm{r}_{i})(\hat{U}_{ij})_{lm}n^{m}(\bm{r}_{j}). (2.7)

Let us write the position of a site jj next to a site ii as 𝒓j=𝒓i+aϵ𝒆k\bm{r}_{j}=\bm{r}_{i}+a\epsilon\bm{e}_{k} where 𝒆k\bm{e}_{k} is the unit vector in the kk-th direction, ϵ=±1\epsilon=\pm 1, and aa stands for the lattice constant. For a1a\ll 1, the field U^ij\hat{U}_{ij} can be approximated by the exponential expansion

U^ijeiaϵAkm(𝒓i)l^m=𝟙+iaϵAkm(𝒓i)l^ma22Akm(𝒓i)Akn(𝒓i)l^ml^n+𝒪(a3),\hat{U}_{ij}\approx e^{ia\epsilon A^{m}_{k}(\bm{r}_{i})\hat{l}_{m}}={\mathbb{1}}+ia\epsilon A_{k}^{m}(\bm{r}_{i})\hat{l}_{m}-\frac{a^{2}}{2}A_{k}^{m}(\bm{r}_{i})A_{k}^{n}(\bm{r}_{i})\hat{l}_{m}\hat{l}_{n}+{\cal O}(a^{3}), (2.8)

where 𝟙{\mathbb{1}} is the unit matrix and l^m\hat{l}_{m} are the generators of SU(3)SU(3) in the adjoint representation, i.e., (l^m)pq=ifmpq(\hat{l}_{m})_{pq}=if_{mpq}. In addition, since the model (2.1) is ferromagnetic, it is natural to assume that nearest-neighbor spins are oriented in the almost same direction, which allows us to use the Taylor expansion

nm(𝒓j)=nm(𝒓i)+aϵknm(𝒓i)+𝒪(a2).n^{m}(\bm{r}_{j})=n^{m}(\bm{r}_{i})+a\epsilon{\partial}_{k}n^{m}(\bm{r}_{i})+{\cal O}(a^{2}). (2.9)

Replacing the sum over the lattice sites in Eq. (2.7) by the integral a2d2x\displaystyle{a^{-2}\int{\mathrm{d}}^{2}x}, we obtain a continuum Hamiltonian, except for a constant term, of the form

H=J8d2x[Tr(k𝔫k𝔫)2iTr(Ak[𝔫,k𝔫])Tr([Ak,𝔫]2)],\displaystyle H=\frac{J}{8}\int{\mathrm{d}}^{2}x\left[\mathrm{Tr}\left({\partial}_{k}\mathfrak{n}{\partial}_{k}\mathfrak{n}\right)-2i\mathrm{Tr}\left(A_{k}\left[\mathfrak{n},{\partial}_{k}\mathfrak{n}\right]\right)-\mathrm{Tr}\left(\left[A_{k},\mathfrak{n}\right]^{2}\right)\right], (2.10)

where Ak=AkmλmA_{k}=A^{m}_{k}\lambda_{m} and 𝔫=nmλm\mathfrak{n}=n^{m}\lambda_{m}. Similar to its SU(2)SU(2) counterpart expressed as Eq. (1.2), this Hamiltonian can also be written as the static energy of an SU(3)SU(3) gauged CP2CP^{2} NLσ\sigma-model

H=J8d2xTr(Dk𝔫Dk𝔫),H=\frac{J}{8}\int{\mathrm{d}}^{2}x\mathrm{Tr}\left(D_{k}\mathfrak{n}\leavevmode\nobreak\ D_{k}\mathfrak{n}\right), (2.11)

where Dk𝔫=k𝔫i[Ak,𝔫]D_{k}\mathfrak{n}={\partial}_{k}\mathfrak{n}-i\left[A_{k},\mathfrak{n}\right] is the SU(3)SU(3) covariant derivative. Since the Hamiltonian is given by the SU(3)SU(3) covariant derivative, Eq. (2.11) is invariant under the SU(3)SU(3) gauge transformation

𝔫g𝔫g1,AkgAkg1+igkg1,\mathfrak{n}\to g\mathfrak{n}g^{-1},\qquad A_{k}\to gA_{k}g^{-1}+ig{\partial}_{k}g^{-1}, (2.12)

where gSU(3)g\in SU(3). Note that, however, since the Hamiltonian (2.11) does not include kinetic terms for the gauge field, like the Yang-Mills term, or the Chern-Simons term, the gauge potential is just a background field, not the dynamical one. We suppose that the gauge field is fixed beforehand by the structure of a sample and give the value by hand, like the SU(2)SU(2) case. The gauge fixing allows us to recognize the second term in Eq. (2.10) as a Lifshitz invariant term.

We would like to emphasize that we do not deal with Eq. (2.11) as a gauge theory. Rather, we deem it the CP2CP^{2} NLσ\sigma-model with a Lifshitz invariant, and show the existence of the exact and the numerical solutions. For the baby Skyrmion solutions we shall obtain, the color field 𝔫\mathfrak{n} approaches to a constant value 𝔫\mathfrak{n}_{\infty} at spatial infinity so that the physical space 2\mathbb{R}^{2} can be topologically compactified to S2S^{2}. Therefore, they are characterized by the topological degree of the map 𝔫:2S2CP2\mathfrak{n}:\mathbb{R}^{2}\sim S^{2}\mapsto CP^{2} given by

Q=i32πd2xεjkTr(𝔫[j𝔫,k𝔫]).Q=-\frac{i}{32\pi}\int{\mathrm{d}}^{2}x\leavevmode\nobreak\ \varepsilon_{jk}\mathrm{Tr}\left(\mathfrak{n}\left[{\partial}_{j}\mathfrak{n},{\partial}_{k}\mathfrak{n}\right]\right). (2.13)

Combining with the assumption that the gauge is fixed, it is reasonable to identify this quantity (2.13) with the topological charge in our model111 If one extends the model (2.11) with a dynamical gauge field, the topological charge is defined by the SU(3)SU(3) gauge invariant quantity which is directly obtained by replacing the partial difference in Eq. (2.13) with the covariant derivative..

3 Exact solutions of the 𝐒𝐔(𝟑)\bm{SU(3)} gauged 𝐂𝐏𝟐\bm{CP^{2}} NL𝛔\bm{\sigma}-model

In this section, we derive exact solutions of the model with the Hamiltonian (2.11) supplemented by a potential term. We first remark on the validity of the variational problem. As discussed in Refs. [20, 25] for the SU(2)SU(2) case, a surface term, which appears in the process of variation, cannot be ignored if the physical space is non-compact and the gauge potential AkA_{k} does not vanish at the spatial infinity like the DM term. This problem can be cured by introducing an appropriate boundary term, like [20]

HBoundary=4ρd2xεjkjTr(𝔫Ak),H_{\rm Boundary}=\mp 4\rho\int{\mathrm{d}}^{2}x\leavevmode\nobreak\ \varepsilon_{jk}{\partial}_{j}\mathrm{Tr}(\mathfrak{n}A_{k})\,, (3.1)

where ρ=J/8\rho=J/8. Here the gauge potential AkA_{k} satisfies

[𝔫,Aj]±i2εij[𝔫,[𝔫,Ak]]=0,\left[\mathfrak{n}_{\infty},A_{j}\right]\pm\frac{i}{2}\varepsilon_{ij}\left[\mathfrak{n}_{\infty},\left[\mathfrak{n}_{\infty},A_{k}\right]\right]=0, (3.2)

where 𝔫\mathfrak{n}_{\infty} is the asymptotic value of 𝔫\mathfrak{n} at spatial infinity. Note that Eq. (3.2) corresponds to the asymptotic form of the BPS equation, which we shall discuss in the next subsection. Hence, all field configurations we consider in this paper satisfy this equation automatically.

Since (3.1) is a surface term, it does not contribute to the Euler-Lagrange equation, i.e., the classical Heisenberg equation. Note that the solutions derived in the following sections satisfy Derrick’s scaling relation with the boundary term, which is obtained by keeping the background field AkA_{k} intact under the scaling, i. e., E1+2E0=0E_{1}+2E_{0}=0 where E1E_{1} denotes the energy contribution from the first derivative terms including the boundary term (3.1) and E0E_{0} from no derivative terms.

