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

\UseRawInputEncoding

Type-II Nodal Line Fermions in New 2\mathbb{Z}_{2} Topological Semimetals A𝐕𝟔𝐒𝐛𝟔\mathbf{V_{6}Sb_{6}} (A=K, Rb, and Cs) with Kagome Bilayer

Y. Yang Hefei National Laboratory for Physical Sciences at Microscale and Department of Physics, and Key Laboratory of Strongly-Couple Quantum Matter Physics, Chinese Academy of Sciences University of Science and Technology of China, Hefei, Anhui 230026, China    R. Wang [email protected] Institute for Structure and Function &\& Department of physics &\& Center for Quantum Materials and Devices, Chongqing University, Chongqing 400044, China Chongqing Key Laboratory for Strongly Coupled Physics, Chongqing 400044, China    M.-Z. Shi Hefei National Laboratory for Physical Sciences at Microscale and Department of Physics, and Key Laboratory of Strongly-Couple Quantum Matter Physics, Chinese Academy of Sciences University of Science and Technology of China, Hefei, Anhui 230026, China    Z. Wang Hefei National Laboratory for Physical Sciences at Microscale and Department of Physics, and Key Laboratory of Strongly-Couple Quantum Matter Physics, Chinese Academy of Sciences University of Science and Technology of China, Hefei, Anhui 230026, China CAS Center for Excellence in Superconducting Electronics (CENSE), Shanghai 200050, China    Z. Xiang Hefei National Laboratory for Physical Sciences at Microscale and Department of Physics, and Key Laboratory of Strongly-Couple Quantum Matter Physics, Chinese Academy of Sciences University of Science and Technology of China, Hefei, Anhui 230026, China    X.-H. Chen Hefei National Laboratory for Physical Sciences at Microscale and Department of Physics, and Key Laboratory of Strongly-Couple Quantum Matter Physics, Chinese Academy of Sciences University of Science and Technology of China, Hefei, Anhui 230026, China CAS Center for Excellence in Superconducting Electronics (CENSE), Shanghai 200050, China CAS Center for Excellence in Quantum Information and Quantum Physics, Hefei, Anhui 230026, China Collaborative Innovation Center of Advanced Microstructures, Nanjing University, Nanjing 210093, China
Abstract

The recently discovered layered kagome metals AV3Sb5\mathrm{V_{3}Sb_{5}} (A=K, Rb, and Cs) attract intensive interest due to their intertwining with superconductivity, charge-density-wave state, and nontrivial band topology. In this work, we show by first-principles calculations and symmetry arguments that unconventional type-II Dirac nodal line fermions close to the fermi level are present in another latest class of experimentally synthesized kagome compounds AV6Sb6\mathrm{V_{6}Sb_{6}} (A=K, Rb, and Cs). These compounds possess a unique kagome (V3Sb)2(\mathrm{V_{3}Sb})_{2} bilayer that dominates their electronic and topological properties, instead of the kagome V3Sb monolayer in AV3Sb5\mathrm{V_{3}Sb_{5}}. Crystal symmetry guarantees that the type-II Dirac nodal lines with quantized Berry phase lie in reflection-invariant planes of the Brillouin zone. We further reveal that the type-II Dirac nodal lines remain nearly intact in the presence of spin-orbital coupling and can be categorized as a 2\mathbb{Z}_{2} classification. The findings establish AV6Sb6\mathrm{V_{6}Sb_{6}} as a class of new fascinating prototypes, which will extend the knowledge of interplay between unconventionally topological fermions and exotic quantum ordered states in kagome systems.

pacs:
73.20.At, 71.55.Ak, 74.43.-f

Interplay of symmetry, quantum ordered states, and nontrivial topology attracts intensive interest in condensed-matter physics. Over the past decade, topological insulators and superconductors have been classified as a function of symmetry class Hasan and Kane (2010), and various symmetry-protected nontrivial quasiparticles have been predicted or verified in gapless topological semimetals or metals Burkov et al. (2011); Wan et al. (2011); Liu et al. (2014); Weng et al. (2015); Armitage et al. (2018). Meanwhile, strict symmetries of free space are not necessarily preserved in crystalline solids, which means that unconventional topological quasiparticles without counterpart in standard model can emerge Bradlyn et al. (2016); Zhu et al. (2016); Soluyanov et al. (2015); Autès et al. (2016); Wang et al. (2016); Zhang et al. (2018). As a representative example, two types of low-energy quasiparticles based on their band profiles have been identified in crystalline materials. The first type is the conventional type-I quasiparticles associated with a closed pointlike Fermi surface. The second type is the unconventional type-II quasiparticles, whose nodal points appear at the boundary of electron and hole pockets, leaving open trajectories Soluyanov et al. (2015). The low-energy excitations around a type-II nodal point do not satisfy the Lorentz symmetry and can lead to exotic quantum transport phenomena such as anisotropic chiral anomaly Soluyanov et al. (2015); Autès et al. (2016); Wang et al. (2016). Considering the diversity of crystal symmetries in crystalline solids, the realization of intertwining between unconventional topological quasiparticles and different quantum ordered states is intriguing.

