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

Fragility of 𝒵2\mathcal{Z}_{2} topological invariant characterizing triplet excitations in a bilayer kagome magnet

Andreas Thomasen Theory of Quantum Matter Unit, Okinawa Institute of Science and Technology Graduate University, Onna-son, Okinawa 904-0395, Japan    Karlo Penc Institute for Solid State Physics and Optics, Wigner Research Centre for Physics, H-1525 Budapest, P.O.B. 49, Hungary    Nic Shannon Theory of Quantum Matter Unit, Okinawa Institute of Science and Technology Graduate University, Onna-son, Okinawa 904-0395, Japan    Judit Romhányi Theory of Quantum Matter Unit, Okinawa Institute of Science and Technology Graduate University, Onna-son, Okinawa 904-0395, Japan Department of Physics and Astronomy, University of California, Irvine, California 92697, USA
(January 12, 2025)
Abstract

The discovery by Kane and Mele of a model of spinful electrons characterized by a 𝒵2\mathcal{Z}_{2} topological invariant had a lasting effect on the study of electronic band structures. Given this, it is natural to ask whether similar topology can be found in the band-like excitations of magnetic insulators, and recently models supporting 𝒵2\mathcal{Z}_{2} topological invariants have been proposed for both magnon [Kondo et al. Phys. Rev. B 99, 041110(R) (2019)] and triplet [D. G. Joshi and A. P. Schnyder, Phys. Rev. B 100, 020407 (2019)] excitations. In both cases, magnetic excitations form time-reversal (TR) partners, which mimic the Kramers pairs of electrons in the Kane-Mele model but do not enjoy the same type of symmetry protection. In this paper, we revisit this problem in the context of the triplet excitations of a spin model on the bilayer kagome lattice. Here the triplet excitations provide a faithful analog of the Kane–Mele model as long as the Hamiltonian preserves the TR×\timesU(1) symmetry. We find that exchange anisotropies, allowed by the point group and typical in realistic models, break the required TR×\timesU(1) symmetry and instantly destroy the 𝒵2\mathcal{Z}_{2} band topology. We further consider the effects of TR breaking by an applied magnetic field. In this case, the lifting of spin–degeneracy leads to a triplet Chern insulator, which is stable against the breaking of TR×\timesU(1) symmetry. Kagome bands realize both a quadratic and a linear band touching, and we provide a thorough characterization of the Berry curvature associated with both cases. We also calculate the triplet–mediated spin Nernst and thermal Hall signals which could be measured in experiments. These results suggest that the 𝒵2\mathcal{Z}_{2} topology of band–like excitations in magnets may be intrinsically fragile compared to their electronic counterparts.

I Introduction

The discovery, in 1980, of an integer quantization in the Hall response of a two–dimensional electron gas in high magnetic field Klitzing et al. (1980), marked a new beginning for condensed matter physics. It was quickly realised that this quantization implied a new form of universality Laughlin (1981), enjoying protection against both interactions and disorder, with the quantized Hall response reflecting the integer values of the Chern indices characterizing the topology of the underlying electron bands Thouless et al. (1982); Avron et al. (1983); Kohmoto (1985). In a celebrated paper, Haldane noted that these conditions could also be met in a simple model of spinless electrons on a honeycomb lattice, with time–reversal symmetry broken by complex hoping integrals, but no magnetic field Haldane (1988). And the generalization of Haldane’s model to spinful electrons, by Kane and Mele Kane and Mele (2005a, b), set the stage for the burgeoning field of topological insulators (TI’s) and superconductors Hasan and Kane (2010); Qi and Zhang (2011), with current estimates suggesting that as many as 27% of materials may have a topological band structure Vergniory et al. (2019). Moreover, since the exotic properties of TI’s follow from the single–particle properties of a band, analagous effects can also be found in a wide range of other systems, including photonic metastructures Onoda et al. (2004); Ozawa et al. (2019), electronic circuits Lee et al. (2018), and both accoustic Peri et al. (2019) and mechanical lattices Fruchart et al. (2020).

Another natural places to look for nontrivial topology is in the band-like integer spin excitations of insulating magnets. These may take the form of magnon (spin–wave) excitations of ordered phases, or triplet excitations of quantum paramagnets. Such excitations are Bosonic, and acquire Berry phases as a consequence of spin–orbit coupling, usually in the form of Dzyaloshniski–Moriya interactions Onose et al. (2010). As a result, both magnon Shindou et al. (2013); Mook et al. (2014a, b); Owerre (2016); Chernyshev and Maksimov (2016); Zyuzin and Kovalev (2016); Kim et al. (2016); Nakata et al. (2017); McClarty et al. (2018); Kondo et al. (2019a); McClarty and Rau (2019); Kondo et al. (2020); McClarty (2021), and triplon Romhányi et al. (2015); McClarty et al. (2017); Joshi and Schnyder (2019), bands can exhibit non-trivial Chern indices, in direct analogue with TI’s Kumar et al. (2020). These systems exhibit exactly the same topologically–protected edge modes as their electronic counterparts Shindou et al. (2013); Mook et al. (2014b); Romhányi et al. (2015); Nakata et al. (2017), and can be indexed in the same way, even in the presence of disorder Akagi et al. (2017); Yoshioka et al. (2018). They also support thermal Hall Mook et al. (2014a); Romhányi et al. (2015); Owerre (2016); Nakata et al. (2017) and spin–Nernst Zyuzin and Kovalev (2016); Kim et al. (2016); Cheng et al. (2016); Owerre (2016); Nakata et al. (2017); Kondo et al. (2019a) effects, in correspondence to the Hall Haldane (1988) and spin–Hall effects seen in electronic systems Kane and Mele (2005a, b).

However, since these topological bands are a feature of excitations, rather than of a ground state, the quantized Hall effect found in Chernful band of electrons Thouless et al. (1982); Avron et al. (1983); Kohmoto (1985), is superseded by a non–integer, temperature–dependent response, coming from thermally–excited Bosons Katsura et al. (2010); Matsumoto and Murakami (2011). Moreover, the fact that interactions between Bosons can be relevant Chernyshev and Zhitomirsky (2015); Chernyshev and Maksimov (2016) also creates a new opportunities to study non–Hermitian aspects of their dynamics McClarty and Rau (2019); Kondo et al. (2020). For a recent review of this, and other related issues see McClarty (2021).

Given the seminal role of the models of Haldane Haldane (1988), and Kane and Mele Kane and Mele (2005a, b) in the understanding of electronic TI’s, it is natural to look for corresponding systems in magnets. The route to a Haldane model for magnons turns out to be both simple and elegant: the Heisenberg ferromagnet (FM) on a honeycomb lattice realises magnons with a graphene–like dispersion Fransson et al. (2016), and the symmetry of this lattice permits DM interactions on second–neighbour bonds. These supply the complex hopping integral invoked by Haldane, opening a gap in the magnon dispersion, and endowing the bands with Chern numbers Kim et al. (2016); Kim and Kee (2017). By extension, an exact analogue of the Kane–Mele model can be realised in a bilayer honeycomb magnet, with interlayer interactions chosen such that it forms two copies of a Haldane model, with magnon bands related by time–reversal symmetry Kondo et al. (2019a), an approach which can be extended to the Fu–Kane–Mele model in three dimensions Kondo et al. (2019b). It is also possible to achieve triplon bands which mirror the Kane–Mele model, in a quantum paramagnet on a bilayer honeycomb lattice Joshi and Schnyder (2019).

While the route to topological bands in magnets is now well established, a number of important questions remain. In particular, most work to date has taken a “top–down” approach, emphasizing how topological effects found in electronic systems can be recreated within the band–like excitations of magnets. Less attention has been paid to features which may be unique to magnets, or to building models of topological phases in magnets from the “ground up”, starting from the most general spin interactions allowed by symmetry of a given lattice. One risk inherent in the “top down” approach is that the symmetries which protect topological phases formed by electrons, need not protect those formed by spins. For example, while the Sz=±1S^{z}=\pm 1 triplet excitations of a quantum paramagnet form a doublet under time–reversal symmetry, they do not satisfy Kramer’s theorem. This means that the Sz=±1S^{z}=\pm 1 doublets are much more “fragile” than the electronic doublets considered by Kane and Mele, and the consequences of any symmetry–allowed terms in the Hamiltonian which mix triplets with Sz=±1S^{z}=\pm 1, therefore need to be considered explicitly. It is also of interest to ask what such a topological quantum paramagnet would look like in experiment, and how this physics might generalize to structures more complicated than the honeycomb lattice.

In this Article, we address these questions in the context of a model a spin–1/2 Kagome bilayer, which provides an analogue to the 𝒵2\mathcal{Z}_{2} topological insulators considered by Kane and Mele Kane and Mele (2005a, b); Fu and Kane (2007). We take a limit in which the ground state is a quantum paramagnet, formed by inter–layer dimers, with nine distinct bands of triplet excitations. We consider first the case where the spin of triplet excitations is conserved, and show that in zero magnetic field the Sz=±1S^{z}=\pm 1 triplet bands can realize an analogue to the quantum spin Hall insulator Kane and Mele (2005a); Fu and Kane (2007); Moore and Balents (2007); Roy (2009). In this case, triplon bands are characterized by a nonzero 𝒵2\mathcal{Z}_{2} invariant, and in open geometries we find corresponding helical triplet edge modes. Furthermore, when the time reversal symmetry is broken by applied magnetic field, the system evolves into a Chern insulator Haldane (1988) characterized by chiral triplon edge modes appearing in a finite sample. We compute the spin Nernst, and thermal Hall responses marking the nontrivial topology in these phases.

We explore the consequences of the spin–mixing terms allowed by the symmetry of the lattice, and discuss their effect on the 𝒵2\mathcal{Z}_{2} and Chern bands. Such spin–mixing terms are also present in the original model of Kane and Mele, in the form of the Rashba coupling Kane and Mele (2005a, b). In that case, the “up” and “down” spin states of an electron form a Kramers pair, enforcing the twofold degeneracy of the bands at certain points in the Brillouin zone. This guarantees the perturbative stability of spin–Hall state against small values of Rashba coupling. However such a protection is not guaranteed by time–reversal symmetry in the case of the triplons, where the representation of the time–reversal operator squares to one. We examine this difference closely, and find that even infinitesimal nematic interactions can eliminate the 𝒵2\mathcal{Z}_{2} topological invariant, opening a gap to the associated helical edge modes.

We further identify an operator Θ\Theta=TR×\timesU(1) which, within a Bogoliubov-de Gennes approach, encodes the symmetry needed to protect a 𝒵2\mathcal{Z}_{2} topological phase in either a quantum paramagnet, or an ordered phase with topological magnon bands. We give a detailed analysis on the commutation of the various terms in the triplet Hamiltonian with Θ\Theta, confirming that the nematic interaction breaks this symmetry, compromising the 𝒵2\mathcal{Z}_{2} band-topology.

We also discuss the implications of these results for the bilayer models considered in Refs. Joshi and Schnyder, 2019 and Kondo et al., 2019a, where identical symmetry considerations apply.

The remainder of the Article is organized as follows. In Section II we give a detailed analysis on the symmetry allowed intra-dimer and first neighbor inter-dimer interactions and introduce the bilayer kagome model. Sec. III is devoted to the Bogoliubov–de Gennes Hamiltonian describing the triplet dynamics, and provides a detailed discussion on the band touching topological transitions appearing as the anisotropies change. In Sec. VI we show that the triplet excitations can provide an analog to the Kane and Mele model Kane and Mele (2005a), characterized by nonzero 𝒵2\mathcal{Z}_{2} topological invariant. We calculate the Nernst effect of triplets, a transverse spin current arising as a response to a temperature gradient. Sec. VII is devoted to the protection of the 𝒵2\mathcal{Z}_{2} topology. We show that the TR×\timesU(1) symmetry corresponds to a pseudo time reversal operation, which, if present, protects the 𝒵2\mathcal{Z}_{2} bands. However, symmetric exchange anisotropies, arising in the form of nematic interactions, break this symmetry, together with the fragile 𝒵2\mathcal{Z}_{2} phase of a non-Kramers pairs. In Sec. VIII we consider the time-reversal symmetry breaking case in the presence of magnetic field. We examine the stability of the triplet Chern bands. Additionally, we compute the thermal Hall signal of the Chern-ful triplon bands. Sec. IX provides a brief summary of our results.

II Symmetry-allowed Hamiltonian

Refer to caption
Figure 1: (a) Top view of the bilayer kagome lattice with the translation vectors, 𝜹1=(1/2,3/2)\bm{\delta}_{1}=\left(1/2,\sqrt{3}/2\right), 𝜹2=(1,0)\bm{\delta}_{2}=\left(-1,0\right), and 𝜹3=(1/2,3/2)\bm{\delta}_{3}=\left(1/2,-\sqrt{3}/2\right). Bottom panel: Hexagonal Brillouin zone with time reversal invariant momenta (TRIM) Γ\Gamma, 𝖬𝟣\mathsf{M_{1}}, 𝖬𝟤\mathsf{M_{2}}, and 𝖬𝟥\mathsf{M_{3}} (red), as well as the Dirac points K, and K’ (blue). (b) Symmetry operations of the bilayer-kagome lattice. The smallest unit exhibiting the full symmetry of the bilayer kagome lattice is a bilayer hexagon with six inter–layer dimers. The indices AA, BB, and CC and following the notation of the main text corresponding to the sublattice flavor, while the AA^{\prime} denotes the dimer AA shifted with the lattice translation vector 𝜹3\bm{\delta}_{3}, BB^{\prime} is dimer BB shifted with 𝜹1\bm{\delta}_{1} and CC^{\prime} corresponds to dimer CC translated by 𝜹2\bm{\delta}_{2}.

The model we consider is the spin–1/2 magnet on the bilayer Kagome lattice, shown in Fig. 1(a). We first establish the most general form of first–neighbour and second–neighbour interactions allowed by the D6h symmetry of this lattice. The resulting model,

D6h1,2=XXZ+DM+Nematic,\mathcal{H}^{\text{1,2}}_{\text{D${}_{6h}$}}=\mathcal{H}_{\text{XXZ}}+\mathcal{H}_{\text{DM}}+\mathcal{H}_{\text{Nematic}}\;, (1)

has fifteen adjustable parameters, and contains terms which we can group as symmetric XXZ exchanges, XXZ\mathcal{H}_{\text{XXZ}}; antisymmetric Dzyaloshinskii-Moriya (DM) interactions, DM\mathcal{H}_{\text{DM}}; and symmetric diagonal and off-diagonal exchange anisotropies, which we name the "bond–nematic" interactions (defined below), Nematic\mathcal{H}_{\text{Nematic}}.

To determine the different contributions to Eq. (1), we consider the hexagonal prism formed by six interlayer spin–dimers [Fig. 1(b)], which forms the smallest building block with the full symmetry of lattice. In what follows, we refer to these interlayer dimers simply as “dimers”, and we will ultimately build topological bands from the triplet excitations of a quantum paramgnet formed by singlets on these dimer bonds. The bilayer hexagon contains three such dimers AA, BB, and CC that lie within the kagome unit cell, as well as three more, AA^{\prime}, BB^{\prime}, and CC^{\prime}, which correspond to neighboring dimers translated by 𝜹3\bm{\delta}_{3}, 𝜹1\bm{\delta}_{1}, and 𝜹2\bm{\delta}_{2}, respectively [Fig. 1(a)].

We proceed by analysing all possible interactions bilinear in spins, starting from the transformation properties of individual spin components under three generators of the D6h point group, C6, C2{}^{\prime}_{2}, and σh\sigma_{h} [Table 1]. All remaining group elements can be constructed as a combinations of these three operations. The different types of term which arise are considered, bond by bond, below.

Table 1: Transformation of the spin (axial vector) components, the dimers, and the site indices under the generators of the D6h point group.
Generators E C6 C2{}^{\prime}_{2} σh\sigma_{h}
spin component SxS^{x} 12Sx+32Sy\frac{1}{2}S^{x}+\frac{\sqrt{3}}{2}S^{y} SxS^{x} Sx-S^{x}
SyS^{y} 32Sx+12Sy-\frac{\sqrt{3}}{2}S^{x}+\frac{1}{2}S^{y} Sy-S^{y} Sy-S^{y}
SzS^{z} SzS^{z} Sz-S^{z} SzS^{z}
dimer label AA CC^{\prime} BB AA
BB AA^{\prime} AA BB
CC BB^{\prime} CC CC
AA^{\prime} CC BB^{\prime} AA^{\prime}
BB^{\prime} AA AA^{\prime} BB^{\prime}
CC^{\prime} BB CC^{\prime} CC^{\prime}
layer index 11 11 22 22
22 22 11 11

II.1 Intra-dimer interactions

The symmetry classification of the intra-dimer interactions according to the D6h symmetry group yields three invariant terms. Two correspond to the Heisenberg exchange anisotropy distinguishing the in-plane and out-of-plane components

XXZ=Jj(Sj1xSj2x+Sj1ySj2y)+JjSj1zSj2z.\mathcal{H}^{\leavevmode\hbox to3.08pt{\vbox to8.77pt{\pgfpicture\makeatletter\hbox{\hskip 229.15997pt\lower-1.5379pt\hbox to0.0pt{\pgfsys@beginscope\pgfsys@invoke{ }\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\pgfsys@setlinewidth{0.4pt}\pgfsys@invoke{ }\nullfont\hbox to0.0pt{\pgfsys@beginscope\pgfsys@invoke{ }] {}{{}}{}{{{}}{}{}{}{}{}{}{}{}}\pgfsys@beginscope\pgfsys@invoke{ }\definecolor[named]{tikz@color}{rgb}{0,0,0}\definecolor[named]{.}{rgb}{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@invoke{ }\pgfsys@color@gray@fill{0}\pgfsys@invoke{ }\definecolor[named]{pgffillcolor}{rgb}{0,0,0}\pgfsys@color@gray@fill{0}\pgfsys@invoke{ }\pgfsys@setlinewidth{0.8pt}\pgfsys@invoke{ }{}\pgfsys@moveto{-227.62207pt}{0.0pt}\pgfsys@moveto{-226.48416pt}{0.0pt}\pgfsys@curveto{-226.48416pt}{0.62845pt}{-226.99362pt}{1.13791pt}{-227.62207pt}{1.13791pt}\pgfsys@curveto{-228.25052pt}{1.13791pt}{-228.75998pt}{0.62845pt}{-228.75998pt}{0.0pt}\pgfsys@curveto{-228.75998pt}{-0.62845pt}{-228.25052pt}{-1.13791pt}{-227.62207pt}{-1.13791pt}\pgfsys@curveto{-226.99362pt}{-1.13791pt}{-226.48416pt}{-0.62845pt}{-226.48416pt}{0.0pt}\pgfsys@closepath\pgfsys@moveto{-227.62207pt}{0.0pt}\pgfsys@fillstroke\pgfsys@invoke{ } \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope{}{{}}{}{{{}}{}{}{}{}{}{}{}{}}\pgfsys@beginscope\pgfsys@invoke{ }\definecolor[named]{tikz@color}{rgb}{0,0,0}\definecolor[named]{.}{rgb}{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@invoke{ }\pgfsys@color@gray@fill{0}\pgfsys@invoke{ }\definecolor[named]{pgffillcolor}{rgb}{0,0,0}\pgfsys@color@gray@fill{0}\pgfsys@invoke{ }\pgfsys@setlinewidth{0.8pt}\pgfsys@invoke{ }{}\pgfsys@moveto{-227.62207pt}{5.69046pt}\pgfsys@moveto{-226.48416pt}{5.69046pt}\pgfsys@curveto{-226.48416pt}{6.31891pt}{-226.99362pt}{6.82837pt}{-227.62207pt}{6.82837pt}\pgfsys@curveto{-228.25052pt}{6.82837pt}{-228.75998pt}{6.31891pt}{-228.75998pt}{5.69046pt}\pgfsys@curveto{-228.75998pt}{5.06201pt}{-228.25052pt}{4.55255pt}{-227.62207pt}{4.55255pt}\pgfsys@curveto{-226.99362pt}{4.55255pt}{-226.48416pt}{5.06201pt}{-226.48416pt}{5.69046pt}\pgfsys@closepath\pgfsys@moveto{-227.62207pt}{5.69046pt}\pgfsys@fillstroke\pgfsys@invoke{ } \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope{}{{}}{} {}{{}}{}{}{}{}{{}}{}\pgfsys@beginscope\pgfsys@invoke{ }\definecolor[named]{tikz@color}{rgb}{0,0,0}\definecolor[named]{.}{rgb}{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@invoke{ }\pgfsys@color@gray@fill{0}\pgfsys@invoke{ }\definecolor[named]{pgffillcolor}{rgb}{1,1,1}\pgfsys@color@gray@fill{1}\pgfsys@invoke{ }\pgfsys@setlinewidth{0.8pt}\pgfsys@invoke{ }{}\pgfsys@moveto{-227.62207pt}{1.13791pt}\pgfsys@moveto{-227.62207pt}{1.13791pt}\pgfsys@lineto{-227.62207pt}{6.82881pt}\pgfsys@lineto{-227.62207pt}{6.82881pt}\pgfsys@lineto{-227.62207pt}{1.13791pt}\pgfsys@closepath\pgfsys@moveto{-227.62207pt}{6.82881pt}\pgfsys@fillstroke\pgfsys@invoke{ } \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope{{ {}{}{}{}{}}{{{}}{{}}}}{}{}\hss}\pgfsys@discardpath\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope\hss}}\lxSVG@closescope\endpgfpicture}}}_{\rm XXZ}=J_{\|}\sum_{j}(S^{x}_{j_{1}}S^{x}_{j_{2}}+S^{y}_{j_{1}}S^{y}_{j_{2}})+J_{\bot}\sum_{j}S^{z}_{j_{1}}S^{z}_{j_{2}}\;. (2)