3.1 BPS solutions

Recently, it has been proved that the SU(2)SU(2) gauged CP1CP^{1} NLσ\sigma-model (1.2) possesses BPS solutions in the presence of a particular potential term [20, 24]. Here, we show that BPS solutions also exist in the SU(3)SU(3) gauged CP2CP^{2} model with a special choice of the potential term, which is given by

Hpot=±4ρd2xTr(𝔫F12),H_{\rm pot}=\pm 4\rho\int{\mathrm{d}}^{2}x\mathrm{Tr}\left(\mathfrak{n}F_{12}\right), (3.3)

where Fjk=jAkkAji[Aj,Ak]F_{jk}={\partial}_{j}A_{k}-{\partial}_{k}A_{j}-i\left[A_{j},A_{k}\right]. As we shall see in the next subsection, the potential term can possess a natural physical interpretation for some background gauge field. It follows that the Hamiltonian we study here reads

H=ρd2xTr(Dk𝔫Dk𝔫)±4ρd2xTr(𝔫F12)4ρd2xεjkjTr(𝔫Ak),H=\rho\int{\mathrm{d}}^{2}x\mathrm{Tr}\left(D_{k}\mathfrak{n}\leavevmode\nobreak\ D_{k}\mathfrak{n}\right)\pm 4\rho\int{\mathrm{d}}^{2}x\mathrm{Tr}\left(\mathfrak{n}F_{12}\right)\mp 4\rho\int{\mathrm{d}}^{2}x\leavevmode\nobreak\ \varepsilon_{jk}{\partial}_{j}\mathrm{Tr}(\mathfrak{n}A_{k}), (3.4)

where the double-sign corresponds to that of Eq. (3.1).

First, let us show that the lower energy bound of Eq. (3.4) is given by the topological charge (2.13). The first term in Eq. (3.4) can be written as

ρd2xTr(Dk𝔫Dk𝔫)\displaystyle\rho\int{\mathrm{d}}^{2}x\leavevmode\nobreak\ \mathrm{Tr}\left(D_{k}\mathfrak{n}\leavevmode\nobreak\ D_{k}\mathfrak{n}\right) =ρ2d2x[Tr(Dk𝔫Dk𝔫)+(i2)2Tr([𝔫,Dk𝔫]2)]\displaystyle=\frac{\rho}{2}\int{\mathrm{d}}^{2}x\left[\mathrm{Tr}\left(D_{k}\mathfrak{n}\leavevmode\nobreak\ D_{k}\mathfrak{n}\right)+\left(\frac{i}{2}\right)^{2}\mathrm{Tr}\left(\left[\mathfrak{n},D_{k}\mathfrak{n}\right]^{2}\right)\right]
=ρ2d2xTr(Dj𝔫±i2εjk[𝔫,Dk𝔫])2±iρ2d2xεjkTr(𝔫[Dj𝔫,Dk𝔫])\displaystyle=\frac{\rho}{2}\int{\mathrm{d}}^{2}x\mathrm{Tr}\left(D_{j}\mathfrak{n}\pm\frac{i}{2}\varepsilon_{jk}\left[\mathfrak{n},D_{k}\mathfrak{n}\right]\right)^{2}\pm\frac{i\rho}{2}\int{\mathrm{d}}^{2}x\leavevmode\nobreak\ \varepsilon_{jk}\mathrm{Tr}\left(\mathfrak{n}\left[D_{j}\mathfrak{n},D_{k}\mathfrak{n}\right]\right)
±iρ2d2xεjkTr(𝔫[Dj𝔫,Dk𝔫]).\displaystyle\geq\pm\frac{i\rho}{2}\int{\mathrm{d}}^{2}x\leavevmode\nobreak\ \varepsilon_{jk}\mathrm{Tr}\left(\mathfrak{n}\left[D_{j}\mathfrak{n},D_{k}\mathfrak{n}\right]\right). (3.5)

It follows that the equality is satisfied if

Dj𝔫±i2εjk[𝔫,Dk𝔫]=0,D_{j}\mathfrak{n}\pm\frac{i}{2}\varepsilon_{jk}\left[\mathfrak{n},D_{k}\mathfrak{n}\right]=0, (3.6)

which reduces to Eq. (3.2) at the spatial infinity. Therefore, one obtains the lower bound of the form

H\displaystyle H ±ρ2d2x[iεjkTr(𝔫[Dj𝔫,Dk𝔫])+8Tr(𝔫F12)8εjkjTr(𝔫Ak)]\displaystyle\geq\pm\frac{\rho}{2}\int{\mathrm{d}}^{2}x\left[i\varepsilon_{jk}\mathrm{Tr}\left(\mathfrak{n}\left[D_{j}\mathfrak{n},D_{k}\mathfrak{n}\right]\right)+8\mathrm{Tr}\left(\mathfrak{n}F_{12}\right)-8\varepsilon_{jk}{\partial}_{j}\mathrm{Tr}\left(\mathfrak{n}A_{k}\right)\right]
=±iρ2d2xεjkTr(𝔫[j𝔫,k𝔫])\displaystyle=\pm\frac{i\rho}{2}\int{\mathrm{d}}^{2}x\leavevmode\nobreak\ \varepsilon_{jk}\mathrm{Tr}\left(\mathfrak{n}\left[{\partial}_{j}\mathfrak{n},{\partial}_{k}\mathfrak{n}\right]\right)
=16πρQ,\displaystyle=\mp 16\pi\rho\leavevmode\nobreak\ Q, (3.7)

where the corresponding BPS equation is given by Eq. (3.6). Note that, unlike the energy bound of the CPNCP^{N} self-dual solutions [7, 27], the energy bound (3.7) can be negative, and it is not proportional to the absolute value of the topological charge.

As is often the case in two-dimensional BPS equations [7, 20], solutions can be best described in terms of the complex coordinates z±=x1±ix2z_{\pm}=x^{1}\pm ix^{2}. Further, we make use of the associated differential operator and background field defined as ±=12(1i2){\partial}_{\pm}=\frac{1}{2}\left({\partial}_{1}\mp i{\partial}_{2}\right) and A±=12(A1iA2)A_{\pm}=\frac{1}{2}(A_{1}\mp iA_{2}). Then, the BPS equation (3.6) can be written as

D±𝔫12[𝔫,D±𝔫]=0.D_{\pm}\mathfrak{n}-\frac{1}{2}\left[\mathfrak{n},D_{\pm}\mathfrak{n}\right]=0. (3.8)

Similar to the SU(2)SU(2) case [20], Eq.  (3.8) with a plus sign can be solved if the background field has the form

A+=ig1+g,A_{+}=ig^{-1}{\partial}_{+}g, (3.9)

where gSL(3,)g\in SL(3,\mathbb{C}). Note that Eq. (3.9) is not necessarily a pure gauge. Similarly, Eq. (3.8) with the minus sign on the right-hand side can be solved if A=ig1gA_{-}=ig^{-1}{\partial}_{-}g. For the background field (3.9), one finds that the BPS equation (3.8) is equivalent to

+𝔫~12[𝔫~,+𝔫~]=0,𝔫~=g𝔫g1,{\partial}_{+}\tilde{\mathfrak{n}}-\frac{1}{2}\left[\tilde{\mathfrak{n}},{\partial}_{+}\tilde{\mathfrak{n}}\right]=0,\leavevmode\nobreak\ \leavevmode\nobreak\ \tilde{\mathfrak{n}}=g\mathfrak{n}g^{-1}, (3.10)

because, under the SL(3,)SL(3,\mathbb{C}) gauge transformation, the fields are changed as 𝔫𝔫~=g𝔫g1\mathfrak{n}\to\tilde{\mathfrak{n}}=g\mathfrak{n}g^{-1} and A+A~±=gA+g1+ig±g1=0A_{+}\to\tilde{A}_{\pm}=gA_{+}g^{-1}+ig{\partial}_{\pm}g^{-1}=0. In the following, we only consider Eq. (3.9) to simplify our discussion.