The kagome lattice, possessing a two-dimensional network of tiled triangles and hexagons, is known as one of the most versatile models for studying various exotic quantum phenomena Zhou et al. (2017). Due to its unique lattice geometry, the kagome lattice hosts linear Dirac bands and flat bands, naturally generating cooperation of many-body correlated effects and nontrivial band topology, such as quantum spin liquid Zhou et al. (2017); Yan et al. (2011), charge density waves (CDW) Guo and Franz (2009), superconductivity Ko et al. (2009), as well as beyond Ramirez (1994); Ko et al. (2009); Isakov et al. (2006); Wang et al. (2013); Liu et al. (2019); Belopolski et al. (2019); Morali et al. (2019); Yin et al. (2020). Recently, a new family of nonmagnetic kagome metals AV3Sb5\mathrm{V_{3}Sb_{5}} (A=K, Rb, and Cs) has been discovered to host fascinating integration of multiple quantum phenomena Ortiz et al. (2019); Yang et al. (2020); Yu et al. (2021); Li et al. (2021); Liang et al. (2021); Tan et al. (2021); Ortiz et al. (2020, 2021); Jiang et al. (2021); Chen et al. (2021); Zhao et al. (2021). These materials possess a layered structure that crystallizes in space group P6/mmm (No. 191), with a prototypical V3Sb plane composed of the V kagome sublattice and Sb hexagonal sublattice. Due to the presence of vanadium kagome net, immediate studies of collective ordered states including superconductivity and CDW have been intensively carried out in these AV3Sb5\mathrm{V_{3}Sb_{5}} kagome metals Li et al. (2021); Liang et al. (2021); Tan et al. (2021). Moreover, the theoretical calculations and angle-resolved photoemission spectroscopy measurement identified that electronic band structures of AV3Sb5\mathrm{V_{3}Sb_{5}} feature nontrivial 2\mathbb{Z}_{2} band topology Ortiz et al. (2020, 2021). While these advancements of intimate correlations between superconductivity, charge order, and nontrivial band topology in AV3Sb5\mathrm{V_{3}Sb_{5}} have been very encouraging Liang et al. (2021); Jiang et al. (2021), unconventional topological quasiparticles and their interplay with various quantum ordered states still remains largely unexplored. In addition, multiple crossing points and relative complicated Fermi surface in the electronic band structures of AV3Sb5\mathrm{V_{3}Sb_{5}} make their understanding by theoretical and experimental difficult. Therefore, it is highly desirable to explore kagome materials with unconventional and ideal topological band structures.

Very recently, we have discovered another class of layered kagome compounds AV6Sb6\mathrm{V_{6}Sb_{6}} with space group R3¯mR\bar{3}m (No. 166) Shi et al. (2021), which would satisfy the above criteria. These new kagome compounds AV6Sb6\mathrm{V_{6}Sb_{6}} have the same chemical compositions but different stoichiometric ratios with AV3Sb5\mathrm{V_{3}Sb_{5}}. Compared with AV3Sb5\mathrm{V_{3}Sb_{5}}, AV6Sb6\mathrm{V_{6}Sb_{6}} can be considered as interposing an additional V3Sb plane; that is, a bilayer (V3Sb)2(\mathrm{V_{3}Sb})_{2} stacking network is present in AV6Sb6\mathrm{AV_{6}Sb_{6}}. Meanwhile, it is worth noting that the AV6Sb6\mathrm{V_{6}Sb_{6}} compounds as well as AV3Sb5\mathrm{V_{3}Sb_{5}} can be viewed as a unified A-V-Sb family with a generic chemical formula Am-1Sb2m(V3Sb)n (i.e., AV3Sb5\mathrm{V_{3}Sb_{5}} with m=2m=2, n=1n=1 , and AV6Sb6\mathrm{V_{6}Sb_{6}} with m=2m=2, n=2n=2), exhibiting common features of this family. Importantly, the experimental results indicates that the AV6Sb6\mathrm{V_{6}Sb_{6}} compounds exhibit superconductivity under pressure Shi et al. (2021). In this work, based on first-principles calculations and symmetry analysis, we identify that the AV6Sb6\mathrm{V_{6}Sb_{6}} compounds are ideal 2\mathbb{Z}_{2} topological semimetals and possess the symmetry-protected type-II Dirac nodal line fermions close to the Fermi level, implying that AV6Sb6\mathrm{V_{6}Sb_{6}} could be expected as a promising platform for investigating exotic quantum phenomena with unconventional topological quasipariticles beyond standard model.

Refer to caption
Figure 1: (a) The crystal structure of AV6Sb6\mathrm{V_{6}Sb_{6}} (A = K, Rb, and Cs) with A-Sb-V3Sb-V3Sb-Sb quintuple layers stacking along the crystallographic cc-axis. 𝐭1,2,3\mathbf{t}_{1,2,3} is the primitive lattice vectors of a rhombohedral structure with space group R3¯mR\bar{3}m (No. 166) SM . (b) The graphitelike honeycomb sublattice of the Sb3 atom and hexagonal sublattice of A-site alkali metal atom. (c) The V3Sb bilayer network composed of two V1 and V2 kagome sublattices and two Sb1 and Sb2 hexagonal sublattices. (d) The rhombohedral BZ and the corresponding (111) surface BZ. High-symmetry points (solid red points) and a high-symmetry middle plane are marked.

To elucidate topological features of the novel kagome compounds AV6Sb6\mathrm{V_{6}Sb_{6}}, we performed first-principles calculations based on the density functional theory Kohn and Sham (1965) as implemented in the Vienna ab initio simulation package Kresse and Furthmüller (1996). The exchange-correlation functional was described by generalized gradient approximation with Perdew-Burke-Ernzerhof formalism Perdew et al. (1996). The core-valence interactions were treated by projector augmented-wave potentials Kresse and Joubert (1999), and a plane-wave basis set with a kinetic-energy cutoff of 450 eV was used. The Brillouin zone (BZ) was sampled by a 12×12×1212\times 12\times 12 Monkhorst-Pack grid Monkhorst and Pack (1976). The crystal structures were fully relaxed by minimizing forces of each atom smaller than 1.0×1031.0\times 10^{3} eV/Å, and van der Waals interactions along the cc-layer stacking direction was considered by the Crimme (DFT-D3) method Grimme et al. (2011). The topological classification was confirmed by the 2\mathbb{Z}_{2} invariants Fu et al. (2007), which are calculated from the parity eigenvalues at the time-reversal invariant momenta (TRIM) points using IRVSP package Gao et al. (2021). We construct a Wannier tight-binding (TB) Hamiltonian based on maximally localized Wannier functions to reveal the topological features as implemented in the WANNIER90 package A. A. Mostofi et al. (2008).