The third intra-dimer term is a symmetric exchange anisotropy, which we refer to as the bond-nematic interaction:

Nematic=Kj𝐧j𝐐j1,j2.\mathcal{H}^{\leavevmode\hbox to3.08pt{\vbox to8.77pt{\pgfpicture\makeatletter\hbox{\hskip 229.15997pt\lower-1.5379pt\hbox to0.0pt{\pgfsys@beginscope\pgfsys@invoke{ }\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\pgfsys@setlinewidth{0.4pt}\pgfsys@invoke{ }\nullfont\hbox to0.0pt{\pgfsys@beginscope\pgfsys@invoke{ }] {}{{}}{}{{{}}{}{}{}{}{}{}{}{}}\pgfsys@beginscope\pgfsys@invoke{ }\definecolor[named]{tikz@color}{rgb}{0,0,0}\definecolor[named]{.}{rgb}{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@invoke{ }\pgfsys@color@gray@fill{0}\pgfsys@invoke{ }\definecolor[named]{pgffillcolor}{rgb}{0,0,0}\pgfsys@color@gray@fill{0}\pgfsys@invoke{ }\pgfsys@setlinewidth{0.8pt}\pgfsys@invoke{ }{}\pgfsys@moveto{-227.62207pt}{0.0pt}\pgfsys@moveto{-226.48416pt}{0.0pt}\pgfsys@curveto{-226.48416pt}{0.62845pt}{-226.99362pt}{1.13791pt}{-227.62207pt}{1.13791pt}\pgfsys@curveto{-228.25052pt}{1.13791pt}{-228.75998pt}{0.62845pt}{-228.75998pt}{0.0pt}\pgfsys@curveto{-228.75998pt}{-0.62845pt}{-228.25052pt}{-1.13791pt}{-227.62207pt}{-1.13791pt}\pgfsys@curveto{-226.99362pt}{-1.13791pt}{-226.48416pt}{-0.62845pt}{-226.48416pt}{0.0pt}\pgfsys@closepath\pgfsys@moveto{-227.62207pt}{0.0pt}\pgfsys@fillstroke\pgfsys@invoke{ } \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope{}{{}}{}{{{}}{}{}{}{}{}{}{}{}}\pgfsys@beginscope\pgfsys@invoke{ }\definecolor[named]{tikz@color}{rgb}{0,0,0}\definecolor[named]{.}{rgb}{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@invoke{ }\pgfsys@color@gray@fill{0}\pgfsys@invoke{ }\definecolor[named]{pgffillcolor}{rgb}{0,0,0}\pgfsys@color@gray@fill{0}\pgfsys@invoke{ }\pgfsys@setlinewidth{0.8pt}\pgfsys@invoke{ }{}\pgfsys@moveto{-227.62207pt}{5.69046pt}\pgfsys@moveto{-226.48416pt}{5.69046pt}\pgfsys@curveto{-226.48416pt}{6.31891pt}{-226.99362pt}{6.82837pt}{-227.62207pt}{6.82837pt}\pgfsys@curveto{-228.25052pt}{6.82837pt}{-228.75998pt}{6.31891pt}{-228.75998pt}{5.69046pt}\pgfsys@curveto{-228.75998pt}{5.06201pt}{-228.25052pt}{4.55255pt}{-227.62207pt}{4.55255pt}\pgfsys@curveto{-226.99362pt}{4.55255pt}{-226.48416pt}{5.06201pt}{-226.48416pt}{5.69046pt}\pgfsys@closepath\pgfsys@moveto{-227.62207pt}{5.69046pt}\pgfsys@fillstroke\pgfsys@invoke{ } \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope{}{{}}{} {}{{}}{}{}{}{}{{}}{}\pgfsys@beginscope\pgfsys@invoke{ }\definecolor[named]{tikz@color}{rgb}{0,0,0}\definecolor[named]{.}{rgb}{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@invoke{ }\pgfsys@color@gray@fill{0}\pgfsys@invoke{ }\definecolor[named]{pgffillcolor}{rgb}{1,1,1}\pgfsys@color@gray@fill{1}\pgfsys@invoke{ }\pgfsys@setlinewidth{0.8pt}\pgfsys@invoke{ }{}\pgfsys@moveto{-227.62207pt}{1.13791pt}\pgfsys@moveto{-227.62207pt}{1.13791pt}\pgfsys@lineto{-227.62207pt}{6.82881pt}\pgfsys@lineto{-227.62207pt}{6.82881pt}\pgfsys@lineto{-227.62207pt}{1.13791pt}\pgfsys@closepath\pgfsys@moveto{-227.62207pt}{6.82881pt}\pgfsys@fillstroke\pgfsys@invoke{ } \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope{{ {}{}{}{}{}}{{{}}{{}}}}{}{}\hss}\pgfsys@discardpath\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope\hss}}\lxSVG@closescope\endpgfpicture}}}_{\rm Nematic}=K_{\|}\sum_{j}\mathbf{n}_{j}\cdot\mathbf{Q}^{\|}_{j_{1},j_{2}}\;. (3)

The index jj runs over the dimers, and 11 and 22 denote the layer indices of dimer-jj. The vectors 𝐧j\mathbf{n}_{j} appearing in the nematic term have the form 𝐧A=(12,32)\mathbf{n}_{A}=\left(\frac{1}{2},\frac{\sqrt{3}}{2}\right), 𝐧B=(12,32)\mathbf{n}_{B}=\left(\frac{1}{2},-\frac{\sqrt{3}}{2}\right), and 𝐧C=(1,0)\mathbf{n}_{C}=\left(-1,0\right), and 𝐐j1,j2\mathbf{Q}^{\|}_{j_{1},j_{2}} denotes the vector (Qj1,j2x2y2,Qj1,j2xy)(Q^{x^{2}\!-\!y^{2}}_{j_{1},j_{2}},Q^{xy}_{j_{1},j_{2}}) made of the nematic interactions

Qi,jx2y2=SixSjxSiySjy,\displaystyle Q^{x^{2}\!-\!y^{2}}_{i,j}=S^{x}_{i}S^{x}_{j}-S^{y}_{i}S^{y}_{j}\;, (4a)
Qi,jxy=SixSjy+SiySjx.\displaystyle Q^{xy}_{i,j}=S^{x}_{i}S^{y}_{j}+S^{y}_{i}S^{x}_{j}\;. (4b)

We illustrate the nematic operators, Qx2y2Q^{x^{2}\!-\!y^{2}} and QxyQ^{xy}, in the fashion of the d-orbitals, dx2y2d^{x^{2}\!-\!y^{2}} and dxyd^{xy}, as they transform in the same way. Fig. 2 introduces our schematic illustration of the in-plane nematic operators, Qx2y2Q^{x^{2}\!-\!y^{2}} and QxyQ^{xy}, and their linear combinations as appear in the intra-dimer interactions, together with the directions of the 𝐧\mathbf{n} vectors.

Refer to caption
Figure 2: (a) Illustration of the components, Qx2y2Q^{x^{2}\!-\!y^{2}} and QxyQ^{xy}, of the intra-dimer nematic interactions. The pink arrows indicates the direction in the 2-dimensional vector-spaces, 𝐐=(Qx2y2,Qxy)\mathbf{Q}^{\|}=(Q^{x^{2}\!-\!y^{2}},Q^{xy}). (b) The linear combinations appearing in the intra-dimer in-plane nematic terms. We plot the vectors nn and the schematic representation of the operators in the same figure. The sketches in the top right corner show the simplified illustration for a given linear combination. (c) Intra-dimer nematic operators on the bonds, and (d) the vectors 𝐧A\mathbf{n}_{A}, 𝐧B\mathbf{n}_{B}, and 𝐧C\mathbf{n}_{C}, corresponding to the linear combination on each dimer.

Due to the bond-inversion, the antisymmetric exchange anisotropy, i.e. the Dzyaloshinskii–Moriya interaction is not allowed on the dimers.

II.2 First neighbor inter-dimer interactions

Classifying the first neighbor inter-dimer interactions, we find six independent operators that transform as the fully symmetric irreducible representation. Beside the anisotropy in the Heisenberg exchange, distinguishing the in-plane and out-of-plane components

XXZ1st\displaystyle\mathcal{H}^{\rm 1st}_{\rm XXZ} =\displaystyle= Ji,jl=1,2(SilxSjlx+SilySjly)+Ji,jl=1,2SilzSjlz,\displaystyle J^{\prime}_{\|}\sum_{\begin{subarray}{c}\langle i,j\rangle\\ l=1,2\end{subarray}}(S^{x}_{i_{l}}S^{x}_{j_{l}}+S^{y}_{i_{l}}S^{y}_{j_{l}})+J^{\prime}_{\bot}\sum_{\begin{subarray}{c}\langle i,j\rangle\\ l=1,2\end{subarray}}S^{z}_{i_{l}}S^{z}_{j_{l}}, (5)

there are additional four operators, namely the in-plane and out-of-plane components of the nematic and DM interactions. The DM interaction has the form

DM1st\displaystyle\mathcal{H}^{\rm 1st}_{\text{DM}} =\displaystyle= i,j𝐃(𝐒i×𝐒j),\displaystyle\sum_{\langle i,j\rangle}\mathbf{D}^{\prime}\cdot(\mathbf{S}_{i}\times\mathbf{S}_{j})\;, (6)

and the in-plane (DD^{\prime}_{\|}) and out-of-plane (DD^{\prime}_{\bot}) components of the vector 𝐃\mathbf{D}^{\prime} are shown in Fig. 3.

Refer to caption
Figure 3: (a-b) DM vector components on the first, and (c-d) on the second neighbor inter-dimer bonds in the top and bottom layers, respectively. For both, the first and second neighbor DM interactions the in-plane components have opposite signs in the top and bottom layer, while the out-of-plane components are the same.

DD^{\prime}_{\bot} is uniform in the top and bottom layers, but the DD^{\prime}_{\|} components change sign under exchanging the layers. As discussed in Section III, only the uniform out-of-plane DM component DD^{\prime}_{\bot} appears in Bogoliubov–de Gennes Hamiltonian describing the triplet dynamics.

In addition, there are in-plane and out-of-plane bond-nematic terms. The in-plane component has the form of

1st=Ki,jl=1,2𝐧ij𝐐il,jl,\mathcal{H}^{\rm 1st}_{\|}=K^{\prime}_{\|}\sum_{\langle i,j\rangle}\sum_{l=1,2}\mathbf{n}_{ij}\cdot\mathbf{Q}^{\|}_{il,jl}\;, (7)

where 𝐐il,jl\mathbf{Q}^{\|}_{il,jl} is defined the same way as in Eq 4, the index ll takes the value 11 for the top, and 22 for the bottom layer, while ii and jj denote first neighbor sites within the layers. The vectors 𝐧ij\mathbf{n}_{ij} have the form

𝐧AB\displaystyle\mathbf{n}_{A^{\prime}B} =\displaystyle= 𝐧AB=(1,0),\displaystyle\mathbf{n}_{AB^{\prime}}=\left(-1,0\right)\;, (8)
𝐧BC\displaystyle\mathbf{n}_{B^{\prime}C} =\displaystyle= 𝐧BC=(12,32),\displaystyle\mathbf{n}_{BC^{\prime}}=\left(\frac{1}{2},\frac{\sqrt{3}}{2}\right)\;, (9)
𝐧CA\displaystyle\mathbf{n}_{C^{\prime}A} =\displaystyle= 𝐧CA=(12,32).\displaystyle\mathbf{n}_{CA^{\prime}}=\left(\frac{1}{2},-\frac{\sqrt{3}}{2}\right)\;. (10)

The in-plane component of the inter-dimer nematic interactions on the first neighbors are shown in Fig. 4, where we use the same notation introduced in Fig. 2 (a) and (b).

As with the out–of–plane DM components, the in–plane nematic terms are uniform in the top and bottom layers, and so will give a finite contribution to a Hamiltonian for triplets.

Refer to caption
Figure 4: (a) Inter-dimer in-plane symmetric exchange anisotropy 𝐐i,j\mathbf{Q}^{\|}_{i,j} in the top (left) and bottom (right) layers. (b) Components of the inter-dimer nematic operators between the first neighbor bonds.

The first-neighbor out-of-plane bond-nematic interaction is

1st=Ki,j𝐧ij𝐐il,jl,\mathcal{H}^{\rm 1st}_{\bot}=K^{\prime}_{\bot}\sum_{\langle i,j\rangle}\mathbf{n}^{\bot}_{ij}\cdot\mathbf{Q}^{\bot}_{il,jl}\;, (11)

where the vectors 𝐧ij\mathbf{n}^{\bot}_{ij} are shown in Fig. 5(b), and the components of 𝐐i,j=(Qi,jzx,Qi,jyz)\mathbf{Q}^{\bot}_{i,j}=(Q^{zx}_{i,j},Q^{yz}_{i,j}) correspond to the bond-nematic operators

Qi,jxz=SizSjx+SixSjz,\displaystyle Q^{xz}_{i,j}=S^{z}_{i}S^{x}_{j}+S^{x}_{i}S^{z}_{j}\;, (12a)
Qi,jyz=SiySjz+SizSjy.\displaystyle Q^{yz}_{i,j}=S^{y}_{i}S^{z}_{j}+S^{z}_{i}S^{y}_{j}\;. (12b)

We utilize the representation of the dyzd^{yz} and dzxd^{zx} orbitals, to illustrate the out-of-plane bond-nematic terms as shown in Fig. 5.

Refer to caption
Figure 5: (a) Illustration of the out-of-plane nematic interaction components, QzxQ^{zx} and QyzQ^{yz}. The panels on the right show the top-view. (b) Some representative out-of-plane nematic terms and the corresponding 𝐧ij\mathbf{n}^{\bot}_{ij} vectors (pink arrow), defining the linear combination of QzxQ^{zx} and QyzQ^{yz}. (c) Inter-dimer out-of-plane symmetric exchange anisotropy in the top (left) and bottom layer (right). (d) Directions of the first neighbor out-of-plane nematic vectors in the top layer. The directions in the bottom layer are the opposite.

Once again, the out-of-plane inter-dimer symmetric exchange anisotropy term is the opposite in the top and bottom layers. For this reason, as with the in-plane DM vectors, it will cancel in the triplet hopping Hamiltonian discussed in Section III.

II.3 Second neighbor inter–dimer interactions

We also consider the effect of second–neighbour interactions within the planes of the Kagome lattice. As with the first–neighbor interactions, there are six different terms: XXZ exchange J′′J^{\prime\prime}_{\|} and J′′J^{\prime\prime}_{\bot}; in–plane and out–of–plane DM interactions, D′′D^{\prime\prime}_{\|} and D′′D^{\prime\prime}_{\bot}; and the in–plane and out–of–plane bond-nematic terms K′′K^{\prime\prime}_{\|} and K′′K^{\prime\prime}_{\bot}. XXZ interactions are defined through

XXZ2nd\displaystyle\mathcal{H}^{\rm 2nd}_{\rm XXZ} =\displaystyle= J′′i,jl=1,2SilxSjlx+SilySjly\displaystyle J^{\prime\prime}_{\|}\sum_{\langle\langle i,j\rangle\rangle}\sum_{l=1,2}S^{x}_{i_{l}}S^{x}_{j_{l}}+S^{y}_{i_{l}}S^{y}_{j_{l}} (13)
+J′′i,jl=1,2SilzSjlz,\displaystyle+J^{\prime\prime}_{\bot}\sum_{\langle\langle i,j\rangle\rangle}\sum_{l=1,2}S^{z}_{i_{l}}S^{z}_{j_{l}}\;,

where the sum i,j\langle\langle i,j\rangle\rangle runs over second–neighbor bonds within the Kagome layers. Similarly, DM interactions are defined through

DM2nd\displaystyle\mathcal{H}^{\rm 2nd}_{\text{DM}} =\displaystyle= i,j𝐃′′(𝐒i×𝐒j),\displaystyle\sum_{\langle\langle i,j\rangle\rangle}\mathbf{D}^{\prime\prime}\cdot(\mathbf{S}_{i}\times\mathbf{S}_{j})\;, (14)

where the components of the associated DM vectors in the top and bottom layers are shown in Fig. 3 (d) and (e), respectively.

The additional bond–nematic–type interactions have the form

2nd=K′′i,jl=1,2𝐧ij𝐐il,jl,\mathcal{H}^{\rm 2nd}_{\|}=K^{\prime\prime}_{\|}\sum_{\langle\langle i,j\rangle\rangle}\sum_{l=1,2}\mathbf{n}_{ij}\cdot\mathbf{Q}^{\|}_{il,jl}\;, (15)

where the vector operator, 𝐐il,jl\mathbf{Q}^{\|}_{il,jl} is defined in Eq. (4); ll takes the values 11 and 22 for top and bottom layers, respectively and the vectors 𝐧ij\mathbf{n}_{ij} have the form

𝐧AB\displaystyle\mathbf{n}_{AB} =\displaystyle= 𝐧AB=(1,0),\displaystyle\mathbf{n}_{A^{\prime}B^{\prime}}=\left(1,0\right)\;, (16)
𝐧BC\displaystyle\mathbf{n}_{BC} =\displaystyle= 𝐧BC=(12,32),\displaystyle\mathbf{n}_{B^{\prime}C^{\prime}}=\left(-\frac{1}{2},-\frac{\sqrt{3}}{2}\right)\;, (17)
𝐧CA\displaystyle\mathbf{n}_{CA} =\displaystyle= 𝐧CA=(12,32).\displaystyle\mathbf{n}_{C^{\prime}A^{\prime}}=\left(-\frac{1}{2},\frac{\sqrt{3}}{2}\right)\;. (18)
Refer to caption
Figure 6: (a) Inter-dimer in-plane symmetric exchange anisotropy 𝐐i,j\mathbf{Q}^{\|}_{i,j} between the second neighbors in the top (left) and bottom (right) layers. (b) Components of the inter-dimer nematic operators.

The out-of-plane bond-nematic term is

2nd=K′′i,jl=1,2𝐧ij𝐐il,jl,\mathcal{H}^{\rm 2nd}_{\bot}=K^{\prime\prime}_{\bot}\sum_{\langle\langle i,j\rangle\rangle}\sum_{l=1,2}\mathbf{n}^{\bot}_{ij}\cdot\mathbf{Q}^{\bot}_{il,jl}\;, (19)

where the components of 𝐐i,j=(Qil,jlzx,Qil,jlyz)\mathbf{Q}^{\bot}_{i,j}=(Q^{zx}_{il,jl},Q^{yz}_{il,jl}) are introduced in Eqs. 12 and the associated vectors 𝐧ij\mathbf{n}^{\bot}_{ij} are shown in Fig. 7(b).

Refer to caption
Figure 7: (a) Inter-dimer out-of-plane symmetric exchange anisotropy on the second neighbor bonds in the top (left) and bottom layer (right). We use the notation for the linear combinations of QzxQ^{zx} and QyzQ^{yz} bond-nematic operators as shown in Fig. 5. (b) Directions of the second neighbor out-of-plane nematic vectors in the top layer. The directions in the bottom layer are the opposite.

III Model of triplon bands

Having established the most general form of interactions allowed by symmetry, we now show how this determines the Hamiltonian for triplet excitations of a quantum paramagnet on the bilayer Kagome lattice. We begin by defining the model [Section III.1], before proceeding to a Hamiltonian expressed in terms of a bond–wave formalism, with explicit Bogoliubov–de Gennes form [Section III.2]. Finally, we set up the framework for subsequent discussion of topological bands, by solving for the (topologically–trivial) triplon bands in the limit of vanishing anisotropic exchange [Section III.3]. Throughout this Article, we work with the full Bogoliubov–de Gennes Hamiltonian for triplet excitations, except where otherwise stated.

III.1 Microscopic Model

We start from a model defined by

\displaystyle\mathcal{H} =\displaystyle= D6h1,2+Zeeman\displaystyle\mathcal{H}^{\text{1,2}}_{\text{D${}_{6h}$}}+\mathcal{H}_{\text{Zeeman}} (20)
=\displaystyle= XXZ+DM+Nematic+Zeeman,\displaystyle\mathcal{H}_{\text{XXZ}}+\mathcal{H}_{\text{DM}}+\mathcal{H}_{\text{Nematic}}+\mathcal{H}_{\text{Zeeman}}\;, (21)

where D6h1,2\mathcal{H}^{\text{1,2}}_{\text{D${}_{6h}$}} [Eq. (1)] is the most general Hamiltonian for a spin–1/2 magnet on a bilayer Kagome lattice with first– and second–neighbour interactions [cf. Section II], and

Zeeman=gzhziSiz,\displaystyle\mathcal{H}_{\text{Zeeman}}=-g_{z}h^{z}\sum_{i}S^{z}_{i}\;, (22)

encodes the effect of a magnetic field perpendicular to the plane of the Kagome lattice.

