Next Article in Journal
Measurement of Reverse-Light-Induced Excited Spin State Trapping in Spin Crossover Systems: A Study Case with Zn1−xFex(6-mepy)3tren(PF6)2·CH3CN; x = 0.5%
Previous Article in Journal
Heterogeneous Nucleation Mechanism of Potassium Iodide on Graphene Surface in Water
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

LaZn1−xBi2 as a Candidate for Dirac Nodal-Line Intermetallic Systems

by
Piotr Ruszała
,
Maciej J. Winiarski
* and
Małgorzata Samsel-Czekała
Institute of Low Temperature and Structure Research, Polish Academy of Sciences, Okólna 2, 50-422 Wrocław, Poland
*
Author to whom correspondence should be addressed.
Crystals 2024, 14(3), 209; https://doi.org/10.3390/cryst14030209
Submission received: 31 January 2024 / Revised: 16 February 2024 / Accepted: 19 February 2024 / Published: 22 February 2024

Abstract

:
The complex theoretical analysis of the density of states, band structures, and Fermi surfaces, based on predictions of the density functional theory methods, unveils the unique electronic properties of the LaZn1−xBi2 system. In this paper, the Zn vacancies (for x = 0.5 ) were modeled using a modified unit cell of lower symmetry than that for a fully stoichiometric one (for x = 0 ). The existence of several Dirac-like features in the electronic band structures was found. Some of them were found to be intimately associated with the nonsymmorphic symmetry of the system, and these were investigated in detail. The calculated Fermi surface shapes, as well as the Fermi velocity values (up to ∼1.2 × 10 6 m/s), are in good agreement with other analogous square-net Dirac semimetals. The combination of charge-carrier uncompensation, relatively small band splitting, and the tolerance factor for square-net semimetals t 0.95 for LaZn0.5Bi2, constitutes a very promising indicator of the topological features of this system, warranting further experimental studies.

1. Introduction

Ternary rare-earth zinc pnictides REZnPn2 ( R E = rare-earth; P n = pnictogen atom) belong to a large family of the 112-type intermetallics, which are excellent subjects for studies on various physical phenomena, including the topological aspects of the electronic band structure [1]. The 112-type materials exhibit the presence of superconductivity (SC), e.g., RENi1−xBi2 [2] and LaPd1−xBi2Bi2 [3], heavy fermions (CeTSb2 [4,5], CeAgSb2 [6,7,8], CeCuBi2 [9]), Dirac fermions in EuMnBi2, LaAg(Sb;Bi)2, and YbMn(Bi;Sb)2, as well as a linear magnetoresistance and charge density wave (CDW) [10,11,12,13,14,15,16]. The tetragonal structure of HfCuSi2-type (including its derivatives) [17,18] and some orthorhombic distortions are characteristic of REZnPn2 compounds [19]. The crystal structure of such systems exhibits strong anisotropy with quasi-two-dimensional (quasi-2D) metallic-like sheets of pnictogens forming atomic square nets; thus, these compounds are so-called square-net semimetals [20].
The REZn1−xBi2 ( R E = La, Ce, Pr) compounds belong to the square-net materials mentioned above. Single crystals of these systems were synthesized within a reaction of pure metals and characterized in a previous work by O. Y. Zelinska and A. Mar [21]. The single-crystal X-ray diffraction showed that the zinc 2b sites in these compounds are half-occupied ( x 0.5), without any signs of a superstructure, which may be a hint of statistically distributed vacancies [21]. The resistivity measurements generally revealed a metallic character of the above REZn1−xBi2 materials, but a shallow minimum around 10 K and an abrupt drop in resistivity below ∼5 K were observed in CeZn1−xBi2, which suggests Kondo-lattice behavior similar to that of Ce(Ni;Cu)Sb2 [22]. The REZn1−xBi2 systems do not exhibit a transition to SC down to 2 K [21], whereas isostructural Pd- and Ni-based compounds, i.e., CeNi1−xBi2 ( x 0.2 ) [23] and LaPd1−x(Sb;Bi)2 ( x 0.15 ) [3,4,24], are superconductors. The electronic structures of the REZn1−xBi2 compounds have not yet been studied.