As illustrated in Fig. 1(a), the presence of kagome V3Sb bilayer net makes the AV6Sb6\mathrm{V_{6}Sb_{6}} structure hosting A-Sb-V3Sb-V3Sb-Sb quintuple layers, which stack along the crystallographic cc-axis. As a result, the AV6Sb6\mathrm{V_{6}Sb_{6}} compounds crystallize in a rhombohedral structure with space group R3¯mR\bar{3}m (No. 166). As shown in Fig. 1(c), the bilayer V3Sb stacking network contains two V1 and V2 kagome sublattices, which share two equivalent Sb1 and Sb2 hexagonal sublattices. The Sb3 atom respectively forms upper and lower graphitelike honeycomb lattices that encapsulate the V3Sb kagome bilayer, and the hexagonal sublattice of A-site alkali metal atom spontaneously fills the space between the Sb3 sublattice [see Fig. 1(b)]. As a result, the atom positions of AV6Sb6\mathrm{V_{6}Sb_{6}} structure are Wyckoff 1a1a of A atom, 6h6h of V1 and V2 atoms, and 3c3c of Sb1, Sb2, and Sb3 atoms. The optimized lattice constants are in excellent agreement with the experimental values (see details in Supplemental Material (SM) SM ). The rhombohedral BZ and the corresponding (111) surface BZ of AV6Sb6\mathrm{V_{6}Sb_{6}} are shown in Fig. 1(d), with high-symmetry points indicated. Since these three novel kagome materials share the similar features, we only present the results of CsV6Sb6\mathrm{CsV_{6}Sb_{6}} in the main text. The results of KV6Sb6\mathrm{KV_{6}Sb_{6}} and RbV6Sb6\mathrm{RbV_{6}Sb_{6}} are included in the SM SM .

The calculated electronic band structures of CsV6Sb6\mathrm{CsV_{6}Sb_{6}} along high-symmetry lines are shown in Fig. 2(a). The bands show that a set of linear band crossings are present along some specific high-symmetry lines. These high-symmetry lines are coplanar with the kzk_{z} axis, forming a high-symmetry middle plane of the BZ [e.g., the kyk_{y}-kzk_{z} plane with kx=0k_{x}=0 in Fig. 1(d)]. Due to the existence of inversion (\mathcal{I}) and time-reversal (𝒯\mathcal{T}) symmetries, the crossing points in CsV6Sb6\mathrm{CsV_{6}Sb_{6}} belong to Dirac points. All the Dirac points are slightly above the Fermi level and their maximum energy at \sim0.09 eV occurs along FF-ZZ, forming hole pockets. A deeper inspection reveals that the Dirac points can actually exist along an arbitrary line that connects the Γ\Gamma/ZZ point and the boundary point of the BZ in the kyk_{y}-kzk_{z} plane [see Fig. 2(b)]. We also plot an energy difference map between the valence band and conduction band in Fig. 2(c). A zero energy difference denoted as the white feature indicates that there are two nodal lines threading the BZ in the kyk_{y}-kzk_{z} plane. By checking the little group of the kyk_{y}-kzk_{z} plane, we find that all Dirac points are with respect to the mirror-reflection symmetry CsC_{s}. Two crossing bands belong to opposite mirror eigenvalues Γ1\Gamma_{1} (+1) and Γ2\Gamma_{2} (-1). By the way, we should point out that the symmetry forbids the existence of Dirac points along Γ\Gamma-KK though the conduction and valence bands seems to nearly touch [see Fig. 2(a)]. A three-dimensional plot of two crossing bands in the kyk_{y}-kzk_{z} plane with kx=0k_{x}=0 is present in Fig. 2(d). This figure depicts that the two inverted bands form two continuous Dirac nodal lines close to the Fermi level with tiny energy dispersion. Clearly, it can be found that band profiles near all Dirac points along the nodal lines are tilted, indicating that low-energy excitations in CsV6Sb6\mathrm{CsV_{6}Sb_{6}} may belong to type-II Dirac nodal line fermions.

Refer to caption
Figure 2: Electronic band structures of CsV6Sb6\mathrm{CsV_{6}Sb_{6}} along high-symmetry lines in the absence of SOC. Three bands close to the Fermi level are highlighted in colors. (b) The enlarged view of band crossings along an arbitrary line in the kyk_{y}-kzk_{z} plane, with irreducible representations of CsC_{s} indicated. (c) Energy difference map between the conduction band and valence band in the kyk_{y}-kzk_{z} plane. (d) Two inverted bands that form two nodal lines, indicating that band profiles along the nodal line are tilted and thus belong to type-II.

We further examine isoenergy contours of Fermi surface to verify the presence of type-II Dirac nodal lines in CsV6Sb6\mathrm{CsV_{6}Sb_{6}}. The cross section of bulk isoenergy surface in the kxk_{x}-kyk_{y} plane with kz=0k_{z}=0 is shown in Figs. 3(a) and 3(b). We can see that there are six symmetry-distributed Dirac points in the first BZ, which are generated from nodal lines across the kz=0k_{z}=0 plane, are present in the first BZ. When tuning the chemical potential of the isoenergy surface, hole and electron pockets only touch at the energy of Dirac points EDE_{D} and exhibit open isoenergy contours [see Fig. 3(b)]. Therefore, CsV6Sb6\mathrm{CsV_{6}Sb_{6}} indeed hosts type-II Dirac nodal lines as its preservation of \mathcal{I} and 𝒯\mathcal{T} symmetries. Considering the three-fold rotational symmetry C3C_{3} in space group R3¯mR\bar{3}m, there are three equivalent mirror planes and six type-II Dirac nodal lines in the BZ.

Refer to caption
Figure 3: (a) The cross section of bulk isoenergy surface in the kxk_{x}-kyk_{y} plane with kz=0k_{z}=0. (b) The enlarged view of isoenergy contours [marked by the black box in (a)] of Δ=4\Delta=4 meV below, at, and above the Dirac point EDE_{D} are shown in the upper, middle, and lower panels, respectively. (c) The electronic band structures as well as Berry curvature along W1W_{1}-Γ\Gamma-W2W_{2} in the kyk_{y}-kzk_{z} plane. (d) A variation of Berry phase along W1W_{1}-Γ\Gamma-W2W_{2} in the kyk_{y}-kzk_{z} plane. The points W1W_{1} and W2W_{2} are positioned at (0.0, -0.7, 0.0) and (0.0, 0.7, 0.0) Å-1, respectively.