In order to solve the equation (3.10), we introduce a tractable parameterization of the color field

𝔫=23Uλ8U,\mathfrak{n}=-\frac{2}{\sqrt{3}}U\lambda_{8}U^{\dagger}, (3.11)

with U=(𝒀1,𝒀2,𝒁)SU(3)U=\left(\bm{Y}_{1},\bm{Y}_{2},\bm{Z}\right)\in SU(3), where 𝒁\bm{Z} is the continuum counter part of the vector 𝒁\bm{Z} in Eq. (2.3) and 𝒀1,𝒀2\bm{Y}_{1},\bm{Y}_{2} are vectors forming an orthonormal basis for 3\mathbb{C}^{3} with 𝒁\bm{Z}. Up to the gauge degrees of freedom, the components 𝒀i\bm{Y}_{i} can be written as

𝒀1=(Z¯3,0,Z¯1)T1|Z2|2,𝒀2=(Z¯2Z1,1|Z2|2,Z¯2Z3)T1|Z2|2.\bm{Y}_{1}=\frac{\left(-\bar{Z}^{3},0,\bar{Z}^{1}\right)^{\rm T}}{\sqrt{1-|Z^{2}|^{2}}},\qquad\bm{Y}_{2}=\frac{\left(-\bar{Z}^{2}Z^{1},1-|Z^{2}|^{2},-\bar{Z}^{2}Z^{3}\right)^{\rm T}}{\sqrt{1-|Z^{2}|^{2}}}. (3.12)

Therefore, the vector 𝒁\bm{Z} fully defines the color field 𝔫\mathfrak{n}. Accordingly, we can write

𝔫~=23Wλ8W1,\tilde{\mathfrak{n}}=-\frac{2}{\sqrt{3}}W\lambda_{8}W^{-1}, (3.13)

with W=gU=(𝑾1,𝑾2,𝑾3)SL(3,)W=gU=\left(\bm{W}_{1},\bm{W}_{2},\bm{W}_{3}\right)\in SL(3,\mathbb{C}). It follows that the field 𝒁\bm{Z}, which is the fundamental field of the model, is given by 𝒁=g1𝑾3\bm{Z}=g^{-1}\bm{W}_{3}. Substituting the field (3.13) into the equation (3.10), one finds that Eq. (3.10) reduces to the coupled equation

{𝑾11+𝑾3=0𝑾21+𝑾3=0,\left\{\begin{aligned} \bm{W}_{1}^{-1}{\partial}_{+}\bm{W}_{3}=0\\ \bm{W}_{2}^{-1}{\partial}_{+}\bm{W}_{3}=0\end{aligned}\right., (3.14)

where 𝑾l1=𝒀lg1\bm{W}_{l}^{-1}=\bm{Y}_{l}^{\dagger}g^{-1} (l=1,2)(l=1,2). Since the three vectors {𝒀1,𝒀2,𝒁}\{\bm{Y}_{1},\bm{Y}_{2},\bm{Z}\} form an orthonormal basis, Eq. (3.14) implies +𝑾3=β𝑾3{\partial}_{+}\bm{W}_{3}=\beta\bm{W}_{3} where the function β\beta is given by β=β𝑾31𝑾3=𝑾31+𝑾3.\beta=\beta\bm{W}^{-1}_{3}\bm{W}_{3}=\bm{W}^{-1}_{3}{\partial}_{+}\bm{W}_{3}. Therefore, the equation (3.10) is solved by any configuration satisfying

𝒟+𝑾3=0,{\cal D}_{+}\bm{W}_{3}=0, (3.15)

where 𝒟+𝚽=+𝚽(𝚽1+𝚽)𝚽{\cal D}_{+}{\bm{\Phi}}={\partial}_{+}{\bm{\Phi}}-({\bm{\Phi}}^{-1}{\partial}_{+}{\bm{\Phi}}){\bm{\Phi}} for arbitrary non-zero vector 𝚽\bm{\Phi}. Moreover, we write

𝑾3=|𝑾3|2𝒘,\bm{W}_{3}=\sqrt{|\bm{W}_{3}|^{2}}\leavevmode\nobreak\ \bm{w}, (3.16)

where 𝒘\bm{w} is a three component unit vector, i.e. |𝒘|2=𝒘𝒘=1|\bm{w}|^{2}=\bm{w}^{\dagger}\bm{w}=1. Then, Eq. (3.15) can be reduced to

𝒟+𝒘μ𝒘(𝒘μ𝒘)𝒘=0,{\cal D}_{+}\bm{w}\equiv{\partial}_{\mu}\bm{w}-\left(\bm{w}^{\dagger}{\partial}_{\mu}\bm{w}\right)\bm{w}=0, (3.17)

which is the very BPS equation of the standard CP2CP^{2} NLσ\sigma-model. Thus, a general solution of Eq. (3.15), up to the gauge degrees of freedom, is given by

𝒘=𝑷|𝑷|,𝑷=(P1(z),P2(z),P3(z))T,\bm{w}=\frac{{\bm{P}}}{|{\bm{P}}|},\qquad{\bm{P}}=\left(P_{1}(z_{-}),P_{2}(z_{-}),P_{3}(z_{-})\right)^{\rm T}, (3.18)

where 𝑷{\bm{P}} has no overall factor, and PaP_{a} is a polynomial in zz_{-}. Therefore, we finally obtain the solution for the 𝒁\bm{Z} field

𝒁=g1𝑾3=χg1𝒘=χg1𝑷,\bm{Z}=g^{-1}\bm{W}_{3}=\chi g^{-1}\bm{w}=\chi g^{-1}{\bm{P}}, (3.19)

where χ\chi is a normalization factor.

3.2 Properties of the BPS solutions

As the BPS bound (3.7) indicates, the lowest energy solution among Eq. (3.19) with a given background function gg possesses the highest topological charge. In terms of the explicit calculation of the topological charge, we discuss the conditions for the lowest energy solutions.

The topological charge (2.13) can be written in terms of 𝒁\bm{Z} as

Q=i2πd2xεij(𝒟i𝒁)𝒟j𝒁.Q=-\frac{i}{2\pi}\int{\mathrm{d}}^{2}x\leavevmode\nobreak\ \varepsilon^{ij}\left({\cal D}_{i}\bm{Z}\right)^{\dagger}{\cal D}_{j}\bm{Z}. (3.20)

We employ the constant background gauge field A+A_{+} for simplicity. Then, the matrix gg in Eq. (3.9) becomes

g=exp(iA+z+),g=\exp\left(-iA_{+}z_{+}\right)\,, (3.21)

so that the components of g1g^{-1} are given by power series in z+z_{+}. It allows us to write Eq. (3.20) as a line integral along the circle at spatial infinity

Q=12πS1C,Q=\frac{1}{2\pi}\int_{S^{1}_{\infty}}C, (3.22)

with C=i𝒁d𝒁C=-i\bm{Z}^{\dagger}{\mathrm{d}}\bm{Z} [27, 38], since the one-form CC becomes globally well-defined. To evaluate the integral in Eq. (3.22), we write explicitly

𝒁=χ|P1|2+|P2|2+|P3|2a(g1a1(z+)Pa(z)g2a1(z+)Pa(z)g3a1(z+)Pa(z)),\bm{Z}=\frac{\chi}{\sqrt{|P_{1}|^{2}+|P_{2}|^{2}+|P_{3}|^{2}}}\sum_{a}\left(\begin{array}[]{c}g^{-1}_{1a}(z_{+})P_{a}(z_{-})\\ g^{-1}_{2a}(z_{+})P_{a}(z_{-})\\ g^{-1}_{3a}(z_{+})P_{a}(z_{-})\end{array}\right), (3.23)

where gab1g^{-1}_{ab} is the (a,b)(a,b) component of the inverse matrix g1g^{-1}.

Let NaN_{a} (KabK_{ab}) be the highest power in PaP_{a} (gab1g^{-1}_{ab}). Note that though gab1g^{-1}_{ab} are formally represented as power series in z+z_{+}, the integers KbaK_{ba} are not always infinite; especially, if a positive integer power of A+A_{+} is zero, all of KbaK_{ba} become finite because g1g^{-1} reduces to a polynomial of finite degree in z+z_{+}. Using the plane polar coordinates {r,θ}\{r,\theta\}, one can write gba1(z+)Pa(z)rNa+Kbaexp[i(NaKba)θ]g^{-1}_{ba}(z_{+})P_{a}(z_{-})\sim r^{N_{a}+K_{ba}}\exp[-i(N_{a}-K_{ba})\theta] at the spatial boundary and find that only the components of the highest power in rr contribute to the integral (3.22). Since we are interested in constructing topological solitons, we consider the case when the physical space 2\mathbb{R}^{2} can be compactified to the sphere S2S^{2}, i.e., the field 𝒁\bm{Z} takes some fixed value on the spatial boundary. Such a compactification is possible if there is only one pair {Na,Kba}\{N_{a},K_{ba}\} giving the largest sum Na+KbaN_{a}+K_{ba} or any pairs {Na,Kba}\{N_{a},K_{ba}\}, sharing the largest sum, have the same value of the difference. For such configurations, the topological charge is given by

Q=Na+Kba,Q=-N_{a}+K_{ba}, (3.24)

where the combination {Na,Kba}\{N_{a},K_{ba}\} yields the largest sum among any pairs {Nc,Kdc}\{N_{c},K_{dc}\}. This equation (3.24) indicates that the highest topological charge configuration is given by the choice Na=0N_{a}=0 for a particular value of aa which gives the biggest KbaK_{ba}.

Refer to caption
Figure 1: Topological charge density of the axial symmetric solution (3.28) with κ=1\kappa=1.