The largest term in this model is taken to be (approximately) Heisenberg exchange JJJ_{\parallel}\approx J_{\perp} on intra–dimer bonds [Section II.1]. Where these interactions are antiferromagnetic, and sufficiently large compared with other terms, the ground state of XXZ\mathcal{H}_{\text{XXZ}} is a quantum paramagnet formed by a product of singlets on all dimer bonds

Ψ0=j|sj,\displaystyle\mid\Psi_{0}\rangle=\prod_{j\in\leavevmode\hbox to3.08pt{\vbox to8.77pt{\pgfpicture\makeatletter\hbox{\hskip 229.15997pt\lower-1.5379pt\hbox to0.0pt{\pgfsys@beginscope\pgfsys@invoke{ }\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\pgfsys@setlinewidth{0.4pt}\pgfsys@invoke{ }\nullfont\hbox to0.0pt{\pgfsys@beginscope\pgfsys@invoke{ }] {}{{}}{}{{{}}{}{}{}{}{}{}{}{}}\pgfsys@beginscope\pgfsys@invoke{ }\definecolor[named]{tikz@color}{rgb}{0,0,0}\definecolor[named]{.}{rgb}{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@invoke{ }\pgfsys@color@gray@fill{0}\pgfsys@invoke{ }\definecolor[named]{pgffillcolor}{rgb}{0,0,0}\pgfsys@color@gray@fill{0}\pgfsys@invoke{ }\pgfsys@setlinewidth{0.8pt}\pgfsys@invoke{ }{}\pgfsys@moveto{-227.62207pt}{0.0pt}\pgfsys@moveto{-226.48416pt}{0.0pt}\pgfsys@curveto{-226.48416pt}{0.62845pt}{-226.99362pt}{1.13791pt}{-227.62207pt}{1.13791pt}\pgfsys@curveto{-228.25052pt}{1.13791pt}{-228.75998pt}{0.62845pt}{-228.75998pt}{0.0pt}\pgfsys@curveto{-228.75998pt}{-0.62845pt}{-228.25052pt}{-1.13791pt}{-227.62207pt}{-1.13791pt}\pgfsys@curveto{-226.99362pt}{-1.13791pt}{-226.48416pt}{-0.62845pt}{-226.48416pt}{0.0pt}\pgfsys@closepath\pgfsys@moveto{-227.62207pt}{0.0pt}\pgfsys@fillstroke\pgfsys@invoke{ } \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope{}{{}}{}{{{}}{}{}{}{}{}{}{}{}}\pgfsys@beginscope\pgfsys@invoke{ }\definecolor[named]{tikz@color}{rgb}{0,0,0}\definecolor[named]{.}{rgb}{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@invoke{ }\pgfsys@color@gray@fill{0}\pgfsys@invoke{ }\definecolor[named]{pgffillcolor}{rgb}{0,0,0}\pgfsys@color@gray@fill{0}\pgfsys@invoke{ }\pgfsys@setlinewidth{0.8pt}\pgfsys@invoke{ }{}\pgfsys@moveto{-227.62207pt}{5.69046pt}\pgfsys@moveto{-226.48416pt}{5.69046pt}\pgfsys@curveto{-226.48416pt}{6.31891pt}{-226.99362pt}{6.82837pt}{-227.62207pt}{6.82837pt}\pgfsys@curveto{-228.25052pt}{6.82837pt}{-228.75998pt}{6.31891pt}{-228.75998pt}{5.69046pt}\pgfsys@curveto{-228.75998pt}{5.06201pt}{-228.25052pt}{4.55255pt}{-227.62207pt}{4.55255pt}\pgfsys@curveto{-226.99362pt}{4.55255pt}{-226.48416pt}{5.06201pt}{-226.48416pt}{5.69046pt}\pgfsys@closepath\pgfsys@moveto{-227.62207pt}{5.69046pt}\pgfsys@fillstroke\pgfsys@invoke{ } \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope{}{{}}{} {}{{}}{}{}{}{}{{}}{}\pgfsys@beginscope\pgfsys@invoke{ }\definecolor[named]{tikz@color}{rgb}{0,0,0}\definecolor[named]{.}{rgb}{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@invoke{ }\pgfsys@color@gray@fill{0}\pgfsys@invoke{ }\definecolor[named]{pgffillcolor}{rgb}{1,1,1}\pgfsys@color@gray@fill{1}\pgfsys@invoke{ }\pgfsys@setlinewidth{0.8pt}\pgfsys@invoke{ }{}\pgfsys@moveto{-227.62207pt}{1.13791pt}\pgfsys@moveto{-227.62207pt}{1.13791pt}\pgfsys@lineto{-227.62207pt}{6.82881pt}\pgfsys@lineto{-227.62207pt}{6.82881pt}\pgfsys@lineto{-227.62207pt}{1.13791pt}\pgfsys@closepath\pgfsys@moveto{-227.62207pt}{6.82881pt}\pgfsys@fillstroke\pgfsys@invoke{ } \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope{{ {}{}{}{}{}}{{{}}{{}}}}{}{}\hss}\pgfsys@discardpath\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope\hss}}\lxSVG@closescope\endpgfpicture}}}|s\rangle_{j}\;, (23)

where

|sj=12(|12|12)j.\displaystyle\left|s\right>_{j}=\frac{1}{\sqrt{2}}(\left|\uparrow_{1}\downarrow_{2}\right>-\left|\downarrow_{1}\uparrow_{2}\right>)_{j}\;. (24)

The low–lying excitations of this quantum paramagnet will be spin-1 triplet excitations,

|t1j=|12j\displaystyle\left|t_{1}\right>_{j}=\left|\uparrow_{1}\uparrow_{2}\right>_{j} (25a)
|t0j=12(|12+|12)j\displaystyle\left|t_{0}\right>_{j}=\frac{1}{\sqrt{2}}(\left|\uparrow_{1}\downarrow_{2}\right>+\left|\downarrow_{1}\uparrow_{2}\right)_{j} (25b)
|t1j=|12j,\displaystyle\left|t_{-1}\right>_{j}=\left|\downarrow_{1}\downarrow_{2}\right>_{j}\;, (25c)

where the indices 11 and 22 denote the top and bottom sites of the dimer, and jj labels the dimer. In the presence of terms connecting different dimers, these triplet excitations will form dispersing bands of excitations, usually referred to as “triplons”.

The spin operators can be expressed using the singlet-triplet basis

Sj1+\displaystyle S^{+}_{j_{1}} =\displaystyle\!\!=\!\! 12(t1,jt0,jt1,jsj+t0,jt1,j+sjt1,j)\displaystyle\frac{1}{\sqrt{2}}\left(t^{\dagger}_{1,j}t^{\phantom{\dagger}}_{0,j}\!-\!t^{\dagger}_{1,j}s^{\phantom{\dagger}}_{j}+t^{\dagger}_{0,j}t^{\phantom{\dagger}}_{-1,j}\!+\!s^{\dagger}_{j}t^{\phantom{\dagger}}_{-1,j}\right) (26a)
Sj1z\displaystyle S^{z}_{j_{1}} =\displaystyle\!\!=\!\! 12(t1,jt1,j+t0,jsj+sjt0,jt1,jt1,j)\displaystyle\frac{1}{2}\left(t^{\dagger}_{1,j}t^{\phantom{\dagger}}_{1,j}+t^{\dagger}_{0,j}s^{\phantom{\dagger}}_{j}+s^{\dagger}_{j}t^{\phantom{\dagger}}_{0,j}-t^{\dagger}_{-1,j}t^{\phantom{\dagger}}_{-1,j}\right) (26b)
Sj1\displaystyle S^{-}_{j_{1}} =\displaystyle\!\!=\!\! 12(t0,jt1,jsjt1,j+t1,jt0,j+t1,jsj)\displaystyle\frac{1}{\sqrt{2}}\left(t^{\dagger}_{0,j}t^{\phantom{\dagger}}_{1,j}\!-\!s^{\dagger}_{j}t^{\phantom{\dagger}}_{1,j}+t^{\dagger}_{-1,j}t^{\phantom{\dagger}}_{0,j}\!+\!t^{\dagger}_{-1,j}s^{\phantom{\dagger}}_{j}\right) (26c)
Sj2+\displaystyle S^{+}_{j_{2}} =\displaystyle\!\!=\!\! 12(t1,jt0,j+t1,jsj+t0,jt1,jsjt1,j)\displaystyle\frac{1}{\sqrt{2}}\left(t^{\dagger}_{1,j}t^{\phantom{\dagger}}_{0,j}\!+\!t^{\dagger}_{1,j}s^{\phantom{\dagger}}_{j}+t^{\dagger}_{0,j}t^{\phantom{\dagger}}_{-1,j}\!-\!s^{\dagger}_{j}t^{\phantom{\dagger}}_{-1,j}\right) (26d)
Sj2z\displaystyle S^{z}_{j_{2}} =\displaystyle\!\!=\!\! 12(t1,jt1,jt0,jsjsjt0,jt1,jt1,j)\displaystyle\frac{1}{2}\left(t^{\dagger}_{1,j}t^{\phantom{\dagger}}_{1,j}-t^{\dagger}_{0,j}s^{\phantom{\dagger}}_{j}-s^{\dagger}_{j}t^{\phantom{\dagger}}_{0,j}-t^{\dagger}_{-1,j}t^{\phantom{\dagger}}_{-1,j}\right) (26e)
Sj2\displaystyle S^{-}_{j_{2}} =\displaystyle\!\!=\!\! 12(t0,jt1,j+sjt1,j+t1,jt0,jt1,jsj)\displaystyle\frac{1}{\sqrt{2}}\left(t^{\dagger}_{0,j}t^{\phantom{\dagger}}_{1,j}\!+\!s^{\dagger}_{j}t^{\phantom{\dagger}}_{1,j}+t^{\dagger}_{-1,j}t^{\phantom{\dagger}}_{0,j}\!-\!t^{\dagger}_{-1,j}s^{\phantom{\dagger}}_{j}\right) (26f)

where the indices 11 and 22 denote the top and bottom site (layer) of the vertical dimer, tm,j|0=|tmjt^{\dagger}_{m,j}\left|0\right>=\left|t_{m}\right>_{j}, sj|0=|sjs^{\dagger}_{j}\left|0\right>=\left|s\right>_{j} with m=1,0,1m=-1,0,1 and the xx and yy components of the spin operators are Sjx=12(Sj++Sj)S^{x}_{j}=\frac{1}{2}(S^{+}_{j}+S^{-}_{j}) and Sjy=i2(Sj+Sj)S^{y}_{j}=-\frac{i}{2}(S^{+}_{j}-S^{-}_{j}).

In general, all components of the different exchange interactions within the Kagome planes will contribute to triplon dispersion. For small to moderate spin–orbit coupling, the largest contributions will come from (approximately) Heisenberg interactions on first–neighbour bonds JJJ^{\prime}_{\parallel}\approx J^{\prime}_{\perp}. However the Berry phase underpinning a topological bandstructure originates in the DM interactions, DM\mathcal{H}_{\text{DM}}. And the fate of these topological bands will in turn depend on the nematic interactions, Nematic\mathcal{H}_{\text{Nematic}}. A few comments are therefore due on how these enter the problem.

Our model allows for DM interactions on first–neighbour bonds, 𝐃\mathbf{D}^{\prime} [Eq. (6)], and second–neighbour bonds, 𝐃′′\mathbf{D}^{\prime\prime} [Eq. (14)], illustrated in Fig. 3. The inter-layer bond–inversion symmetry precludes any DM interaction acting on the local dimers, and consequently the mixing between the odd singlet and even triplet states. Furthermore, as long as the σh\sigma_{h} mirror symmetry is preserved, only the uniform out-of-plane DM components, 𝐃=(0,0,D)\mathbf{D}^{\prime}=(0,0,D^{\prime}) and 𝐃′′=(0,0,D′′)\mathbf{D}^{\prime\prime}=(0,0,D^{\prime\prime}), survive in the triplet Hamiltonian to linear order.

Similar considerations apply to nematic interactions Nematic\mathcal{H}_{\text{Nematic}}. In this case, the sign of the out-of-plane nematic terms is opposite in the layers, leaving only the in-plane components in the triplet Hamiltonian. Nematic\mathcal{H}_{\text{Nematic}} thus simplifies to only the intra–dimer and inter–dimer in–plane nematic interactions, introduced in Eq. (3), Eq. (7), and (15).

Without loss of generality, we consider only the first neighbour interactions. The further neighbor terms do not change the overall shape (and generality) of the triplet Hamiltonian, only add to the complexity of the coefficients. Though, we keep the second neighbour DM interaction (D′′D^{\prime\prime}) to discuss band touching topological transitions as the functions of DD^{\prime} and D′′D^{\prime\prime}.

In the remainder of this paper we will slowly build up a complete picture of the topological physics of a quantum paramagnet described by Eq. (21), starting in Section IV from a simplified model with only Heisenberg and DM interactions, before progressively restoring the complexity of the full Hamiltonian, including nematic terms. Before doing so, in what follows, we set up the necessary technical framework for evaluating triplon bands.

III.2 Bond-wave Hamiltonian

To describe the dynamics of the triplet excitations at momentum 𝐤{\bf k} , we rely on the conventional bond wave theorySachdev and Bhatt (1990); Collins et al. (2006) resulting in the Bogoliubov–de Gennes Hamiltonian

k=(𝐭~𝐤𝐭~𝐤)T(M~𝐤N~𝐤N~𝐤M~𝐤)(𝐭~𝐤𝐭~𝐤).\mathcal{H}_{k}=\begin{pmatrix}\tilde{\mathbf{t}}^{\dagger}_{\mathbf{k}}\\ \tilde{\mathbf{t}}^{\phantom{\dagger}}_{-\!\bf k}\end{pmatrix}^{T}\begin{pmatrix}\tilde{M}_{\mathbf{k}}&\tilde{N}_{\mathbf{k}}\\ \tilde{N}^{\dagger}_{\mathbf{-k}}&\tilde{M}^{\dagger}_{\mathbf{-k}}\end{pmatrix}\begin{pmatrix}\tilde{\mathbf{t}}^{\phantom{\dagger}}_{\mathbf{k}}\\ \tilde{\mathbf{t}}^{\dagger}_{-\!\bf k}\end{pmatrix}. (27)

The vector 𝐭~𝐤\tilde{\mathbf{t}}^{\dagger}_{\mathbf{k}} contains the 9 different triplets excitations, corresponding to the three spin states and the three sublattices;

𝐭~𝐤=(𝐭1,𝐤,𝐭0,𝐤,𝐭1,𝐤),\tilde{\mathbf{t}}^{\dagger}_{\mathbf{k}}=\left(\mathbf{t}_{1,{\bf k}}^{\dagger},\mathbf{t}_{0,{\bf k}}^{\dagger},\mathbf{t}_{-1,{\bf k}}^{\dagger}\right), (28)

where 𝐭m,𝐤\mathbf{t}_{m,{\bf k}}^{\dagger} (m=1,0,1m=-1,0,1) is

𝐭m,𝐤=(tA,m,𝐤,tB,m,𝐤,tC,m,𝐤).\mathbf{t}_{m,{\bf k}}^{\dagger}=(t_{A,m,{\bf k}}^{\dagger},t_{B,m,{\bf k}}^{\dagger},t_{C,m,{\bf k}}^{\dagger}). (29)

M𝐤M_{\mathbf{k}} and N𝐤N_{\mathbf{k}} are 9-by-9 matrices, corresponding to the hopping hamiltonian, and the pairing terms, respectively. As it will turn out below, the matrices M~𝐤\tilde{M}_{\mathbf{k}} and N~𝐤\tilde{N}_{\mathbf{k}} are Hermitian and contain only cosines of the wave vector, so they are even in 𝐤\mathbf{k}: M~𝐤=M~𝐤=M~𝐤\tilde{M}^{\phantom{\dagger}}_{\mathbf{k}}=\tilde{M}^{\dagger}_{\mathbf{k}}=\tilde{M}^{\phantom{\dagger}}_{-\mathbf{k}} and N~𝐤=N~𝐤=N~𝐤\tilde{N}^{\phantom{\dagger}}_{\mathbf{k}}=\tilde{N}^{\dagger}_{\mathbf{k}}=\tilde{N}^{\phantom{\dagger}}_{-\mathbf{k}}. This will simplify the formulas, also the action of the time-reversal operator.

In the presence of the σh\sigma_{h} mirror plane, the |t0\left|t_{0}\right> triplet decouples from the time-reversal pair |t1\left|t_{1}\right> and |t1\left|t_{-1}\right> and the most general form of the Bogoliubov–de Gennes Hamiltonian, describing the triplet dynamics, decomposes into two blocks,

𝐤(1,1)\displaystyle\mathcal{H}_{\mathbf{k}}^{(1,-1)} =(𝐭𝐤𝐭𝐤)T(M𝐤+M𝐤ZeemanN𝐤N𝐤M𝐤M𝐤Zeeman)(𝐭𝐤𝐭𝐤)\displaystyle=\begin{pmatrix}\mathbf{t}_{{\bf k}}^{\dagger}\\ \mathbf{t}_{-\!{\bf k}}^{\phantom{\dagger}}\end{pmatrix}^{T}\begin{pmatrix}M_{\mathbf{k}}\!+\!M^{\text{Zeeman}}_{\mathbf{k}}&N_{\mathbf{k}}\\ N_{\mathbf{k}}&M_{\mathbf{k}}\!-\!M^{\text{Zeeman}}_{\mathbf{k}}\end{pmatrix}\begin{pmatrix}\mathbf{t}_{{\bf k}}^{\phantom{\dagger}}\\ \mathbf{t}_{-\!{\bf k}}^{\dagger}\end{pmatrix} (30a)
𝐤(0)\displaystyle\mathcal{H}_{\mathbf{k}}^{(0)} =(𝐭0,𝐤𝐭0,𝐤)T(M0,𝐤N0,𝐤N0,𝐤M0,𝐤)(𝐭0,𝐤𝐭0,𝐤)\displaystyle=\begin{pmatrix}\mathbf{t}_{0,{\bf k}}^{\dagger}\\ \mathbf{t}_{0,-\!{\bf k}}^{\phantom{\dagger}}\end{pmatrix}^{T}\begin{pmatrix}M_{0,\mathbf{k}}&N_{0,\mathbf{k}}\\ N_{0,\mathbf{k}}&M_{0,\mathbf{k}}\end{pmatrix}\begin{pmatrix}\mathbf{t}_{0,{\bf k}}^{\phantom{\dagger}}\\ \mathbf{t}_{0,-\!{\bf k}}^{\dagger}\end{pmatrix} (30b)

where the M𝐤M_{\mathbf{k}} and N𝐤N_{\mathbf{k}} are 6-by-6, and M0,𝐤M_{0,\mathbf{k}}, and N0,𝐤N_{0,\mathbf{k}} are 3-by-3 matrices. The spinful subspace in Eq. (30a) is spanned by

𝐭𝐤=(𝐭1,𝐤𝐭1,𝐤)and𝐭𝐤=(𝐭1,𝐤𝐭1,𝐤).\mathbf{t}_{{\bf k}}^{\phantom{\dagger}}=\begin{pmatrix}\mathbf{t}_{1,{\bf k}}^{\phantom{\dagger}}\\ \mathbf{t}_{-1,{\bf k}}^{\phantom{\dagger}}\end{pmatrix}\quad\text{and}\quad\mathbf{t}_{-{\bf k}}^{\dagger}=\begin{pmatrix}\mathbf{t}_{-1,-{\bf k}}^{\dagger}\\ \mathbf{t}_{1,-{\bf k}}^{\dagger}\end{pmatrix}. (31)

Using this basis, both the diagonal and off-diagonal matrices are Hermitian, M𝐤=M𝐤M^{\dagger}_{\mathbf{k}}=M^{\phantom{\dagger}}_{\mathbf{k}}, M0,𝐤=M0,𝐤M^{\dagger}_{0,\mathbf{k}}=M^{\phantom{\dagger}}_{0,\mathbf{k}}, N𝐤=N𝐤N^{\dagger}_{\mathbf{k}}=N^{\phantom{\dagger}}_{\mathbf{k}}, and N0,𝐤=N0,𝐤N^{\dagger}_{0,\mathbf{k}}=N^{\phantom{\dagger}}_{0,\mathbf{k}}. The σh\sigma_{h} acts on the individual S=1/2S=1/2 spins as C2=C63C_{2}=C_{6}^{3} combined with swapping the layer indices. As a consequence, the |t0\left|t_{0}\right> tranforms differently from the |t±1\left|t_{\pm 1}\right> and |s\left|s\right> under the σh\sigma_{h} operation: while |t0|t0\left|t_{0}\right>\to-\left|t_{0}\right>, the |t±1|t±1\left|t_{\pm 1}\right>\to\left|t_{\pm 1}\right> and |s|s\left|s\right>\to\left|s\right>. This happens for example in the terms containing SixSjzS^{x}_{i}S^{z}_{j} or SiySjzS^{y}_{i}S^{z}_{j}. Such terms are present in the in-plane DM and in the out-of-plane nematic interactions, resulting in t0,i()t±1,j()t^{(\dagger)}_{0,i}t^{(\dagger)}_{\pm 1,j} and t±1,i()t0,j()t^{(\dagger)}_{\pm 1,i}t^{(\dagger)}_{0,j} terms. These terms are odd under the σh\sigma_{h} reflection, and therefore cancel in the triplet Hamiltonian 111Let us mention, however, that a magnetic field applied in the xyxy-plane induces mixing between the spinful (|t1\left|t_{1}\right> and |t1\left|t_{-1}\right>) and the spinless (|t0\left|t_{0}\right>) subspaces.. Due to the cancellation of the in-plane DM terms in the Hamiltonian, the only terms that do not conserve the total STzS^{z}_{T} are the the in-plane nematic interactions, which change the STzS^{z}_{T} by ±2\pm 2 by creating an |t1\left|t_{1}\right> from |t1\left|t_{-1}\right> and vice versa.