Next, we carry out a symmetry argument to understand the existence of six symmetry-protected type-II Dirac nodal lines in CsV6Sb6\mathrm{CsV_{6}Sb_{6}}. As shown in Fig. 1(b), the nonmagnetic space group R3¯mR\bar{3}m contains three-fold rotational symmetry 𝒞3z\mathcal{C}_{3z} along the zz (or cc) direction, three twofold rotational symmetries (i.e., one is the 𝒞2x\mathcal{C}_{2x} along the xx (or aa) direction, and the other two are with respect to 𝒞3z\mathcal{C}_{3z}), \mathcal{I}-symmetry, and 𝒯\mathcal{T}-symmetry. In general, two inverted bands can be described by a two-band effective 𝐤𝐩\mathbf{k}\cdot\mathbf{p} Hamiltonian as

(𝐤)=f0(𝐤)σ0+fx(𝐤)σx+fy(𝐤)σy+fz(𝐤)σz,\mathcal{H}(\mathbf{k})=f_{0}(\mathbf{k})\sigma_{0}+f_{x}(\mathbf{k})\sigma_{x}+f_{y}(\mathbf{k})\sigma_{y}+f_{z}(\mathbf{k})\sigma_{z}, (1)

where f0(𝐤)f_{0}(\mathbf{k}) is the kinetic term, σ0\sigma_{0} is the identity matrix, σi\sigma_{i} (i=x,y,zi=x,y,z) are three Pauli matrices, fi(𝐤)f_{i}(\mathbf{k}) are real functions, and wavevector 𝐤\mathbf{k} is relative to the inverted point. The kinetic term f0(𝐤)f_{0}(\mathbf{k}) only shifts Dirac points and thus can be ignored in Eq. (1) in the following. The two inverted bands have the opposite eigenvalues of little group CsC_{s}, indicating that the \mathcal{I}-symmetry can be represented as =σz\mathcal{I}=\sigma_{z}, which constrains the Hamiltonian as

(𝐤)1=(𝐤).\mathcal{I}\mathcal{H}(\mathbf{k})\mathcal{I}^{-1}=\mathcal{H}(-\mathbf{k}). (2)

In the absence of SOC, the 𝒯\mathcal{T} symmetry can be represented by 𝒯=K\mathcal{T}=K, where KK is the complex conjugate operator. The 𝒯\mathcal{T}-symmetry indicates

𝒯(𝐤)𝒯1=(𝐤).\mathcal{T}\mathcal{H}(\mathbf{k})\mathcal{T}^{-1}=\mathcal{H}(-\mathbf{k}). (3)

Based on Eqs. (2) and (3), the existence of \mathcal{I} and 𝒯\mathcal{T} symmetries requires

fx(𝐪)0,fy(𝐤)=fy(𝐤),fz(𝐤)=fz(𝐤).f_{x}(\mathbf{q})\equiv 0,\ \ f_{y}(\mathbf{k})=-f_{y}(\mathbf{-k}),\ \ f_{z}(\mathbf{k})=f_{z}(\mathbf{-k}). (4)

The band crossings require fy(𝐤)=0f_{y}(\mathbf{k})=0 and fz(𝐤)=0f_{z}(\mathbf{k})=0, which indicate that there is a codimension one, and thus allowing nodal-lines in momentum space. The mirror-reflection symmetry x\mathcal{M}_{x} can be considered as the product of \mathcal{I} and 𝒞2x\mathcal{C}_{2x} (i.e., x=𝒞2x\mathcal{M}_{x}=\mathcal{I}\mathcal{C}_{2x}). In the kyk_{y}-kzk_{z} plane with kx=0k_{x}=0, the commutation relation [x,]=0[\mathcal{M}_{x},\mathcal{H}]=0 indicates fy(0,ky,kz)0f_{y}(0,k_{y},k_{z})\equiv 0. Then, we can expand fz(0,ky,kz)f_{z}(0,k_{y},k_{z}) and remain the lowest orders as

fz(0,ky,kz)=Az+Bzky2+Czkz2+Dzkykz,f_{z}(0,k_{y},k_{z})=A_{z}+B_{z}k_{y}^{2}+C_{z}k_{z}^{2}+D_{z}k_{y}k_{z}, (5)

in which the parameters AzA_{z}, BzB_{z}, CzC_{z}, and DzD_{z} can be fitted from first-principles calculations. The condition of open nodal lines in the kyk_{y}-kzk_{z} plane requires that Az<0A_{z}<0, 4BzCz<Dz24B_{z}C_{z}<D_{z}^{2} and Bz>0B_{z}>0 or 4BzCz>Dz24B_{z}C_{z}>D_{z}^{2} and Bz<0B_{z}<0. Similarly, the other two mirror planes can also possess two same type-II nodal lines due to 𝒞3z\mathcal{C}_{3z}. It should be mentioned that the stability of type-II Dirac nodal lines in CsV6Sb6\mathrm{CsV_{6}Sb_{6}} is topologically protected by the coexistence of \mathcal{I} and 𝒯\mathcal{T} symmetries, and the additional rotational and mirror symmetries just force nodal lines to locate in the mirror planes of the BZ.