In this paper, we construct a simplified unit cell model for the LaZn1−xBi2 intermetallic with a realistic value of x = 0.5 (and compare it to that for x = 0 ). Our approach is justified by earlier works, in which a similar simplification was assumed, e.g., for familiar compounds like LaAg1−xZnyAs2 [25], as well as for Cu-doped CaAuAs [26]. Based on this assumption, subsequently, the electronic structure is studied using the density functional theory (DFT) methods. The density of states (DOS), band structures, and Fermi surfaces (FSs), determined theoretically, are presented and discussed, with a special focus on the disorder effect. A careful analysis of the electronic band structures reveals conic-like bands and some Dirac points (DP) near the Fermi level ( E F ) in the studied systems. The obtained results are compared to those reported for well-known Dirac square-net materials such as ZrSiS and LaAgSb 2 (both crystallized in the same tetragonal space group P4/nmm as the one considered here). The possible existence of topological semimetal (TSM) or Dirac nodal-line semimetal (DNLSM) states characteristic of the square-net pnictide intermetallics is discussed.

2. Materials and Methods

The electronic structure calculations were performed with the full-potential local-orbital (FPLO-14) code [27]. The Perdew–Burke–Ernzerhof parametrization (PBE96 [28]) of the generalized gradient approximation (GGA) was used in both the scalar and fully relativistic [29] calculations, i.e., both without and with spin-orbit coupling (SOC). The experimental values of the lattice parameters, a = 0.4599 nm and c = 0.9987 nm, of the tetragonal unit cell (P4/nmm, space group, No. 129) for LaZn1−xBi2 were used [21]. The band structures for LaZn1−xBi2 were compared to those for well-known square-net materials, i.e., ZrSiS and LaAgSb 2 (adopting the P4/nmm structure). The cell parameters of ZrSiS (a = 0.3545 nm; c = 0.8058 nm) and LaAgSb 2 (a = 0.439 nm; c = 1.084 nm) were taken from Refs. [30,31]. In turn, the simplified deficient unit cell of LaZn0.5Bi2 was constructed by removing one Zn atom from the basic crystal structure (see Figure 1a), resulting in a unit cell with the P 4 ¯ m2 space group (No. 115) and containing artificially ordered Zn vacancies.
It is worth noting that the unit cell of space group No. 129 is centrosymmetric and nonsymmorphic, which is opposite to the unit cell of space group No. 115, which is noncentrosymmetric and symmorphic. Valence-basis sets were automatically selected by the internal FPLO procedure. The total energy values of the considered systems were converged for the 16 × 16 × 16 k-point meshes corresponding to 405 ( x = 0 ) and 657 ( x = 0.5 ) points (from a total of 4096 k-points) in the irreducible part of the Brillouin zone (BZ).

3. Results and Discussion

The DOS plots of the LaZn1−xBi2 compounds, calculated for x = 0 and 0.5, are presented in Figure 1b and Figure 1c, respectively. The overall shapes of the total DOS plots reflect the metallic character of both Zn compositions. One can see, roughly speaking, a downshift of the Fermi level ( E F ) for LaZn0.5Bi2 by ∼0.7–0.8 eV with respect to that of LaZnBi 2 , which is caused by a partial deficiency of Zn ions. However, the subpeak structures also seem to be quite different depending on the contents of x. Additionally, with the increase in the Zn content, the relative contribution of the partial Bi states to the total DOS at E F decreases compared to that of the other partial orbitals.