We are looking for the lowest energy solutions with an explicit background field. As a particular example, let us consider

A1=κ(λ1+λ4+λ5),A2=κ(λ2+λ4λ5),A_{1}=\kappa\left(\lambda_{1}+\lambda_{4}+\lambda_{5}\right),\qquad A_{2}=\kappa\left(\lambda_{2}+\lambda_{4}-\lambda_{5}\right), (3.25)

where κ\kappa is a constant. Clearly, this choice yields the potential term

V=4Tr(𝔫F12)=163κ2n8=16κ2(23(S3)2),V=4\mathrm{Tr}\left(\mathfrak{n}F_{12}\right)=-16\sqrt{3}\kappa^{2}n^{8}=16\kappa^{2}\left(2-3\langle{(S^{3})^{2}}\rangle\right), (3.26)

which can be interpreted as an easy-axis anisotropy, or quadratic Zeeman term, which naturally appears in condensed matter physics. In this case, the solution (3.19) can be written as

𝒁=χΔ(P1(z)+2κz+eπi4P3(z)P2(z)+iκz+P1(z)+κ2z+22e3πi4P3(z)P3(z)).\bm{Z}=\frac{\chi}{\sqrt{\Delta}}\left(\begin{array}[]{c}P_{1}(z_{-})+\sqrt{2}\kappa z_{+}e^{\frac{\pi i}{4}}P_{3}(z_{-})\\ P_{2}(z_{-})+i\kappa z_{+}P_{1}(z_{-})+\frac{\kappa^{2}z_{+}^{2}}{\sqrt{2}}e^{\frac{3\pi i}{4}}P_{3}(z_{-})\\ P_{3}(z_{-})\end{array}\right). (3.27)

Therefore, the solution with the highest topological charge is given by P1=α1P_{1}=\alpha_{1}, P2=α2z+α3P_{2}=\alpha_{2}z_{-}+\alpha_{3} with αi\alpha_{i}\in\mathbb{C}, and P3P_{3} being a nonzero constant. Choosing P1=P2=0P_{1}=P_{2}=0, one can obtain the axially-symmetric solution

𝒁=1Δ(2κz+eπi4κ2z+22e3πi41),Δ=1+2κ2z+z+κ42z+2z2,\bm{Z}=\frac{1}{\sqrt{\Delta}}\left(\begin{array}[]{c}\sqrt{2}\kappa z_{+}e^{\frac{\pi i}{4}}\\ \frac{\kappa^{2}z_{+}^{2}}{\sqrt{2}}e^{\frac{3\pi i}{4}}\\ 1\end{array}\right),\qquad\Delta=1+2\kappa^{2}z_{+}z_{-}+\frac{\kappa^{4}}{2}z_{+}^{2}z_{-}^{2}, (3.28)

which possesses the topological charge Q=2Q=2. Note that this configuration also satisfies the BPS equation of the pure CP2CP^{2} NLσ\sigma-model [26, 27, 31]. Figure 1 shows the distribution of the topological charge (3.20) of this solution (3.28) with κ=1\kappa=1. We find that the topological charge density has a single peak, although higher charge topological solitons with axial symmetry are likely to possess a volcano structure, see e.g., Ref. [39]. These highest charge solutions give the asymptotic values at spatial infinity of the color field

(n1,n2,n3,n4,n5,n6,n7,n8)=(0,0,1,0,0,0,0,1/3).\left(n_{\infty}^{1},n_{\infty}^{2},n_{\infty}^{3},n_{\infty}^{4},n_{\infty}^{5},n_{\infty}^{6},n_{\infty}^{7},n_{\infty}^{8}\right)=(0,0,-1,0,0,0,0,1/\sqrt{3})\leavevmode\nobreak\ . (3.29)

It indicates that 𝔫\mathfrak{n} takes the vacuum value in the Cartan subalgebra of SU(3)SU(3). Hence, the vacuum of the model corresponds to a spin nematic, i.e., S1=S2=S3=0\langle{S^{1}}\rangle=\langle{S^{2}}\rangle=\langle{S^{3}}\rangle=0 and (S2)2=0,(S1)2=(S3)2=1\langle{\left(S^{2}\right)^{2}}\rangle=0,\langle{\left(S^{1}\right)^{2}}\rangle=\langle{\left(S^{3}\right)^{2}}\rangle=1. Unlike the pure CP2CP^{2} model, there is no degeneracy between the spin nematic state and ferromagnetic state in our model because the SU(3)SU(3) global symmetry is broken. As shown in Fig. 2, the spin nematic state is partially broken around the soliton because the expectation values Sa\langle{S^{a}}\rangle become finite. Fig. 3 shows that (Sa)2\langle{\left(S^{a}\right)^{2}}\rangle of the solution (3.28) are axially symmetric, although the expectation values Sa\langle{S^{a}}\rangle have angular dependence.

Refer to caption
Figure 2: The expectation values Sa\langle{S^{a}}\rangle for the solution (3.28) with κ=1\kappa=1.
Refer to caption
Figure 3: The expectation values (Sa)2\langle{\left(S^{a}\right)^{2}}\rangle for the solution (3.28) with κ=1\kappa=1.

3.3 Exact solutions off the BPS point

Note that the Hamiltonian (1.1) with B=2AB=2A admits closed-form analytical solutions [40]. Further, the CP1CP^{1} BPS truncation corresponds to the restricted choice of the parameters, B=2A=κ2B=2A=\kappa^{2}. The relation B=2AB=2A is referred to as the solvable line, whereas the restriction B=2A=κ2B=2A=\kappa^{2} is called the BPS point [25]. Here we show that similar restrictions occur in our model. For this purpose, we consider the generalized Hamiltonian

H=HD+HL+HBoundary+ν2Hani+μ2Hpot,H=H_{\rm D}+H_{\rm L}+H_{\rm Boundary}+\nu^{2}H_{\rm ani}+\mu^{2}H_{\rm pot}, (3.30)

where ν\nu and μ\mu are real coupling constants. Here, HDH_{\rm D} indicates the CP2CP^{2} Dirichlet term, i.e., the first term in the r.h.s of Eq. (2.10), and HLH_{\rm L} does the Lifshitz invariant term which is the second term of that. Explicitly, these and other terms read

HD=ρd2xTr(k𝔫k𝔫),\displaystyle H_{\rm D}=\rho\int{\mathrm{d}}^{2}x\mathrm{Tr}\left({\partial}_{k}\mathfrak{n}\leavevmode\nobreak\ {\partial}_{k}\mathfrak{n}\right), (3.31)
HL=2iρd2xTr(Ak[𝔫,k𝔫]),\displaystyle H_{\rm L}=-2i\rho\int{\mathrm{d}}^{2}x\mathrm{Tr}\left(A_{k}\left[\mathfrak{n},{\partial}_{k}\mathfrak{n}\right]\right), (3.32)
Hani=ρd2x[Tr([Ak,𝔫]2)Tr([Ak,𝔫]2)],\displaystyle H_{\rm ani}=-\rho\int{\mathrm{d}}^{2}x\left[\mathrm{Tr}\left(\left[A_{k},\mathfrak{n}\right]^{2}\right)-\mathrm{Tr}\left(\left[A_{k},\mathfrak{n}_{\infty}\right]^{2}\right)\right], (3.33)
Hpot=4ρd2x[Tr(𝔫F12)Tr(𝔫F12)],\displaystyle H_{\rm pot}=4\rho\int{\mathrm{d}}^{2}x\left[\mathrm{Tr}\left(\mathfrak{n}F_{12}\right)-\mathrm{Tr}\left(\mathfrak{n}_{\infty}F_{12}\right)\right],\, (3.34)

where AkA_{k} is a constant background field, as before. Finally, the boundary term HBoundaryH_{\rm Boundary} is defined by Eq. (3.1) with the negative sign in the r.h.s., the same as before. Note that we also introduced constant terms in Eqs. (3.33) and (3.34) in order to guarantee the finiteness of the total energy. Clearly, the Hamiltonian (3.30) is reduced to Eq. (3.4) as we set ν2=μ2=1\nu^{2}=\mu^{2}=1.