We use the 8 Gell-Mann matrices as the basis for the sublattice degrees of freedom, corresponding to the three dimers, AA, BB, and CC in the unit cell:

λ1=(010100000),λ2=(0i0i00000),λ3=(100010000),\displaystyle\lambda_{1}\!=\!\!\!\left(\begin{array}[]{ccc}0&1&0\\ 1&0&0\\ 0&0&0\end{array}\right)\!\!,\lambda_{2}\!=\!\!\!\left(\begin{array}[]{ccc}0&-i&0\\ i&0&0\\ 0&0&0\end{array}\right)\!\!,\lambda_{3}=\!\!\!\left(\begin{array}[]{ccc}1&0&0\\ 0&-1&0\\ 0&0&0\end{array}\right)\!\!, (41)
λ4=(001000100),λ5=(00i000i00),λ6=(000001010),\displaystyle\lambda_{4}\!=\!\!\!\left(\begin{array}[]{ccc}0&0&1\\ 0&0&0\\ 1&0&0\end{array}\right)\!\!,\lambda_{5}\!=\!\!\!\left(\begin{array}[]{ccc}0&0&i\\ 0&0&0\\ -i&0&0\end{array}\right)\!\!,\lambda_{6}=\!\!\!\left(\begin{array}[]{ccc}0&0&0\\ 0&0&1\\ 0&1&0\end{array}\right)\!\!, (51)
λ7=(00000i0i0),λ8=13(100010002).\displaystyle\lambda_{7}\!=\!\!\!\left(\begin{array}[]{ccc}0&0&0\\ 0&0&-i\\ 0&i&0\end{array}\right)\!\!,\lambda_{8}\!=\!\frac{1}{\sqrt{3}}\!\!\left(\begin{array}[]{ccc}1&0&0\\ 0&1&0\\ 0&0&-2\end{array}\right)\!\!. (58)

The Gell-Mann matrices, and the 3-by-3 identity matrix, I3I_{3} sufficiently characterize the spinless m=0m=0 subspace. The hopping Hamiltonian has the form

M0,𝐤=M0,𝐤XXZ\displaystyle M_{0,\mathbf{k}}=M^{\text{XXZ}}_{0,\mathbf{k}} =\displaystyle= JI3+J[cos𝜹1𝐤2λ4\displaystyle J_{\|}I_{3}+J^{\prime}_{\bot}\left[\cos\frac{\bm{\delta}_{1}\mathbf{k}}{2}\lambda_{4}\right. (59)
+cos𝜹2𝐤2λ1+cos𝜹3𝐤2λ6],\displaystyle\left.+\cos\frac{\bm{\delta}_{2}\mathbf{k}}{2}\lambda_{1}+\cos\frac{\bm{\delta}_{3}\mathbf{k}}{2}\lambda_{6}\right]\;,

and the pairing terms are the same as M0,𝐤M_{0,\mathbf{k}} without the diagonal elements.

N0,𝐤=M0,𝐤JI3.\displaystyle N_{0,\mathbf{k}}=M_{0,\mathbf{k}}-J_{\|}I_{3}\;. (60)

Note that the Hamiltonian in the m=0m=0 subspace contains only the Heisenberg terms, the symmetric nematic exchange and the antisymmetric DM interaction do not affect the |t0\left|t_{0}\right> triplet.

The triplet hopping of the spinful subspace,

M𝐤=M𝐤XXZ+M𝐤DM+M𝐤NematicM_{\mathbf{k}}=M^{\text{XXZ}}_{\mathbf{k}}+M^{\text{DM}}_{\mathbf{k}}+M^{\text{Nematic}}_{\mathbf{k}} (61)

in Eq. (30a) is a 6×66\times 6 matrix. The spin degree of freedom provided by the Sz=±1S^{z}=\pm 1 triplets is represented by the spin operators sxs^{x}, sys^{y}, and szs^{z}, corresponding to the Pauli matrices times 12\frac{1}{2}. The 6-dimensional local Hilbert space for the spinful triplets is constructed as the tensor product of the 2-by-2 matrices {I2,sx,sy,sz}\{I_{2},s^{x},s^{y},s^{z}\} acting on the spin-space, and the 3-by-3 Gell-Mann matrices extended with the identity I3I_{3}, acting in the sublattice space. We discuss the various contributions separately. The Heisenberg interaction only contains the identity operator, I2I_{2} in the spin-space, i.e. it does not affect the spin degrees of freedom, and is the same for the |t1\left|t_{1}\right> and |t1\left|t_{-1}\right> triplets.

M𝐤XXZ\displaystyle M^{\text{XXZ}}_{\mathbf{k}} =\displaystyle\!=\! J+J2I2I3+Jcos𝜹1𝐤2I2λ4\displaystyle\frac{J_{\|}\!+\!J_{\bot}}{2}I_{2}\!\otimes\!I_{3}+J^{\prime}_{\|}\cos\frac{\bm{\delta}_{1}\mathbf{k}}{2}I_{2}\!\otimes\!\lambda_{4} (62)
+Jcos𝜹2𝐤2I2λ1+Jcos𝜹3𝐤2I2λ6.\displaystyle+J^{\prime}_{\|}\cos\frac{\bm{\delta}_{2}\mathbf{k}}{2}I_{2}\!\otimes\!\lambda_{1}+J^{\prime}_{\|}\cos\frac{\bm{\delta}_{3}\mathbf{k}}{2}I_{2}\!\otimes\!\lambda_{6}.

The pairing terms from the Heisenberg interaction are similar to M𝐤XXZM^{\text{XXZ}}_{\mathbf{k}}, but have opposite sign and no diagonal elements

N𝐤XXZ\displaystyle N^{\text{XXZ}}_{\mathbf{k}} =\displaystyle= M𝐤XXZ+J+J2I2I3.\displaystyle-M^{\text{XXZ}}_{\mathbf{k}}+\frac{J_{\|}+J\bot}{2}I_{2}\otimes I_{3}\;. (63)

The DM interaction has the form

M𝐤DM\displaystyle M^{\text{DM}}_{\mathbf{k}} =\displaystyle\!=\! [Dcos𝜹1𝐤2+D′′cos(𝜹2𝜹3)𝐤2]szλ5\displaystyle\left[\!D^{\prime}\!\cos\frac{\bm{\delta}_{1}\mathbf{k}}{2}\!+\!D^{\prime\prime}\cos\!\frac{(\bm{\delta}_{2}\!-\!\bm{\delta}_{3})\mathbf{k}}{2}\!\right]\!s^{z}\!\otimes\!\lambda_{5} (64)
+\displaystyle+ [Dcos𝜹2𝐤2+D′′cos(𝜹3𝜹1)𝐤2]szλ2\displaystyle\left[\!D^{\prime}\!\cos\frac{\bm{\delta}_{2}\mathbf{k}}{2}\!+\!D^{\prime\prime}\cos\!\frac{(\bm{\delta}_{3}\!-\!\bm{\delta}_{1})\mathbf{k}}{2}\!\right]\!s^{z}\!\otimes\!\lambda_{2}
+\displaystyle+ [Dcos𝜹3𝐤2+D′′cos(𝜹1𝜹2)𝐤2]szλ7.\displaystyle\left[\!D^{\prime}\!\cos\frac{\bm{\delta}_{3}\mathbf{k}}{2}\!+\!D^{\prime\prime}\cos\!\frac{(\bm{\delta}_{1}\!-\!\bm{\delta}_{2})\mathbf{k}}{2}\!\right]\!s^{z}\!\otimes\!\lambda_{7}\;.

The intra-dimer DM interaction is forbidden by the bond-inversion of the dimers, thus there are no diagonal elements in M𝐤DMM^{\text{DM}}_{\mathbf{k}} and the form of the pairing terms simply correspond to

N𝐤DM\displaystyle N^{\text{DM}}_{\mathbf{k}} =\displaystyle= M𝐤DM.\displaystyle-M^{\text{DM}}_{\mathbf{k}}\;. (65)

Here, the only operator acting in the spin-space is szs^{z}, leaving the spin degrees of freedom unchanged, and introducing a sign difference for the DM interaction in the up and down-spin sector.

Let us make some comments on the time-reversal (TR) properties of the Gell-Mann matrices and the pseudo-spin-half operators. The Gell-Mann matrices act in the sublattice space, i.e. account for changing the dimer indices, AA, BB, and CC. Time-reversal symmetry leaves such indices invariant, therefore the real Gell-Mann matrices are TR invariant. As TR symmetry contains a complex conjugation, the imaginary Gell-Matrices are TR breaking, changing sign under TR. A complete analysis on the TR symmetry of the pseudo-spin-half operators is provided in the Appendix A, where we show that while the szs^{z} is TR-breaking, as one would expect from a spin-operator, the sxs^{x} and sys^{y} components are TR invariant operators. This is a consequential difference with respect to the original Kane and Mele model, where the Pauli matrices describe a physical spin-half Kramers doublet and thus all break TR symmetry.

Coming back to the DM terms in our triplet hopping Hamiltonian in Eq. (64), the appearing cross-product operators are all TR invariant, as both the szs^{z} spin operator, and the λ5\lambda_{5}, λ2\lambda_{2}, and λ7\lambda_{7} complex Gell-Mann matrices are odd under TR symmetry.

The nematic interactions couple the subspaces of the different mm subspaces. They have the form

M𝐤Nematic\displaystyle M_{\mathbf{k}}^{\text{Nematic}} =\displaystyle\!=\! 3K4[sxλ8syλ3]\displaystyle\frac{\sqrt{3}K_{\|}}{4}\left[s^{x}\otimes\lambda_{8}-s^{y}\otimes\lambda_{3}\right] (66)
+Kcos𝜹3𝐤2(12sx32sy)λ6\displaystyle+K^{\prime}_{\|}\cos\frac{\bm{\delta}_{3}\mathbf{k}}{2}\left(\!\!-\frac{1}{2}s^{x}\!-\!\frac{\sqrt{3}}{2}s^{y}\!\!\right)\!\otimes\!\lambda_{6}
+Kcos𝜹1𝐤2(12sx+32sy)λ4\displaystyle+K^{\prime}_{\|}\cos\frac{\bm{\delta}_{1}\mathbf{k}}{2}\left(\!\!-\frac{1}{2}s^{x}\!+\!\frac{\sqrt{3}}{2}s^{y}\!\!\right)\!\otimes\!\lambda_{4}
+Kcos𝜹2𝐤2sxλ1,\displaystyle+K^{\prime}_{\|}\cos\frac{\bm{\delta}_{2}\mathbf{k}}{2}\cdot s^{x}\!\otimes\!\lambda_{1}\;,

The Gell-Mann matrices in the nematic interaction are the real ones, preserving TR symmetry. The nematic interactions exclusively consist of spin-mixing operators, sxs^{x} and sys^{y} that are TR invariant terms themselves. (For details see App. A). When the nematic terms are present, the total STzS^{z}_{T} ceases to be a good quantum number, and the spin up and down components mix.

Such spin-mixing term is present in the Kane and Mele model too, in the form of a Rashba spin-orbit coupling, permitted by the breaking of the mirror-symmetry σh\sigma_{h}. Here, the spin-mixing nematic terms are allowed without breaking σh\sigma_{h}. The TR symmetry does not protect the degeneracy of the spin-up and down triplets at the TR-invariant points in the momentum space, as it would for the Kramers pair electron-spins. Therefore, the nematic term hybridizes the spins, ending the fragile 𝒵2\mathcal{Z}_{2} topology of the bands, as discussed in Sec. VII.

The nematic interaction is allowed on the dimers too. Note that the operators λ8\lambda_{8} and λ3\lambda_{3} are diagonal in the sublattice space. The sxs^{x} and sys^{y} operators, however mix the spins, placing the intra-dimer nematic interaction KK_{\|} in the diagonal of the block connecting the +1+1 and 1-1 triplets. The N1,𝐤NematicN^{\text{Nematic}}_{1,\mathbf{k}} matrix of the pairing terms corresponds again to the M1,𝐤Nematic-M^{\text{Nematic}}_{1,\mathbf{k}} minus the ‘diagonal’ elements

N𝐤Nematic=M𝐤Nematic+3K4[sxλ8syλ3].\displaystyle N^{\text{Nematic}}_{\mathbf{k}}\!=\!-M^{\text{Nematic}}_{\mathbf{k}}\!+\!\frac{\sqrt{3}K_{\|}}{4}\left[s^{x}\!\otimes\!\lambda_{8}\!-\!s^{y}\!\otimes\!\lambda_{3}\right]\;. (67)

Lastly, the out-of plane magnetic field appears in the diagonal of the Hamiltonian as

M𝐤Zeeman\displaystyle M_{\mathbf{k}}^{\text{Zeeman}} =\displaystyle\!=\! gzhz2szI3.\displaystyle-g_{z}h_{z}2s^{z}\otimes I_{3}\;. (68)

We conclude this tour of the terms in the quadratic Bogoliubov–de Gennes Hamiltonian with a brief comment on what it neglects, namely interactions between triplon modes occurring at higher order in bond operators. The effect of electron–electron interactions on topological insulators and superconductors remains an open problem. And the effect of triplon–triplon interactions on band topology in quantum paramagnets is even less explored.

None the less, some work has been done the characterize the effect of magnon–magnon interactions in magnetically–ordered systems with topological bands Chernyshev and Zhitomirsky (2015); Chernyshev and Maksimov (2016); McClarty et al. (2018); McClarty and Rau (2019); Kondo et al. (2020); McClarty (2021). One example which has been been quite well characterized in the Kitaev model in high magnetic field, where the non–interacting theory of topological magnon bands can be compared with both interacting spin–wave calculations and DMRG results McClarty et al. (2018). And here the effect of interactions is relatively benign, being chiefly limited to a finite broadening of magnon modes, and renormalisation of band their dispersion.

Its is reasonable to expect the same will be true in quantum paramagnets, since the form of Bogoliubov–de Gennes Hamiltonian is identical. And this particulary true where the potential for triplon decay is restricted by a substantial band gap. It is also plausible that the topological excitations of quantum paramagnets will exhibit some of the same interesting, non–Hermitian features, as topological magnons, a topic reviewed in McClarty (2021). The role of interactions within a 𝒵2\mathcal{Z}_{2} topological phase of a magnetic insulator is clearly worthy of further investigation. However this lies outside the scope of this Article, which has the more limited goal of characterizing band topology in the non–interacting limit.

III.3 Bond-wave dispersions in the absence of anisotropies

Let us start with the time reversal symmetric case, when the magnetic field is zero. In the isotropic Heisenberg limit, J=J=JJ_{\bot}=J_{\|}=J, and J=J=JJ^{\prime}_{\bot}=J^{\prime}_{\|}=J^{\prime}, D=0D^{\prime}=0, D′′=0D^{\prime\prime}=0, K=0K_{\|}=0, and K=0K^{\prime}_{\|}=0 the model has SU(2) symmetry, the M𝐤M_{\mathbf{k}} and N𝐤N_{\mathbf{k}} matrix in Eq. (30a) becomes block diagonal,

M𝐤=(M1,𝐤00M1,𝐤),N𝐤=(N1,𝐤00N1,𝐤),M_{\mathbf{k}}=\begin{pmatrix}M_{1,\mathbf{k}}&0\\ 0&M_{-1,\mathbf{k}}\end{pmatrix}\;,\quad N_{\mathbf{k}}=\begin{pmatrix}N_{1,\mathbf{k}}&0\\ 0&N_{-1,\mathbf{k}}\end{pmatrix}\;, (69)

and the Hamiltonians is identical in each of the m=1,0,1m=-1,0,1 subspaces (see Eqs. (30a) and (30b)) so that Mm,𝐤=Mm,𝐤SU(2)M_{m,\mathbf{k}}=M^{\text{SU(2)}}_{m,\mathbf{k}}, and Nm,𝐤=Nm,𝐤SU(2)N_{m,\mathbf{k}}=N^{\text{SU(2)}}_{m,\mathbf{k}}, with mm-independent

Mm,𝐤SU(2)\displaystyle M^{\text{SU(2)}}_{m,\mathbf{k}} =\displaystyle\!\!=\!\! JI3+J[cos𝜹1𝐤2λ4+cos𝜹2𝐤2λ1+cos𝜹3𝐤2λ6],\displaystyle JI_{3}\!+\!J^{\prime}\!\!\left[\cos\frac{\bm{\delta}_{1}\mathbf{k}}{2}\lambda_{4}\!+\!\cos\frac{\bm{\delta}_{2}\mathbf{k}}{2}\lambda_{1}\!+\!\cos\frac{\bm{\delta}_{3}\mathbf{k}}{2}\lambda_{6}\right]\!,
Nm,𝐤SU(2)\displaystyle N^{\text{SU(2)}}_{m,\mathbf{k}} =\displaystyle\!\!=\!\! JI3M𝐤SU(2).\displaystyle JI_{3}-M^{\text{SU(2)}}_{\mathbf{k}}\;. (70)

matrices (we keep the mm index only for bookkeeping purposes). The mm subspaces are spanned by the basis

(𝐭m,𝐤,𝐭m,𝐤),(\mathbf{t}_{m,{\bf k}}^{\dagger},\mathbf{t}_{-m,-{\bf k}}^{\phantom{\dagger}}), (71)

defined in Eq. (28). Each of these operators is going to change the total STzS^{z}_{T} by mm, either by creating an mm triplon with 𝐭m,𝐤\mathbf{t}_{m,{\bf k}}^{\dagger} or by annihilating a m-m triplon with 𝐭m,𝐤\mathbf{t}_{-m,-{\bf k}}. A rotation eiφSTze^{-i\varphi S^{z}_{T}} by an angle φ\varphi about the zz axis in the spin space manifests in a phase factor

(𝐭m,𝐤,𝐭m,𝐤)eimφ(𝐭m,𝐤,𝐭m,𝐤).(\mathbf{t}_{m,{\bf k}}^{\dagger},\mathbf{t}_{-m,-{\bf k}}^{\phantom{\dagger}})\to e^{-im\varphi}(\mathbf{t}_{m,{\bf k}}^{\dagger},\mathbf{t}_{-m,-{\bf k}}^{\phantom{\dagger}}). (72)

Thereby in the Hamiltonian we encounter normal terms of the form 𝐭m,𝐤Mm,𝐤𝐭m,𝐤\mathbf{t}_{m,{\bf k}}^{\dagger}M_{m,\mathbf{k}}^{\phantom{\dagger}}\mathbf{t}_{m,{\bf k}}^{\phantom{\dagger}} and 𝐭m,𝐤Mm,𝐤𝐭m,𝐤\mathbf{t}_{-m,-{\bf k}}^{\phantom{\dagger}}M_{m,\mathbf{k}}^{\phantom{\dagger}}\mathbf{t}_{-m,-{\bf k}}^{\dagger}, and anomalous terms in the form of 𝐭m,𝐤Nm,𝐤𝐭m,𝐤\mathbf{t}_{m,{\bf k}}^{\dagger}N_{m,\mathbf{k}}^{\phantom{\dagger}}\mathbf{t}_{-m,-{\bf k}}^{\dagger} and 𝐭m,𝐤Nm,𝐤𝐭m,𝐤\mathbf{t}_{-m,{\bf k}}^{\phantom{\dagger}}N_{m,\mathbf{k}}^{\phantom{\dagger}}\mathbf{t}_{m,-{\bf k}}^{\phantom{\dagger}} which are invariant with respect to the phase transformation of Eq. (72). These all commute with the

STz=𝐭1,𝐤𝐭1,𝐤𝐭1,𝐤𝐭1,𝐤,S^{z}_{T}=\mathbf{t}_{1,{\bf k}}^{\dagger}\mathbf{t}_{1,{\bf k}}^{\phantom{\dagger}}-\mathbf{t}_{-1,-{\bf k}}^{\dagger}\mathbf{t}_{-1,-{\bf k}}^{\phantom{\dagger}}, (73)

so that the total STzS^{z}_{T} is conserved, generating a U(1) symmetry. The higher SU(2) symmetry is exemplified by the matrices being independent from mm.

In this case, each of the three bands, coming from the three sublattices, are threefold degenerate, as the m=1,0,1m=1,0,-1 have the same energies in the entire Brillouin zone. The dispersion computed from the SU(2) symmetric Bogoliubov–de Gennes Hamiltonian has the form

ω1,m\displaystyle\omega_{1,m} =J(J2J)\displaystyle=\sqrt{J(J-2J^{\prime})} (74a)
ω2,m\displaystyle\omega_{2,m} =J(J+J)J~(𝐤)\displaystyle=\sqrt{J(J+J^{\prime})-\tilde{J}({\mathbf{k}})} (74b)
ω3,m\displaystyle\omega_{3,m} =J(J+J)+J~(𝐤).\displaystyle=\sqrt{J(J+J^{\prime})+\tilde{J}({\mathbf{k}})}\;. (74c)

with J~(𝐤)=JJ3+2αcos𝜹α𝐤\tilde{J}({\mathbf{k}})=JJ^{\prime}\sqrt{3+2\sum_{\alpha}\cos{\bm{\delta}_{\alpha}}\cdot\mathbf{k}}. We show these bands in Fig. 8(a). Additionally, discrete lattice symmetries give rise to degeneracies between the sublattice bands in the form of a linear band touching at the KK and KK^{\prime} points, and a quadratic band touching at the zone center, Γ\Gamma.

Refer to caption
Figure 8: (a) Triplet bands in the SU(2) symmetric case, J=1J=1, J=0.2J^{\prime}=0.2, h=0h=0, and the DM and nematic terms are zero. All three bands are 3-fold degenerate due to the spin-rotation symmetry. (b) Triplet band structure for finite second neighbor DM interaction, D′′=0.01D^{\prime\prime}=0.01. (c) The m=±1m=\pm 1 bands omitting the trivial m=0m=0. All three bands are twofold degenerate, corresponding to the +1+1 and 1-1 spin of the triplets.