The major contributions to the total DOS near E F in the systems come from the Bi p and La 5d states. The unoccupied La 4f orbitals are present in a hybridization-like tail at E F . Similar DOS plots have been published for hypothetical LaZnAs 2 [32]. The non-metallic character of LaZn0.67As2, in contrast to the metallic-like LaZn0.5Bi2 [21,32], indicates that the position of E F in the former system should be close to a DOS pseudo-gap. The roughly estimated shift of E F in LaZn0.5Bi2 with respect to that of LaZnBi 2 also leads to a pronounced decrease in the DOS( E F ).
The valence-electron count in the LaZn1−xPn2 compounds for isoelectronic Pn = As, Sb, and Bi series depends on the Zn deficiency, x [20,21,25]. Furthermore, this deficiency in the 112-type systems is enhanced with the decreasing electronegativity of pnictogen ions ( x 0.33 for As; x 0.4 for Sb; x 0.5 for Bi) [21,32,33]. The electrical resistivity of LaZn0.6Sb2 single crystals suggests a weakly metallic character [34], which appears to be in between the nonmetallic LaZn0.67As2 [32] and metallic-like LaZn0.5Bi2 [21]. On one hand, the electron excess caused by a possibly high Zn content weakens the stability of the system through the antibonding of the Zn–Pn states pinned close to E F [20,25,35,36]. The Zn–Bi2 bonds might be responsible for a decrease in the stability of the 112 systems, as described above, and a significant Zn deficiency is preferred. On the other hand, hypervalent PnPn bonds in the square net could be a stabilizing factor [20,32]. Interestingly, for some intermetallics (e.g., LnAuSb, Ln = lanthanide), it has been shown that Au–Au interactions can stabilize the phase through the localization of one electron in the bond and the Au–Sb antibonding states left empty [36]. However, the Zn–Zn distance of 3.25 Å (which is also equal to the Bi1–Bi12 distance) in LaZnBi2 is longer than the Au–Au distance (3.12 Å) in LaAuSb [21,36]. The average Zn–Zn interatomic distance increases with decreasing Zn content, and for LaZn0.5Bi2, it is equal to the lattice parameter a 4.6 Å [21]. Furthermore, the Zn–Zn interactions in LaZn1−xAs2 are rather weak and non-bonding [32]; hence, in the case of Zn-based 112-type bismuthides, such an effect seems to be less pronounced.
The overall shapes of the total DOS plots, as well as the partial contributions obtained here for LaZn0.5Bi2, are quite similar to those reported for other compounds of a general formula LaMPn2, where M is a transition metal with a completely filled d-shell and in the M1+ or M2+ oxidation state, e.g., LaAgBi2 [11], LaAgSb2 [12,13], and LaCuSb2 [37]. There are also some similarities to the DOS plots obtained for other zinc bismuthides, e.g., BaZnBi2 [38,39]. The total DOS for LaZnBi2 is similar to the results reported for arsenides with a hypothetical (higher than the measured one) Zn content such as LaAg0.5Zn0.5As2 [25], LaZnAs2 [32], and LaZn2As3 [35].
For square-net semimetals, one can define the so-called tolerance factor, t, as the ratio of the in-plane atomic distance Bi1–Bi1 in the square net and the distance to the nearest out-of-plane atom (La–Bi1) [18,20]. For values of t 0.95 , it is predicted that the system can be topologically non-trivial and the band structure can be inverted [20]. The parameter t = 0.941 0.95 for both compositions of LaZn1−xBi2 suggests a possible topological order in the system, especially expected for the more realistic x = 0.5 composition. This is also in line with the quasi-two-dimensionality of some FS sheets obtained for LaZn1−xBi2 (as discussed later). The values of the total DOS at E F , N ( E F ) are 1.4 and 1.8 states eV−1 f.u.−1 for LaZn0.5Bi2 and LaZnBi2, respectively, and are close to those reported for similar compounds such as LaCuSb2 (1.3 states eV−1 f.u.−1) and LaAgSb2 (0.9 states eV−1 f.u.−1) [37,40]. One may also estimate the Sommerfeld coefficients, γ , based on the simple free-electron formula [ γ 2.36 × N ( E F ) ]. Thus, the estimated values of γ are equal to 3.32 and 4.24 mJ mol−1 K−2 for LaZn0.5Bi2 and LaZnBi2, respectively, indicating rather weak electronic correlations and electron–phonon coupling, which requires experimental verification. Moreover, these values are close to those measured for LaCuSb2 ( γ e x p = 2.92 mJ mol−1 K−2) and LaAgSb2 ( γ e x p = 2.62 mJ mol−1 K−2) [4]. Due to the very small contributions of the transition metal d-states to the total DOS at E F and weak electron–phonon coupling, compounds like LaMPn2 (for M = Ag, Au, Cu, Zn) usually do not exhibit transitions to SC (at least down to 1.5 K), which is also consistent with the relatively small values of the Sommerfeld coefficients [4,17,41,42,43,44].