The existence of exact solutions of the Hamiltonian (3.30) with ν2=μ2\nu^{2}=\mu^{2} can be easily shown if we rescale the space coordinates as xr0x\vec{x}\to r_{0}\vec{x}, where r0r_{0} is a positive constant, while the background gauge field AkA_{k} remains intact. By rescaling, the Hamiltonian (3.30) becomes

H=HD+r0(HL+HBoundary)+r02(ν2Hani+μ2Hpot).H=H_{\rm D}+r_{0}\left(H_{\rm L}+H_{\rm Boundary}\right)+r_{0}^{2}\left(\nu^{2}H_{\rm ani}+\mu^{2}H_{\rm pot}\right). (3.35)

Setting ν2=μ2\nu^{2}=\mu^{2} and choosing the scale parameter r0=ν2r_{0}=\nu^{-2}, one gets

Hν2=μ2r0=ν2=HD+ν2(HL+HBoundary+Hani+Hpot).H^{r_{0}=\nu^{-2}}_{\nu^{2}=\mu^{2}}=H_{\rm D}+\nu^{-2}\left(H_{\rm L}+H_{\rm Boundary}+H_{\rm ani}+H_{\rm pot}\right). (3.36)

Notice that since the solutions (3.19) with PiP_{i} being arbitrary constants are holomorphic maps from S2S^{2} to CP2CP^{2}, they satisfy not only the variational equations δHν2=μ2=1=0\delta H_{\nu^{2}=\mu^{2}=1}=0 but also the equations δHD=0\delta H_{\rm D}=0, where δ\delta denotes the variation with respect to 𝔫\mathfrak{n} with preserving the constraint (2.6). Therefore, the solutions also satisfy the equations δHν2=μ2r0=ν2=0\delta H^{r_{0}=\nu^{-2}}_{\nu^{2}=\mu^{2}}=0. This implies that, in the limit μ2=ν2\mu^{2}=\nu^{2}, the Hamiltonian (3.30) supports a family of exact solutions of the form

𝒁(ν2)=exp[iν2A+z+]𝒄,\bm{Z}(\nu^{2})=\exp\left[i\nu^{2}A_{+}z_{+}\right]\bm{c}\,, (3.37)

where 𝒄\bm{c} is a three-component complex unit vector.

Since the solution (3.37) is a BPS solution of the pure CP2CP^{2} model with the positive topological charge QQ, one gets HD[𝒁(ν2)]=16πρQH_{\rm D}[\bm{Z}(\nu^{2})]=16\pi\rho Q. In addition, the lower bound at the BPS point (3.7) indicates that Hν2=μ2=1[𝒁(ν2=1)]=16πρQH_{\nu^{2}=\mu^{2}=1}[\bm{Z}(\nu^{2}=1)]=-16\pi\rho Q. Combining these bounds, we find that the total energy of the solution (3.37) is given by

Hν2=μ2[𝒁(ν2)]=16πρ(12ν2)Q.\displaystyle H_{\nu^{2}=\mu^{2}}[\bm{Z}(\nu^{2})]=16\pi\rho\left(1-\frac{2}{\nu^{2}}\right)Q. (3.38)

Since the energy becomes negative if ν2<2\nu^{2}<2, we can expect that for small values of the coupling ν2\nu^{2}, the homogeneous vacuum state becomes unstable, and then separated 2D Skyrmions (or a Skyrmion lattice) emerges as a ground state.

4 Numerical solutions

4.1 Axial symmetric solutions

In this section, we study baby Skyrmion solutions of the Hamiltonian (3.30) with various combinations of the coupling constants. Apart from the solvable line, no exact solutions could find analytically, and then we have to solve the equations numerically. Here, we restrict ourselves to the case of the background field given by Eq. (3.25).

For the background field (3.25), by analogy with the case of the single CP1CP^{1} magnetic Skyrmion solution, we can look for a configuration described by the axially symmetric ansatz

𝒁=(sinF(r)cosG(r)eiΦ1(θ),sinF(r)sinG(r)eiΦ2(θ),cosF(r)),\bm{Z}=\left(\sin F(r)\cos G(r)e^{i\Phi_{1}(\theta)},\sin F(r)\sin G(r)e^{i\Phi_{2}(\theta)},\cos F(r)\right), (4.1)

where FF and GG (Φ1\Phi_{1} and Φ2\Phi_{2}) are real functions of the plane polar coordinates rr (θ\theta).

The exact solution on the solvable line ν2=μ2\nu^{2}=\mu^{2} with axial symmetry can be written in terms of the ansatz with the functions

F=tan12ν4κ2r2+ν8κ4r42,G=tan1(ν2κr2),Φ1=θ+π4,Φ2=2θ+3π4.F=\tan^{-1}\sqrt{2\nu^{4}\kappa^{2}r^{2}+\frac{\nu^{8}\kappa^{4}r^{4}}{2}},\qquad G=\tan^{-1}\left(\frac{\nu^{2}\kappa r}{2}\right),\qquad\Phi_{1}=\theta+\frac{\pi}{4},\qquad\Phi_{2}=2\theta+\frac{3\pi}{4}. (4.2)

Further, the solution (3.28) is given by Eq. (4.2) with ν2=1\nu^{2}=1. This configuration is a useful reference point in the configuration space as we discuss below some properties of numerical solutions in the extended model (3.30).

For our numerical study, it is convenient to introduce the energy unit 8ρ8\rho and the length unit κ1\kappa^{-1}, in order to scale the coupling constants. Then, the rescaled components of the Hamiltonian with the ansatz (4.1) become

HD=d2x[F2+sin2FG2+\displaystyle H_{\rm D}=\int{\mathrm{d}}^{2}x\left[F^{\prime 2}+\sin^{2}FG^{\prime 2}+\right.
+sin2Fr2{Φ˙12cos2G+Φ˙22sin2G}sin4Fr2(Φ˙1cos2G+Φ˙2sin2G)2],\displaystyle\left.\qquad\qquad\quad+\frac{\sin^{2}F}{r^{2}}\left\{\dot{\Phi}_{1}^{2}\cos^{2}G+\dot{\Phi}_{2}^{2}\sin^{2}G\right\}-\frac{\sin^{4}F}{r^{2}}\left(\dot{\Phi}_{1}\cos^{2}G+\dot{\Phi}_{2}\sin^{2}G\right)^{2}\right], (4.3)
HL=2d2xr[2cos(θ+π4Φ1){r(cosGFsin2FsinGG2)+sin2FcosGΦ˙12\displaystyle H_{\rm L}=-2\int\frac{{\mathrm{d}}^{2}x}{r}\Bigg{[}\sqrt{2}\cos\left(\theta+\frac{\pi}{4}-\Phi_{1}\right)\left\{r\left(\cos GF^{\prime}-\sin 2F\sin G\frac{G^{\prime}}{2}\right)+\sin 2F\cos G\frac{\dot{\Phi}_{1}}{2}\right.
sin2Fsin2FcosG(cos2GΦ˙1+sin2GΦ˙2)}\displaystyle\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad-\sin 2F\sin^{2}F\cos G\left(\cos^{2}G\dot{\Phi}_{1}+\sin^{2}G\dot{\Phi}_{2}\right)\Big{\}}
sin(θ+Φ1Φ2){rsin2FG+12sin2Fsin2G(Φ˙1+Φ˙2)\displaystyle\qquad\qquad\qquad\qquad\quad-\sin\left(\theta+\Phi_{1}-\Phi_{2}\right)\left\{r\sin^{2}FG^{\prime}+\frac{1}{2}\sin^{2}F\sin 2G\left(\dot{\Phi}_{1}+\dot{\Phi}_{2}\right)\right.
sin4Fsin2G(cos2GΦ˙1+sin2GΦ˙2)}],\displaystyle\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad\quad\qquad\qquad-\sin^{4}F\sin 2G\left(\cos^{2}G\dot{\Phi}_{1}+\sin^{2}G\dot{\Phi}_{2}\right)\Big{\}}\Bigg{]}, (4.4)
Hani=12d2x[16sin2Fcos2G{cos2F12cos(2Φ1Φ2+π4)sin2FsinG+sin2FsinG2}\displaystyle H_{\rm ani}=\frac{1}{2}\int{\mathrm{d}}^{2}x\Bigg{[}16\sin^{2}F\cos^{2}G\left\{\cos^{2}F-\frac{1}{\sqrt{2}}\cos\left(2\Phi_{1}-\Phi_{2}+\frac{\pi}{4}\right)\sin 2F\sin G+\sin^{2}F\sin G^{2}\right\}
+sin22F(1+2sin2G)+8(cos2Fcos2Gsin2F)2+4cos22Gsin4F4],\displaystyle\qquad\qquad\qquad\qquad+\sin^{2}2F(1+2\sin^{2}G)+8(\cos^{2}F-\cos^{2}G\sin^{2}F)^{2}+4\cos^{2}2G\sin^{4}F-4\Bigg{]}, (4.5)
Hpot=2d2x(13n8)=6d2xcos2F,\displaystyle H_{\rm pot}=2\int{\mathrm{d}}^{2}x\left(1-\sqrt{3}n^{8}\right)=6\int{\mathrm{d}}^{2}x\cos^{2}F, (4.6)

where the prime and the dot ˙\dot{} stands for the derivatives with respect to the radial coordinate rr and angular coordinate θ\theta, respectively. The system of corresponding Euler-Lagrange equations for Φi\Phi_{i} can be solved algebraically for an arbitrary set of the coupling constants, and the solutions are