The intra-dimer Heisenberg coupling term, JJ, is nothing but the singlet–triplet gap, separating the triplet bands from the ground state, while the inter-dimer Heisenberg interaction, JJ^{\prime} gives dispersion to the triplets. The form of dispersion is immediately familiar from studies of graphene Haldane (1988); Kane and Mele (2005a), with the additional feature of a flat band just above the singlet–triplet gap, reflecting the frustration of the Kagome lattice.

From the band dispersion, Eq. (74), we can also read off the instabilities of the quantum paramagnet. For the ferromagnetic J=J/4J^{\prime}\!=\!-J/4, the ω3,m\omega_{3,m} softens at the Γ\Gamma point, signaling a transition to a long-range ordered time-reversal symmetry breaking state, where the spins on each layer are aligned ferromagneticaly, while the two layers are aligned antiferromagneticaly. Similarly, we may notice that the energy of the ω1,m\omega_{1,m} flat band becomes 0 for antiferromagnetic J=J/2J^{\prime}=J/2, indicating a transition to a time-reversal symmetry breaking state, where the ordering is selected by the DM interactions or quantum fluctuations.

IV DM interaction induced band topology

In this Section, we explore the topologically non–trivial triplon bands which arise as result of DM interactions in a quantum paramagnet described by Eq. (21), setting both the nematic terms and the magnetic field to zero. For simplicity, we will also set J=J=JJ_{\parallel}=J_{\perp}=J in XXZ terms, such that they reduce to isotropic Heisenberg interactions. The analysis is further simplified by the fact that only the zz component of the DM interaction survives (up to first order). It follows that total STzS^{z}_{T} remains a good quantum number, and excitations with different m=1,0,1m=1,0,-1 decouple from one another.

The out–of–plane DM interaction lowers the SU(2) symmetry to U(1), and splits the degeneracy at the corners and center of the BZ for the m=±1m=\pm 1 triplets, as shown in Fig. 8(c). While JJ^{\prime} provides a real hopping amplitude, the inter-dimer DM interactions, DD^{\prime} and D′′D^{\prime\prime} couple to the complex Gell-Mann matrices, and are responsible for the non-trivial topology, generating finite Berry curvature via the complex triplet hopping amplitude. The mm-independent Heisenberg Hamiltonian (70) is extended with the DM interaction (64). As the DM is diagonal with respect to the spin degrees of freedom, we can write the U(1)-symmetric Hamiltonian in a block-diagonal form, with decoupled m=1,0,1m=1,0,-1 subspaces. We account for the szs^{z} operator in the DM interaction (64) with the factor mm:

Mm,𝐤U(1)\displaystyle M^{\text{U(1)}}_{m,\mathbf{k}} =Mm,𝐤SU(2)+m[Dcos𝜹1𝐤2+D′′cos(𝜹2𝜹3)𝐤2]λ5\displaystyle=M^{\text{SU(2)}}_{m,\mathbf{k}}+m\left[\!D^{\prime}\!\cos\frac{\bm{\delta}_{1}\mathbf{k}}{2}\!+\!D^{\prime\prime}\cos\!\frac{(\bm{\delta}_{2}\!-\!\bm{\delta}_{3})\mathbf{k}}{2}\!\right]\!\lambda_{5}
+m[Dcos𝜹2𝐤2+D′′cos(𝜹3𝜹1)𝐤2]λ2\displaystyle\phantom{=}+m\left[\!D^{\prime}\!\cos\frac{\bm{\delta}_{2}\mathbf{k}}{2}\!+\!D^{\prime\prime}\cos\!\frac{(\bm{\delta}_{3}\!-\!\bm{\delta}_{1})\mathbf{k}}{2}\!\right]\!\lambda_{2}
+m[Dcos𝜹3𝐤2+D′′cos(𝜹1𝜹2)𝐤2]λ7,\displaystyle\phantom{=}+m\left[\!D^{\prime}\!\cos\frac{\bm{\delta}_{3}\mathbf{k}}{2}\!+\!D^{\prime\prime}\cos\!\frac{(\bm{\delta}_{1}\!-\!\bm{\delta}_{2})\mathbf{k}}{2}\!\right]\!\lambda_{7}\;,
Nm,𝐤U(1)\displaystyle N^{\text{U(1)}}_{m,\mathbf{k}} =JI3Mm,𝐤U(1).\displaystyle=JI_{3}-M^{\text{U(1)}}_{m,\mathbf{k}}\;. (75)

To determine the DM gap, we compute the energies at the Γ\Gamma, KK, and KK^{\prime} points from the full Bogoliubov–de Gennes equations.

At the Γ\Gamma point the bands have the energies

ω1,m(Γ)=J(J2JmΔΓ),\displaystyle\omega_{1,m}(\Gamma)=\sqrt{J(J-2J^{\prime}-m\Delta_{\Gamma})}\;, (76a)
ω2,m(Γ)=J(J2J+mΔΓ),\displaystyle\omega_{2,m}(\Gamma)=\sqrt{J(J-2J^{\prime}+m\Delta_{\Gamma})}\;, (76b)
ω3,m(Γ)=J(J+4J),\displaystyle\omega_{3,m}(\Gamma)=\sqrt{J(J+4J^{\prime})}\;, (76c)

where ΔΓ=23(D+D′′)\Delta_{\Gamma}=2\sqrt{3}(D^{\prime}+D^{\prime\prime}). A finite ΔΓ\Delta_{\Gamma} opens a gap J(J2J+ΔΓ)J(J2JΔΓ)\sqrt{J(J-2J^{\prime}+\Delta_{\Gamma})}-\sqrt{J(J-2J^{\prime}-\Delta_{\Gamma})} between the bands 1 and 2 at the Γ\Gamma point. For ΔΓJ,J\Delta_{\Gamma}\ll J,J^{\prime}, the gap becomes J/(J2J)ΔΓ\approx\sqrt{J/(J-2J^{\prime})}\Delta_{\Gamma}, thus proportional to the DM interactions.

At the KK and KK^{\prime} points the bands have energies

ω1,m(K)=J(J2J),\displaystyle\omega_{1,m}(K)=\sqrt{J(J-2J^{\prime})}\;, (77a)
ω2,m(K)=J(J+J+mΔK),\displaystyle\omega_{2,m}(K)=\sqrt{J(J+J^{\prime}+m\Delta_{K})}\;, (77b)
ω3,m(K)=J(J+JmΔK),\displaystyle\omega_{3,m}(K)=\sqrt{J(J+J^{\prime}-m\Delta_{K})}\;, (77c)

where ΔK=3J(D2D′′)\Delta_{K}=\sqrt{3}J(D^{\prime}-2D^{\prime\prime}). This results in a gap of J(J+J+ΔK)J(J+JΔK)\sqrt{J(J+J^{\prime}+\Delta_{K})}-\sqrt{J(J+J^{\prime}-\Delta_{K})} at the KK (and KK^{\prime}) point between the top bands with nonzero mm. Note that the m=0m=0 triplet band is not affected by the DM interaction, retaining their degeneracies at KK, KK^{\prime}, and Γ\Gamma points. Furthermore, the m=±1m=\pm 1 triplets experience an opposite effect, reflecting the action of the spin-orbit coupling on the up and down spins, lending them an opposite torque.

Refer to caption
Figure 9: Chern numbers as the function of first (DD^{\prime}) and second (D′′D^{\prime\prime}) neighbor inter-dimer DM interaction, listed in order of ascending band energy for a given mm-subspace. The lines D=2D′′D^{\prime}=2D^{\prime\prime} and D=D′′D^{\prime}=-D^{\prime\prime} denote the boundaries of the band touching topological transitions. For D=2D′′D^{\prime}=2D^{\prime\prime} the gap between the upper two bands closes at the corners in the form of Dirac cones. Crossing the D=D′′D^{\prime}=-D^{\prime\prime} line, the lower bands go through a quadratic touching at the Γ\Gamma point.
Refer to caption
Figure 10: Distribution of Berry curvature in the vicinity of the transition points for the m=1m=1 bands. The Berry curvature distribution for m=1m=-1 is reversed, and not well definied for m=0m=0. At D=D′′D^{\prime}=-D^{\prime\prime} the gap closes at the Γ\Gamma point between the middle and lower bands. We chose the values J=1J=1, J=0.2J^{\prime}=0.2, D=0.01D^{\prime}=0.01 (a) D′′=0.01ϵD^{\prime\prime}=-0.01-\epsilon and (b) D′′=0.01+ϵD^{\prime\prime}=-0.01+\epsilon, where ϵ=0.0025\epsilon=0.0025. At D=2D′′D^{\prime}=2D^{\prime\prime} the gap at KK and KK^{\prime} closes. Berry curvature for D′′=0.005ϵD^{\prime\prime}=0.005-\epsilon (c) and D′′=0.005+ϵD^{\prime\prime}=0.005+\epsilon (d).

The three triplet bands within the m=1m=1 (or m=1m=-1) subspace become fully gapped. Although the bands of m=1m=1 are overlapping with the bands of m=1m=-1 everywhere in the BZ, there are no matrix elements between the two, and (their canceling) Chern numbers can be computed independently for them Simon (1983); Berry (1984); Kohmoto (1985); Fukui et al. (2005). The triplet bands in the case of a finite DM interaction are plotted in Fig. 8(b) and (c). The Chern number of the nn-th band of triplet mm,

Cn,m=12πiBZ𝑑kx𝑑kyFn,mxy,\displaystyle C_{n,m}=\frac{1}{2\pi i}\int_{\rm BZ}dk_{x}dk_{y}F^{xy}_{n,m}\;, (78)

is the integral of the Berry curvature

Fn,mxy(𝐤)=kxn(𝐤)|Σzkyn(𝐤)kyn(𝐤)|Σzkxn(𝐤)\displaystyle F^{xy}_{n,m}({\bf k})\!\!=\!\!\left<\partial_{k_{x}}n({\bf k})|\Sigma^{z}\partial_{k_{y}}n({\bf k})\right>-\left<\partial_{k_{y}}n({\bf k})|\Sigma^{z}\partial_{k_{x}}n({\bf k})\right> (79)

over the Brillouin zoneShindou et al. (2013); Mook et al. (2014a). The |n(𝐤)\left|n({\bf k})\right> eigenfunctions are the solutions of the Bogoliubov-de Gennes equations,

(Mm,𝐤U(1)Nm,𝐤U(1)Nm,𝐤U(1)Mm,𝐤U(1))|n(𝐤)=ωn,m(𝐤)Σz|n(𝐤),\begin{pmatrix}M^{\text{U(1)}}_{m,\mathbf{k}}&N^{\text{U(1)}}_{m,\mathbf{k}}\\ N^{\text{U(1)}}_{m,\mathbf{k}}&M^{\text{U(1)}}_{m,\mathbf{k}}\end{pmatrix}\left|n({\bf k})\right>=\omega_{n,m}(\mathbf{k})\Sigma^{z}\left|n({\bf k})\right>\;, (80)

with

Σz=(I3𝟎𝟎I3).\Sigma^{z}=\begin{pmatrix}I_{3}&\mathbf{0}&\\ \mathbf{0}&-I_{3}\\ \end{pmatrix}\;. (81)

To map out a band touching phase diagram, we compute the Chern numbers as the function of the first and second neighbor DM interactions, DD^{\prime} and D′′D^{\prime\prime} in Fig. 9. The numbers in Fig. 9 represent the Chern numbers in the mm subspace, going from the bottom band to the top band. The m=1m=1 and m=1m=-1 triplet bands have opposite Chern numbers, reflecting their opposite chirality.

As the Chern number can only be changed via closing the gap, we can analytically determine the phase boundaries by examining when ΔΓ\Delta_{\Gamma} or ΔK\Delta_{K} become zero.

When D=2D′′D^{\prime}=2D^{\prime\prime}, the ΔK\Delta_{K} is zero closing the gap between the upper two bands at the KK and KK^{\prime} points in a linear Dirac-cone-like touching. At each corner points, the Chern number of the bands is changed locally by one as the gap closes and reopens. The contribution from KK and KK^{\prime} adds up to a +2+2 and 2-2 change in the Chern numbers of the upper and middle bands (see Fig. 9). The lowest band remains unaltered.

ΔΓ\Delta_{\Gamma} becomes zero for D=D′′D^{\prime}=-D^{\prime\prime}, closing the gap between the lower two bands in a quadratic touching at the zone center. The Chern numbers of the lower and middle bands change by ±2\pm 2. The topmost band is unaffected.

To illustrate the exchange of topological charge across the band touching transitions, we plotted the Berry curvature distribution on the m=1m=1 triplet bands in the vicinity of D=D′′D^{\prime}=-D^{\prime\prime} and D=2D′′D^{\prime}=2D^{\prime\prime} in Fig. 10. The Berry curvature has opposite sign for the m=1m=-1 triplet bands.

When computing the Chern numbers shown in Fig. 9, and the distribution of the Berry curvature in Fig. 10, we use the numerical method introduced in Ref. Fukui et al. (2005) equipped with the structure for particle-particle terms present in the Bogoliubov–de Gennes type of Hamiltonians. The numerical computation, however, does not provide us a deeper insight. Therefore, in Sec. V, we give a detailed discussion on the topological transitions and the changing of the Chern numbers at the linear and quadratic touching points, restricting our analysis to the hopping part of the Bogoliubov–de Gennes Hamiltonian. The topological properties of the reduced hopping Hamiltonian and the Bogoliubov–de Gennes Hamiltonian are the same, as the MM and NN matrices are similar, i.e. they only differ in their diagonal.

V Linear versus quadratic band touchings

The transitions between bands with different Chern indices, shown in the phase diagram Fig. 9, occur where the gap between bands closes. The point where bands touch, may have a linear or quadratic form, which we characterize in detail in this section. We provide analytic expressions for both the Berry curvature, and the associated Chern numbers, before and after the transitions, where they are well defined Oshikawa (1994). We focus on how topological charge is exchanged between bands. In both cases, two units of Chern number are exchanged. For linear band touching this occurs through the exchange of a single unit at two different points in the BZ, KK and KK^{\prime}, while for quadratic band touchings two units are exchanged at the Γ\Gamma point.

We continue with the U(1) symmetric model of Eq. (75) including the isotropic Heisenberg interaction and the DM interactions, but keeping the nematic terms zero.

Our approach will be to reduce the description of each touching point to a 2×22\times 2 matrix describing only those bands which touch

M𝐤eff=d0(𝐤)I2+𝐝(𝐤)𝝈,M^{\rm eff}_{\mathbf{k}}=d_{0}(\mathbf{k})I_{2}+\mathbf{d}(\mathbf{k})\cdot\bm{\sigma}\;, (82)

where 𝝈\bm{\sigma} is a vector of Pauli matrices, and the linear or quadratic form of the band touching is encoded in the coefficients d0(𝐤)d_{0}(\mathbf{k}) and 𝐝(𝐤)\mathbf{d}(\mathbf{k}). These coefficients can be expanded about the relevant wave vector, thereby allowing the analytic calculation of the Chern number

C=12πBZΩ(kx,ky)𝑑kx𝑑kyC=\frac{1}{2\pi}\int_{\text{BZ}}\Omega(k_{x},k_{y})dk_{x}dk_{y}\; (83)

through the associated Berry curvature

Ω(kx,ky)=12𝐝(𝐝kx×𝐝ky)(𝐝𝐝)3/2,\Omega(k_{x},k_{y})=\frac{1}{2}\frac{\mathbf{d}\cdot\left(\frac{\partial\mathbf{d}}{\partial k_{x}}\times\frac{\partial\mathbf{d}}{\partial k_{y}}\right)}{(\mathbf{d}\cdot\mathbf{d})^{3/2}}, (84)

which, in this case, has the interpretation of a Skyrmion density.

For simplicity, the derivation we present is restricted to the hopping matrix Mm,𝐤U(1)M^{\text{U(1)}}_{m,\mathbf{k}}, and neglects the effect of the pair creation and annihilation terms. This restriction can however be relaxed, at the expense of more lengthly expressions. If we were to consider the full Hamiltonian, the unitary transformations would correspond to I2UΛI_{2}\otimes U_{\Lambda}, (Λ=K,K,Γ\Lambda=K,K^{\prime},\Gamma), and instead of the 2×22\times 2 matrix describing the bands in question, we would get a 4×44\times 4 problem that can be solved using the Bogoliubov transformation. The topological properties of the bands are generally not affected by the pairing terms (an exception is provided in Ref. McClarty et al. (2018), where the diagonal (MM) and off-diagonal (NN) matrices were not similar). The equivalency of the topology of MM and the full Hamiltonian has been discussed in Ref. [Nawa et al., 2019].

Ultimately, our goal is to project the three-level problem (of each mm separately) onto a two-level one, involving the bands that touch at the corners and the center of the hexagonal Brillouin zone. To do this, we introduce unitary transformations, UKU_{K} and UΓU_{\Gamma} that diagonalize M𝐤SU(2)M^{\text{SU(2)}}_{\mathbf{k}} (and N𝐤SU(2)N^{\text{SU(2)}}_{\mathbf{k}} too) at the band touching points. Naturally, we can perform the expansion about the touching point without projecting onto the two levels involved. Nonetheless, we restrict ourselves to a two-level problem to be able to represent the sublattice degrees of freedom with only three Pauli matrices instead of eight Gell-Mann matrices, which would give an eight-dimensional 𝐝\mathbf{d} vector.

V.1 Linear band touching at KK

We begin with discussing the linear touchings that occur at the corners 𝐤K=(4π3,0)\mathbf{k}_{K}\!=\!(\frac{4\pi}{3},0) and 𝐤K=(4π3,0)\mathbf{k}_{K^{\prime}}\!=\!(-\frac{4\pi}{3},0) of the Brillouin zone for D=2D′′D^{\prime}=2D^{\prime\prime}. The matrix UKU_{K} that diagonalizes M𝐤SU(2)M^{\text{SU(2)}}_{\mathbf{k}} at both KK-points is

UK=(13121613121613026).U_{K}=\left(\begin{array}[]{ccc}-\frac{1}{\sqrt{3}}&\frac{1}{\sqrt{2}}&\frac{1}{\sqrt{6}}\\ -\frac{1}{\sqrt{3}}&-\frac{1}{\sqrt{2}}&\frac{1}{\sqrt{6}}\\ \frac{1}{\sqrt{3}}&0&\frac{2}{\sqrt{6}}\\ \end{array}\right)\;. (85)

The column vectors in UKU_{K} correspond to the eigenvectors at K and also form a basis for a one- and two-dimensional irreducible representation of the three-fold symmetry at these points in the absence of the DM interactions. The band that belongs to the symmetric representation is the bottom band, which is well separated from the upper two bands in the vicinity of the zone-corners. The double representation stretches the subspace that we keep in the linearization. Using UKU_{K}, we transform Mm,𝐤U(1)M^{\text{U(1)}}_{m,\mathbf{k}} into the form

UKMm,𝐤U(1)UK=(JJ1432Jky+im342Dkx1432Jkx+im342Dky1432Jkyim342DkxJ+J2+34Jkx34Jky+im3(D2D′′)1432Jkxim342Dky34Jkyim3(D2D′′)J+J234Jkx),U_{K}^{\dagger}\cdot M^{\text{U(1)}}_{m,{\bf k}}\cdot U_{K}=\left(\begin{array}[]{ccc}J-J^{\prime}&-\frac{1}{4}\sqrt{\frac{3}{2}}J^{\prime}k_{y}+im\frac{3}{4\sqrt{2}}D^{\prime}k_{x}&\frac{1}{4}\sqrt{\frac{3}{2}}J^{\prime}k_{x}+im\frac{3}{4\sqrt{2}}D^{\prime}k_{y}\\ -\frac{1}{4}\sqrt{\frac{3}{2}}J^{\prime}k_{y}-im\frac{3}{4\sqrt{2}}D^{\prime}k_{x}&J+\frac{J^{\prime}}{2}+\frac{\sqrt{3}}{4}J^{\prime}k_{x}&-\frac{\sqrt{3}}{4}J^{\prime}k_{y}+im\sqrt{3}\left(\frac{D^{\prime}}{2}-D^{\prime\prime}\right)\\ \frac{1}{4}\sqrt{\frac{3}{2}}J^{\prime}k_{x}-im\frac{3}{4\sqrt{2}}D^{\prime}k_{y}&-\frac{\sqrt{3}}{4}J^{\prime}k_{y}-im\sqrt{3}\left(\frac{D^{\prime}}{2}-D^{\prime\prime}\right)&J+\frac{J^{\prime}}{2}-\frac{\sqrt{3}}{4}J^{\prime}k_{x}\\ \end{array}\right), (86)

where mm is the spin-index of the triplets taking the values 1,0,1-1,0,1, and kxk_{x} and kyk_{y} are measured from the KK point, i.e. KK corresponds to kx=ky=0k_{x}=k_{y}=0. At the corners of the Brillouin zone, this matrix is block-diagonal and has eigen-energies JJJ-J^{\prime}, and J+J/2±m3(D′′D/2)J+J^{\prime}/2\pm m\sqrt{3}(D^{\prime\prime}-D^{\prime}/2). As we go away from the KK points, small off-diagonal matrix elements appear that are linear in momentum. Projecting (86) on the relevant subspace, we can write the two-band matrix as

Klin=(J+J2)I2+𝐝K𝝈.\mathcal{H}^{\text{lin}}_{K}=\left(J+\frac{J^{\prime}}{2}\right)I_{2}+\mathbf{d}_{K}\cdot\bm{\sigma}\;. (87)

The vector 𝐝K\mathbf{d}_{K} has the form

dKx\displaystyle d^{x}_{K} =34Jky,\displaystyle=-\frac{\sqrt{3}}{4}J^{\prime}k_{y}\;, (88a)
dKy\displaystyle d^{y}_{K} =m32(D2D′′),\displaystyle=-m\frac{\sqrt{3}}{2}\left(D^{\prime}-2D^{\prime\prime}\right)\;, (88b)
dKz\displaystyle d^{z}_{K} =34Jkx.\displaystyle=\frac{\sqrt{3}}{4}J^{\prime}k_{x}\;. (88c)

The 𝐝K\mathbf{d}_{K^{\prime}} vector for the KK^{\prime}-point is given by (dKx,dKy,dKz)=(dKx,dKy,dKz)(d^{x}_{K^{\prime}},d^{y}_{K^{\prime}},d^{z}_{K^{\prime}})=(-d^{x}_{K},d^{y}_{K},-d^{z}_{K}).

