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

Anomalies of kagome antiferromagnets on magnetization plateaus

Shunsuke C. Furuya Department of Physics, Ibaraki University, Mito, Ibaraki 310-8512, Japan Condensed Matter Theory Laboratory, RIKEN, Wako, Saitama 351-0198, Japan    Yusuke Horinouchi RIKEN Center for Emergent Matter Science (CEMS), Wako, Saitama 351-0198, Japan    Tsutomu Momoi Condensed Matter Theory Laboratory, RIKEN, Wako, Saitama 351-0198, Japan RIKEN Center for Emergent Matter Science (CEMS), Wako, Saitama 351-0198, Japan
Abstract

We discuss the ground-state degeneracy of spin-1/21/2 kagome-lattice quantum antiferromagnets on magnetization plateaus by employing two complementary methods: the adiabatic flux insertion in closed boundary conditions and a ’t Hooft anomaly argument on inherent symmetries in a quasi-one-dimensional limit. The flux insertion with a tilted boundary condition restricts the lower bound of the ground-state degeneracy on 1/91/9, 1/31/3, 5/95/9, and 7/97/9 magnetization plateaus under the U(1)\mathrm{U(1)} spin-rotation and the translation symmetries: 33, 11, 33, and 33, respectively. This result motivates us further to develop an anomaly interpretation of the 1/31/3 plateau. Taking advantage of the insensitivity of anomalies to spatial anisotropies, we examine the existence of the unique gapped ground state on the 1/31/3 plateau from a quasi-one-dimensional viewpoint. In the quasi-one-dimensional limit, kagome antiferromagnets are reduced to weakly coupled three-leg spin tubes. Here, we point out the following anomaly description of the 1/31/3 plateau. While a simple S=1/2S=1/2 three-leg spin tube cannot have the unique gapped ground state on the 1/31/3 plateau because of an anomaly between a 3×3\mathbb{Z}_{3}\times\mathbb{Z}_{3} symmetry and the translation symmetry at the 1/31/3 filling, the kagome antiferromagnet breaks explicitly one of the 3\mathbb{Z}_{3} symmetries related to a 3\mathbb{Z}_{3} cyclic transformation of spins in the unit cell. Hence the kagome antiferromagnet can have the unique gapped ground state on the 1/31/3 plateau.

I Introduction

The last decade has witnessed the bloom of topological classification of gapped quantum phases Schnyder et al. (2008); Chen et al. (2011, 2013). Topology, in a broad sense, has made possible the universal classification of gapped topological phases outside the reach of the order-parameter paradigm. By contrast, much less is known about the classification of gapless quantum phases protected by symmetries Furuya and Oshikawa (2017); Yao et al. (2019). One milestone in the classification of gapless quantum phases is the Lieb-Schultz-Mattis (LSM) theorem originally proven for spin-1/21/2 XXZ chains stating that they cannot have a unique gapped ground state Lieb et al. (1961). The LSM theorem is related to the U(1)\mathrm{U(1)} flux insertion to quantum many-body systems, where the filling of particles plays an essential role. A fractional filling is necessary, not sufficient, though, for excluding the possibility of the unique gapped ground state in the periodic boundary condition. On the basis of the filling argument, the LSM theorem was extended to quantum many-body systems in two or higher dimensions Oshikawa (2000); Hastings (2004).

In quantum spin systems, the LSM theorem is immediately related to magnetization plateaus, as exemplified by the Oshikawa-Yamanaka-Affleck (OYA) condition Oshikawa et al. (1997). The OYA condition successfully explains the possible emergence of plateaus and their ground-state degeneracy in (quasi-)one-dimensional quantum spin systems with the filling argument in analogy with that of the LSM theorem. Later the OYA condition was extended to two or higher dimensional quantum spin systems by Oshikawa himself Oshikawa (2000). Interestingly, the adiabatic flux insertion argument in Ref. Oshikawa (2000) enables us to deal with the OYA condition on equal footing with the LSM theorem. However, the flux-insertion argument in Ref. Oshikawa (2000) leads to a system-size dependent ground-state degeneracy, which makes the thermodynamic limit ambiguous. The system-size dependence originates from the periodic boundary condition in one direction, where the dd-dimensional system can be regarded as a product S1×Md1S^{1}\times M_{d-1} of a ring S1S^{1} and a (d1)(d-1)-dimensional “cross-section” Md1M_{d-1}. When the argument of Ref. Oshikawa (2000) is applied to quantum spin systems, the number of spins on Md1M_{d-1} must not be any integral multiple of the filling fraction to exclude the possibility of the unique gapped ground state, leading to the inconvenient system-size dependence of the ground-state degeneracy. Still, the flux insertion argument is advantageous even today for its intuitive picture and its direct connection to ’t Hooft anomalies Cho et al. (2017); Yao and Oshikawa (2020); Furuya and Horinouchi (2019), which would possibly be a counterpart of the topology in the classification of gapless phases.

The importance of anomalies in condensed matter physics has been well recognized, for example, in the context of symmetry-protected surface states of topological phases Ryu and Zhang (2012); Sule et al. (2013). It was also pointed out that the LSM-type argument of dd-dimensional bulk phases is related to surface anomalies of (dd+1)-dimensional weak SPT phases Cheng et al. (2016); Jian et al. (2018); Thorngren and Else (2018). However, such an anomaly description of magnetization plateaus in two or higher dimensions is yet to be established. Unambiguous flux insertion argument of magnetization plateaus will offer the first step toward their anomaly description.

Recently, the authors Furuya and Horinouchi (2019) derived the OYA condition in frustrated quantum magnets on the checkerboard lattice by adapting the U(1)\mathrm{U(1)} flux insertion argument avoiding the problem in the periodic boundary condition. The key idea in this approach is to use a closed boundary condition accompanied by a spatial twist that preserves the checkerboard symmetries instead of the simple periodic one. This argument is based on a simple assumption that we can choose arbitrary boundary conditions if they keep the symmetries in question. Reference Furuya and Horinouchi (2019) provides us with a viewpoint that an appropriate symmetric boundary condition gives shape to a relation between the magnetization plateau and the anomaly. However, the spatially twisted boundary condition stands on characteristics of the checkerboard and is thus inapplicable to a broad class of geometrically frustrated quantum spin systems with triangle-based lattices. Most importantly, it is inapplicable to kagome-lattice quantum antiferromagnets, which are famous for their fertility of magnetization plateaus Hida (2001); Schulenburg et al. (2002); Nishimoto et al. (2013); Capponi et al. (2013); Nakano and Sakai (2014, 2015); Picot et al. (2016).

In this paper, we revisit magnetization plateaus of S=1/2S=1/2 quantum spin systems on the kagome lattice and investigate them from the viewpoint of the OYA condition in a symmetric closed boundary condition such as a tilted boundary condition Yao and Oshikawa (2020). Some closed boundary conditions, including the tilted one, enable us to regard kagome quantum antiferromagnets as one-dimensional quantum antiferromagnets with long-range interactions. This viewpoint gives us some insight into the 1/31/3 magnetization plateau. If we consider a simple three-leg spin tube with short-range interactions Sakai et al. (2008); Okunishi et al. (2012); Plat et al. (2012); Yonaga and Shibata (2015), we find a ’t Hooft anomaly in a (1+1)-dimensional quantum field theory as an effective description of the S=1/2S=1/2 three-leg spin tube on the 1/31/3 magnetization plateau. The ’t Hooft anomaly of this field theory excludes the possibility of the unique gapped ground state on the 1/31/3 plateau of the spin tube under a certain on-site symmetry and the translation symmetry. However, when considered as a three-leg spin tube the kagome antiferromagnet has a long-range interaction, which hinders the direct inheritance of the anomaly of the simple three-leg spin tube.

To discuss how the anomaly of the three-leg spin tube is broken in the kagome antiferromagnet on the 1/31/3 plateau, we take another approach from a quasi-one-dimensional limit. We can see the absence of the anomaly more explicitly by introducing a spatial anisotropy to the Hamiltonian because the anisotropy preserves the required symmetries. In the quasi-one-dimensional limit, the kagome antiferromagnet turns into a weakly coupled three-leg spin tubes with short-range interactions. Here, we can find that the kagome lattice’s symmetry explicitly breaks one of these symmetries involved with the anomaly unless the Hamiltonian is fine-tuned. The quasi-one-dimensional viewpoint tells us that the unique gapped ground state on the 1/31/3 plateau is permitted by the kagome geometry thanks to the resolution of the anomaly.

The paper is organized as follows. We define a symmetric closed boundary condition called the tilted boundary condition Yao and Oshikawa (2020) in Sec. II and argue the flux insertion with this boundary condition. In Sec. III, we develop an anomaly argument on magnetization plateaus of kagome antiferromagnets as coupled spin tubes. After these sections, we summarize this paper in Sec. IV.

II Tilted boundary condition

Refer to caption
Figure 1: The kagome lattice. The location of the unit cell is specified by a two-dimensional vector 𝑹=n1𝒆1+n2𝒆2\bm{R}=n_{1}\bm{e}_{1}+n_{2}\bm{e}_{2} with n1,n2n_{1},n_{2}\in\mathbb{Z} and two-dimensional unit vectors 𝒆1\bm{e}_{1} and 𝒆2\bm{e}_{2}.

II.1 Assumption

Let us begin by clarifying an assumption about the effect of boundary conditions that we rely on in this paper. The assumption is that if a system has the unique gapped ground state in the periodic boundary condition under certain symmetries, it also does in any other closed boundary conditions that respect the symmetries. This assumption is natural since the closed boundary condition is just an artificial condition of theories to minimize the boundary effect. The bulk properties should be independent of a specific choice of symmetric boundary conditions.