The presence of type-II Dirac nodal lines corresponds to a quantized Berry phase or winding number with π\pi mod 2π2\pi, which is defined as Fang et al. (2015)

γ=𝒞𝒜(𝐤)𝑑𝐤,\gamma=\oint_{\mathcal{C}}\mathcal{A}(\mathbf{k})\cdot d\mathbf{k}, (6)

where 𝒜(𝐤)\mathcal{A}(\mathbf{k}) is the Berry connection, and 𝒞\mathcal{C} is a closed loop in three-dimensional momentum space. The corresponding Berry curvature is 𝛀(𝐤)=×𝒜(𝐤)\mathbf{\Omega}(\mathbf{k})=\nabla\times\mathcal{A}(\mathbf{k}). As shown in Fig. 3(c), we calculate the Berry curvature 𝛀(𝐤)\mathbf{\Omega}(\mathbf{k}) along an arbitrary line W1W_{1}-Γ\Gamma-W2W_{2} in the kyk_{y}-kzk_{z} plane. This figure shows that Ωx(𝐤)\mathrm{\Omega}_{x}(\mathbf{k}) has opposite peaks near the momentum positions of crossing points along the opposite directions, which is in accordance with the coexistence of \mathcal{I} and 𝒯\mathcal{T} symmetries. To further reveal the nontrivial type-II Dirac nodal lines in CsV6Sb6\mathrm{CsV_{6}Sb_{6}}, we calculate a variation of Berry phase γ\gamma in the kyk_{y}-kzk_{z} plane, which corresponds to the one-dimensional system along the W1W_{1}-Γ\Gamma-W2W_{2}. As shown in Fig. 3(d), there is a jump of π\pi (π-\pi) when the closed path 𝒞\mathcal{C} encircles the nodal lines from W1W_{1} (W2W_{2}) to Γ\Gamma, indicating its nontrivial topology.

Refer to caption
Figure 4: Surface states of CsV6Sb6\mathrm{CsV_{6}Sb_{6}}. (a) The calculated LDOS and (b) Fermi surface projected on the semi-infinite (111) surface of CsV6Sb6\mathrm{CsV_{6}Sb_{6}}. In panel (b), there are two Fermi arcs well separated from the bulk bands enclosed to the Γ¯\bar{\Gamma} point. The high-symmetry lines and the first BZ of (111) surface are marked by yellow-dashed lines.

For the nodal lines protected by \mathcal{I} and 𝒯\mathcal{T} symmetries, the spin-orbital coupling (SOC) can always gap out the nodal lines and drive the system to other topological phases, such as topological insulators, Dirac semimetals, and even topologically trivial phases Kim et al. (2015); Jin et al. (2017). Indeed, the calculated band structures within SOC of CsV6Sb6\mathrm{CsV_{6}Sb_{6}} indicate that two crossing bands without SOC now belong to the same mirror eigenvalues Γ4\Gamma_{4} of CsC_{s} and thus the band crossings are avoided (see Fig. S2 in the SM SM ). However, the rather weak SOC effect of CsV6Sb6\mathrm{CsV_{6}Sb_{6}} only leads to a band gap width less than \sim1 meV. That is to say, its type-II Dirac nodal line fermions are nearly intact. The nodal line phases usually have the same 2\mathbb{Z}_{2} invariant as its gapped topological phases Kim et al. (2015). To verify the topological classification of CsV6Sb6\mathrm{CsV_{6}Sb_{6}}, we calculate the parity eigenvalues of occupied bands at the time-reversal invariant momenta points (see the details in the SM SM ). The results reveal that the band gaps opened by SOC in CsV6Sb6\mathrm{CsV_{6}Sb_{6}} are topologically nontrivial with a (0;111) weak 2\mathbb{Z}_{2} index. Moreover, we also find that other nontrivial bands are present near the Fermi level SM , exhibiting entangled topological features. The type-II Dirac nodal line fermions categorized as a 2\mathbb{Z}_{2} classification are expected to give rise to exotic quantum transport phenomena.

The type-II Dirac nodal lines combining with 2\mathbb{Z}_{2}-type nontrivial band topology in close proximity to the Fermi level can give rise to topological surface states. To illustrate this, we calculate the surface states of CsV6Sb6\mathrm{CsV_{6}Sb_{6}} using the iterative Green’s function method based on Wannier TB Hamiltonian as implemented in the WANNIERTOOLS package Sancho et al. (1984); Wu et al. (2018). The calculated local density of states (LDOS) and Fermi surface projected on the semi-infinite (111) surface of CsV6Sb6\mathrm{CsV_{6}Sb_{6}} are shown in Fig. 4. As depicted in Fig. 4(a), we can see that the projected type-II Dirac cones along Γ¯\bar{\Gamma}-M¯\bar{M} exhibit a band width due to the tiny energy dispersion of nodal-line. The surface states along Γ¯\bar{\Gamma}-M¯\bar{M} and Γ¯\bar{\Gamma}-K¯\bar{K} exhibit apparent anisotropy but are degenerate at the Γ¯\bar{\Gamma} point, indicating 2\mathbb{Z}_{2}-type band topology of CsV6Sb6\mathrm{CsV_{6}Sb_{6}}. As shown in Fig. 4(b), the Fermi surface projected to (111) surface shows that there are two closed nontrivial Fermi arcs crossing along Γ¯\bar{\Gamma}-M¯\bar{M}. This is in accordance with Dirac nodal lines lying in the mirror planes. Importantly, the nontrivial Fermi arcs are well separated from the bulk projected bands, facilitating their detection in experiments and further exploration.

In conclusion, we have studied the electronic and topological properties of a latest class of experimentally synthesized layered compounds AV6Sb6\mathrm{V_{6}Sb_{6}} with a kagome (V3Sb)2(\mathrm{V_{3}Sb})_{2} bilayer Shi et al. (2021). Based on first-principles calculations and topological analysis, we reveal that these novel AV6Sb6\mathrm{V_{6}Sb_{6}} compounds possess type-II Dirac nodal line fermions close to the fermi level and exhibit 2\mathbb{Z}_{2}-type nontrivial band topology. Importantly, it is worth noting that AV6Sb6\mathrm{V_{6}Sb_{6}} as well as AV3Sb5\mathrm{V_{3}Sb_{5}} can be represented by a generic chemical formula Am-1Sb2m(V3Sb)n. Considering that superconductivity in AV6Sb6\mathrm{V_{6}Sb_{6}} was realized under pressure Shi et al. (2021), our findings are expected to extend the knowledge toward understanding quantum ordered states and unconventional topological fermions beyond standard model in the A-V-Sb kagome family.