Φ1=θ+π4,Φ2=2θ+3π4+mπ,\Phi_{1}=\theta+\frac{\pi}{4},\qquad\Phi_{2}=2\theta+\frac{3\pi}{4}+m\pi\,, (4.7)

where mm is an integer. Without loss of generality, we choose m=0m=0 by transferring the corresponding multiple windings of the phase Φ2\Phi_{2} to the sign of the profile function GG. Then, the system of the Euler-Lagrange equations for the profile functions with the phase factor (4.7) reads

δHDδF+δHLδF+ν2δHaniδF+μ2δHpotδF=0,δHDδG+δHLδG+ν2δHaniδG+μ2δHpotδG=0,\begin{split}&\frac{\delta H_{\rm D}}{\delta F}+\frac{\delta H_{\rm L}}{\delta F}+\nu^{2}\frac{\delta H_{\rm ani}}{\delta F}+\mu^{2}\frac{\delta H_{\rm pot}}{\delta F}=0,\\ &\frac{\delta H_{\rm D}}{\delta G}+\frac{\delta H_{\rm L}}{\delta G}+\nu^{2}\frac{\delta H_{\rm ani}}{\delta G}+\mu^{2}\frac{\delta H_{\rm pot}}{\delta G}=0,\end{split} (4.8)

with

δHDδF=[2rF′′+2Fsin2F{rG2+1+3sin2Gr2sin2Fr(1+sin2G)2}],\displaystyle\frac{\delta H_{\rm D}}{\delta F}=\left[2rF^{\prime\prime}+2F^{\prime}-\sin 2F\left\{rG^{\prime 2}+\frac{1+3\sin^{2}G}{r}-\frac{2\sin^{2}F}{r}\left(1+\sin^{2}G\right)^{2}\right\}\right], (4.9)
δHLδF=2[22sin2F{rsinGG+cosG+cosG(1+sin2G)(4cos2F1)}\displaystyle\frac{\delta H_{\rm L}}{\delta F}=-2\left[2\sqrt{2}\sin^{2}F\{-r\sin GG^{\prime}+\cos G+\cos G\left(1+\sin^{2}G\right)\left(4\cos^{2}F-1\right)\}\right.
rsin2FG32sin2Fsin2G+4cosFsin3Fsin2G(1+sin2G)],\displaystyle\left.\qquad\qquad\qquad\qquad-r\sin 2FG^{\prime}-\frac{3}{2}\sin 2F\sin 2G+4\cos F\sin^{3}F\sin 2G\left(1+\sin^{2}G\right)\right], (4.10)
δHaniδF=2r[42sinGcos2Gsin2F(34sin2F)4cosFsin3Fcos22G\displaystyle\frac{\delta H_{\rm ani}}{\delta F}=2r\left[4\sqrt{2}\sin G\cos^{2}G\sin^{2}F\left(3-4\sin^{2}F\right)-4\cos F\sin^{3}F\cos^{2}2G\right.
+4sin2F{cos2Fsin2Fcos2G(1+sin2G)}sin2Fcos2F(1+2sin2G)],\displaystyle\left.\qquad\qquad\qquad+4\sin 2F\left\{\cos^{2}F-\sin^{2}F\cos^{2}G\left(1+\sin^{2}G\right)\right\}-\sin 2F\cos 2F(1+2\sin^{2}G)\right], (4.11)
δHpotδF=6rsin2F,\displaystyle\frac{\delta H_{\rm pot}}{\delta F}=6r\sin 2F, (4.12)
δHDδG=[2rsinF2G′′+2rsin2FFG+2sin2FGsin2Fsin2Gr{32sin2F(1+sin2G)}],\displaystyle\frac{\delta H_{\rm D}}{\delta G}=\left[2r\sin F^{2}G^{\prime\prime}+2r\sin 2FF^{\prime}G^{\prime}+2\sin^{2}FG^{\prime}-\frac{\sin^{2}F\sin 2G}{r}\left\{3-2\sin^{2}F\left(1+\sin^{2}G\right)\right\}\right], (4.13)
δHLδG=2[2sin2FsinG{2rF+sin2F(13sin2G)}\displaystyle\frac{\delta H_{\rm L}}{\delta G}=-2\left[\sqrt{2}\sin^{2}F\sin G\left\{2rF^{\prime}+\sin 2F\left(1-3\sin^{2}G\right)\right\}\right.
+rsin2FF+sin2F(13cos2G)+sin4F(1+3cos2G2cos22G)],\displaystyle\left.\qquad\qquad\qquad\qquad+r\sin 2FF^{\prime}+\sin^{2}F(1-3\cos 2G)+\sin^{4}F\left(1+3\cos 2G-2\cos^{2}2G\right)\right], (4.14)
δHaniδG=r[82cosFsin3FcosG(13sin2G)+16sin4Fcos3GsinGsin22Fsin2G],\displaystyle\frac{\delta H_{\rm ani}}{\delta G}=r\left[8\sqrt{2}\cos F\sin^{3}F\cos G\left(1-3\sin^{2}G\right)+16\sin^{4}F\cos^{3}G\sin G-\sin^{2}2F\sin 2G\right], (4.15)
δHpotδG=0.\displaystyle\frac{\delta H_{\rm pot}}{\delta G}=0. (4.16)

We solve the equations for ν2μ2\nu^{2}\neq\mu^{2} numerically with the boundary condition

F(0)=G(0)=0,limrF(r)=limrG(r)=π/2,F(0)=G(0)=0,\qquad\lim_{r\to\infty}F(r)=\lim_{r\to\infty}G(r)=\pi/2, (4.17)

which the exact solution (4.2) satisfies. This vacuum corresponds to the spin nematic state (3.29).

Refer to caption
Figure 4: Plot of the profile functions {F,G}\{F,G\} (left) and the topological charge density (right) of numerical solutions for changing the coupling constant ν2\nu^{2} at μ2=1.5\mu^{2}=1.5. The gray line indicates the quantities of the exact solution (4.2) on the solvable line.
ν2\nu^{2} HH HDH_{\rm D} HLH_{\rm L} ν2Hani\nu^{2}H_{\rm ani} μ2Hpot\mu^{2}H_{\rm pot} HBoundaryH_{\rm Boundary} Derrick QQ
0.1 -117.47 13.51 -136.48 125.49 5.67 -125.67 -2.00 2.00
0.3 -34.02 13.41 -53.60 41.37 6.69 -41.89 -1.99 2.00
0.8 -8.46 13.06 -29.37 14.73 8.82 -15.71 -1.91 2.00
 
1.5 -4.19 12.57 -16.76 1.09 15.66 -16.76 -2 2
Table 1: The Hamiltonian and topological charge for the numerical solutions with μ2=1.5\mu^{2}=1.5 where ”Derrick” denotes the value (HL+HBoundary)/(ν2Hani+μ2Hpot)(H_{\rm L}+H_{\rm Boundary})/(\nu^{2}H_{\rm ani}+\mu^{2}H_{\rm pot}), which is expected to be 2-2 by the scaling argument. For ν2=1.5\nu^{2}=1.5, we used the exact solution (4.2) so that the ”Derrick” and topological charge for ν2=1.5\nu^{2}=1.5 are exact values.