Taking the contraposition of the assumption, we can state that if the unique gapped ground state is forbidden in one symmetric closed boundary condition, it is also forbidden in the periodic boundary condition. This statement motivates us to search for an appropriate symmetric closed boundary condition that clarifies a universal constraint forbidding the unique gapped ground state. Among such symmetric closed boundary conditions is the tilted boundary condition Yao and Oshikawa (2020) that we consider in this section.

II.2 Symmetries

Before defining the tilted boundary condition, we define the model and its symmetries that we use in this paper. We discuss spin-1/21/2 kagome antiferromagnets with translation symmetries and the U(1)\mathrm{U(1)} spin-rotation symmetry. We assume translation symmetries of a unit cell by one unit in the 𝒆1\bm{e}_{1} and the 𝒆2\bm{e}_{2} directions of Fig. 1, where the smallest upward triangle is considered as the unit cell. All the results derived in this section also hold for spin-SS models with S>1/2S>1/2.

Let us specify the location of the unit cell by a two-dimensional vector 𝑹=n1𝒆1+n2𝒆2\bm{R}=n_{1}\bm{e}_{1}+n_{2}\bm{e}_{2} with 𝒆1=(1,0)\bm{e}_{1}=(1,0), 𝒆2=(12,32)\bm{e}_{2}=(-\frac{1}{2},\frac{\sqrt{3}}{2}) (Fig. 1), and n1,n2n_{1},n_{2}\in\mathbb{Z}. We employed a unit of the lattice spacing a=1a=1. Accordingly, the spin operator can be denoted as 𝑺μ(n1,n2)\bm{S}_{\mu}(n_{1},n_{2}). In the figures, the indices μ=1,2\mu=1,2, and 33 are distinguished visually by red, white, and blue circles, respectively. Then the Hamiltonian of the spin-SS Heisenberg antiferromagnet is represented as

\displaystyle\mathcal{H} =Jn1,n2[𝑺1(n1,n2)𝑺2(n1,n2)\displaystyle=J\sum_{n_{1},n_{2}}[\bm{S}_{1}(n_{1},n_{2})\cdot\bm{S}_{2}(n_{1},n_{2})
+𝑺2(n1,n2)𝑺3(n1,n2)+𝑺3(n1,n2)𝑺1(n1,n2)]\displaystyle+\bm{S}_{2}(n_{1},n_{2})\cdot\bm{S}_{3}(n_{1},n_{2})+\bm{S}_{3}(n_{1},n_{2})\cdot\bm{S}_{1}(n_{1},n_{2})]
+Jn1,n2[𝑺3(n1,n2)𝑺2(n1+1,n2)\displaystyle+J\sum_{n_{1},n_{2}}[\bm{S}_{3}(n_{1},n_{2})\cdot\bm{S}_{2}(n_{1}+1,n_{2})
+𝑺1(n1,n2){𝑺3(n1,n2+1)+𝑺2(n1+1,n2+1)}]\displaystyle+\bm{S}_{1}(n_{1},n_{2})\cdot\{\bm{S}_{3}(n_{1},n_{2}+1)+\bm{S}_{2}(n_{1}+1,n_{2}+1)\}]
hn1,n2,μSμz(n1,n2),\displaystyle-h\sum_{n_{1},n_{2},\mu}S_{\mu}^{z}(n_{1},n_{2}), (1)

with J>0J>0 and h0h\geq 0. This uniform kagome Heisenberg antiferromagnet is merely a specific example that satisfies the U(1)\mathrm{U(1)} and the T1T_{1} and T2T_{2} translation symmetries, where TnT_{n} represents the translation by the one unit in the 𝒆n\bm{e}_{n} direction:

T1𝑺μ(n1,n2)T11\displaystyle T_{1}\bm{S}_{\mu}(n_{1},n_{2})T_{1}^{-1} =𝑺μ(n1+1,n2),\displaystyle=\bm{S}_{\mu}(n_{1}+1,n_{2}), (2)
T2𝑺μ(n1,n2)T21\displaystyle T_{2}\bm{S}_{\mu}(n_{1},n_{2})T_{2}^{-1} =𝑺μ(n1,n2+1).\displaystyle=\bm{S}_{\mu}(n_{1},n_{2}+1). (3)

We can add any symmetric interactions to the Hamiltonian whenever we want.

To make the translation symmetries well-defined, we need to specify the boundary condition. The Hamiltonian (1) possesses the U(1)\mathrm{U}(1) spin-rotation symmetry and translation symmetries if the periodic boundary condition is imposed on the xx and the yy directions (Fig. 1). When we adopt the flux insertion argument Oshikawa (2000) to this system with the periodic boundary condition, we face the previously mentioned problem of the ambiguous thermodynamic limit Pal et al. (2020). In the case of the dd-dimensional hyper cubic lattice, this problem can be resolved in Ref. Yao and Oshikawa (2020) by another closed boundary condition that respects the symmetries, the tilted boundary condition.

II.3 Tilted boundary condition

Refer to caption
Figure 2: A rhombus finite-size cluster of the kagome lattice with 3636 sites in the tilted boundary condition. The label on each upward triangle expresses the one-dimensional coordinate r1r_{1} of the unit cell in the tilted boundary condition [Eq. (8)].
Refer to caption
Figure 3: A schematic picture of the tilted boundary condition in a rombus finite-size cluster of the kagome lattice. Details of the kagome lattice are omitted, and the seam ABCD of the tilted boundary condition on which the boundary condition is imposed is shown. In the tilted boundary condition, a point 22 on the right seam is identified with another point 22^{\prime} on the left seam, where the latter is dislocated from the point 11 by the unit vector 𝒆2\bm{e}_{2}.

We define the tilted boundary condition on the kagome lattice. The kagome lattice is a non-Bravais lattice whose unit cell consists of three sites in the periodic boundary condition. Likewise, the unit cell contains three sites in the tilted boundary condition. The one-unit translation operator T1T_{1} in the 𝒆1\bm{e}_{1} direction acts on the spin operator 𝑺μ(n1,n2)\bm{S}_{\mu}(n_{1},n_{2}) (μ=1,2,3)\mu=1,2,3) as Eq. (2). To define the tilted boundary condition, we first consider a finite-size cluster of the kagome lattice of the rhombic shape (Figs. 2 and 3) and next take the thermodynamic limit by making the system size infinite. The rhombic cluster breaks some of the symmetries that the infinite-size kagome lattice possesses, for example, a C6C_{6} rotation symmetry whose rotation axis pierces the center of a hexagon. The rhombic cluster is chosen because this paper is focused on an anomaly between the U(1)\mathrm{U(1)} spin-rotation symmetry and the translation symmetry. The rhombic shape excludes the effects of the C6C_{6} and other symmetries on the ground-state degeneracy.

Let us define the origin 𝑹=0\bm{R}=0 as the left bottom corner of the rhombic finite-size cluster (the center of a triangle with a label 1 in Fig. 2). Suppose 0n1<N10\leq n_{1}<N_{1} and 0n2<N20\leq n_{2}<N_{2} for positive integers N1,N2N_{1},N_{2}. On the rhombic finite-size cluster, the tilted boundary condition is defined as

T1𝑺μ(N11,n2)T11\displaystyle T_{1}\bm{S}_{\mu}\bigl{(}N_{1}-1,\,n_{2}\bigr{)}T_{1}^{-1} =𝑺μ(0,n2+1),\displaystyle=\bm{S}_{\mu}\bigl{(}0,\,n_{2}+1\bigr{)}, (4)

for n2[0,N2)n_{2}\in[0,N_{2}) and

T1𝑺μ(N11,N21)T11\displaystyle T_{1}\bm{S}_{\mu}\bigl{(}N_{1}-1,N_{2}-1\bigr{)}T_{1}^{-1} =𝑺μ(0,0).\displaystyle=\bm{S}_{\mu}\bigl{(}0,0\bigr{)}. (5)

When we reach the right seam of the system, we reenter the system from the left seam with the dislocation by the unit vector 𝒆2\bm{e}_{2} (Fig. 3). It immediately follows from Eqs. (4) and (5) that the T2T_{2} translation symmetry, that is, the translation symmetry in the 𝒆2\bm{e}_{2} direction, depends on the T1T_{1} one in the tilted boundary condition for a relation T2=(T1)N1T_{2}=(T_{1})^{N_{1}}. The aspect ratio N2/N1N_{2}/N_{1} of the rhombus can be arbitrary. The total number of sites is given by

V\displaystyle V =3N1N2\displaystyle=3N_{1}N_{2} (6)

We can sweep all the upward triangles on the kagome lattice one dimensionally by applying T1T_{1} of Eqs. (2), (4), and (5) repeatedly. This path allows us to relabel the spin 𝑺μ(n1,n2)\bm{S}_{\mu}(n_{1},n_{2}) with a one-dimensional coordinate along that path as

𝑺μ(n1,n2)=𝑺r1,μ,\displaystyle\bm{S}_{\mu}(n_{1},n_{2})=\bm{S}_{r_{1},\mu}, (7)

where the one-dimensional coordinate r1[1,V/3]r_{1}\in[1,V/3] of the unit cell is related to 𝑹=n1𝒆1+n2𝒆2\bm{R}=n_{1}\bm{e}_{1}+n_{2}\bm{e}_{2} through

r1\displaystyle r_{1} =1+n1+n2N1\displaystyle=1+n_{1}+n_{2}N_{1} (8)

The Hamiltonian (1) in the tilted boundary condition is T1T_{1}-symmetric, that is, [,T1]=0[\mathcal{H},T_{1}]=0.