This work is supported by the National Natural Science Foundation of China (11888101 and 11974062), the National Key Research and Development Program of the Ministry of Science and Technology of China (2017YFA0303001, 2019YFA0704901, and 2016YFA0300201), the Anhui Initiative in Quantum Information Technologies (AHY160000), the Strategic Priority Research Program of the Chinese Academy of Sciences (XDB25000000), the Science Challenge Project of China (TZ2016004), and the Key Research Program of Frontier Sciences, CAS, China (QYZDYSSWSLH021). The DFT calculations in this work are supported by the Supercomputing Center of University of Science and Technology of China.

References

  • Hasan and Kane (2010) M. Z. Hasan and C. L. Kane, Rev. Mod. Phys. 82, 3045 (2010).
  • Burkov et al. (2011) A. A. Burkov, M. D. Hook, and L. Balents, Phys. Rev. B 84, 235126 (2011).
  • Wan et al. (2011) X. Wan, A. M. Turner, A. Vishwanath, and S. Y. Savrasov, Phys. Rev. B 83, 205101 (2011).
  • Liu et al. (2014) Z. K. Liu, B. Zhou, Y. Zhang, Z. J. Wang, H. M. Weng, D. Prabhakaran, S.-K. Mo, Z. X. Shen, Z. Fang, X. Dai, Z. Hussain, and Y. L. Chen, Science 343, 864 (2014).
  • Weng et al. (2015) H. Weng, C. Fang, Z. Fang, B. A. Bernevig, and X. Dai, Phys. Rev. X 5, 011029 (2015).
  • Armitage et al. (2018) N. P. Armitage, E. J. Mele, and A. Vishwanath, Rev. Mod. Phys. 90, 015001 (2018).
  • Bradlyn et al. (2016) B. Bradlyn, J. Cano, Z. Wang, M. G. Vergniory, C. Felser, R. J. Cava, and B. A. Bernevig, Science 353, 6299 (2016).
  • Zhu et al. (2016) Z. Zhu, G. W. Winkler, Q. Wu, J. Li, and A. A. Soluyanov, Phys. Rev. X 6, 031003 (2016).
  • Soluyanov et al. (2015) A. A. Soluyanov, D. Gresch, Z. Wang, Q. Wu, and M. Troyer, Nature 527, 495 (2015).
  • Autès et al. (2016) G. Autès, D. Gresch, M. Troyer, A. A. Soluyanov, and O. V. Yazyev, Phys. Rev. Lett. 117, 066402 (2016).
  • Wang et al. (2016) Z. Wang, D. Gresch, A. A. Soluyanov, W. Xie, S. Kushwaha, X. Dai, M. Troyer, R. J. Cava, and B. A. Bernevig, Phys. Rev. Lett. 117, 056805 (2016).
  • Zhang et al. (2018) T. Zhang, Z. Song, A. Alexandradinata, H. Weng, C. Fang, L. Lu, and Z. Fang, Phys. Rev. Lett. 120, 016401 (2018).
  • Zhou et al. (2017) Y. Zhou, K. Kanoda, and T.-K. Ng, Rev. Mod. Phys. 89, 025003 (2017).
  • Yan et al. (2011) S. Yan, D. A. Huse, and S. R. White, Science 332, 1173 (2011).
  • Guo and Franz (2009) H.-M. Guo and M. Franz, Phys. Rev. B 80, 113102 (2009).
  • Ko et al. (2009) W.-H. Ko, P. A. Lee, and X.-G. Wen, Phys. Rev. B 79, 214502 (2009).
  • Ramirez (1994) A. P. Ramirez, Annual Review of Materials Science 24, 453 (1994).
  • Isakov et al. (2006) S. V. Isakov, S. Wessel, R. G. Melko, K. Sengupta, and Y. B. Kim, Phys. Rev. Lett. 97, 147202 (2006).
  • Wang et al. (2013) W.-S. Wang, Z.-Z. Li, Y.-Y. Xiang, and Q.-H. Wang, Phys. Rev. B 87, 115135 (2013).
  • Liu et al. (2019) D. F. Liu, A. J. Liang, E. K. Liu, Q. N. Xu, Y. W. Li, C. Chen, D. Pei, W. J. Shi, S. K. Mo, P. Dudin, T. Kim, C. Cacho, G. Li, Y. Sun, L. X. Yang, Z. K. Liu, S. S. P. Parkin, C. Felser, and Y. L. Chen, Science 365, 1282 (2019).
  • Belopolski et al. (2019) I. Belopolski, K. Manna, D. S. Sanchez, G. Chang, B. Ernst, J. Yin, S. S. Zhang, T. Cochran, N. Shumiya, H. Zheng, B. Singh, G. Bian, D. Multer, M. Litskevich, X. Zhou, S.-M. Huang, B. Wang, T.-R. Chang, S.-Y. Xu, A. Bansil, C. Felser, H. Lin, and M. Z. Hasan, Science 365, 1278 (2019).
  • Morali et al. (2019) N. Morali, R. Batabyal, P. K. Nag, E. Liu, Q. Xu, Y. Sun, B. Yan, C. Felser, N. Avraham, and H. Beidenkopf, Science 365, 1286 (2019).
  • Yin et al. (2020) J.-X. Yin, M. Wenlong, T. A. Cochran, X. Xu, S. S. Zhang, H.-J. Tien, N. Shumiya, C. Guangming, K. Jiang, L. Biao, S. Zhida, G. Chang, I. Belopolski, M. Daniel, M. Litskevich, Z.-J. Cheng, X. P. Yang, B. Swidler, H. Zhou, H. Lin, T. Neupert, Z. Wang, N. Yao, T.-R. Chang, S. Jia, and M. Z. Hasan, Nature 583, 533 (2020).
  • Ortiz et al. (2019) B. R. Ortiz, L. C. Gomes, J. R. Morey, M. Winiarski, M. Bordelon, J. S. Mangum, I. W. H. Oswald, J. A. Rodriguez-Rivera, J. R. Neilson, S. D. Wilson, E. Ertekin, T. M. McQueen, and E. S. Toberer, Phys. Rev. Materials 3, 094407 (2019).