Using Eqs. (84) we get

ΩK(k)=mJ2(D2D′′)(J2k2+4(D2D′′)2)3/2,\Omega_{K}(k)=\frac{mJ^{\prime 2}(D^{\prime}-2D^{\prime\prime})}{(J^{\prime 2}k^{2}+4(D^{\prime}-2D^{\prime\prime})^{2})^{3/2}}\;, (89)

where we substituted kx2+ky2=k2k^{2}_{x}+k^{2}_{y}=k^{2}. We note that the Berry curvature does not depend on the valley index, i.e. it is the same at the KK and KK^{\prime} points. To obtain a simpler form for Ω(k)\Omega(k), we introduce the dimensionless parameter

k0=2(D2D′′)J,\displaystyle k_{0}=\frac{2(D^{\prime}-2D^{\prime\prime})}{J^{\prime}}\;, (90)

so that

ΩK(k)=mk02(k2+k02)3/2.\displaystyle\Omega_{K}(k)=m\frac{k_{0}}{2(k^{2}+k_{0}^{2})^{3/2}}\;. (91)

Ω(k)\Omega(k) has maximum at k=0k=0, i.e. the Berry curvature is concentrated at the KK and KK^{\prime} points, as shown in Figs. 10 and 11. Then ΩK=msgn(k0)2k02msgn(D2D′′)J2(D2D′′)2\Omega_{K}=m\frac{{\rm sgn}(k_{0})}{2k_{0}^{2}}\propto m\operatorname{sgn}(D^{\prime}-2D^{\prime\prime})\frac{J^{\prime 2}}{(D^{\prime}-2D^{\prime\prime})^{2}}, which diverges as the D2D′′D^{\prime}\to 2D^{\prime\prime}. The D=2D′′D^{\prime}=2D^{\prime\prime} line marks the band touching transition at the two KK-points as shown in Fig. 9. At this point the dKy=0d^{y}_{K}=0 and the 𝐝K\mathbf{d}_{K} vector is restricted to the xx-zz plane. When we go around the KK point in the (kx,ky)(k_{x},k_{y}) plane, the vector (dKx,dKz)(d^{x}_{K},d^{z}_{K}) winds once around the origin.

Using Eq. (83), we integrate the Berry curvature Eq. (91) in a disk around the KK and KK^{\prime} points that has radius kk

CK(k)\displaystyle C_{K}(k) =12π0kΩK(k)2πk𝑑k=\displaystyle=\frac{1}{2\pi}\int_{0}^{k}\Omega_{K}(k^{\prime})2\pi k^{\prime}dk^{\prime}=
=m2sgn(k0)(1|k0|k02+k2).\displaystyle=\frac{m}{2}{\rm sgn}(k_{0})\left(1-\frac{|k_{0}|}{\sqrt{k_{0}^{2}+k^{2}}}\right)\;. (92)

In the vicinity of the transition point the second term goes to zero, and the KK and KK^{\prime} points both contribute m2\frac{m}{2} to the total Chern number of the bands. As we cross the transition line D=2D′′D^{\prime}=2D^{\prime\prime}, i.e. as the k0k_{0} changes sign, the m2\frac{m}{2} Berry charge is exchanged and the Chern number is changed by ±1\pm 1 both at K and K’ – the transferred charge is determined by the winding number at the touching point. The total change of the Chern number through the linear touching at the zone-corners is the sum of the contribution of KK and KK^{\prime}, corresponding to ±2\pm 2, as indicated in Fig. 9.

Let us note that CKC_{K} does not give the total Chern number of the band calculated numerically in Fig. 9. CKC_{K} only accounts for the Berry curvature concentrated around the KK (and KK^{\prime}) points, and does not include the contribution from the vicinity of the Γ\Gamma point, which is significant for the middle band as shown in Fig. 10. We use CKC_{K} to discuss the exchange of topological charges through the band-touching transition at the corners of the BZ. For obtaining the total Chern number, one needs to consider the contribution of the Berry curvature at the zone corner too, which we discuss next.

V.2 Quadratic band touching at Γ\Gamma

The quadratic touching is a little different. In the following, we will show that the total Berry charge is exchanged at a single point, where the Chern number changes by ±2\pm 2. Thus, at the quadratic touching the bands have twice as much Berry charge as at the linear touching point. Furthermore, the Berry curvature in the case of the quadratic touching is not centered at a single point, as was the case with the linear touching. Instead, it is concentrated on a ring around the touching point. As the bands approach each other, the radius of the ring decreases, shrinking into a point when the bands touch. To see how this happens, we follow the procedure described above, expanding the rotated Mm,𝐤U(1)M^{\text{U(1)}}_{m,{\bf k}} about the Γ\Gamma point. The transformation matrix that diagonalizes the Heisenberg model Mm,𝐤SU(2)M^{\text{SU(2)}}_{m,{\bf k}} at the zone center has the form

UΓ=(13121613121613026).U_{\Gamma}=\left(\begin{array}[]{ccc}\frac{1}{\sqrt{3}}&\frac{1}{\sqrt{2}}&\frac{1}{\sqrt{6}}\\ \frac{1}{\sqrt{3}}&-\frac{1}{\sqrt{2}}&\frac{1}{\sqrt{6}}\\ \frac{1}{\sqrt{3}}&0&-\frac{2}{\sqrt{6}}\\ \end{array}\right)\;. (93)

Rotating Mm,𝐤U(1)M^{\text{U(1)}}_{m,{\bf k}} with UΓU_{\Gamma}, it becomes block-diagonal at the Γ\Gamma point, decoupling the touching bands from the top band. Moving away from the zone center, additional small matrix elements appear between the subspaces,

UΓMΓmUΓ=(JJ8k2+2JJ82kxkyJ162(kx2ky2)J82kxkyJJ+J8kx2im3(D+D′′)+J8kxkyJ162(kx2ky2)im3(D+D′′)+J8kxkyJJ+J8ky2).U_{\Gamma}^{\dagger}\cdot M^{m}_{\Gamma}\cdot U_{\Gamma}\!=\!\left(\begin{array}[]{ccc}J-\frac{J^{\prime}}{8}k^{2}+2J^{\prime}&-\frac{J^{\prime}}{8\sqrt{2}}k_{x}k_{y}&-\frac{J^{\prime}}{16\sqrt{2}}(k_{x}^{2}-k_{y}^{2})\\ -\frac{J^{\prime}}{8\sqrt{2}}k_{x}k_{y}&J-J^{\prime}+\frac{J^{\prime}}{8}k_{x}^{2}&-im\sqrt{3}(D^{\prime}+D^{\prime\prime})+\frac{J^{\prime}}{8}k_{x}k_{y}\\ -\frac{J^{\prime}}{16\sqrt{2}}(k_{x}^{2}-k_{y}^{2})&im\sqrt{3}(D^{\prime}+D^{\prime\prime})+\frac{J^{\prime}}{8}k_{x}k_{y}&J-J^{\prime}+\frac{J^{\prime}}{8}k_{y}^{2}\\ \end{array}\right)\;. (94)

where kx2+ky2=k2k^{2}_{x}+k^{2}_{y}=k^{2}. We keep only the leading terms, taking D,D′′JD^{\prime},D^{\prime\prime}\ll J^{\prime} we neglect terms as Dk2D^{\prime}k^{2} and D′′k2D^{\prime\prime}k^{2}. Diagonalizing this matrix at the Γ\Gamma point we find the energies ω1=J+2J\omega_{1}=J+2J^{\prime} for the upper band, and ω2,3=JJ±m3(D+D′′)\omega_{2,3}=J-J^{\prime}\pm m\sqrt{3}(D^{\prime}+D^{\prime\prime}) for the lower bands split by the DM-interactions [cf. Eq. (76)]. The effective 2×22\times 2 Hamiltonian describing the splitting in the vicinity of the Γ\Gamma point is the bottom right corner of the matrix (94),

HΓeff=JJ(1k216)I2+𝐝Γ𝝈,H_{\Gamma}^{\text{eff}}=J-J^{\prime}\left(1-\frac{k^{2}}{16}\right)I_{2}+\mathbf{d}_{\Gamma}\cdot\bm{\sigma}\;, (95)

where the first term is an energy shift, and the 𝐝Γ\mathbf{d}_{\Gamma} reads

dΓx\displaystyle d^{x}_{\Gamma} =\displaystyle= 116J2kxky,\displaystyle\frac{1}{16}J^{\prime}2k_{x}k_{y}\;, (96a)
dΓy\displaystyle d^{y}_{\Gamma} =\displaystyle= m3(D+D′′),\displaystyle m\sqrt{3}(D^{\prime}+D^{\prime\prime})\;, (96b)
dΓz\displaystyle d^{z}_{\Gamma} =\displaystyle= 116J(kx2ky2).\displaystyle\frac{1}{16}J^{\prime}(k_{x}^{2}-k_{y}^{2})\;. (96c)

In the absence of the DM interactions dΓy=0d^{y}_{\Gamma}=0, and we can recognize the mark of the quadratic band touching: as we go around the Γ\Gamma point in the (kx,ky)(k_{x},k_{y}) plane, the vector (dΓx,dΓz)(d^{x}_{\Gamma},d^{z}_{\Gamma}) winds twice around the origin Chong et al. (2008). Turning on the DD^{\prime} and/or the D′′D^{\prime\prime}, a gap Δ~Γ(𝐤)=2𝐝Γ𝐝Γ\tilde{\Delta}_{\Gamma}(\mathbf{k})=2\sqrt{\mathbf{d}_{\Gamma}\cdot\mathbf{d}_{\Gamma}} opens between the two bands of the size

Δ~Γ(𝐤)=12(D+D′′)2+164J2k4,\tilde{\Delta}_{\Gamma}(\mathbf{k})=\sqrt{12(D^{\prime}+D^{\prime\prime})^{2}+\frac{1}{64}{J^{\prime}}^{2}k^{4}}, (97)

[c.f. Eq. (76)]. Inserting 𝐝Γ(𝐤)\mathbf{d}_{\Gamma}(\mathbf{k}) into Eq. (84) for the Berry curvature of the lower band, we get

ΩΓ(k)=4m3(D+D′′)(Jk8)2[12(D+D′′)2+(Jk28)2]3/2,\Omega_{\Gamma}(k)=\frac{4m\sqrt{3}(D^{\prime}+D^{\prime\prime})\left(\frac{J^{\prime}k}{8}\right)^{2}}{\left[12(D^{\prime}+D^{\prime\prime})^{2}+\left(\frac{J^{\prime}k^{2}}{8}\right)^{2}\right]^{3/2}}\;, (98)

while ΩΓ(k)\Omega_{\Gamma}(k) changes sign for the upper band. It is convenient to rewrite the curvature as

ΩΓ(k)=m2k2k02(k04+k4)3/2sgn(D+D′′),\Omega_{\Gamma}(k)=m\frac{2k^{2}k_{0}^{2}}{\left(k_{0}^{4}+k^{4}\right)^{3/2}}\operatorname{sgn}(D^{\prime}+D^{\prime\prime})\;, (99)

where

k02=163|D+D′′|J.k_{0}^{2}=16\sqrt{3}\frac{|D^{\prime}+D^{\prime\prime}|}{J^{\prime}}\;. (100)

ΩΓ(k)\Omega_{\Gamma}(k) has a maximum for k=21/4k0k=2^{-1/4}k_{0}, where it diverges as

ΩΓ(21/4k0)mk02sgn(D+D′′)mJD+D′′\Omega_{\Gamma}(2^{-1/4}k_{0})\propto\frac{m}{k_{0}^{2}}\operatorname{sgn}(D^{\prime}+D^{\prime\prime})\propto m\frac{J^{\prime}}{D^{\prime}+D^{\prime\prime}} (101)

for D+D′′0D^{\prime}+D^{\prime\prime}\to 0. Furthermore, the ΩΓ(k)\Omega_{\Gamma}(k) vanishes for both kk0k\ll k_{0} and kk0k\gg k_{0}:

ΩΓ(k)={msgn(D+D′′)2k2k04+,if kk0;msgn(D+D′′)2k02k4+,if kk0.\Omega_{\Gamma}(k)=\begin{cases}m\operatorname{sgn}(D^{\prime}+D^{\prime\prime})\frac{2k^{2}}{k_{0}^{4}}+\cdots,&\mbox{if }k\ll k_{0};\\ m\operatorname{sgn}(D^{\prime}+D^{\prime\prime})\frac{2k_{0}^{2}}{k^{4}}+\cdots,&\mbox{if }k\gg k_{0}.\end{cases} (102)

The maximum of the Berry curvature forms a ring around the Γ\Gamma point. The ring is nicely seen for the lowest two bands in Fig. 10. This behavior is unlike the linear band touching, where the Berry curvature is concentrated at the touching KK-points.

Refer to caption
Figure 11: The radial distribution of the Berry curvature Ω(k)\Omega(k) (solid lines) and the contribution to the Chern number C(k)C(k) (dashed lines) for the linear band touching around the K points [Eqs. (91) and (92)] and for the quadratic band touching around the Γ\Gamma point [Eqs. (99) and (103)] in the Brillouin zone. The Berry curvature is maximal at the KK point for linear band touching, while in the case of the quadratic band touching it forms a ring–like structure around the Γ\Gamma point. In both cases, their integrals over a disk of radius kk centered at the touching points saturates quickly at the 1/21/2 and 11 values (C(k)C(k)).

Using Eq. (83) and integrating the curvature around the Γ\Gamma point within a disk of radius kk, we can check that the ring has enough curvature to collect a contribution ±1\pm 1 to the Chern number of the bands:

CΓ(k)\displaystyle C_{\Gamma}(k) =msgn(D+D′′)(1k02k04+k4)\displaystyle=m\operatorname{sgn}(D^{\prime}+D^{\prime\prime})\left(1-\frac{k_{0}^{2}}{\sqrt{k_{0}^{4}+k^{4}}}\right) (103)
msgn(D+D′′)(1k02k2),\displaystyle\approx m\operatorname{sgn}(D^{\prime}+D^{\prime\prime})\left(1-\frac{k_{0}^{2}}{k^{2}}\right), if kk0.\displaystyle\mbox{if }k\gg k_{0}. (104)

Here m=1,0,1m=-1,0,1, and the sign depends on the band as well as the sign of the DM interaction. As the DM is continuously tuned across the the quadratic band touching transition, the +1+1 and 1-1 Berry charges are exchanged between the two bands, leading to the ΔC=2\Delta C=2 transition for the two lowest bands, as seen along the D=D′′D^{\prime}=-D^{\prime\prime} line in Fig. 9 and 10.

Again, we emphasize that investigating CΓC_{\Gamma} characterizes how much the topological charge changes through band touching transition, and cannot produce the total Chern number of the band. For example, the middle band would have Berry curvature contribution from the corners of the BZ too, this however does not change when the gap closes at the Γ\Gamma point.

Refer to caption
Figure 12: (a)–(c) Schematic figures representing periodic boundaries along the xx direction and a finite size in the yy direction with various edges. (d)–(f) The projected m=±1m=\pm 1 triplet bands and helical triplet edge modes for finite second neighbor DM interaction, D′′=0.01D^{\prime\prime}=0.01 in the open geometries shown in (a)–(c), respectively. The edge modes are colored according to the spin degrees of freedom and the edges, with red colors representing the up-spin and the blue colors denoting the down-spin. The open arrows correspond to the bottom edge, and the filled ones to the top edge.
Refer to caption
Figure 13: (a) The spin Nernst coefficient, αxy\alpha_{xy} along the blue dashed line of Fig. 9 as a function of temperature and D′′D^{\prime\prime} (J/J=0.2J^{\prime}/J=0.2, D/J=0.1D^{\prime}/J=0.1). The white dashed lines indicate the boundaries of the band touching topological transitions, where the Berry curvatures of the touching bands are exchanged. This change in the topological charge is also signaled by the anomaly of spin Nernst effect at the transition lines. (b) The contributions of the Sz=+1S^{z}=+1 and 1-1 triplet bands to the Nernst coefficient as the function of D′′D^{\prime\prime} at given temperatures. The αxy\alpha_{xy} has the same size and an opposite sign for the up and down spins, corresponding to their degenerate energies but opposite chirality. (c) The total αxy=α1,xyα1,xy\alpha_{xy}=\alpha_{1,xy}-\alpha_{-1,xy} Nernst coefficient as the function of D′′D^{\prime\prime} at various temperature values.

VI Triplet 𝒵2\mathcal{Z}_{2} topological insulator

We now turn to the problem of classifying the topological phases of the model, and characterizing their experimental characteristics. We start with the gapped, topological phase found in the simplified model of Section IV, which we show to be a spin–Hall state characterized by a 𝒵2\mathcal{Z}_{2} topological invarient, in direct analogue with the model of Kane and Mele Kane and Mele (2005a).

Within each mm subspace, we can think about the DM interaction term as an effective magnetic field in momentum space that acts on a pseudo spin-1 corresponding to the sublattice degree of freedom. The m=0m=0 subspace is clearly no subject to any kind of magnetic field, while the spinful m=1m=1 and 1-1 triples are affected in the opposite way, the effective field (DM) splits the "spin-1", both m=1m=1 and m=1m=-1 corresponding to the Haldane model Haldane (1988) with opposite chirality. Therefore, the time-reversal pair m=1m=1 and m=1m=-1 together realize an analog of the Kane and Mele model. Let us emphasize that the triplets, being components of an integer spin, do not form Kramer’s pairs, and their degeneracy may be lifted even when the TR symmetry is preserved. When the nematic exchange anisotropies are present, for example, the STzS^{z}_{T} ceases to be a good quantum number, and the m=1m=1 and m=1m=-1 subspaces hybridize, destroying the 𝒵2\mathcal{Z}_{2} phase. We discuss this scenario in Sec. VII

In the case when the nematic terms are zero, the complete overlap between the m=1m=1 and m=1m=-1 bands renders the net Chern number zero. The m=±1m=\pm 1 triplets realize an analog of the spin Hall insulator state and are characterized by a 𝒵2\mathcal{Z}_{2} invariant Kane and Mele (2005a, b); Hasan and Kane (2010); Qi and Zhang (2011).

As shown in Sec. IV, the Chern numbers are multiples of mm (see Fig: 9), and the bands with opposite spin have opposite Chern number. Therefore, the total Chern number of each degenerate band, formed by the time-reversal partners, m=±1m=\pm 1, vanishes: Cn,1+Cn,1=0C_{n,1}+C_{n,-1}=0. Similar to electronic systems with conserved STzS^{z}_{T}, the 𝒵2\mathcal{Z}_{2} index can be understood as the “spin Chern number” Kondo et al. (2019a) and computed as the staggered quantity 12(Cn,1Cn,1)mod2\frac{1}{2}(C_{n,1}-C_{n,-1})\mod 2.

For the bottom and top bands the Chern numbers in each phase are ±m\pm m, thus the spin Chern number is ±22mod2=1\pm\frac{2}{2}\mod 2=1. The middle band has either Chern number 0, or ±2m\pm 2m, resulting in a trivial 0 𝒵2\mathcal{Z}_{2} index. This also shows that the 𝒵2\mathcal{Z}_{2} topological invariant does not depend on the DM anisotropy, but only on the conservation of STzS^{z}_{T}. Even when we close the band gaps at the band touching transitions, the 𝒵2\mathcal{Z}_{2} indices will not change. The 𝒵2\mathcal{Z}_{2} index can be computed using the eigenvalues of the parity operator too, which we discuss in detail in the Appendix B.

As a consequence of the 𝒵2\mathcal{Z}_{2} topology, the system with open boundaries has helical triplet edge-modes, as shown in Fig. 12. Note that one helical edge state is made of two chiral edge states going in opposite directions. We chose three different edge geometries, illustrated in Fig. 12 (a)–(c), and computed the bands for each of those (see Fig. 12 (d)–(f)). The spin degree of freedom of the edge-modes is denoted with red and blue colors, while the filled and open arrows corresponds to the top and bottom edges, respectively.