Let us consider the asymptotic behavior of the solutions of the equations (4.8). Near the origin, the leading terms in the power series expansion are

FcFr,GcGr,F\approx c_{F}\leavevmode\nobreak\ r,\qquad G\approx c_{G}\leavevmode\nobreak\ r, (4.18)

where cFc_{F} and cGc_{G} are some constants implicitly depending on the coupling constants of the model. To see the behavior of solutions at large rr, we shift the profile functions as

F=π2,G=π2𝒢.F=\frac{\pi}{2}-{\cal F},\qquad G=\frac{\pi}{2}-{\cal G}. (4.19)

Then, one obtains linearized asymptotic equations on the functions {\cal F} and 𝒢{\cal G} of the forms

(′′+r4r2)+22(𝒢𝒢r)2(ν2+3μ2)=0,(𝒢′′+𝒢r𝒢r2)22(+2r)=0.\begin{split}&\left({\cal F}^{\prime\prime}+\frac{{\cal F}^{\prime}}{r}-\frac{4{\cal F}}{r^{2}}\right)+2\sqrt{2}\left({\cal G}^{\prime}-\frac{{\cal G}}{r}\right)-2\left(\nu^{2}+3\mu^{2}\right){\cal F}=0\leavevmode\nobreak\ ,\\ &\left({\cal G}^{\prime\prime}+\frac{{\cal G}^{\prime}}{r}-\frac{{\cal G}}{r^{2}}\right)-2\sqrt{2}\left({\cal F}^{\prime}+\frac{2{\cal F}}{r}\right)=0\leavevmode\nobreak\ .\end{split} (4.20)

Unfortunately, the equations (4.20) may not support an analytical solution. However, these equations imply that the asymptotic behavior of the profile functions is similar to that of the functions (4.2), by a replacement ν2κ\nu^{2}\kappa with (ν2+3μ2)/4(\nu^{2}+3\mu^{2})/4. Indeed, the asymptotic equations (4.20) depend on such a combination of the coupling constants, and there may exist an exact solution on the solvable line with the same character of asymptotic decay as the localized soliton solution of the equation (4.8).

To implement a numerical integration of the coupled system of ordinary differential equations (4.8), we introduce the normalized compact coordinate X(0,1]X\in(0,1] via

r=1XX.r=\frac{1-X}{X}. (4.21)

The integration was performed by the Newton-Raphson method with the mesh point NMESH=2000N_{\rm MESH}=2000.

In Fig. 4, we display some set of numerical solutions for different values of the coupling ν2\nu^{2} at μ2=1.5\mu^{2}=1.5 and their topological charge density 𝒬{\cal Q} defined through Q=2πr𝒬drQ=2\pi\int r{\cal Q}{\mathrm{d}}r. The solutions enjoy Derrick’s scaling relation and possess a good approximated value of the topological charge, as shown in Table 1. One observes that as the value of the coupling ν2\nu^{2} becomes relatively small, the function GG is delocalizing while the profile function FF is approaching its vacuum value everywhere in space except for the origin. This is an indication that any regular non-trivial solution does not exist ν2=0\nu^{2}=0.

4.2 Asymptotic behavior

Asymptotic interaction of solitons is related to the overlapping of the tails of the profile functions of well-separated single solitons [3]. Bounded multi-soliton configurations may exist if there is an attractive force between two isolated solitons.

Considering the above-mentioned soliton solutions of the gauged CP2CP^{2} NLσ\sigma-model, we have seen that the exact solution (4.2) has the same type of asymptotic decay as any solution of the general system (4.8). Therefore, it is enough to examine the asymptotic force between the solutions on the solvable line (4.2) to understand whether or not the Hamiltonian (3.30) supports multi-soliton solutions of higher topological degrees. Thus, without loss of generality, we can set μ2=ν2\mu^{2}=\nu^{2}.

Following the approach discussed in Ref. [3], let us consider a superposition of two exact solutions above. This superposition is no longer a solution of the Euler-Lagrange equation, except for in the limit of infinite separation, because there is a force acting on the solitons. The interaction energy of two solitons can be written as

Eint(R)=Hsp(R)2Hexact,E_{\rm int}(R)=H_{\rm sp}(R)-2H_{\rm exact}, (4.22)

where Hsp(R)H_{\rm sp}(R) is the energy of two BPS solitons separated by some large but finite distance RR from each other, and HexactH_{\rm exact} stands for the static energy of a single exact solution. Notice that the lower bound of the Hamiltonian (3.30) with μ2=ν2\mu^{2}=\nu^{2} is given

H=ν2Hν2=μ2=1+(1ν2)HD2π(12ν2)Q,H=\nu^{-2}H_{\nu^{2}=\mu^{2}=1}+(1-\nu^{2})H_{\rm D}\geq 2\pi(1-2\nu^{-2})Q, (4.23)

where the equality is enjoyed only by holomorphic solutions. Therefore, we immediately conclude

Hsp(R)2Hexact,H_{\rm sp}(R)\geq 2H_{\rm exact}, (4.24)

where the equality is satisfied only at the limit RR\to\infty. It follows that the interaction energy is always positive for finite separation, and the interaction is repulsive. Since the exact solution has the topological charge Q=2Q=2, it implies that there are no isolated soliton solutions with the topological charge Q4Q\geq 4 in this model. Note that, however, as the BPS solution (3.19) suggests, there can exist soliton solutions with an arbitrary negative charge, which are topological excited states on top of the homogeneous vacuum state.

5 Conclusion

In this paper, we have studied two-dimensional Skyrmions in the CP2CP^{2} NLσ\sigma-model with a Lifshitz invariant term which is an SU(3)SU(3) generalization of the DM term. We have shown that the SU(3)SU(3) tilted FM Heisenberg model turns out to be an SU(3)SU(3) gauged CP2CP^{2} NLσ\sigma-model in which the term linear in a background gauge field can be viewed as a Lifshitz invariant. We have found exact BPS-type solutions of the gauged CP2CP^{2} model in the presence of a potential term with a specific value of the coupling constant. The least energy configuration among the BPS solutions has been discussed. We have reduced the gauged CP2CP^{2} model to the (ungauged) CP2CP^{2} model with a Lifshitz invariant by choosing a background gauge field. In the reduced model, we have constructed an exact solution for a special combination of coupling constants called the solvable line and numerical solutions for a wider range of them.

For numerical study, we chose the background field, generating a potential term that can be interpreted as the quadratic Zeeman term or uniaxial anisotropic term. One can also choose a background field generating the Zeeman term; if the background field is chosen as A1=κλ7A_{1}=-\kappa\lambda_{7} and A2=κλ5A_{2}=\kappa\lambda_{5}, the associated potential term is proportional to S3\langle{S^{3}}\rangle. The Euler-Lagrange equation for the extended CP2CP^{2} model with this background field is not compatible with the axial symmetric ansatz (4.1). Therefore, a two-dimensional full simulation is required to obtain a solution with this background field. This problem, numerical simulation for non-axial symmetric solutions in the CP2CP^{2} model with a Lifshitz invariant, is left to future study. In addition, the construction of a CP2CP^{2} Skyrmion lattice is a challenging problem. The physical interpretation of the Lifshitz invariants is also an important future task. The microscopic derivation of the SU(3)SU(3) tilted Heisenberg model [21] may enable us to understand the physical interpretation and physical situation where the Lifshitz invariant appears. Other future work would be the extension of the present study to the SU(3)SU(3) antiferromagnetic Heisenberg model where soliton/sphaleron solutions can be constructed [41, 42, 43].

We restricted our analysis on the case that the additional potential term μ2Hpot\mu^{2}H_{\rm pot} is balanced or dominant against the anisotropic potential term ν2Hani\nu^{2}H_{\rm ani}, i.e., ν2μ2\nu^{2}\leq\mu^{2}. We expect that a classical phase transition occurs outside of the condition, and it causes instability of the solution. At the moment, the phase structure of the model (3.30) is not clear, and we will discuss it in our subsequent work.

Moreover, it has been reported that in some limit of a three-component Ginzburg-Landau model [44, 45], and of a three-component Gross-Pitaevskii model [46, 47], their vortex solutions can be well-described by planar CP2CP^{2} Skyrmions. We believe that our result provides a hint to introduce a Lifshitz invariant to the models, and that our solutions find applications not only in SU(3)SU(3) spin systems but also in superconductors and Bose-Einstein condensates described by the extended models, including the Lifshitz invariant.

Acknowledgments
This work was supported by JSPS KAKENHI Grant Nos. JP17K14352, JP20K14411, and JSPS Grant-in-Aid for Scientific Research on Innovative Areas “Quantum Liquid Crystals” (KAKENHI Grant No. JP20H05154). Ya.S. gratefully acknowledges support by the Ministry of Education of Russian Federation, project FEWF-2020-0003. Y. Amari would like to thank Tokyo University of Science for its kind hospitality.