In Figure 2a,b, the same BZs are displayed but with different paths highlighted. Red indicates the sketch of the cage-like path made of connected square-like nodal-line loops (meaning at least two electronic bands are fairly close to degenerating along these red lines). This kind of nodal-line loop is characteristic of square-net semimetals with the tetragonal P4/nmm space group, e.g., it was discovered previously in ZrSiS, which is the reference compound with small SOC (∼20 meV) [30,45,46]. It is worth mentioning that this kind of nodal line is the consequence of the enlarged (doubled) square-net and nonsymmorphic symmetry of the system (offsite fourfold axis, two screw axes, and glide z-mirror plane), which results in the corresponding band structure with the nodal line related to the band folding (encircling the Γ point) [46,47,48,49,50]. In highly 2D systems, the k z -dispersion is small, and every k z -cut exhibits a similar square-shaped nodal-like structure (e.g., ZrSiS [51]). However, it is not robust against SOC caused by the hybridization effect (even when small in magnitude), with states coming from the out-of-plane atoms (here, La 5d and Bi2 6p) in many known materials [46,47,48]. Band degeneracy can occur due to the nonsymmorphic symmetry along the A–R and M–X lines without SOC [30,46]. However, when SOC is present, there are some possible crossings of these bands due to locally vanishing SOC [46,52,53]. The straight lines at the BZ boundary (marked in cyan and green in Figure 2b) are symmetry-enforced nodal lines connected with the nonsymmorphic and time-reversal T symmetries [46,48,49,54]. The difference between the green (R–X–R) and cyan (A–M–A) lines is clear: the electronic bands at the M and A points are distant from E F in LaZn1−xBi2, as well as in the other compounds mentioned [30,52,53]. Note that the R–X–R nodal line in the vicinity of E F is the most interesting.
The electronic band structures calculated for LaZn1−xBi2 (Figure 2c,d) exhibit characteristic features found in p-block square-net materials [18,20]. Similar band structures with highly dispersive and linear electronic bands have also been reported for LaCuSb 2 [37], LaAgSb 2 [13], LaAgBi 2 [11], ZrSiS [30], and SmSbTe [55]. The reference band structures for Dirac systems such as ZrSiS and LaAgSb 2 , calculated here using the same method, are also presented in Figure 2e and Figure 2f, respectively. When compared to scalar relativistic band structures (not shown), the SOC changes the electronic bands to some extent and opens gaps between a few bands, e.g., along the M– Γ and A–Z directions in the BZ, enforcing the decoupling of bands along the X–M and R–A lines. Considering the impact of SOC related to the heavy Bi atoms, the band structures of LaZnBi 2 exhibit conic-like features common with the ZrSiS system (Figure 2e), where the Dirac states come from the Si 3p electrons and are only negligibly influenced by SOC and other electronic bands [30,46]. The linear bands along the M– Γ and A–Z lines in the BZ of LaZnBi 2 are strongly influenced by the neighboring states compared to the case of ZrSiS. For the Bi square net, the nodal-line loops are distorted in LaZn1−xBi2, as seen in the isostructural Weyl semimetal of type-II—YbMnBi2 [14]. Meanwhile, the bands around the X and R points in LaZnBi 2 resemble those in LaAgSb 2 (Figure 2f). The cones placed around E F exactly at the X and R points belong to the symmetry-enforced nodal lines on the BZ boundaries, marked by thick green lines in Figure 2b. Besides LaAgSb 2 , similar bands can also be found in LaAgBi 2 and partially in InBi [52,53,54]. In this manner, possible Dirac bands along the R–X line ( D X and D R ) are displayed for both compositions in Figure 3. For x = 0.5, the two bands cross E F along that line due to visible small splitting. Note that in the case of our model P 4 ¯ m2 structure of LaZn0.5Bi2, the inversion symmetry (P) is absent but the time-reversal