II.4 Flux insertion

m/Sm/S 0 1/91/9 1/31/3 5/95/9 7/97/9
θ/2π\theta/2\pi 3/23/2 4/34/3 11 2/32/3 1/31/3
dmd_{\rm m} 22 33 11 33 33
Table 1: The angle θ\theta of Eq. (12) for zero magnetization and for fractional magnetizations m/S=(2n1)/9m/S=(2n-1)/9 with n=1,2,3,4n=1,2,3,4 is listed for S=1/2S=1/2. The third row refers to the lower bound dmd_{\rm m} of the ground-state degeneracy of the S=1/2S=1/2 kagome antiferromagnet on those magnetization plateaus.

We insert the flux adiabatically into the Hamiltonian (1) by replacing transverse exchange interactions,

Sr1,μ+Sr1,μ\displaystyle S_{r_{1},\mu}^{+}S_{r^{\prime}_{1},\mu^{\prime}}^{-} +H.c.\displaystyle+\mathrm{H.c.}
ei(r1r1)ϕ/VSr1,μ+Sr1,μ+H.c.,\displaystyle\to e^{i(r_{1}-r^{\prime}_{1})\phi/V}S_{r_{1},\mu}^{+}S_{r^{\prime}_{1},\mu^{\prime}}^{-}+\mathrm{H.c.}, (9)

and increase the flux amount ϕ\phi\in\mathbb{R} slowly from zero to the unit amount, 2π2\pi. The 2π2\pi flux can be absorbed by a U(1)\mathrm{U(1)} large gauge transformation,

U\displaystyle U =exp(i2πV/3r1=1V/3r1n(r1)),\displaystyle=\exp\biggl{(}i\frac{2\pi}{V/3}\sum_{r_{1}=1}^{V/3}r_{1}n(r_{1})\biggr{)}, (10)

where n(r1)n(r_{1}) is a number density of magnons at the unit cell r1r_{1}

n(r1)=3SSr1z,\displaystyle n(r_{1})=3S-S_{r_{1}}^{z}, (11)

with Sr1z=μ=13Sr1,μzS_{r_{1}}^{z}=\sum_{\mu=1}^{3}S_{r_{1},\mu}^{z}. The Hamiltonian with the flux, whatever the amount of the inserted flux is, keeps the T1T_{1} symmetry and the global U(1)\mathrm{U(1)} spin-rotation symmetry at the same time. On the other hand, the translation T1T_{1} and the U(1)\mathrm{U(1)} large gauge transformation UU satisfy a relation,

T1UT11\displaystyle T_{1}UT_{1}^{-1} =Ueiθ,\displaystyle=Ue^{i\theta}, (12)

with a nontrivial angle,

θ2π=3(Sm),\displaystyle\frac{\theta}{2\pi}=3(S-m), (13)

with the magnetization density per a site, m=r1Sr1z/(V/3)m=\sum_{r_{1}}S_{r_{1}}^{z}/(V/3). When θ\theta is trivial (i.e. θ=0mod2π\theta=0\mod 2\pi), the two operators T1T_{1} and UU are commutative with each other. Then nothing prevents the ground state from being unique and gapped. When the system has nontrivial θ0mod2π\theta\not=0\mod 2\pi, it is forbidden to have any unique gapped ground state as follows Oshikawa (2000). Let us denote a ground state of the kagome antiferromagnet without the flux as |Ψ0\ket{\Psi_{0}}. If the adiabatic flux insertion deforms smoothly |Ψ0\ket{\Psi_{0}} to |Ψ0\ket{\Psi^{\prime}_{0}}, the latter is a ground state of the kagome antiferromagnet with the unit flux. The large-gauge-transformed U|Ψ0U\ket{\Psi^{\prime}_{0}} is a ground state of the kagome antiferromagnet without the flux. Note that both |Ψ0\ket{\Psi_{0}} and |Ψ0\ket{\Psi^{\prime}_{0}} have the same eigenvalue of T1T_{1} because the adiabatic flux insertion is compatible with the translation symmetry. If θ\theta satisfies θ0mod2π\theta\not=0\mod 2\pi, U|Ψ0U\ket{\Psi^{\prime}_{0}} is orthogonal to |Ψ0\ket{\Psi_{0}} thanks to Eqs. (12) and (13), in other words, U|Ψ0U\ket{\Psi^{\prime}_{0}} is a degenerate ground state or a gapless excited state of the kagome antiferromagnet without the flux.

Because the angle (13) depends on neither N1N_{1} nor N2N_{2}, we can take the thermodynamic limit, V+V\to+\infty, without any ambiguity. This well-defined thermodynamic limit is a great advantage of the tilted boundary condition over the periodic boundary condition Oshikawa (2000); Pal et al. (2020).

In the absence of the magnetic field, the model (1) possesses the time-reversal symmetry that imposes m=0m=0 unless the spontaneous ferromagnetic order is generated, which is unlikely. Therefore, θ/2π=3S\theta/2\pi=3S follows at zero magnetic field. When S+1/2S\in\mathbb{Z}+1/2, the ground state of the spin-SS kagome Heisenberg antiferromagnet has either the gapless ground state or (at least) doubly degenerate gapped ground states. The former is consistent with the U(1)\mathrm{U}(1) Dirac spin liquid scenario Ran et al. (2007); Iqbal et al. (2011); He et al. (2017) and the latter is consistent with the gapped 2\mathbb{Z}_{2} spin liquid Yan et al. (2011); Depenbrock et al. (2012); Jiang et al. (2012). The degeneracy predicted by the relation (12) refers only to that by the intrinsic anomaly between the U(1)\mathrm{U}(1) spin-rotation symmetry and the translation symmetry for a fixed filling. The ground state can, in principle, be more degenerate than Eq. (12) tells. Therefore, the ground-state degeneracy predicted by Eq. (12), which we denote as dmd_{\rm m}, gives the minimum possible value of the actual ground-state degeneracy.

In the presence of the magnetic field, the spin-1/21/2 Heisenberg antiferromagnet on the kagome lattice (1) is believed to have 1/91/9, 1/31/3, 5/95/9, and 7/97/9 magnetization plateaus. The angles of Eq. (13) for those fractions of the magnetization are listed in Table 1. When θ=2πp/q\theta=2\pi p/q with coprime integers pp and qq, the following qq states, |Ψ0\ket{\Psi_{0}}, U|Ψ0U\ket{\Psi^{\prime}_{0}}, U2|Ψ0,U^{2}\ket{\Psi^{\prime}_{0}},\cdots, and Uq1|Ψ0U^{q-1}\ket{\Psi^{\prime}_{0}} have the same eigenenergy in the thermodynamic limit but have different eigenvalues of T1T_{1}. If the ground state is gapped, these qq states are qq-fold degenerate gapped ground states. Given this possible minimum ground-state degeneracy, of particular interest is the 1/31/3 plateau where the unique gapped ground state is allowed. In fact, Ref. Parameswaran et al. (2013) constructed the unique gapped ground state of a model explicitly on the 1/31/3 plateau without breaking any symmetry of the kagome lattice. This is consistent with the relation (12). However, it remains obscure what allows for the unique gapped ground state on the 1/31/3 plateau because the condition (12) with the angle (13) only tells that the minimum number of the ground-state degeneracy allowed by the translation and the U(1)\mathrm{U}(1) spin-rotation symmetries is 11. In the subsequent section, we propose one interpretation of the unique gapped ground state’s appearance possible on the 1/31/3 plateau from the viewpoint of an anomaly in a quasi-one-dimensional limit.

III Anomaly and spatial anisotropy

In this section, we employ the periodic boundary condition with avoiding the known problem of the size-dependent ground-state degeneracy. The key idea is the insensitivity of a ’t Hooft anomaly to spatial anisotropies.

III.1 insensitivity of anomalies to spatial anisotropy

The most significant advantage of the tilted boundary condition is that it reduces the number of spins in the cross-section Md1M_{d-1} down to O(1)O(1). In the periodic boundary condition, it is O(V(d1)/d)O(V^{(d-1)/d}). The system with O(1)O(1) spins on each cross-section seems like a one-dimensional system. In fact, we can view the kagome antiferromagnet in the tilted boundary condition as a one-dimensional quantum spin system in the periodic boundary condition where the upward triangle is one-dimensionally aligned. The flux insertion argument is independent of whether the system is viewed as a dd-dimensional one or a one-dimensional one. Since the flux insertion argument picks up the anomaly between the U(1)\mathrm{U(1)} symmetry and the translation symmetry, the anomaly is also independent of the viewpoint. This observation about dimensionality motivates us to describe the anomaly of kagome quantum antiferromagnets by using one-dimensional theoretical tools.

In general, relating dd-dimensional quantum many-body systems to one-dimensional ones is a useful idea (e.g. the coupled wire construction of topological phases Kane et al. (2002); Lu and Vishwanath (2012)). This is partly because the latter is usually much better equipped with theoretical tools than the former Giamarchi (2004). The anomaly is no exception Furuya and Oshikawa (2017); Cho et al. (2017); Yao et al. (2019); Tanizaki and Sulejmanpasic (2018). In our case, however, the effective one-dimensional system inevitably contains long-range interactions that are extremely inconvenient for the anomaly argument, in particular, for the anomaly matching Hooft (1980). For example, a nearest-neighbor exchange interaction 𝑺1(n1,n2)𝑺2(n1,n2+1)\bm{S}_{1}(n_{1},n_{2})\cdot\bm{S}_{2}(n_{1},n_{2}+1) of the Hamiltonian (1) can be seen as an exchange interaction, 𝑺r1,1𝑺r1+N1,2\bm{S}_{r_{1},1}\cdot\bm{S}_{r_{1}+N_{1},2} over the long distance N1N_{1}. This is a disadvantage of the tilted boundary condition.