References

  • [1] T. Skyrme, “A Nonlinear field theory”, Proc. Roy. Soc. Lond. A 260, 127–138 (1961) .
  • [2] T. Skyrme, “A Unified Field Theory of Mesons and Baryons”, Nucl. Phys. 31, 556–569 (1962) .
  • [3] N. Manton, and P. Sutcliffe, “Topological solitons”, Cambridge Monographs on Mathematical Physics, Cambridge University Press, 2004.
  • [4] A. Bogolubskaya, and I. Bogolubsky, “Stationary Topological Solitons in the Two-dimensional Anisotropic Heisenberg Model With a Skyrme Term”, Phys. Lett. A 136, 485–488 (1989) .
  • [5] A. Bogolyubskaya, and I. Bogolyubsky, “ON STATIONARY TOPOLOGICAL SOLITONS IN TWO-DIMENSIONAL ANISOTROPIC HEISENBERG MODEL”, Lett. Math. Phys. 19, 171–177 (1990) .
  • [6] R. A. Leese, M. Peyrard, and W. J. Zakrzewski, “Soliton Scatterings in Some Relativistic Models in (2+1)-dimensions”, Nonlinearity 3, 773–808 (1990) .
  • [7] A. M. Polyakov, and A. Belavin, “Metastable States of Two-Dimensional Isotropic Ferromagnets”, JETP Lett. 22, 245–248 (1975) .
  • [8] A. N. Bogdanov, and D. Yablonskii, “Thermodynamically stable “vortices” in magnetically ordered crystals. The mixed state of magnets”, Zh. Eksp. Teor. Fiz 95 (1), 178 (1989) .
  • [9] A. Neubauer, C. Pfleiderer, B. Binz, A. Rosch, R. Ritz, P. Niklowitz, and P. Böni, “Topological Hall effect in the A phase of MnSi”, Phys. Rev. Lett. 102 (18), 186602 (2009) .
  • [10] N. Nagaosa, and Y. Tokura, “Topological properties and dynamics of magnetic skyrmions”, Nat. Nanotechnol. 8 (12), 899–911 (2013) .
  • [11] A. Leonov, I. Dragunov, U. Rößler, and A. Bogdanov, “Theory of skyrmion states in liquid crystals”, Phys. Rev. E 90 (4), 042502 (2014) .
  • [12] I. I. Smalyukh, Y. Lansac, N. A. Clark, and R. P. Trivedi, “Three-dimensional structure and multistable optical switching of triple-twisted particle-like excitations in anisotropic fluids”, Nat. Mater. 9 (2), 139–145 (2010) .
  • [13] 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 (5916), 915–919 (2009) .
  • [14] X. Yu, Y. Onose, N. Kanazawa, J. Park, J. Han, Y. Matsui, N. Nagaosa, and Y. Tokura, “Real-space observation of a two-dimensional skyrmion crystal”, Nature 465 (7300), 901–904 (2010) .
  • [15] S. Heinze, K. Von Bergmann, M. Menzel, J. Brede, A. Kubetzka, R. Wiesendanger, G. Bihlmayer, and S. Blügel, “Spontaneous atomic-scale magnetic skyrmion lattice in two dimensions”, Nat. Phys. 7 (9), 713–718 (2011) .
  • [16] C. Back, V. Cros, H. Ebert, K. Everschor-Sitte, A. Fert, M. Garst, T. 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 J. Zang, “The 2020 skyrmionics roadmap”, J. Phys. D: Appl. Phys. 53 (36), 363001 (2020) .
  • [17] I. Dzyaloshinsky, “A thermodynamic theory of “weak” ferromagnetism of antiferromagnetics”, J. Phys. Chem. Solids 4 (4), 241–255 (1958) .
  • [18] T. Moriya, “Anisotropic superexchange interaction and weak ferromagnetism”, Physical review 120 (1), 91 (1960) .
  • [19] Y.-Q. Li, Y.-H. Liu, and Y. Zhou, “General spin-order theory via gauge Landau-Lifshitz equation”, Phys. Rev. B 84 (20), 205123 (2011) .
  • [20] B. J. Schroers, “Gauged Sigma Models and Magnetic Skyrmions”, SciPost Phys. 7 (3), 030 (2019) .
  • [21] S. Zhu, Y.-Q. Li, and C. D. Batista, “Spin-orbit coupling and electronic charge effects in Mott insulators”, Phys. Rev. B 90 (19), 195107 (2014) .
  • [22] A. Sparavigna, “Role of Lifshitz invariants in liquid crystals”, Materials 2 (2), 674–698 (2009) .
  • [23] P. Yudin, and A. Tagantsev, “Fundamentals of flexoelectricity in solids”, Nanotechnology 24 (43), 432001 (2013) .
  • [24] B. Barton-Singer, C. Ross, and B. J. Schroers, “Magnetic Skyrmions at Critical Coupling”, Commun. Math. Phys. 375 (3), 2259–2280 (2020) .
  • [25] C. Ross, N. Sakai, and M. Nitta, “Skyrmion Interactions and Lattices in Solvable Chiral Magnets”arXiv:2003.07147.
  • [26] V. Golo, and A. Perelomov, “Solution of the Duality Equations for the Two-Dimensional SU(N) Invariant Chiral Model”, Phys. Lett. B 79, 112–113 (1978) .
  • [27] A. D’Adda, M. Luscher, and P. Di Vecchia, “A 1/n Expandable Series of Nonlinear Sigma Models with Instantons”, Nucl. Phys. B 146, 63–76 (1978) .
  • [28] A. Din, and W. Zakrzewski, “General Classical Solutions in the CP(n-1) Model”, Nucl. Phys. B 174, 397–406 (1980) .
  • [29] L. Ferreira, and P. Klimas, “Exact vortex solutions in a CPNCP^{N} Skyrme-Faddeev type model”, JHEP 10, 008 (2010) .
  • [30] Y. Amari, P. Klimas, N. Sawado, and Y. Tamaki, “Potentials and the vortex solutions in the CPNCP^{N} Skyrme-Faddeev model”, Phys. Rev. D 92 (4), 045007 (2015) .
  • [31] B. Ivanov, R. Khymyn, and A. Kolezhuk, “Pairing of Solitons in Two-Dimensional S=1 Magnets”, Phys. Rev. Lett. 100 (4), 047203 (2008) .
  • [32] A. Smerald, and N. Shannon, “Theory of spin excitations in a quantum spin-nematic state”, Phys. Rev. B 88, 184430 (2013) .
  • [33] R. Hernandez, and E. Lopez, “The SU(3) spin chain sigma model and string theory”, JHEP 2004 (04), 052 (2004) .
  • [34] M. Greiter, S. Rachel, and D. Schuricht, “Exact results for SU(3) spin chains: Trimer states, valence bond solids, and their parent Hamiltonians”, Phys. Rev. B 75 (6), 060401 (2007) .
  • [35] K. Penc, and A. M. Läuchli, “Spin nematic phases in quantum spin systems”, in: Introduction to Frustrated Magnetism, Springer, 2011, pp. 331–362.
  • [36] B. A. Ivanov, and A. K. Kolezhuk, “Effective field theory for the S=1quantum nematic”, Phys. Rev. B 68 (5).
  • [37] K.-I. Kondo, S. Kato, A. Shibata, and T. Shinohara, “Quark confinement: Dual superconductor picture based on a non-Abelian Stokes theorem and reformulations of Yang–Mills theory”, Phys. Rept. 579, 1–226 (2015) .
  • [38] W. J. Zakrzewski, “Low dimensional sigma models”, Hilger, 1989.
  • [39] B. Piette, B. Schroers, and W. Zakrzewski, “Multi-solitons in a two-dimensional Skyrme model”, Z. Phys. C 65, 165–174 (1995) .
  • [40] L. Döring, and C. Melcher, “Compactness results for static and dynamic chiral skyrmions near the conformal limit”, Calc. Var. Partial Differ. Equ. 56 (3), 60 (2017) .
  • [41] D. Bykov, “Classical solutions of a flag manifold σ\sigma-model”, Nucl. Phys. B 902, 292–301 (2016) .
  • [42] H. T. Ueda, Y. Akagi, and N. Shannon, “Quantum solitons with emergent interactions in a model of cold atoms on the triangular lattice”, Phys. Rev. A 93, 021606(R) (2016) .
  • [43] Y. Amari, and N. Sawado, “BPS sphalerons in the F2{F}_{2} nonlinear sigma model”, Phys. Rev. D 97, 065012 (2018) .
  • [44] J. Garaud, J. Carlstrom, and E. Babaev, “Topological solitons in three-band superconductors with broken time reversal symmetry”, Phys. Rev. Lett. 107, 197001 (2011) .
  • [45] J. Garaud, J. Carlström, E. Babaev, and M. Speight, “Chiral CP2CP^{2} skyrmions in three-band superconductors”, Phys. Rev. B 87 (1), 014507 (2013) .
  • [46] M. Eto, and M. Nitta, “Vortex trimer in three-component Bose-Einstein condensates”, Phys. Rev. A 85, 053645 (2012) .
  • [47] M. Eto, and M. Nitta, “Vortex graphs as N-omers and CP(N-1) Skyrmions in N-component Bose-Einstein condensates”, EPL 103 (6), 60006 (2013) .