symmetry (T) is present. However, despite visible orbital contributions from the heavy Bi atoms (and the other ones) to the bands near E F , the anisotropic band splitting induced by asymmetric SOC (ASOC) is relatively small because the Zn square-net breaking inversion symmetry is well separated both spatially (the distance between the Zn plane and the remaining atoms is quite large) and in the band energy scale from the Bi square net. This kind of system may exhibit Weyl states or extra nodal lines around the X(R) points due to breaking the P symmetry [49]. Interestingly, the band structure of LaZn0.5Bi2 (Figure 2c) is very similar to that of LaZn0.5Sb2, obtained with the use of large supercells (56 and 112 atoms) for modeling Zn vacancies [20].
On one hand, in the relatively small unit cells (statistically non-uniform), the broken inversion symmetry introduces additional asymmetric band splitting. On the other hand, in large supercell calculations, the procedure of electronic band unfolding blurs the picture [20]. The results presented here for x = 0.5 seem reasonable. Apart from the asymmetric band splitting in LaZn0.5Bi2, there are still some linear bands along the M– Γ and A–Z lines, which are similar to those in ZrSiS, whereas at the X point, the linearity reminds us of the band structure of LaAgSb 2 . In Figure 3, the symmetry-enforced degeneracy and nodal lines (along the R–X–R direction) are plotted. The electronic bands along the M–X and R–A lines for LaZnBi 2 (Figure 3b) show “eye-mask”-type band connections [50]. The anisotropic band crossing, protected near the R point along the R–A line, is caused by this kind of band connection. Along the X–R and A–M directions, one can see strictly symmetry-enforced “zipper” nodal lines [50]. In the case of LaZn0.5Bi2 (Figure 3a), the band connectivity is different, and is rather like a “double eye mask” along the M–X line and “double hourglass” along the R–A line [50]. As seen in Figure 3b,f,g, the Dirac cone along the X–R direction at k z = 0.43 π / c ( D XR ) in LaZnBi 2 is present exactly at E F and exhibits a 2D character.
A similar Dirac cone is revealed in LaAgBi 2 (located at k z = 0.6 π / c ), and its upper Dirac branch is closer to linear than that in the aforementioned compound [53]. Additionally, for LaZn0.5Bi2 (Figure 3a), two bands cross E F along the X–R line. These bands (Figure 3d,e) exhibit a quadratic dispersion along the perpendicular direction to the X–R line and Rashba-like ASOC splitting [56,57]. In the case of the full stoichiometry (x = 0), the Dirac point ( D ZR ) seen along the Z–R line (Figure 2d) is analyzed in detail in Figure 4. This anisotropic band along the Z–R line (almost at half its length) is only negligibly influenced by SOC (energy gap is just ∼13 meV) in LaZnBi 2 . It belongs to the nodal-line loop (red one in Figure 2b) and is rather unique, e.g., it is gapped in LaAg(Sb;Bi) 2 [11,12,52]. The dispersion of the associated Dirac-like bands in the k z -direction (see Figure 4d) is very small compared to those in other directions. Additionally, along the third perpendicular direction (L– D ZR –L path), the dispersion is quadratic-like (Figure 4c). Based on it, the cone is not a true 3D Dirac cone and, hence, is not very interesting in view of topological features. In general, the above Dirac-like bands are expected to exist along several directions (along Γ –M, Γ –X, and Z–A) extending from the Γ point and associated with the nodal-line loop. It is also predicted both theoretically and experimentally that these bands should be strongly anisotropic [12,14,17,47,58]. An interesting example is LaAgSb 2 , in which Dirac points along the Γ –M and Z–A lines are formed by two almost linear and anisotropic electronic bands [12,59], as seen in Figure 2f. The corresponding dispersions calculated along the other two perpendicular directions are clearly quadratic for cones in both the k z = 0 and k z = π planes [59].