VI.1 Triplet Nernst effect

To obtain an experimentally detectable signature of the 𝒵2\mathcal{Z}_{2} triplet bands, we compute the boson analog of the spin Hall effect. Applying a temperature gradient on the sample induces an energy current of triplet excitations. The m=1m=1 and 1-1 triplets are affected in an opposite way by the DM interaction, due to their opposite chirality, and deflect into opposite directions. The triplet spin separation perpendicular to the temperature gradient leads to the cancelation of the transverse triplet heat current, but gives a finite transverse spin current. The transverse spin current arising in response to an applied temperature gradient is called the Nernst effect. We directly apply the formula of magnon mediated spin Nernst effect Cheng et al. (2016); Zyuzin and Kovalev (2016); Kovalev and Zyuzin (2016); Nakata et al. (2017); Zyuzin and Kovalev (2018) for the triplet excitations,

jSN=αxy𝐳^×T,\displaystyle j_{\text{SN}}=\alpha_{xy}\mathbf{\hat{z}}\times\nabla T\;, (105)

where the spin Nernst coefficient, αxy\alpha_{xy} can be expressed as

αxy=ikBm,nBZmc1(ρn,m)Fn,mxy(𝐤)d2𝐤,\displaystyle\alpha_{xy}=-i\frac{k_{B}}{\hbar}\sum_{m,n}\int_{\rm BZ}m\cdot c_{1}(\rho_{n,m})F^{xy}_{n,m}(\mathbf{k})d^{2}\mathbf{k}\;, (106)

where Fn,mxy(𝐤)F^{xy}_{n,m}(\mathbf{k}) is the Berry curvature of the nn-th band of triplet mm defined in Eq. (79)

c1(ρ)=0ρln(1+t1)𝑑t=(1+ρ)ln(1+ρ)ρlnρc_{1}(\rho)=\int_{0}^{\rho}\ln(1+t^{-1})dt=(1+\rho)\ln(1+\rho)-\rho\ln\rho\, (107)

and

ρn,m=(eωn,mβ1)1\displaystyle\quad\rho_{n,m}=(e^{\omega_{\!n\!,\!m\!}\beta}-1)^{-1} (108)

is the Bose–Einstein distribution function.

A density plot of the numerically computed triplet mediated spin Nernst coefficient, αxy\alpha_{xy} is shown in Fig. 13 as function of temperature and D′′D^{\prime\prime}. We calculated αxy\alpha_{xy} along the blue dashed line in Fig. 9, using the complete Bogoliubov–de Gennes Hamiltonian. At the topological band touching lines D′′=DD^{\prime\prime}=-D^{\prime} and D′′=2DD^{\prime\prime}=2D^{\prime}, the spin Nernst effect has inflection points, corresponding to the exchange of topological charge between the touching bands.

As long as the magnetic field hzh_{z} is zero and the time reversal symmetry is preserved, αxy,1=αxy,1\alpha_{xy,1}=-\alpha_{xy,-1}, and the transverse spin Nernst current, jSN=jSN,1jSN,1j_{\text{SN}}=j_{\text{SN},1}-j_{\text{SN},-1}, can be written as 2jSN,12j_{\text{SN},1}. An applied magnetic field Zeeman-splits the triplets, pushing the m=1m=1 and m=1m=-1 bands in opposite directions (see Sec.VIII). As a consequence, the thermal filling of m=1m=1 and m=1m=-1 becomes different, leading to an imbalanced contribution from the up and down spins but still providing a finite spin Nernst effect. Note that we consider an out-of-plane field direction, that does not harm the U(1) symmetry, preserving STzS^{z}_{T} as good quantum number.

VII Nematic interaction and the fate of the 𝒵2\mathcal{Z}_{2} phase

We now explore the consequences of the terms which mix the triplets with m=±1m=\pm 1, namely the nematic interactions introduced in Section II. Symmetric exchange anisotropies of this type are naturally present in many spin systems, and arise in both the single– and bilayer kagome and honeycomb models, widely discussed as candidates for topological magnon and triplon phases.

In the original Kane and Mele model Kane and Mele (2005a), analagous spin–mixing terms, such as the Rashba spin-orbit coupling, can be present (when the σh\sigma_{h} reflection is broken). Although such a term hybridizes the bands with up and down spins, the spin-degeneracy at the time reversal invariant momenta (TRIM) remains protected by Kramer’s theorem. As a consequence, the 𝒵2\mathcal{Z}_{2} topological phase considered by Kane and Mele is perturbatively stable against the introduction of Rashba interactions.

To explicitly see the difference from the Kane-Mele model, we consider the subspace of the pseudo-spin-half formed by the +1+1 and 1-1 triplets, and investigate the fate of the 𝒵2\mathcal{Z}_{2} band-topology when the spin-mixing nematic interactions are finite. We first examine their effect on the spin-degeneracy at the TRIM [Section VII.1]. Then, taking an open geometry, we examine the consequence of the nematic interactions for the non-trivial edge modes [Section VII.2]. To formalize our findings, in Section VII.3 , we show that the TR×\timesU(1) symmetry corresponds to a pseudo time-reversal operator Θ\Theta, which squares to 1-1 and can protect the 𝒵2\mathcal{Z}_{2} band-topology. We identify the terms in our model that possess this symmetry and those which break it.

VII.1 Gap opening at the TRIM

We consider first the effect of nematic interactions on triplon dispersion at time–reversal invariant momenta (TRIM). The Bogoliubov–de Gennes Hamiltonian of the TR-pair m=±1m=\pm 1 triplets corresponds to the Hamiltonian (30a):

H𝐤(1,1)=(M𝐤N𝐤N𝐤M𝐤).\displaystyle H_{\mathbf{k}}^{(1,-1)}=\left(\!\!\begin{array}[]{cc}M_{\mathbf{k}}&N_{\mathbf{k}}\\ N_{\mathbf{k}}&M_{\mathbf{k}}\end{array}\!\!\right)\;. (111)

We compute the energies of the bands at the TRIM, namely

Γ\displaystyle\Gamma =\displaystyle= (0,0),𝖬1=(π,π3),\displaystyle(0,0)\;,\qquad\mathsf{M}_{1}=\left(\pi,\frac{\pi}{\sqrt{3}}\right)\;,
𝖬2\displaystyle\mathsf{M}_{2} =\displaystyle= (0,2π3),𝖬3=(π,π3)\displaystyle\left(0,\frac{2\pi}{\sqrt{3}}\right)\;,\qquad\mathsf{M}_{3}=\left(-\pi,\frac{\pi}{\sqrt{3}}\right)\, (112)

shown in the bottom panel of Fig. 1(a). Including the nematic terms, the energies at the Γ\Gamma point become

ω1,2(Γ)=\displaystyle\omega_{1,2}(\Gamma)= J+J232(D+D′′)Λ,\displaystyle J\!+\!\frac{J^{\prime}}{2}\!-\!\frac{\sqrt{3}}{2}(D^{\prime}\!+\!D^{\prime\prime})-\Lambda, (113a)
ω3,4(Γ)=\displaystyle\omega_{3,4}(\Gamma)= JJ+3(D+D′′)±12(K4K),\displaystyle J\!-\!J^{\prime}\!+\!\sqrt{3}(D^{\prime}\!+\!D^{\prime\prime})\pm\frac{1}{2}(K_{\|}\!-\!4K^{\prime}_{\|}), (113b)
ω5,6(Γ)=\displaystyle\omega_{5,6}(\Gamma)= J+J232(D+D′′)+Λ,\displaystyle J\!+\!\frac{J^{\prime}}{2}\!-\!\frac{\sqrt{3}}{2}(D^{\prime}\!+\!D^{\prime\prime})\!+\!\Lambda, (113c)

where

Λ=123(D+D′′+3J)2+(K+2K)2\displaystyle\Lambda=\frac{1}{2}\sqrt{3(D^{\prime}+D^{\prime\prime}+\sqrt{3}J^{\prime})^{2}+(K_{\|}+2K^{\prime}_{\|})^{2}} (114)

At the zone center, the top and bottom bands remain degenerate, while the middle two bands split as a result of the nematic interaction.

The effect of the nematic term is more drastic at the less symmetric 𝖬\mathsf{M} points, where the energies of the six triplets are all non-degenerate

ω1,2(Γ)=\displaystyle\omega_{1,2}(\Gamma)= JKΛ±,\displaystyle J\mp K^{\prime}_{\|}-\Lambda_{\pm}\;, (115a)
ω3,4(Γ)=\displaystyle\omega_{3,4}(\Gamma)= JK2,\displaystyle J\mp\frac{K_{\|}}{2}\;, (115b)
ω5,6(Γ)=\displaystyle\omega_{5,6}(\Gamma)= JK+Λ±,\displaystyle J\mp K^{\prime}_{\|}+\Lambda_{\pm}\;, (115c)

with

Λ±=12(2JK2)2+(2(DD′′)±32K)2.\displaystyle\Lambda_{\pm}=\frac{1}{2}\sqrt{(2J^{\prime}-\frac{K_{\|}}{2})^{2}+(2(D^{\prime}-D^{\prime\prime})\pm\frac{\sqrt{3}}{2}K_{\|})^{2}}\;. (116)

The opening of a gaps ΔK\Delta\sim K at these TRIM implies that even infinitesimal nematic interactions are effective in destroying the 𝒵2\mathcal{Z}_{2} band-topology.

We note that, for simplicity, the eigenvalues and band-gaps above have been calculated considering the M𝐤M_{\bf{k}} matrix and not the entire Bogoliubov–de Gennes Hamiltonian. Solving the Bogoliubov–de Gennes problem denies us a simple analytical form, however, the physics remains the same: the infinitesimal nematic terms break spin-degeneracy, opening a gap at the TRIM. This contrasts with the Kane–Mele model, where Rashba coupling does not break the time–reversal symmetry protecting the degeneracy at the TRIM Kane and Mele (2005b). And it follows that infinitesimal Rashba interaction cannot lift the degeneracy of Kramers doublets at TRIM.

We return to this point in Section VII.3, where we analyze the symmetry protecting 𝒵2\mathcal{Z}_{2} band topology in our model, and show that nematic interactions break this symmetry. Before that, we examine another consequence of the loss of 𝒵2\mathcal{Z}_{2} band topology, namely the hybridization of the edge states.

VII.2 Hybridization of the edge states

The band splitting at the TRIM already indicates that the degeneracy at these points are not protected by a symmetry corresponding to the time-reversal in the Kane-Mele model. To elaborate on the effect of the nematic terms on the 𝒵2\mathcal{Z}_{2} topology, we compute the bands of the bilayer kagome stripes for the three different edge types shown in Fig. 12 (a)–(c). In an open system, the time reversal invariant point is k=πk=\pi. The spin-degeneracy at this point will not necessarily be protected by TR symmetry, as it were for a Kramer’s pair. To illustrate this, in Fig. 14, we plot the spinful triplet bands for the various edge-geometries for finite intra-dimer ((a)–(c)) and finite inter-dimer nematic interaction ((d)–(f)). We find that in each case the edge modes hybridize, becoming trivial via an avoided crossing, i.e. they no longer connect the bands and close the gap. This is a clear indication that the fragile 𝒵2\mathcal{Z}_{2} topology is quashed by the nematic terms.

Refer to caption
Figure 14: The effect of the nematic interaction on the edge modes and the topology of the bands. (a), (b), and (c) show the case when J=0.1JJ^{\prime}=0.1J, D′′=0.01JD^{\prime\prime}=0.01J, and K=0.02JK_{\|}=0.02J, for the spiky, flat, and mixed geometries, respectively (see Fig. 12 (a)–(c)). The KK^{\prime}_{\|} is zero. The finite intra-dimer nematic term (KK) leads to the hybridization of the helical edge modes with up and down spin, and a gap is opened at k=πk=\pi between the boundary-modes on the spiky edges, rendering them topologically trivial. The boundary modes at the flat edges do not merge with the middle band at k=πk=\pi, but avoid that and become trivial as well as shown in the insets of (b) and (c). (d), (e), and (f) represent the case when the inter-dimer nematic interaction is finite, K=0.01JK^{\prime}_{\|}=0.01J and J=0.1JJ^{\prime}=0.1J, D′′=0.01JD^{\prime\prime}=0.01J but K=0K_{\|}=0 . Here too the edge-modes with up and down spins hybridize and become gapped at π\pi. In the case of the mixed boundaries, the edge-modes merge hybridization within the gap (see inset of (f)), and thus cannot collapse the band-gap, signaling topologically trivial bands.

We note that the vanishing of 𝒵2\mathcal{Z}_{2} topology is not the consequence of a band-touching transition, which happens in the original Kane and Mele model, where increasing the Rashba term will close the band gap and induce a topological transition from a spin-Hall state to a trivial one. The values of the nematic terms were carefully chosen to be smaller than the critical values at which the band-gaps close (see Appendix C).

VII.3 Pseudo-TR symmetry of the Hamiltonian

In what follows, we give a more formal proof for the instability of the 𝒵2\mathcal{Z}_{2} band-topology against an arbitrary small nematic interaction. For this, we will express the TR×\timesU(1) symmetry as a tensor product of 2×22\times 2 and a 3×33\times 3 matrix comprising the spin and sublattice degrees of freedom, respectively. The TR is an anti-unitary operator and in the basis (𝐭𝐤,𝐭𝐤)(\mathbf{t}_{{\bf k}}^{\dagger},\mathbf{t}_{-{\bf k}}^{\phantom{\dagger}}) defined in Eqs. (31) has the form

T=(I22sxI3)𝒦,T=(I_{2}\otimes 2s^{x}\otimes I_{3})\cdot\mathcal{K}\;, (117)

where the 2-dimensional identity matrix I2I_{2} accounts for the particle-hole space, 2sx=σx2s^{x}=\sigma^{x} acts on the spin space spanned by m=±1m=\pm 1, I3I_{3} acts on the A, B, C sublattice degrees of freedom, and 𝒦\mathcal{K} is the complex conjugation. For details on the action of TT on the +1+1 and 1-1 triplets see Appendix A. We note that since the +1+1 and 1-1 triplets have integer spins the square of the TR is T2=𝟏T^{2}=\mathbf{1}, and it cannot ensure the Kramers degeneracy.

From our numerics it appears that the U(1)U(1) symmetry — the conservation of the STzS^{z}_{T} — is needed for the 𝒵2\mathcal{Z}_{2} topology. In the 2×22\times 2 spin subspace the U(1)U(1) rotation are described by the eiφ2szcos(φ)I2isin(φ)2sze^{-i\varphi 2s^{z}}\propto\cos(\varphi)I_{2}-i\sin(\varphi)2s^{z}, so the commutation of the bond wave Hamiltonian with the

R=I22szI3,R=I_{2}\otimes 2s^{z}\otimes I_{3}\;, (118)

unitary matrix ensures the U(1)U(1) symmetry – it commutes with the M𝐤XXZM^{\text{XXZ}}_{\mathbf{k}}, M𝐤DMM^{\text{DM}}_{\mathbf{k}}, and M𝐤ZeemanM^{\text{Zeeman}}_{\mathbf{k}} bond wave Hamiltonian matrices in Eq. (61): RΣzM𝐤XXZΣzM𝐤XXZR=0R\cdot\Sigma^{z}\cdot M^{\text{XXZ}}_{\mathbf{k}}-\Sigma^{z}\cdot M^{\text{XXZ}}_{\mathbf{k}}\cdot R=0, and so on, where we defined the pseudo-identity for the Bogoliubov–de Gennes formalism,

Σz=2szI2I3.\Sigma_{z}=2s_{z}\otimes I_{2}\otimes I_{3}\;. (119)

The composition of the TR and U(1) symmetry gives the

Θ=(I22isyI3)𝒦\Theta=(I_{2}\otimes 2is^{y}\otimes I_{3})\cdot\mathcal{K}\\ (120)

anti-unitary operator. Since

Θ2=1,\Theta^{2}=-1\;, (121)

Θ\Theta is the desired pseudo-time reversal operator, assuming the role of the TR symmetry and ensuring the protection of the 𝒵2\mathcal{Z}_{2} topology. We note that an analogous pseudo-TR operator was previously introduced in Ref. Kondo et al., 2019a for magnetically ordered spins. However, this operator was associated with the combination of TR and mirror symmetries, which squares to +1 in the present case.

In what follows, we explicitly show that the Hamiltonian containing only the XXZ and DM anisotropies, M𝐤XXZM^{\text{XXZ}}_{\mathbf{k}} and M𝐤DMM^{\text{DM}}_{\mathbf{k}} in Eq. (61), commutes with Θ\Theta and will possess the 𝒵2\mathcal{Z}_{2} topology. The nematic interactions, defined by Eq. (66), however, do not commute neither with STzS^{z}_{T}, nor with Θ\Theta. As a consequence, the 𝒵2\mathcal{Z}_{2} bands will no longer be protected when either of the nematic interactions are finite.

For this purpose we introduce the anti-unitary operator

𝒯=(i2syI3)𝒦,\displaystyle\mathcal{T}=(i2s^{y}\otimes I_{3})\cdot\mathcal{K}\;, (122)

which acts as a pseudo TR operator within the particle (hole) subspace:

Θ=I2𝒯=(𝒯𝟎𝟎𝒯).\displaystyle\Theta=I_{2}\otimes\mathcal{T}=\left(\!\!\begin{array}[]{cc}\mathcal{T}&\mathbf{0}\\ \mathbf{0}&\mathcal{T}\end{array}\!\!\right). (125)

The Θ\Theta antiunitary operator is a symmetry of H𝐤(1,1)H_{\mathbf{k}}^{(1,-1)} if Kondo et al. (2019a)

ΘΣzH𝐤(1,1)ΣzH𝐤(1,1)Θ=0.\Theta\Sigma^{z}H_{\mathbf{k}}^{(1,-1)}-\Sigma^{z}H_{\mathbf{k}}^{(1,-1)}\Theta=0\;. (126)

Using Eq. (122), this can be expressed as commutations between 𝒯\mathcal{T} and the diagonal M𝐤M_{\mathbf{k}} and the off-diagonal N𝐤N_{\mathbf{k}} :

ΘΣzH𝐤(1,1)ΣzH𝐤(1,1)Θ=Σz([𝒯,M𝐤][𝒯,N𝐤][𝒯,N𝐤][𝒯,M𝐤])\Theta\Sigma^{z}H_{\mathbf{k}}^{(1,-1)}-\Sigma^{z}H_{\mathbf{k}}^{(1,-1)}\Theta=\Sigma^{z}\begin{pmatrix}\left[\mathcal{T},M_{\mathbf{k}}\right]&\left[\mathcal{T},N_{\mathbf{k}}\right]\\ \left[\mathcal{T},N_{\mathbf{k}}\right]&\left[\mathcal{T},M_{\mathbf{k}}\right]\end{pmatrix} (127)

Since the matrices M𝐤M_{\mathbf{k}} and N𝐤N_{\mathbf{k}} only differ in their diagonal, it is sufficient to consider the commutation of 𝒯\mathcal{T} and M𝐤M_{\mathbf{k}}. The various terms in M𝐤M_{\mathbf{k}} are expressed in the same basis as 𝒯\mathcal{T} in terms of I2I3I_{2}\otimes I_{3}, sαI3s^{\alpha}\otimes I_{3}, I2λnI_{2}\otimes\lambda_{n}, and sαλns^{\alpha}\otimes\lambda_{n}, with α=x,y,z\alpha=x,y,z, and n=1,,8n=1,\ldots,8, as introduced in Sec. III.2. Let us note that the coefficients of these operators only contain even functions of 𝐤{\bf k}, therefore 𝒯\mathcal{T} does not affect those.

The effect of the 𝒯\mathcal{T} operator on a general term can be written as

𝒯(sαλn)𝒯\displaystyle\mathcal{T}(s^{\alpha}\otimes\lambda_{n})\mathcal{T}^{\dagger} =(i2syI3)(sαλn)(i2syI3)\displaystyle=(i2s^{y}\otimes I_{3})(s^{\alpha}\otimes\lambda_{n})^{*}(i2s^{y}\otimes I_{3})^{\dagger}
=(i2sy)(sα)(i2sy)(I3)(λn)(I3)\displaystyle=(i2s^{y})(s^{\alpha})^{*}(i2s^{y})^{\dagger}\otimes(I_{3})(\lambda_{n})^{*}(I_{3})^{\dagger}
=(sα)λn,\displaystyle=(-s^{\alpha})\otimes\lambda_{n}^{*}, (128a)
𝒯(I2λn)𝒯\displaystyle\mathcal{T}(I_{2}\otimes\lambda_{n})\mathcal{T}^{\dagger} =I2λn,\displaystyle=I_{2}\otimes\lambda_{n}^{*}, (128b)
𝒯(sαI3)𝒯\displaystyle\mathcal{T}(s^{\alpha}\otimes I_{3})\mathcal{T}^{\dagger} =sαI3,\displaystyle=-s^{\alpha}\otimes I_{3}, (128c)

where the stands for complex conjugation. The I2I3I_{2}\otimes I_{3} is trivially invariant. Furthermore, the operators I2λnI_{2}\otimes\lambda_{n} preserve (break) the pseudo TR symmetry when the λn\lambda_{n} Gell-Mann matrices are real (imaginary), while the terms sαλns^{\alpha}\otimes\lambda_{n} are invariant if λn\lambda_{n} is imaginary and break 𝒯\mathcal{T} for λn\lambda_{n}\in\mathbb{R}.

Table 2 contains the transformation of the different terms appearing in the triplet Hamiltonian under the pseudo and real TR operators and the U(1) symmetry. We can readily see that the nematic interactions break the U(1) symmetry and consequently the pseudo TR symmetry too that would guarantee the protection of the 𝒵2\mathcal{Z}_{2} topology. The Zeeman term breaks both the pseudo and the physical TR symmetries.