  • Yang et al. (2020) S.-Y. Yang, Y. Wang, B. R. Ortiz, D. Liu, J. Gayles, E. Derunova, R. Gonzalez-Hernandez, L. ?mejkal, Y. Chen, S. S. P. Parkin, S. D. Wilson, E. S. Toberer, T. McQueen, and M. N. Ali, Sci. Adv. 6, eabb6003 (2020).
  • Yu et al. (2021) F. H. Yu, M. D. H., W. Z. Zhou, S. Q. Liu, X. K. Wen, B. Lei, J. J. Ying, and X. H. Chen, Nat. Commun. 12, 3645 (2021).
  • Li et al. (2021) H. Li, T. T. Zhang, T. Yilmaz, Y. Y. Pai, C. E. Marvinney, A. Said, Q. W. Yin, C. S. Gong, Z. J. Tu, E. Vescovo, C. S. Nelson, R. G. Moore, S. Murakami, H. C. Lei, H. N. Lee, B. J. Lawrie, and H. Miao, Phys. Rev. X 11, 031050 (2021).
  • Liang et al. (2021) Z. Liang, X. Hou, F. Zhang, W. Ma, P. Wu, Z. Zhang, F. Yu, J.-J. Ying, K. Jiang, L. Shan, Z. Wang, and X.-H. Chen, Phys. Rev. X 11, 031026 (2021).
  • Tan et al. (2021) H. Tan, Y. Liu, Z. Wang, and B. Yan, Phys. Rev. Lett. 127, 046401 (2021).
  • Ortiz et al. (2020) B. R. Ortiz, S. M. L. Teicher, Y. Hu, J. L. Zuo, P. M. Sarte, E. C. Schueller, A. M. M. Abeykoon, M. J. Krogstad, S. Rosenkranz, R. Osborn, R. Seshadri, L. Balents, J. He, and S. D. Wilson, Phys. Rev. Lett. 125, 247002 (2020).
  • Ortiz et al. (2021) B. R. Ortiz, P. M. Sarte, E. M. Kenney, M. J. Graf, S. M. L. Teicher, R. Seshadri, and S. D. Wilson, Phys. Rev. Materials 5, 034801 (2021).
  • Jiang et al. (2021) Y.-X. Jiang, J.-X. Yin, M. M. Denner, N. Shumiya, B. R. Ortiz, G. Xu, Z. Guguchia, J. He, M. S. Hossain, X. Liu, J. Ruff, L. Kautzsch, S. S. Zhang, G. Chang, I. Belopolski, Q. Zhang, T. A. Cochran, D. Multer, M. Litskevich, Z.-J. Cheng, X. P. Yang, Z. Wang, R. Thomale, T. Neupert, S. D. Wilson, and M. Z. Hasan, Nat. Mater. 20, 1353 (2021).
  • Chen et al. (2021) H. Chen, H. Yang, B. Hu, Z. Zhao, J. Yuan, Y. Xing, G. Qain, Z. Huang, G. Li, Y. Ye, S. Ma, S. Ni, H. Zhang, Q. Yin, C. Gong, Z. Tu, H. Lei, H. Tan, S. Zhou, C. Shen, X. Dong, B. Yan, Z. Wang, and H.-J. Gao, Nature 10.1038/s41586-021-03983-5 (2021).
  • Zhao et al. (2021) H. Zhao, H. Li, B. R. Ortiz, S. M. L. Teicher, T. Park, M. Ye, Z. Wang, L. Balents, S. D. Wilson, and I. Zeljkovic, Nature 10.1038/s41586-021-03946-w (2021).
  • Shi et al. (2021) M. Shi, F. Yu, Y. Yang, F. Meng, B. Lei, Y. Luo, Z. Sun, J. He, R. Wang, T. Wu, Z. Wang, Z. Xiang, J. Ying, and X. Chen, A new class of bilayer kagome lattice compounds with dirac nodal lines and pressure-induced superconductivity (2021), arXiv:2110.09782 [cond-mat.mtrl-sci] .
  • (36) Supplemental Material.
  • Kohn and Sham (1965) W. Kohn and L. J. Sham, Phys. Rev. 140, A1133 (1965).
  • Kresse and Furthmüller (1996) G. Kresse and J. Furthmüller, Phys. Rev. B 54, 11169 (1996).
  • Perdew et al. (1996) J. P. Perdew, K. Burke, and M. Ernzerhof, Phys. Rev. Lett. 77, 3865 (1996).
  • Kresse and Joubert (1999) G. Kresse and D. Joubert, Phys. Rev. B 59, 1758 (1999).
  • Monkhorst and Pack (1976) H. J. Monkhorst and J. D. Pack, Phys. Rev. B 13, 5188 (1976).
  • Grimme et al. (2011) S. Grimme, S. Ehrlich, and L. Goerigk, J. Comput. Chem. 32, 1456 (2011).
  • Fu et al. (2007) L. Fu, C. L. Kane, and E. J. Mele, Phys. Rev. Lett. 98, 106803 (2007).
  • Gao et al. (2021) J. Gao, Q. Wu, C. Persson, and Z. Wang, Comput. Phys. Commun. 261, 107760 (2021).
  • A. A. Mostofi et al. (2008) Y. S. L. I. S. A. A. Mostofi, J. R. Yates, D. Vanderbilt, and N. Marzari, Comput. Phys. Commun. 178, 685 (2008).
  • Fang et al. (2015) C. Fang, Y. Chen, H.-Y. Kee, and L. Fu, Phys. Rev. B 92, 081201 (2015).
  • Kim et al. (2015) Y. Kim, B. J. Wieder, C. L. Kane, and A. M. Rappe, Phys. Rev. Lett. 115, 036806 (2015).
  • Jin et al. (2017) Y.-J. Jin, R. Wang, J.-Z. Zhao, Y.-P. Du, C.-D. Zheng, L.-Y. Gan, J.-F. Liu, H. Xu, and S. Y. Tong, Nanoscale 9, 13112 (2017).
  • Sancho et al. (1984) M. P. L. Sancho, J. M. L. Sancho, and J. Rubio, J. Phys. F: Metal Phys. 14, 1205 (1984).
  • Wu et al. (2018) Q. Wu, S. Zhang, H.-F. Song, M. Troyer, and A. A. Soluyanov, Comput. Phys. Commun. 224, 405 (2018).