To relate anomalies of kagome antiferromagnets to one-dimensional physics with avoiding this inconvenience, we notice the insensitivity of anomalies to spatial anisotropies implied by the flux-insertion argument with the tilted boundary condition. The large-gauge transformation operator (10) depends explicitly on the number of spins inside the unit cell and the lattice structure through r1r_{1} but is independent of the strength of coupling constants. In general, ’t Hooft anomalies are robust against continuous variation of the Hamiltonian as long as the symmetries in question are preserved since an anomaly is identified with a surface term of a topological action of a symmetry-protected topological phase, which is manifestly topologically invariant Cheng et al. (2016); Jian et al. (2018); Thorngren and Else (2018). We can weaken or strengthen interactions in a specific spatial direction without interfering with the flux insertion argument. More generally, if an anomaly involves translation symmetries and on-site symmetries and is unrelated to spatial rotation symmetries, we can weaken interaction strengths in the directions perpendicular to the 𝒆1\bm{e}_{1} direction with keeping the symmetries. The insensitivity to the spatial anisotropy opens a way to access the accumulated knowledge about anomalies of (1+1)-dimensional quantum field theories.

This section gives particular attention to the 1/31/3 plateau where the unique gapped ground state is allowed in the flux-insertion argument and is indeed constructed in a specific model Parameswaran et al. (2013). We discuss that kagome antiferromagnets on the 1/31/3 plateau, such as that of Ref. Parameswaran et al. (2013), can have the unique gapped ground state as a result of an explicit breaking of a symmetry that involves an anomaly. This section first discusses an anomaly of a one-dimensional quantum spin system, an S=1/2S=1/2 three-leg spin tube. Next, using this knowledge of one-dimensional physics, we give an anomaly interpretation of the unique gapped ground state of the kagome antiferromagnet on the 1/31/3 plateau.

III.2 Quasi-one-dimensional limit

In order to bridge physics in the one dimension to that on the kagome lattice, we consider a deformed kagome antiferromagnet with the following Hamiltonian:

δ\displaystyle\mathcal{H}_{\delta} =1d+δ,\displaystyle=\mathcal{H}_{\rm 1d}+\delta\mathcal{H}^{\prime}, (14)
1d\displaystyle\mathcal{H}_{\rm 1d} =J1n1,n2[𝑺1(n1,n2)𝑺2(n1,n2)+𝑺2(n1,n2)𝑺3(n1,n2)+𝑺3(n1,n2)𝑺1(n1,n2)]\displaystyle=J_{1}\sum_{n_{1},n_{2}}[\bm{S}_{1}(n_{1},n_{2})\cdot\bm{S}_{2}(n_{1},n_{2})+\bm{S}_{2}(n_{1},n_{2})\cdot\bm{S}_{3}(n_{1},n_{2})+\bm{S}_{3}(n_{1},n_{2})\cdot\bm{S}_{1}(n_{1},n_{2})]
+J3n1,n2,μ𝑺μ(n1,n2)𝑺μ(n1+1,n2)hn1,n2,μSμz(n1,n2),\displaystyle\quad+J_{3}\sum_{n_{1},n_{2},\mu}\bm{S}_{\mu}(n_{1},n_{2})\cdot\bm{S}_{\mu}(n_{1}+1,n_{2})-h\sum_{n_{1},n_{2},\mu}S_{\mu}^{z}(n_{1},n_{2}), (15)
\displaystyle\mathcal{H}^{\prime} =J1n1,n2[𝑺1(n1,n2)𝑺2(n1+1,n2+1)+𝑺2(n1+1,n2)𝑺3(n1,n2)+𝑺3(n1,n2+1)𝑺1(n1,n2)]\displaystyle=J_{1}\sum_{n_{1},n_{2}}[\bm{S}_{1}(n_{1},n_{2})\cdot\bm{S}_{2}(n_{1}+1,n_{2}+1)+\bm{S}_{2}(n_{1}+1,n_{2})\cdot\bm{S}_{3}(n_{1},n_{2})+\bm{S}_{3}(n_{1},n_{2}+1)\cdot\bm{S}_{1}(n_{1},n_{2})]
+J3n1,n2μ=1,2,3[𝑺μ(n1,n2){𝑺μ(n1,n2+1)+𝑺μ(n1+1,n2+1)}],\displaystyle\quad+J_{3}\sum_{n_{1},n_{2}}\sum_{\mu=1,2,3}[\bm{S}_{\mu}(n_{1},n_{2})\cdot\{\bm{S}_{\mu}(n_{1},n_{2}+1)+\bm{S}_{\mu}(n_{1}+1,n_{2}+1)\}], (16)

where the parameter δ[0,1]\delta\in[0,1] controls the spatial anisotropy. The coupling constants J1J_{1} and J3J_{3} represent the nearest-neighbor and the third-neighbor exchange interactions when we view the model (14) as a two-dimensional system.

Let us impose the periodic boundary condition in the 𝒆1\bm{e}_{1} and 𝒆2\bm{e}_{2} directions. For δ=1\delta=1, all the nearest-neighbor bonds have the same strength of the exchange interaction and so do the third-neighbor ones. For δ=0\delta=0, the model (14) is a set of mutually independent three-leg spin tubes (Fig. 4). Note that the smooth change of δ\delta keeps the T1T_{1} and T2T_{2} translation symmetries and the U(1)\mathrm{U}(1) spin-rotation symmetry. We use the periodic boundary condition to investigate the model (14).

In this section, we set δ=0\delta=0 for a while and discuss its ’t Hooft anomaly on the 1/31/3 plateau. Later in Sec. III.5, we will resurrect δ\delta and discuss the anomaly in the quasi-one-dimensional limit, 0<δ10<\delta\ll 1.

Refer to caption
Figure 4: (a) The three-leg spin tube (15) in the periodic boundary condition. The periodic boundary condition is imposed on the leg direction. Thin and thick lines represent J1J_{1} and J3J_{3} interactions of Eq. (15), respectively. (b) The flux is inserted into the spin tube. The spin tube is regarded as a ring with a triangular cross-section.

III.3 Oshikawa-Yamanaka-Affleck condition in spin tubes

It is now a good occasion to review the original derivation of the OYA condition in the specific case of the three-leg spin tube. Let us consider a situation where three spin chains are weakly coupled to each other.

Each spin chain is equivalent to an interacting spinless fermion chain thanks to the Jordan-Wigner transformation Giamarchi (2004). Let us denote an annihilation operator of the spinless fermion at a location xx as ψμ(x)\psi_{\mu}(x), where μ=1,2,3\mu=1,2,3 are degrees of freedom that specify the leg. In the continuum limit, μ\mu represent internal degrees of freedom, which we call “color” in this paper: μ=1,2,3\mu=1,2,3 for red, white, and blue circles, respectively in Figs. 2 and 4. At low energies, the spinless fermion operator can further be split into two species:

ψμ(x)\displaystyle\psi_{\mu}(x) eikFxψR,μ(x)+eikFxψL,μ(x),\displaystyle\approx e^{ik_{F}x}\psi_{R,\mu}(x)+e^{-ik_{F}x}\psi_{L,\mu}(x), (17)

where kF>0k_{F}>0 is the Fermi wave number and ψR,μ(x)\psi_{R,\mu}(x) and ψL,μ(x)\psi_{L,\mu}(x) are annihilation operators of the right-moving and left-moving spinless fermions, respectively.

The magnetization process of the three-leg spin tube is described by a U(1)\mathrm{U(1)} conformal field theory (CFT) of compactified bosons. The spinless fermion operators ψR,μ(x)\psi_{R,\mu}(x) and ψL,μ(x)\psi_{L,\mu}(x) can be bosonized as eiϕμ/RψL,μψR,μe^{i\phi_{\mu}/R}\propto\psi_{L,\mu}^{\dagger}\psi_{R,\mu} and e2πiRθμψL,μψR,μe^{2\pi iR\theta_{\mu}}\propto\psi_{L,\mu}\psi_{R,\mu}. Here, R>0R>0 is the compactification radius of those bosons. The U(1)\mathrm{U(1)} CFT is written in two compactified U(1)\mathrm{U(1)} bosons ϕ=μ=13ϕμ\phi=\sum_{\mu=1}^{3}\phi_{\mu} and θ=μ=13θμ\theta=\sum_{\mu=1}^{3}\theta_{\mu}.

The Fermi wave number kFk_{F} is related to the magnetization per site mm as

kF=πa(12m),\displaystyle k_{F}=\frac{\pi}{a}\biggl{(}\frac{1}{2}-m\biggr{)}, (18)

where aa is the lattice spacing of the spin chain. The T1T_{1} translation xx+ax\to x+a along the leg effectively turns into an on-site symmetry, ψR,μ(x)eikFaψR,μ(x)\psi_{R,\mu}(x)\to e^{ik_{F}a}\psi_{R,\mu}(x) and ψL,μ(x)eikFaψL,μ(x)\psi_{L,\mu}(x)\to e^{-ik_{F}a}\psi_{L,\mu}(x) in the continuum limit a0a\to 0. In the boson language, the T1T_{1} translation xx+ax\to x+a affects only ϕμ\phi_{\mu}:

ϕμ\displaystyle\phi_{\mu} ϕμ+2akFR,\displaystyle\to\phi_{\mu}+2ak_{F}R, (19)
ϕ\displaystyle\phi ϕ+6akFR.\displaystyle\to\phi+6ak_{F}R. (20)

When 6akF2π6ak_{F}\in 2\pi\mathbb{Z}, a perturbation cos(ϕ/R)\cos(\phi/R) of the U(1)\mathrm{U(1)} CFT is permitted by the translation symmetry. Note that cosines and sines of θ\theta are forbidden by the U(1)\mathrm{U(1)} spin-rotation symmetry. If cos(ϕ/R)\cos(\phi/R) is relevant in the sense of the renormalization group, this cosine interaction opens the gap without any spontaneous symmetry breaking. If it is irrelevant, the ground state is gapless because it is the most relevant operator allowed by the symmetries. The condition 6akF2π6ak_{F}\in 2\pi\mathbb{Z} is rephrased as

3(Sm),\displaystyle 3(S-m)\in\mathbb{Z}, (21)

with S=1/2S=1/2. The relation (21) is the OYA condition in the case of the three-leg spin tube. It is straightforward to derive the same condition (21) for higher spin quantum numbers S>1/2S>1/2.