Table 2: Invariance of the various terms appearing in the Hamiltonian for the m=±1m=\pm 1 triplets under the pseudo time-reversal operator, the physical time-reversal symmetry, and the U(1) symmetry. Note that the pseudo TR symmetry corresponds to TR×\timesU(1).
Pseudo TR Physical TR U(1)
(i2syI3)𝒦(i2s^{y}\!\otimes\!I_{3})\mathcal{K} (2sxI3)𝒦(2s^{x}\!\otimes\!I_{3})\mathcal{K} 2szI32s^{z}\otimes I_{3}
XXZ I2I3I_{2}\otimes I_{3} \checkmark \checkmark \checkmark
I2λnI_{2}\otimes\lambda_{n} \checkmark \checkmark \checkmark
(for real λn\lambda_{n})
DM szλns^{z}\otimes\lambda_{n} \checkmark \checkmark \checkmark
(for imaginary λn\lambda_{n})
Nematic sxλns^{x}\otimes\lambda_{n} - \checkmark -
(for real λn\lambda_{n})
syλns^{y}\otimes\lambda_{n} - \checkmark -
(for real λn\lambda_{n})
Zeeman szI3s^{z}\otimes I_{3} - - \checkmark
Refer to caption
Figure 15: The finite magnetic field (h=0.2Jh=0.2J, gz=2g_{z}=2) splits the triplets according to their spin degree of freedom mm. (a) Triplet bands for the D=D′′=0D^{\prime}=D^{\prime\prime}=0 isotropic case. The bands of each mm sector have the same dispersion shifted by the Zeeman energy (b)D=0D^{\prime}=0, D′′=0.01JD^{\prime\prime}=0.01J. The m=±1m=\pm 1 bands become fully gapped with well defined and finite Chern numbers, while the m=0m=0 modes remain unaffected. (c) 1D dispersion in the open geometry (corresponding to Fig. 12(a)) for the D=0D^{\prime}=0, D′′=0.01JD^{\prime\prime}=0.01J case. Nontrivial edge states collapse the anisotropy gaps for m=±1m=\pm 1, signaling nontrivial band topology.
Refer to caption
Figure 16: (a) The thermal Hall coefficient for h=0.2Jh=0.2J along the blue dashed line depicted in the phase diagram of Fig. 9 using the same parameters as in Fig. 13. (b) We show the contributions from the +1+1 and 1-1 triplets separately. The Sz=1S^{z}=-1 triplet moves up in energy, due to the Zeeman splitting, and thus at low temperatures these bands are less filled than those with Sz=1S^{z}=1, giving a smaller contribution. At higher temperature this difference shrinks, and the sum of the signals, having opposite sign, cancels as TT\to\infty. (c) Thermal Hall coefficient at various temperatures as the function of D′′D^{\prime\prime}. The dashed vertical lines indicate the phase transitions at D′′=DD^{\prime\prime}=-D^{\prime} and D′′=D/2D^{\prime\prime}=D^{\prime}/2 respectively, where the thermal Hall coefficient has an inflection point.

VIII Time-reversal symmetry breaking and thermal Hall effect

So far, we have only considered states found in the absence of magnetic field. However the breaking of time–reversal symmetry by magnetic field also has interesting consequences. In Fig. 15, we show how the triplon band structure of a model with Heisenberg and DM interactions changes as a function of magnetic field.

An immediate consequence of magnetic field is that the time–reversal pairs m=±1m=\pm 1 split, and their Chern numbers no longer cancel. As a result, we end up with an analog of a Chern insulator state, but realized by the triplets.

A second consequence of the Zeeman splitting of triplets is that the thermal filling of m=1m=1 and m=1m=-1 becomes different, and the imbalance of the up and down spin current produces a finite thermal Hall coefficient. The thermal Hall signal is the transverse energy current in response to an applied temperature gradient (and perpendicular magnetic field). In the TR symmetric case the up and down-spin triplets had overlapping energies and consequently identical thermal filling, providing the same number of excitations moving in opposite directions and giving a zero net thermal Hall response. As the degeneracy is lifted, the up and down-spin contributions become different giving a finite net transverse energy current. The thermal Hall coefficient can be written as Matsumoto and Murakami (2011).

κxy=i1βn,mBZc2(ρn,m)Fn,mxy(𝐤)d2𝐤,\displaystyle\kappa_{xy}=-i\frac{1}{\beta}\sum_{n,m}\int_{\rm BZ}c_{2}(\rho_{n,m})F^{xy}_{n,m}(\mathbf{k})d^{2}\mathbf{k}\;, (129)

where Fn,mxy(𝐤)F^{xy}_{n,m}(\mathbf{k}) is the Berry’s curvature, c2(ρ)=0ρln2(1+t1)𝑑tc_{2}(\rho)=\int_{0}^{\rho}\ln^{2}(1+t^{-1})dt, and ρn,m\rho_{n,m} is the Bose-Einstein distribution.

We computed the thermal Hall coefficient, κxy\kappa_{xy} along the blue dashed line indicated in Fig. 9 as the function of temperature and using the entire Bogoliubov–de Gennes Hamiltonian. κxy\kappa_{xy} is plotted in Fig. 16.

In summary, the parameters DD^{\prime} and D′′D^{\prime\prime} together with the Chern numbers define a topological phase diagram as seen on Fig. 9. The different phases can be clearly distinguished by the thermal Hall coefficient which is largest in the (m,0,m)(m,0,-m) phase, as can be seen on Fig. 16. The effect is equally large in the (m,0,m)(-m,0,m) phase, but of opposite sign, as this phase can be reached simply by inverting the sign of DD^{\prime} and D′′D^{\prime\prime} thereby negating the sign of the mm-dependent term in the Hamiltonian.

Let us point out that including the nematic terms have no effect on the Chern bands, unlike the case of the 𝒵2\mathcal{Z}_{2} bands. As long as the nematic interactions are small enough for the band-gap to remain open, the topology of the triplet bands does not change.

IX Conclusions

In this Article, we investigate some of the novel features which arise in the topology of the triplon bands found in spin–1/2 quantum paramagnet on bilayer kagome lattice. We go beyond the XXZ model extended with the DM interactions, the archetypal analogue of the electronic tight-binding hopping Hamiltonian with spin-orbit coupling. Deriving the most general form of the Hamiltonian allowed by the symmetry of the lattice [Section II] we explore the ramifications of each symmetry-allowed terms. Reducing the Hamiltonian to a model for triplon excitations of the quantum paramagnet [Section III], we characterize these bands for models of increasing complexity, by lowering the symmetries.

The simplest case is when the Hamiltonian is the pure SU(2) symmetric Heisenberg model, discussed in Sec. III.3. In this case, the band structure is trivial, with 3-fold degenerate bands, exhibiting a quadratic band touching at the Γ\Gamma point and linear band touchings at the KK and KK^{\prime} points in the Brillouin zone.

In the TR×\timesU(1) symmetric case, we find that the triplets provide an analog to the 𝒵2\mathcal{Z}_{2} topological insulator considered by Kane and Mele Kane and Mele (2005a, b), with helical triplet modes on open edges [Section IV]. This model supports topological phases with bands having different Chern numbers. We give a detailed description of the exchanged topological charges at the phase transitions, for both the linear and quadratic touchings [Section V]. The behaviour of the 𝒵2\mathcal{Z}_{2} topological phase is also characterized through calculations of its topological invariant, and the associated triplet Nernst effect [Section VI].

Finally, we explore two different mechanisms which can eliminate the 𝒵2\mathcal{Z}_{2} topology of the triplon bands by breaking either the pseudo-TR symmetry with the inclusion of nematic interactions, or by breaking the physical TR symmetry with applied magnetic field. The first route to remove the 𝒵2\mathcal{Z}_{2} phase is the breaking of time–reversal symmetry. In applied magnetic field the triplon bands split and the system becomes a Chern insulator, exhibiting finite thermal Hall response and chiral modes on open edges [Section VIII], in a straightforward analogy with the Kane and Mele model Kane and Mele (2005a, b). The second, less trivial route, is the inclusion of bond–nematic interactions, permitted by the symmetry of the lattice, which breaks a pseudo TR symmetry introduced in Section VII.

In contrast to the electronic model of Kane and Mele, where the mixing of states with Sz=±12S^{z}=\pm\frac{1}{2} by (weak) Rashba interactions is compatible with a spin–Hall state, these terms have a singular effect, immediately changing the topology of the triplon bands. Such bond–nematic interactions, typically referred to as symmetric exchange anisotropies, are naturally present in other spin models too, proposed to exhibit 𝒵2\mathcal{Z}_{2} bands realized by magnetic excitations. Although one can introduce a pseudo TR symmetry that squares to 1-1, in analogy to the physical TR symmetry present in the Kane and Mele model, this symmetry does not prevail in a general model.

The nematic terms and in-plane DM interactions that mix the spins and also break the pseudo TR symmetry (Θ\Theta) can also be present in other models proposed as bosonic analogues of 𝒵2\mathcal{Z}_{2} bands, including bilayer ordered magnets Kondo et al. (2019a), and paramagnets Joshi and Schnyder (2019). Our results call for a detailed investigations of the consequences of the various Θ\Theta symmetry-breaking terms in bosonic systems in general.

Acknowledgements.
The authors are pleased to acknowledge discussions with Christian D. Batista and Yutaka Akagi. The authors also gratefully acknowledge the hospitality of the KITP program “Topological Quantum Matter: Concepts and Realizations”, where a part of the work was carried out. This work was supported by the Theory of Quantum Matter Unit of the Okinawa Institute of Science and Technology Graduate University (OIST), and the Hungarian NKFIH Grant No. K124176. This research was supported in part by the National Science Foundation under Grant No. NSF PHY-1748958.

Appendix A Time-reversal symmetry

The TR operator for a dimer has the from

T=eiπ(S1y+S2y)𝒦=(1000000100100100)𝒦,\displaystyle T=e^{i\pi(S^{y}_{1}+S^{y}_{2})}\cdot\mathcal{K}=\left(\begin{array}[]{cccc}1&0&0&0\\ 0&0&0&1\\ 0&0&-1&0\\ 0&1&0&0\end{array}\right)\cdot\mathcal{K}\;, (134)

where the basis is (|s,|t1,|t0,|t1)(\left|s\right>,\left|t_{-1}\right>,\left|t_{0}\right>,\left|t_{1}\right>) and 𝒦\mathcal{K} denotes a complex conjugation. Thus the TR operator for the spin-1 formed by the triplets is

T=(001010100)𝒦.\displaystyle T=\left(\begin{array}[]{ccc}0&0&1\\ 0&-1&0\\ 1&0&0\end{array}\right)\cdot\mathcal{K}\;. (138)

The state |t0\left|t_{0}\right> changes sign, while the |t1\left|t_{1}\right> and |t1\left|t_{-1}\right> transform into each other. This can also be shown writing the triplets in their usual form |t1=|\left|t_{1}\right>=\left|\uparrow\uparrow\right>, |t1=|\left|t_{-1}\right>=\left|\downarrow\downarrow\right>, and |t0=12(|+|)\left|t_{0}\right>=\frac{1}{\sqrt{2}}(\left|\uparrow\downarrow\right>+\left|\downarrow\uparrow\right>), and using that the TR acts on the spin-half as T:||T:\left|\uparrow\right>\to\left|\downarrow\right>, and T:||T:\left|\downarrow\right>\to-\left|\uparrow\right>.

Restricting ourselves to the up and down triplet states, and considering them as the components of a pseudo spin-half, the Pauli matrices formed by them will transform differently under TR than those of a real spin-half (where all of them break TR). The TR acts on the pseudo-up and down spins as T:|1|1T:\left|1\right>\to\left|-1\right>, and T:|1|1T:\left|-1\right>\to\left|1\right>. Therefore, among the Pauli matrices

σx\displaystyle\sigma^{x} =\displaystyle= |11|+|11|\displaystyle\left|1\right>\left<-1\right|+\left|-1\right>\left<1\right| (139a)
σy\displaystyle\sigma^{y} =\displaystyle= i|11|+i|11|\displaystyle-i\left|1\right>\left<-1\right|+i\left|-1\right>\left<1\right| (139b)
σz\displaystyle\sigma^{z} =\displaystyle= |11||11|,\displaystyle\left|1\right>\left<1\right|-\left|-1\right>\left<-1\right|\;, (139c)

the σx\sigma^{x} and σy\sigma^{y} are invariant under TR, and σz\sigma^{z} remains the only TR breaking operator. This is an important difference between the pseudo-spin-half formed by the triplets |1\left|1\right> and |1\left|-1\right>, and the real spin-half of an electron, for example.

We now write the TT operator in the basis of Eqs. 31: First, we represent the action of TT on the |1\left|1\right> and |1\left|-1\right> triplets with the σx=2sx\sigma_{x}=2s_{x} operator, then we account for the particle-hole space of the Bogoliubov–de Gennes equation by a 2-by-2 identity matrix I2I_{2}, and the three sublattice flavors by a 3-by-3 identity matrix I3I_{3}, since the TR operator leaves the sublattices intact. The TR operator becomes T=(I32sxI3)𝒦T=(I_{3}\otimes 2s^{x}\otimes I_{3})\cdot\mathcal{K} where 𝒦\mathcal{K} is the complex conjugation. In the sublattice subspace represented by the Gell-Mann matrices, the real λ\lambda’s are TR invariant, while the imaginary λ2\lambda_{2}, λ5\lambda_{5}, and λ7\lambda_{7} break time reversal symmetry.

In the main text we discuss the U(1) symmetry operator written in the same basis as TT, as well as the TR×\timesU(1) symmetry corresponding to a pseudo TR operator Θ\Theta. We show that to see the commutation relation of the the various terms in the Bogoliubov–de Gennes Hamiltonian it is sufficient to consider the particle (hole) subspace. In Table 2, we collect the transformation of the different terms entering the triplet Hamiltonian, M𝐤M_{{\bf k}}.

Appendix B 𝒵2\mathcal{Z}_{2} invariant from the parity eigenvalues

In an inversion symmetric system, the 𝒵2\mathcal{Z}_{2} can be easily calculated using the parity eigenvalues at the four time-reversal invariant momenta (TRIM) Fu and Kane (2007).

i=14ξ(Γi)=(1)ν.\prod_{i=1}^{4}\xi(\Gamma_{i})=(-1)^{\nu}\;. (140)

The TRIM, Γi\Gamma_{i} correspond to Γ\Gamma, 𝖬𝟣\mathsf{M_{1}}, 𝖬𝟤\mathsf{M_{2}}, and 𝖬𝟥\mathsf{M_{3}} shown in Fig. 1(a) of the main text. When the product is 1-1, the exponent ν\nu is odd and the system is topologically nontrivial. For even ν\nu values the system is trivial.

We define the parity operator (𝒫\mathcal{P}) as the inversion through the center of the dimer AA. Consequently, the effect of 𝒫\mathcal{P} on dimer AA is the exchange of its sites 11 and 22, while dimer BB will also be shifted by 𝜹y{\bm{\delta}}_{y} and dimer CC by 𝜹x{\bm{\delta}}_{x} from their original positions beside exchanging their sites (see Fig. 1 in the main text).

Changing the site indices, 11 and 22 does not affect the triplets, which are even under permutation, nor has the inversion any effect on the spin degrees of freedom. Thus, 𝒫\mathcal{P} will not mix different mm bands and we can treat each sector separately again. In momentum space 𝒫m\mathcal{P}_{m} becomes diagonal,

𝒫m=(𝐭m(𝐤)𝐭m(𝐤))(P𝐤𝟎𝟎P𝐤)(𝐭m(𝐤)𝐭m(𝐤))\displaystyle\mathcal{P}_{m}\!\!=\!\!\left(\!\!\begin{array}[]{c}\mathbf{t}_{m}^{\dagger}(\mathbf{k})\\ \mathbf{t}_{-\!m\!}^{\phantom{\dagger}}({-\!\bf k})\end{array}\!\right)\!\!\left(\!\!\begin{array}[]{cc}P_{\mathbf{k}}&\bm{0}\\ \bm{0}&P_{\mathbf{k}}\end{array}\!\!\right)\!\left(\!\!\begin{array}[]{c}\mathbf{t}_{m}^{\phantom{\dagger}}(\mathbf{k})\\ \mathbf{t}_{-m}^{\dagger}({-\!\bf k})\end{array}\!\right)\!\!\;\; (147)

where

P𝐤=diag(1,ei𝜹y𝐤,ei𝜹x𝐤)P_{\mathbf{k}}={\rm diag}\left(1,e^{i{\bm{\delta}}_{y}\cdot\mathbf{k}},e^{i{\bm{\delta}}_{x}\cdot\mathbf{k}}\right) (148)

We compute the eigenvectors of m\mathcal{H}_{m} numerically at each TRIM, Γi=(Γ,𝖬𝟣,𝖬𝟤,𝖬𝟥)\Gamma_{i}=(\Gamma,\mathsf{M_{1}},\mathsf{M_{2}},\mathsf{M_{3}}), and determine their eigenvalue, ξm(Γi)\xi_{m}(\Gamma_{i}) with 𝒫m\mathcal{P}_{m} defined in Eq. (147). When inversion symmetry is present, the 𝒫m\mathcal{P}_{m} commutes with m\mathcal{H}_{m} at the TRIM. The parity eigenvalues of the bands are collected in Table 3. The 𝒵2\mathcal{Z}_{2} indices of the bottom and top bands are 1, while it is 0 for the middle band. Let us note that if we chose dimer BB as the center of inversion, the parity eigenvalues for the point 𝖬𝟤\mathsf{M_{2}} would become 1-1, 11, and 1-1, while ξm(𝖬𝟣)\xi_{m}(\mathsf{M_{1}}) and ξm(𝖬𝟥)\xi_{m}(\mathsf{M_{3}}) would be 11, 1-1, and 11. Similarly, setting dimer CC as the inversion center results in a further cyclic permutation of the rows of Table 3. This corresponds to the three-fold symmetry of the ABCABC-triangles.

Table 3: Parity eigenvalues of the bottom (black), middle (purple) and, top (red) bands for each spin degree of freedom, mm at the time reversal invariant momenta, Γi\Gamma_{i}.
band bottom middle top
parity eigenvalue ξm(Γ)\xi_{m}(\Gamma) 1 1 1
ξm(𝖬𝟣)\xi_{m}(\mathsf{M_{1}}) -1 1 -1
ξm(𝖬𝟤)\xi_{m}(\mathsf{M_{2}}) 1 -1 1
ξm(𝖬𝟥)\xi_{m}(\mathsf{M_{3}}) 1 -1 1
𝒵2\mathcal{Z}_{2} index ν\nu 1 0 1

Appendix C Estimates of the gap closing transition induced by the nematic terms

In the K=0K^{\prime}_{\|}=0 case, corresponding to Fig. 14 (a)–(c), the critical value of the intra-dimer nematic interaction at which the band-touching occurs is Kcrit=4(D~23D~J~)3D~J~K^{\rm crit}_{\|}=\frac{4\left(\tilde{D}^{2}-\sqrt{3}\tilde{D}\tilde{J^{\prime}}\right)}{\sqrt{3}\tilde{D}-\tilde{J^{\prime}}}. The parameters D~\tilde{D} and J~\tilde{J^{\prime}} depend on where the gap closes in the Brillouin zone: The gap between two low-lying bands close at the Γ\Gamma point for D~(D+D′′)(JJ+1)\tilde{D}\approx(D^{\prime}+D^{\prime\prime})(\frac{J^{\prime}}{J}+1) and J~JJ22J+(D+D′′)22J\tilde{J^{\prime}}\approx J^{\prime}-\frac{J^{\prime 2}}{2J}+\frac{(D^{\prime}+D^{\prime\prime})^{2}}{2J}. While, the gap between the top bands collapses at the K(K)K(K^{\prime}) point for D~(D2D′′)(J4J12)\tilde{D}\approx(D^{\prime}-2D^{\prime\prime})(\frac{J^{\prime}}{4J}-\frac{1}{2}) and J~J2J28J+(D2D′′)28J\tilde{J^{\prime}}\approx-\frac{J^{\prime}}{2}-\frac{J^{\prime 2}}{8J}+\frac{(D^{\prime}-2D^{\prime\prime})^{2}}{8J}. The values of D~\tilde{D} and J~\tilde{J^{\prime}} were determined perturbatively, assuming that JJ is the leading term. Inserting the parameter values used in Fig. 14 (a)–(c), we get Kcrit(Γ)±0.0889K^{\rm crit}_{\|}(\Gamma)\approx\pm 0.0889 and Kcrit(K)±0.0551K^{\rm crit}_{\|}(K)\approx\pm 0.0551. Both of these values are larger than the K=0.02K=0.02, which we chose to show the hybridization of the edge modes.

ForK=0K_{\|}=0, the critical value of the inter-dimer nematic interaction is Kcrit=J~±3D~±D~2+J~2K^{\prime\rm crit}_{\|}=\mp\tilde{J^{\prime}}\pm\sqrt{3}\tilde{D}\pm\sqrt{\tilde{D}^{2}+\tilde{J^{\prime}}{}^{2}}. The lower gap closes at the Γ\Gamma point when D~(D+D′′)(JJ+1)\tilde{D}\approx(D^{\prime}+D^{\prime\prime})(\frac{J^{\prime}}{J}+1) and J~JJ22J+(D+D′′)22J\tilde{J^{\prime}}\approx J^{\prime}-\frac{J^{\prime 2}}{2J}+\frac{(D^{\prime}+D^{\prime\prime})^{2}}{2J} and the gap between the top bands closes at the K(K)K(K^{\prime}) point when D~(D2D′′)(1J2J)\tilde{D}\approx(D^{\prime}-2D^{\prime\prime})(1-\frac{J^{\prime}}{2J}) and J~J+J24J(D2D′′)24J\tilde{J^{\prime}}\approx J^{\prime}+\frac{J^{\prime 2}}{4J}-\frac{(D^{\prime}-2D^{\prime\prime})^{2}}{4J}. For the parameter values of Fig. 14 (d)–(e), the Kcrit(Γ)±0.0197K^{\prime\rm crit}_{\|}(\Gamma)\approx\pm 0.0197 and Kcrit(K)±0.0312K^{\prime\rm crit}_{\|}(K)\approx\pm 0.0312, both of which are exceeding K=0.01K^{\prime}=0.01 that we chose.

References