Supplemental Material for

“Type-II Nodal Line Fermions in New 2\mathbb{Z}_{2} Topological Metals AV6Sb6\mathbf{V_{6}Sb_{6}} (A=K, Rb, and Cs) with Kagome Bilayer”

In this Supplemental Materials, we present lattice parameters of AV6Sb6\mathrm{V_{6}Sb_{6}} (A=K, Rb, and Cs) compounds from first-principles calculations and experiments, electronic band structures of of RbV6Sb6\mathrm{RbV_{6}Sb_{6}} and KV6Sb6\mathrm{KV_{6}Sb_{6}} in the absence of spin-orbital coupling (SOC), electronic band structures of CsV6Sb6\mathrm{CsV_{6}Sb_{6}} in the presence of SOC, and parity analysis of CsV6Sb6\mathrm{CsV_{6}Sb_{6}} at the time-reversal invariant momenta points.

The AV6Sb6\mathrm{V_{6}Sb_{6}} (A=K, Rb, and Cs) compounds crystallize in a rhombohedral structure with space group R3¯mR\bar{3}m (No. 166). The primitive lattice vectors of the rhombohedral representation can be expressed as

𝐭1=a2𝐢36a𝐣+c3𝐤,𝐭2=a2𝐢36a𝐣+c3𝐤,𝐭3=33a𝐣+c3𝐤,\begin{split}&\mathbf{t}_{1}=-\frac{a}{2}\mathbf{i}-\frac{\sqrt{3}}{6}a\mathbf{j}+\frac{c}{3}\mathbf{k},\\ &\mathbf{t}_{2}=\frac{a}{2}\mathbf{i}-\frac{\sqrt{3}}{6}a\mathbf{j}+\frac{c}{3}\mathbf{k},\\ &\mathbf{t}_{3}=\frac{\sqrt{3}}{3}a\mathbf{j}+\frac{c}{3}\mathbf{k},\end{split} (S1)

where aa and cc are lattice constants of the hexagonal representation as included in Table SI.

Table SI: Lattice parameters of AV6Sb6\mathrm{V_{6}Sb_{6}} (A=K, Rb, and Cs) compounds from first-principles calculations and experiments. aa is the lattice constant in the kagome (or hexagonal) plane and cc is the lattice constant along its stacking direction. The comparison shows a good agreement.
       CsV6Sb6\mathrm{CsV_{6}Sb_{6}}        RbV6Sb6\mathrm{RbV_{6}Sb_{6}}        KV6Sb6\mathrm{KV_{6}Sb_{6}}
       Space group        R3¯mR\bar{3}m (No. 166)
       aa (Å)        this work        5.43        5.42        5.42
       expt.        5.51        5.51        /
       cc (Å)        this work        34.94        34.23        33.69
       expt.        35.28        34.61        /
Refer to caption
Figure S1: Electronic band structures of RbV6Sb6\mathrm{RbV_{6}Sb_{6}} (left panel) and KV6Sb6\mathrm{KV_{6}Sb_{6}} (right panel) along high-symmetry lines in the absence of SOC. Three bands 115, 117, 119 are colored by pink, dark-yellow, and blue, respectively.
Refer to caption
Figure S2: (a) The top views of a rhombohedral Brillouin zone (BZ). There are three high-symmetry middle planes of the BZ with little group CsC_{s}, which are with respect to three-fold rotational symmetry C3zC_{3z}. (b) The enlarged view of bands of CsV6Sb6\mathrm{CsV_{6}Sb_{6}} in the absence of spin-orbital coupling (SOC) along an arbitrary line in a middle plane. Two crossing bands belong to opposite mirror eigenvalues Γ1\Gamma_{1} (+1) and Γ2\Gamma_{2} (-1) of CsC_{s}. (c) In the presence of SOC, two crossing bands without SOC now belong to the same mirror eigenvalues Γ4\Gamma_{4} of CsC_{s} and thus the band crossings are avoided.
Table SII: The irreducible representations (irrep.) and parity products (prod.) of occupied bands of CsV6Sb6\mathrm{CsV_{6}Sb_{6}} at the time-reversal invariant momenta points. The band topology between the lowest conduction band (119) and highest valence band (117) characterized by the (0;111) 2\mathbb{Z}_{2} index. The nontrivial band topology are also present when the occupied band is band 115 or band 113, which is close to the Fermi level.
Occ. band irrep. Parity Prod. Invar.
Γ\Gamma 3×F3\times F 3×L3\times L Z δΓ\delta_{\Gamma} δF\delta_{F} δL\delta_{L} δZ\delta_{Z} 2\mathbb{Z}_{2}
(D3dD3d) (C2h)(C2h) (C2h)(C2h) (D3dD3d)
119 Γ4\Gamma_{4}^{-} Γ3+Γ4+\Gamma_{3}^{+}\oplus\Gamma_{4}^{+} Γ3Γ4\Gamma_{3}^{-}\oplus\Gamma_{4}^{-} Γ4+\Gamma_{4}^{+} - - - - (0;000)
117 Γ4+\Gamma_{4}^{+} Γ3+Γ4+\Gamma_{3}^{+}\oplus\Gamma_{4}^{+} Γ3Γ4\Gamma_{3}^{-}\oplus\Gamma_{4}^{-} Γ4\Gamma_{4}^{-} + - + - (0:111)
115 Γ4\Gamma_{4}^{-} Γ3Γ4\Gamma_{3}^{-}\oplus\Gamma_{4}^{-} Γ3+Γ4+\Gamma_{3}^{+}\oplus\Gamma_{4}^{+} Γ4+\Gamma_{4}^{+} + - - + (0;111)
113 Γ5Γ6\Gamma_{5}^{-}\oplus\Gamma_{6}^{-} Γ3Γ4\Gamma_{3}^{-}\oplus\Gamma_{4}^{-} Γ3+Γ4+\Gamma_{3}^{+}\oplus\Gamma_{4}^{+} Γ5+Γ6+\Gamma_{5}^{+}\oplus\Gamma_{6}^{+} - + - + (0;111)