When 6akF2π(+p/q)6ak_{F}\in 2\pi(\mathbb{Z}+p/q) with coprime positive integers pp and qq, the most relevant symmetric interaction in the U(1)\mathrm{U(1)} CFT is cos(qϕ/R)\cos(q\phi/R). If the ground state is gapped, the ground state is at least qq-fold degenerate by breaking the translation symmetry ϕϕ+2πpR/qmod2πR\phi\to\phi+2\pi pR/q\mod 2\pi R spontaneously. We thus reached the same conclusion as that derived from Eqs. (12) and (13). However, this is not the end of the story.

III.4 Color constraints on the 1/31/3 plateau

S=1/2S=1/2 three-leg spin tubes possess the 1/31/3 plateau Okunishi et al. (2012); Plat et al. (2012); Yonaga and Shibata (2015). According to the OYA condition (21), the S=1/2S=1/2 three-leg spin tube can, in principle, have the unique gapped ground state on the 1/31/3 magnetization plateau since 3(Sm)=13(S-m)=1\in\mathbb{Z}. On the other hand, to the best of our knowledge, the unique gapped ground state has not yet been reported for the simple three-leg spin tube with the Hamiltonian equivalent to Eq. (15) on the 1/31/3 magnetization plateau Okunishi et al. (2012); Yonaga and Shibata (2015). This is not a coincidence. The unique gapped ground state is indeed forbidden, as we show below.

We reuse the spinless fermion picture. When three spinless fermion chains are gapless interacting Dirac ones decoupled from each other, the system has the U(3)\mathrm{U(3)} symmetry. Interchain interactions in the spin tube reduce this symmetry. The charge of the spinless fermion corresponds to the magnetization of the spin tube. On the 1/31/3 magnetization plateau where the total SzS^{z} is frozen, magnetic excitations can be gapped without breaking any symmetry, as the OYA condition (21) implies. The remaining degrees of freedom, which are the color, are effectively described by a perturbed SU(3)\mathrm{SU(3)} WZW theory Affleck and Haldane (1987). At the fixed point, the SU(3)\mathrm{SU(3)} WZW theory has a global SU(3)R×SU(3)L3\frac{\mathrm{SU(3)}_{R}\times\mathrm{SU(3)}_{L}}{\mathbb{Z}_{3}} symmetry Ohmori et al. (2019), where VR/LSU(3)R/LV_{R/L}\in\mathrm{SU(3)}_{R/L} act on the field gSU(3)g\in\mathrm{SU(3)} of the WZW theory as gVLgVRg\to V_{L}gV_{R}^{\dagger}. The symbols RR and LL represent the right-moving and the left-moving parts of particles of the WZW theory. In terms of spinless fermions ψR/L,μ\psi_{R/L,\mu}, a (μ,ν)(\mu,\nu) component, gμνg_{\mu\nu}, of the SU(3)\mathrm{SU(3)} matrix gg can be represented as gμνψL,μψR,νg_{\mu\nu}\propto\psi_{L,\mu}^{\dagger}\psi_{R,\nu} Affleck (1988). Generally, the RR and LL degrees of freedom are coupled to each other by intrachain and interchain interactions when the field theory deviates from the fixed point. SU(3)R×SU(3)L\mathrm{SU(3)}_{R}\times\mathrm{SU(3)}_{L} is reduced to a single SU(3)\mathrm{SU(3)} with VR=VLV_{R}=V_{L}, away from the fixed point. Then, the global symmetry is reduced to PSU(3)\mathrm{PSU(3)}.

We did not include this SU(3)\mathrm{SU(3)} WZW theory in the previous subsection. The SU(3)\mathrm{SU(3)} WZW theory is actually crucial on the 1/31/3 magnetization plateau because of the following reason. The Fermi wave number kFk_{F} is given by kF=π/3ak_{F}=\pi/3a on the 1/31/3 plateau. The field gμνψL,μψR,νg_{\mu\nu}\propto\psi_{L,\mu}^{\dagger}\psi_{R,\nu} transforms under the T1T_{1} translation as

ge2πi/3g.\displaystyle g\to e^{2\pi i/3}g. (22)

The T1T_{1} translation symmetry is turned into the on-site 3\mathbb{Z}_{3} symmetry (22). The SU(3)\mathrm{SU(3)} WZW theory has an anomaly between the 3\mathbb{Z}_{3} symmetry (22) and the PSU(3)\mathrm{PSU(3)} symmetry Yao et al. (2019). According to Ref. Yao et al. (2019), the 3\mathbb{Z}_{3} symmetry (22) and the PSU(3)\mathrm{PSU(3)} symmetry give an LSM index 3\mathcal{I}_{3} that equals the number of the Young-tableau box per unit cell. The LSM index is a quantity related to the ground-state degeneracy. If and only if 3=0mod3\mathcal{I}_{3}=0\mod 3, the ground state can be unique and gapped. Otherwise, the ground state is either gapless or gapped with at least qq-fold degeneracy. Here, qq is given by

q=3gcd(3,3),\displaystyle q=\frac{3}{\operatorname{gcd}(\mathcal{I}_{3},3)}, (23)

where gcd(n1,n2)\operatorname{gcd}(n_{1},n_{2}) is the greatest common divisor of two integers n1n_{1} and n2n_{2}. The 3×33\times 3 matrix gg belongs to either the fundamental or the conjugate representation of SU(3)\mathrm{SU(3)}. The number of the Young tableau of the fundamental (conjugate) representations contains one box (two boxes, respectively). Hence, the three-leg spin tube on the 1/31/3 plateau has the LSM index 3=1\mathcal{I}_{3}=1 or 3=2\mathcal{I}_{3}=2, both of which lead to q=3q=3. The spin tube’s ground state on the 1/31/3 plateau is either gapless or gapped with at least three-fold degeneracy. The gapless ground state is indeed possible on the 1/31/3 magnetization plateau, which is a liquid state of chirality degrees of freedom Okunishi et al. (2012); Plat et al. (2012).

We saw that the translation and the PSU(3)\mathrm{PSU(3)} color symmetries forbid the unique gapped ground state on the 1/31/3 plateau. However, the PSU(3)\mathrm{PSU(3)} color symmetry is too large to be naturally realized in three-leg spin tubes. In fact, in general, spin tubes have a 3\mathbb{Z}_{3} color-rotation symmetry instead of PSU(3)\mathrm{PSU(3)}. It is desirable to reduce the symmetries that forbid the unique gapped ground state as much as possible.

We can reduce the color symmetry relevant to the ground-state degeneracy by referring to another quantum field theory, a (1+1)-dimensional SU(3)/U(1)2\mathrm{SU(3)/U(1)^{2}} nonlinear sigma model Lajkó et al. (2017); Tanizaki and Sulejmanpasic (2018). This nonlinear sigma model, as well as the SU(3)\mathrm{SU(3)} WZW theory, is also an effective field theory of SU(3)\mathrm{SU(3)} spin chains Lajkó et al. (2017); Tanizaki and Sulejmanpasic (2018); Yao et al. (2019). Though the SU(3)/U(1)2\mathrm{SU(3)/U(1)^{2}} model is not directly related to our spin tube (15), we consider this nonlinear sigma model for its relation to the SU(3)\mathrm{SU(3)} WZW theory; the low-energy limit of the (1+1)-dimensional SU(3)/U(1)2\mathrm{SU(3)/U(1)^{2}} nonlinear sigma model perturbed by local interactions is described by the SU(3)\mathrm{SU(3)} WZW theory Ohmori et al. (2019). According to the anomaly matching argument Hooft (1980), the SU(3)/U(1)2\mathrm{SU(3)/U(1)^{2}} nonlinear sigma model and the SU(3)\mathrm{SU(3)} WZW theory share the mixed ’t Hooft anomaly in common. In fact, Ref. Tanizaki and Sulejmanpasic (2018) shows that the SU(3)/U(1)2\mathrm{SU(3)/U(1)^{2}} nonlinear sigma model has an anomaly between the PSU(3)\mathrm{PSU(3)} symmetry and the one-unit translation symmetry. This is consistent with the result of Ref. Yao et al. (2019). Reference Tanizaki and Sulejmanpasic (2018) further argued that the nonlinear sigma model has an anomaly between a finite subgroup 3×3\mathbb{Z}_{3}\times\mathbb{Z}_{3} of PSU(3)\mathrm{PSU(3)} and the translation symmetry. Note that 3×3\mathbb{Z}_{3}\times\mathbb{Z}_{3} acts on the sigma fields projectively Tanizaki and Sulejmanpasic (2018).

The 3×3\mathbb{Z}_{3}\times\mathbb{Z}_{3} group is related to the following matrices,

M1\displaystyle M_{1} =(1000ω000ω2),M2=(010001100).\displaystyle=\begin{pmatrix}1&0&0\\ 0&\omega&0\\ 0&0&\omega^{2}\end{pmatrix},\qquad M_{2}=\begin{pmatrix}0&1&0\\ 0&0&1\\ 1&0&0\end{pmatrix}. (24)

An isomorhpism H:SU(3)PSU(3)H:\mathrm{SU(3)}\to\mathrm{PSU(3)} maps MnM_{n} for n=1,2n=1,2 to H(Mn)H(M_{n}) that generate the 3×3\mathbb{Z}_{3}\times\mathbb{Z}_{3}. Note that M1M_{1} is a gauge symmetry of the SU(3)/U(1)2\mathrm{SU(3)/U(1)^{2}} nonlinear sigma model, which does not forbid any interaction that perturbs the nonlinear sigma model. By contrast, M2M_{2} forbids some interactions. It acts on the gg field as M2gM2M_{2}gM_{2}^{\dagger}. In other words, M2M_{2} represents the 3\mathbb{Z}_{3} color-rotation symmetry, which is essential in what follows.

The 3\mathbb{Z}_{3} color-rotation symmetry forbids interactions that treat the legs of the spin tube unequally. By imposing the 3\mathbb{Z}_{3} color-rotation symmetry, the U(1)\mathrm{U(1)} spin-rotation symmetry, and the translation symmetry on the spin tube, we can forbid the system (15) from having the unique gapped ground state on the 1/31/3 plateau. On the other hand, this anomaly is absent in generic kagome antiferromagnets such as Eq. (1), as we show below.

III.5 Unique gapped ground state on the 1/31/3 plateau

While the anomaly argument in Sec. III.4 is already nontrivial as quantum physics in one dimension, it also gives interesting feedback to the original two-dimensional problem. Let us look back on the kagome antiferromagnet (15). As we mentioned, the anomaly in the two-dimensional system is inherited by the quasi-one-dimensional system where, in our case, three-leg spin tubes are weakly coupled with short-range interactions. Recall that the periodic boundary conditions are imposed on the 𝒆1\bm{e}_{1} and the 𝒆2\bm{e}_{2} directions. Hence, the anomaly argument in the quasi-one-dimensional limit reversely gives us a glimpse of the anomaly in the isotropic two-dimensional case. Let us suppose that these intertube interactions respect the 3\mathbb{Z}_{3} color-rotation symmetry, the U(1)\mathrm{U}(1) spin-rotation symmetry, and the T1T_{1} and T2T_{2} translation symmetries.

The unique gapped ground state is forbidden under those symmetries of intertube interactions. When the single spin tube has degenerate gapped ground states with a spontaneous symmetry breaking, the weak enough symmetric intertube interactions keep the gap open and do not lift the degeneracy.

On the other hand, when the single spin tube has the gapless ground state on the 1/31/3 plateau, that is, the chirality liquid state Okunishi et al. (2012), a weak symmetric intertube interaction can induce a quantum phase transition. This phase transition occurs even if the coupling constant of the intertube interaction is infinitesimal Schulz (1996); Furuya et al. (2016). This is because the chirality liquid state is a Tomonaga-Luttinger liquid. At zero temperature, the Tomonaga-Luttinger liquid state has a divergent susceptibility of the order parameter Giamarchi (2004). This quantum phase transition drives the chirality liquid state into an ordered phase where the 3\mathbb{Z}_{3} symmetry is spontaneously broken. After all, the unique gapped ground state is forbidden in both cases.

Now we go back to the J1J3J_{1}-J_{3} model (14) of the kagome antiferromagnet. We showed that the model with δ=0\delta=0 has the anomaly between the 3\mathbb{Z}_{3} color-rotation symmetry and the T1T_{1} translation symmetry on the 1/31/3 plateau. As soon as δ\delta is turned on, the 3\mathbb{Z}_{3} symmetry breaks down because of the first three terms of Eq. (16) with the coupling J1J_{1} break the 3\mathbb{Z}_{3} color-rotation symmetry. On the basis of the anomaly argument, we can expect that the absence of the 3\mathbb{Z}_{3} color-rotation symmetry makes the unique gapped ground state possible on the 1/31/3 plateau of the kagome antiferromagnet.

In the end, we caught a glimpse of the anomaly of the kagome antiferromagnet on the 1/31/3 plateau with the high 3\mathbb{Z}_{3} color symmetry by fine-tuned parameters. However, we also found no impediment to the possibility of the unique gapped ground state in general kagome geometry because the 3\mathbb{Z}_{3} color-rotation symmetry is broken unless parameters of the Hamiltonian are fine-tuned.

Our argument shows that there is no unique gapped ground state on the 1/31/3 plateau of a model with the 3\mathbb{Z}_{3} color-rotation symmetry, the U(1)\mathrm{U}(1) spin-rotation symmetry, and the T1T_{1} and T2T_{2} translation symmetries beyond the quasi-one-dimensional limit. Unfortunately, it remains challenging to confirm the existence of this anomaly directly in the isotropic two-dimensional case. We leave it for future works.

III.6 Reduction of the unit cell

To close this section, we need to mention an unlikely possibility of reducing the number of spins in the unit cell. In this paper, we took an upward triangle as a unit cell as a natural choice. We can choose, in principle, a unit cell with any other shape. One might expect that regardless of the unit cell’s shape, the unit cell of the kagome lattice will have three spins inside the unit cell.

This statement is actually nontrivial when we take into account general closed boundary conditions. For example, antiferromagnets on the checkerboard lattice have two spins inside the unit cell under the periodic boundary condition or the tilted boundary condition but have only one spin under the spatially twisted boundary condition Furuya and Horinouchi (2019). This reduction of the number of spins in the unit cell was a key to show the impossibility of the unique gapped ground state in the checkerboard antiferromagnet at zero magnetic field in Ref. Furuya and Horinouchi (2019). If we could construct a symmetric closed boundary condition with a unit cell containing only one spin on the kagome lattice, the lower bound dmd_{\rm m} shown in Table. 1 would be raised.

The bottom line is that such closed boundary conditions with the reduced unit cell can indeed be constructed, but they break the T1T_{1} translation symmetry. These closed boundary conditions with the reduced unit cell do not qualify as the symmetric closed boundary condition in the kagome lattice case different from the checkerboard lattice. We construct those closed boundary conditions in Appendix A. We thus close this section by concluding that dmd_{\rm m} in Table. 1 gives the maximum value of the lower bound of the ground-state degeneracy under the U(1)\mathrm{U(1)} spin-rotation symmetry and the translation symmetry.

IV Summary

We discussed magnetization plateaus of geometrically frustrated quantum antiferromagnets on the kagome lattice from the viewpoint of the OYA condition in one symmetric boundary condition, the tilted boundary condition Yao and Oshikawa (2020). Here, the OYA condition was derived in a form independent of the aspect ratio N2/N1N_{2}/N_{1} of the rhombic finite-size cluster of the kagome lattice. Using the flux insertion argument with the tilted boundary condition, we showed that the ground states on the 1/91/9, 5/95/9, and 7/97/9 plateaus of spin-1/21/2 kagome antiferromagnets are at least threefold degenerate. The 1/31/3 plateau, different from the other plateaus, can host a unique gapped ground state, as was explicitly demonstrated before in a specific model Parameswaran et al. (2013).

To foster a better understanding of the 1/31/3 plateau, we gave our attention to the insensitivity of the anomaly to spatial anisotropies when no spatial rotation symmetry is required for the appearance of the anomaly. Investigating the anomaly in the quasi-one-dimensional limit is thus expected to help understand the anomaly of higher-dimensional systems. We exemplified the usefulness of this idea in the kagome antiferromagnet on the 1/31/3 magnetization plateau.

The 1/31/3 magnetization plateau has one special significance in this viewpoint. The ’t Hooft anomaly argument of (1+1)-dimensional quantum field theories clarified one characteristic feature of the 1/31/3 plateau of the kagome lattice: The simple three-leg spin tube (15) with the 3\mathbb{Z}_{3} color symmetry cannot have the unique gapped ground state on the 1/31/3 magnetization plateau because of the anomaly between this symmetry and the translation symmetry. This impossibility of the unique gapped ground state is not predicted in the original OYA condition Oshikawa et al. (1997). This is because the OYA condition refers to the spin degrees of freedom, but our argument here involves the color degrees of freedom. On the other hand, when we adopt this argument to the kagome antiferromagnet, we find that this system can have the unique gapped ground state because the kagome lattice breaks the 3\mathbb{Z}_{3} color rotation symmetry unless the Hamiltonian is fine-tuned. We thus conclude that the absence of the 3\mathbb{Z}_{3} color symmetry enables the unique gapped ground state on the 1/31/3 plateau of kagome antiferromagnets in general.

Acknowledgments

The authors are grateful to Yohei Fuji, Akira Furusaki, and Takuya Furusawa for stimulating discussions. This work was supported by KAKENHI Grant Nos. JP16K05425 and JP20K03778 from the Japan Society for the Promotion of Science. S.C.F is supported by a Grant-in-Aid for Scientific Research on Innovative Areas “Quantum Liquid Crystals” (Grant No. JP19H05825).

Appendix A Closed boundary condition with spatial rotations

A.1 Reduction of the number of spins

We saw that the kagome lattice does not have the 3\mathbb{Z}_{3} color symmetry in the tilted boundary condition without fine-tuning parameters. On the other hand, the absence of the 3\mathbb{Z}_{3} color-rotation symmetry sounds strange because the three colors of the kagome lattice look equivalent in the isotropic thermodynamic limit. In this appendix, we construct a boundary condition to respect this 3\mathbb{Z}_{3} color-rotation symmetry of the kagome lattice. However, the translation symmetry is broken with this boundary condition, as we show later.

We have another motivation to consider such a strange boundary condition, as mentioned in Sec. III.6. We will see that the boundary conditions constructed in what follows contain only one spin inside the unit cell.

Refer to caption
Figure 5: The C3C_{3}-rotated boundary condition. The kagome lattice is put on the complex plane \mathbb{C} whose origin is C3C_{3}-invariant. Translation rules (25) in the bulk and (26) on the seam enables us to follow a closed path 12345611\to 2\to 3\to 4\to 5\to 6\to 1, starting from a point 11 on the left seam AFE.

We can construct two closed boundary conditions with which the unit cell contains only one spin. One of these boundary conditions is referred to as a C3C_{3}-rotated boundary condition in this paper because it goes with the C3C_{3} spatial rotation on the system’s seams (Fig. 5). Another is referred to as a tilted C3C_{3}-rotated boundary condition, a combination of the C3C_{3}-rotated boundary condition and the tilted boundary condition.

A.2 Rotated boundary conditions

The C3C_{3}-rotated boundary condition is defined as follows. Let us consider a hexagonal finite-size cluster of the kagome lattice and denote the location of each site as a single complex variable zz. We define the origin z=0z=0 as the C3C_{3}-rotation invariant point. Accordingly, we relabel the spin operator as 𝑺μ(z)\bm{S}_{\mu}(z) for μ=1,2,3\mu=1,2,3. The translation T1T_{1} in this boundary condition is defined as

T1𝑺μ(z)T11\displaystyle T_{1}\bm{S}_{\mu}(z)T_{1}^{-1} =𝑺μ(z+2ωμ1),\displaystyle=\bm{S}_{\mu}(z+2\omega^{\mu-1}), (25)

with ω=exp(2πi/3)\omega=\exp(2\pi i/3) if a segment that connects zz and z+2ωμ1z+2\omega^{\mu-1} does not overpass the seam ABCDEF of Fig. 5. If it overpasses the seam, we perform a C3C_{3} rotation,

T1𝑺μ(zseam)T11\displaystyle T_{1}\bm{S}_{\mu}(z_{\rm seam})T_{1}^{-1} =𝑺μ+1(ω2μz¯seam),\displaystyle=\bm{S}_{\mu+1}(-\omega^{2-\mu}\bar{z}_{\rm seam}), (26)

where zseamz_{\rm seam} is the complex coordinate on the seam of the system and 𝑺4(z)=𝑺1(z)\bm{S}_{4}(z)=\bm{S}_{1}(z).

Refer to caption
Figure 6: The 4242-site kagome lattice in the tilted C3C_{3}-rotated boundary condition. Every site is labeled by a single number that corresponds to the one-dimensional coordinate r1r^{\prime}_{1}. The T1T_{1} translation increases r1r^{\prime}_{1} by one.

This boundary condition is depicted in Fig. 5. Let us apply the translation repeatedly and return to the starting point, say, the point 11. We denote the complex coordinate of the point nn of Fig. 5 as znz_{n}. These complex coordinates are related through

(z1,z3,z5)\displaystyle(z_{1},z_{3},z_{5}) =ω(z5,z1,z3),\displaystyle=\omega(z_{5},z_{1},z_{3}), (27)
(z2,z4,z6)\displaystyle(z_{2},z_{4},z_{6}) =ω(z6,z2,z4),\displaystyle=\omega(z_{6},z_{2},z_{4}), (28)
(z1,z3,z5)\displaystyle(z_{1},z_{3},z_{5}) =(ω2z¯6,ωz¯2,z¯4),\displaystyle=(-\omega^{2}\bar{z}_{6},-\omega\bar{z}_{2},-\bar{z}_{4}), (29)

where z¯\bar{z} is the complex conjugate of zz. The point 11 is located at the left seam of the system. If we denote the real part of z1z_{1} as -\ell, spin operators 𝑺μ(zn)\bm{S}_{\mu}(z_{n}) are related to each other in accordance with Eqs. (25) and (26):

(T1)𝑺1(z1)(T1)\displaystyle(T_{1})^{\ell}\bm{S}_{1}(z_{1})(T_{1})^{-\ell} =𝑺1(z2),\displaystyle=\bm{S}_{1}(z_{2}), (30)
T1𝑺1(z2)T11\displaystyle T_{1}\bm{S}_{1}(z_{2})T_{1}^{-1} =𝑺2(z3),\displaystyle=\bm{S}_{2}(z_{3}), (31)
(T1)𝑺2(z3)(T1)\displaystyle(T_{1})^{\ell}\bm{S}_{2}(z_{3})(T_{1})^{-\ell} =𝑺2(z4),\displaystyle=\bm{S}_{2}(z_{4}), (32)
T1𝑺2(z4)T11\displaystyle T_{1}\bm{S}_{2}(z_{4})T_{1}^{-1} =𝑺3(z5),\displaystyle=\bm{S}_{3}(z_{5}), (33)
(T1)𝑺3(z5)(T1)\displaystyle(T_{1})^{\ell}\bm{S}_{3}(z_{5})(T_{1})^{-\ell} =𝑺3(z6),\displaystyle=\bm{S}_{3}(z_{6}), (34)
T1𝑺3(z6)T11\displaystyle T_{1}\bm{S}_{3}(z_{6})T_{1}^{-1} =𝑺1(z1).\displaystyle=\bm{S}_{1}(z_{1}). (35)

Note that \ell is a positive integer that depends on z1z_{1}. Repeated operations of T1T_{1}, thus bring us back to the starting point, z1z_{1}, eventually.

To insert the flux, we introduce a tilt to the C3C_{3}-rotated boundary condition. Namely, we modify Eq. (26) to

T1𝑺μ(zseam)T11\displaystyle T_{1}\bm{S}_{\mu}(z_{\rm seam})T_{1}^{-1} =𝑺μ+1(ω2μz¯seam+i3δμ,3),\displaystyle=\bm{S}_{\mu+1}(-\omega^{2-\mu}\bar{z}_{\rm seam}+i\sqrt{3}\delta_{\mu,3}), (36)

where δa,b\delta_{a,b} is the Kronecker delta. We call the boundary condition (36) as a tilted C3C_{3}-rotated boundary condition.

Operations (25) and (36) define a path 𝒞\mathcal{C} along which we can sweep every spin on the kagome lattice once and only once. Let us define the starting point as a site at the left bottom corner (Fig. 6). We can claim that adiabatically inserted unit flux is eliminated by a large gauge transformation UrotU_{\rm rot} defined as

Urot\displaystyle U_{\rm rot} =exp(i2πVr1=1Vr1n(r1)),\displaystyle=\exp\biggl{(}i\frac{2\pi}{V}\sum_{r^{\prime}_{1}=1}^{V}r^{\prime}_{1}n(r^{\prime}_{1})\biggr{)}, (37)

where r1r^{\prime}_{1} is the one-dimensional coordinate corresponding to zz along the path 𝒞\mathcal{C}. However, the Hamiltonian on the kagome lattice in this boundary condition inevitably breaks the translation symmetry along the path 𝒞\mathcal{C}. In other words, the T1T_{1} operator defined as Eqs. (25) and (36) leads to [,T1]0[\mathcal{H},T_{1}]\not=0. The absence of the translation symmetry is evident from, for instance, the inequivalence of sites 11 and site 77 of Fig. 6. The spin on the latter site has nearest-neighbor exchange interaction between four neighboring sites, but the spin on the former site has those between only two neighboring sites. Of course, we did not exclude rigorously at all the possibility of the symmetric closed boundary condition with which the unit cell contains one spin. However, it is also certain that almost no room is left for this possibility. We thus conclude that it is highly unlikely that the number of spins in the unit cell is reduced without any explicit symmetry breaking.

References

  • Schnyder et al. (2008) Andreas P. Schnyder, Shinsei Ryu, Akira Furusaki,  and Andreas W. W. Ludwig, “Classification of topological insulators and superconductors in three spatial dimensions,” Phys. Rev. B 78, 195125 (2008).
  • Chen et al. (2011) Xie Chen, Zheng-Cheng Gu,  and Xiao-Gang Wen, “Classification of gapped symmetric phases in one-dimensional spin systems,” Phys. Rev. B 83, 035107 (2011).
  • Chen et al. (2013) Xie Chen, Zheng-Cheng Gu, Zheng-Xin Liu,  and Xiao-Gang Wen, “Symmetry protected topological orders and the group cohomology of their symmetry group,” Phys. Rev. B 87, 155114 (2013).
  • Furuya and Oshikawa (2017) Shunsuke C. Furuya and Masaki Oshikawa, “Symmetry Protection of Critical Phases and a Global Anomaly in 1+11+1 Dimensions,” Phys. Rev. Lett. 118, 021601 (2017).
  • Yao et al. (2019) Yuan Yao, Chang-Tse Hsieh,  and Masaki Oshikawa, “Anomaly Matching and Symmetry-Protected Critical Phases in SU(N)SU(N) Spin Systems in 1+11+1 Dimensions,” Phys. Rev. Lett. 123, 180201 (2019).
  • Lieb et al. (1961) Elliott Lieb, Theodore Schultz,  and Daniel Mattis, “Two soluble models of an antiferromagnetic chain,” Annals of Physics 16, 407 – 466 (1961).
  • Oshikawa (2000) Masaki Oshikawa, “Commensurability, Excitation Gap, and Topology in Quantum Many-Particle Systems on a Periodic Lattice,” Phys. Rev. Lett. 84, 1535–1538 (2000).
  • Hastings (2004) M. B. Hastings, “Lieb-Schultz-Mattis in higher dimensions,” Phys. Rev. B 69, 104431 (2004).
  • Oshikawa et al. (1997) Masaki Oshikawa, Masanori Yamanaka,  and Ian Affleck, “Magnetization Plateaus in Spin Chains: “Haldane Gap” for Half-Integer Spins,” Phys. Rev. Lett. 78, 1984–1987 (1997).
  • Cho et al. (2017) Gil Young Cho, Chang-Tse Hsieh,  and Shinsei Ryu, “Anomaly manifestation of lieb-schultz-mattis theorem and topological phases,” Phys. Rev. B 96, 195105 (2017).
  • Yao and Oshikawa (2020) Yuan Yao and Masaki Oshikawa, “Generalized Boundary Condition Applied to Lieb-Schultz-Mattis-Type Ingappabilities and Many-Body Chern Numbers,” Phys. Rev. X 10, 031008 (2020).
  • Furuya and Horinouchi (2019) Shunsuke C. Furuya and Yusuke Horinouchi, “Translation constraints on quantum phases with twisted boundary conditions,” Phys. Rev. B 100, 174435 (2019).
  • Ryu and Zhang (2012) Shinsei Ryu and Shou-Cheng Zhang, “Interacting topological phases and modular invariance,” Phys. Rev. B 85, 245132 (2012).
  • Sule et al. (2013) Olabode Mayodele Sule, Xiao Chen,  and Shinsei Ryu, “Symmetry-protected topological phases and orbifolds: Generalized laughlin’s argument,” Phys. Rev. B 88, 075125 (2013).
  • Cheng et al. (2016) Meng Cheng, Michael Zaletel, Maissam Barkeshli, Ashvin Vishwanath,  and Parsa Bonderson, “Translational symmetry and microscopic constraints on symmetry-enriched topological phases: A view from the surface,” Phys. Rev. X 6, 041068 (2016).
  • Jian et al. (2018) Chao-Ming Jian, Zhen Bi,  and Cenke Xu, “Lieb-schultz-mattis theorem and its generalizations from the perspective of the symmetry-protected topological phase,” Phys. Rev. B 97, 054412 (2018).
  • Thorngren and Else (2018) Ryan Thorngren and Dominic V. Else, “Gauging spatial symmetries and the classification of topological crystalline phases,” Phys. Rev. X 8, 011040 (2018).
  • Hida (2001) Kazuo Hida, “Magnetization Process of the S=1 and 1/2 Uniform and Distorted Kagomé Heisenberg Antiferromagnets,” Journal of the Physical Society of Japan 70, 3673–3677 (2001).
  • Schulenburg et al. (2002) J. Schulenburg, A. Honecker, J. Schnack, J. Richter,  and H.-J. Schmidt, “Macroscopic Magnetization Jumps due to Independent Magnons in Frustrated Quantum Spin Lattices,” Phys. Rev. Lett. 88, 167207 (2002).
  • Nishimoto et al. (2013) Satoshi Nishimoto, Naokazu Shibata,  and Chisa Hotta, “Controlling frustrated liquids and solids with an applied field in a kagome heisenberg antiferromagnet,” Nature communications 4, 2287 (2013).
  • Capponi et al. (2013) Sylvain Capponi, Oleg Derzhko, Andreas Honecker, Andreas M. Läuchli,  and Johannes Richter, “Numerical study of magnetization plateaus in the spin-12\frac{1}{2} kagome Heisenberg antiferromagnet,” Phys. Rev. B 88, 144416 (2013).
  • Nakano and Sakai (2014) Hiroki Nakano and Tôru Sakai, “Anomalous Behavior of the Magnetization Process of the S = 1/2 Kagome-Lattice Heisenberg Antiferromagnet at One-Third Height of the Saturation,” Journal of the Physical Society of Japan 83, 104710 (2014).
  • Nakano and Sakai (2015) Hiroki Nakano and Tôru Sakai, “Magnetization Process of the Spin-S Kagome-Lattice Heisenberg Antiferromagnet,” Journal of the Physical Society of Japan 84, 063705 (2015).
  • Picot et al. (2016) Thibaut Picot, Marc Ziegler, Román Orús,  and Didier Poilblanc, “Spin-SS kagome quantum antiferromagnets in a field with tensor networks,” Phys. Rev. B 93, 060407 (2016).
  • Sakai et al. (2008) Tôru Sakai, Masahiro Sato, Kouichi Okunishi, Yuichi Otsuka, Kiyomi Okamoto,  and Chigak Itoi, “Quantum phase transitions of the asymmetric three-leg spin tube,” Phys. Rev. B 78, 184415 (2008).
  • Okunishi et al. (2012) Kouichi Okunishi, Masahiro Sato, Tôru Sakai, Kiyomi Okamoto,  and Chigak Itoi, “Spin-chirality separation and S3{S}_{3} symmetry breaking in the magnetization plateau of the quantum spin tube,” Phys. Rev. B 85, 054416 (2012).
  • Plat et al. (2012) X. Plat, S. Capponi,  and P. Pujol, “Combined analytical and numerical approach to magnetization plateaux in one-dimensional spin tube antiferromagnets,” Phys. Rev. B 85, 174423 (2012).
  • Yonaga and Shibata (2015) Kouki Yonaga and Naokazu Shibata, “Ground State Phase Diagram of Twisted Three-Leg Spin Tube in Magnetic Field,” Journal of the Physical Society of Japan 84, 094706 (2015).
  • Pal et al. (2020) Santanu Pal, Anirban Mukherjee,  and Siddhartha Lal, “Topological approach to quantum liquid ground states on geometrically frustrated heisenberg antiferromagnets,” Journal of Physics: Condensed Matter 32, 165805 (2020).
  • Ran et al. (2007) Ying Ran, Michael Hermele, Patrick A. Lee,  and Xiao-Gang Wen, “Projected-wave-function study of the spin-1/21/2 heisenberg model on the kagomé lattice,” Phys. Rev. Lett. 98, 117205 (2007).
  • Iqbal et al. (2011) Yasir Iqbal, Federico Becca,  and Didier Poilblanc, “Projected wave function study of 𝕫2{\mathbb{z}}_{2} spin liquids on the kagome lattice for the spin-12\frac{1}{2} quantum heisenberg antiferromagnet,” Phys. Rev. B 84, 020407 (2011).
  • He et al. (2017) Yin-Chen He, Michael P. Zaletel, Masaki Oshikawa,  and Frank Pollmann, “Signatures of Dirac Cones in a DMRG Study of the Kagome Heisenberg Model,” Phys. Rev. X 7, 031020 (2017).
  • Yan et al. (2011) Simeng Yan, David A. Huse,  and Steven R. White, “Spin-liquid ground state of the s = 1/2 kagome heisenberg antiferromagnet,” Science 332, 1173–1176 (2011).
  • Depenbrock et al. (2012) Stefan Depenbrock, Ian P. McCulloch,  and Ulrich Schollwöck, “Nature of the spin-liquid ground state of the s=1/2s=1/2 heisenberg model on the kagome lattice,” Phys. Rev. Lett. 109, 067201 (2012).
  • Jiang et al. (2012) Hong-Chen Jiang, Zhenghan Wang,  and Leon Balents, “Identifying topological order by entanglement entropy,” Nature Physics 8, 902–905 (2012).
  • Parameswaran et al. (2013) S. A. Parameswaran, Itamar Kimchi, Ari M. Turner, D. M. Stamper-Kurn,  and Ashvin Vishwanath, “Wannier Permanent Wave Functions for Featureless Bosonic Mott Insulators on the 1/31/3-Filled Kagome Lattice,” Phys. Rev. Lett. 110, 125301 (2013).
  • Kane et al. (2002) C. L. Kane, Ranjan Mukhopadhyay,  and T. C. Lubensky, “Fractional quantum hall effect in an array of quantum wires,” Phys. Rev. Lett. 88, 036401 (2002).
  • Lu and Vishwanath (2012) Yuan-Ming Lu and Ashvin Vishwanath, “Theory and classification of interacting integer topological phases in two dimensions: A chern-simons approach,” Phys. Rev. B 86, 125119 (2012).
  • Giamarchi (2004) T. Giamarchi, Quantum Physics in One Dimension (Oxford University Press, Oxford, 2004).
  • Tanizaki and Sulejmanpasic (2018) Yuya Tanizaki and Tin Sulejmanpasic, “Anomaly and global inconsistency matching: θ\theta angles, SU(3)/U(1)2SU(3)/U{(1)}^{2} nonlinear sigma model, SU(3)SU(3) chains, and generalizations,” Phys. Rev. B 98, 115126 (2018).
  • Hooft (1980) G.’t Hooft, “Naturalness, chiral symmetry, and spontaneous chiral symmetry breaking,” in Recent Developments in Gauge Theories, edited by G.’t Hooft, C. Itzykson, A. Jaffe, H. Lehmann, P. K. Mitter, I. M. Singer,  and R. Stora (Springer US, Boston, MA, 1980) pp. 135–157.
  • Affleck and Haldane (1987) Ian Affleck and F. D. M. Haldane, “Critical theory of quantum spin chains,” Phys. Rev. B 36, 5291–5300 (1987).
  • Ohmori et al. (2019) Kantaro Ohmori, Nathan Seiberg,  and Shu-Heng Shao, “Sigma Models on Flags,” SciPost Phys. 6, 17 (2019).
  • Affleck (1988) Ian Affleck, “Critical behaviour of SU(n) quantum chains and topological non-linear σ\sigma-models,” Nuclear Physics B 305, 582 – 596 (1988).
  • Lajkó et al. (2017) Miklós Lajkó, Kyle Wamer, Frédéric Mila,  and Ian Affleck, “Generalization of the haldane conjecture to su(3) chains,” Nuclear Physics B 924, 508 – 577 (2017).
  • Schulz (1996) H. J. Schulz, “Dynamics of coupled quantum spin chains,” Phys. Rev. Lett. 77, 2790–2793 (1996).
  • Furuya et al. (2016) Shunsuke C. Furuya, Maxime Dupont, Sylvain Capponi, Nicolas Laflorencie,  and Thierry Giamarchi, “Dimensional modulation of spontaneous magnetic order in quasi-two-dimensional quantum antiferromagnets,” Phys. Rev. B 94, 144403 (2016).