The Fermi surface (FS) sheets for the LaZn1−xBi2 systems are illustrated in Figure 5, revealing their strongly anisotropic character. However, the influence of SOC on the FS sheets is rather subtle. Note that in the presence of the inversion symmetry in LaZnBi 2 , the Kramers degeneracy in the fully relativistic mode (i.e., with SOC) is preserved. The pillar-like FS sheet (No. 79 in Figure 5a) is open in the k z -direction but differs from the closed FS pocket (No. 157 in Figure 5b). The four Kramers double-degenerate sheets (Nos. 153, 155, 157, and 159) are associated with the blue, red, green, and orange bands in Figure 2d. The green and orange bands form electron-like Dirac FS pockets (No. 157 and No. 159 in Figure 5b). Moreover, these two electron pockets touch each other along the X–R nodal-line direction, which is very similar to the situation observed in the LaAgBi 2 compound [53]. The corresponding Dirac cone is shown in Figure 3b (as described in the previous paragraph). The position of this point (denoted by D XR ) at k z = 0.43 π c separates the volumes occupied by electrons near the X point and unoccupied (holes) near the R point. The dispersion relations along the two remaining perpendicular directions (Figure 3f,g) reveal the quasi-2D character of these conic bands.
When analyzing the FS of LaZn0.5Bi2, one can see that the lack of inversion symmetry here lifts the Kramers degeneracy, leading to an anisotropic band splitting (ASOC), as seen in Figure 5c,d). The three-dimensional (3D) hole-like pillow (No. 67 in Figure 5c) obtained in the scalar relativistic mode (i.e., without SOC) splits into two different FS sheets (Nos. 133 and 134) in the fully relativistic approach with SOC, as depicted in Figure 5d. In a similar way, the corrugated quasi-2D square-like electron sheet (No. 68 in Figure 5c) splits into two different sheets (Nos. 135 and 136 in Figure 5d). The above-mentioned FS sheet (No. 135) reveals quite a similar shape to that of No. 68, whereas the FS sheet related to the band (No. 136) is smaller and more complex compared to that of No. 68. The FS sheet (No. 136) also exhibits two characteristic features of many other 112-type Dirac semimetals. The electron-like feature around the X point in FS sheet No. 136 can also be found in compounds such as LaAgBi 2 [11,53], CaMnBi 2 [47,58], BaZnBi 2 [38,60,61], EuMnBi 2 , and YbMnBi 2 [14]. The other electron-like FS pieces of No. 136, which resemble lenses and are distributed perpendicular to the Z–A line, also occur in FS sheets Nos. 155 and 156 of LaZnBi 2 but are hole-like and closer to the Z point (Figure 5b). These “lenses” are clearly associated with the Dirac band (red one) in the corresponding band structures (Figure 2c,d). The electron- and hole-like character of these “lenses” depends on the position of E F and the details of the dispersion of the corresponding bands. It is also strongly dependent on a given chemical composition of the 112-type system. In isostructural YbMnBi 2 , which hosts type-II Weyl states, these “lenses” are hole-like and connected to the electron-like pockets toward the X points. Here, SOC introduces a small energy gap (∼10 meV), and the Weyl dispersion is not substantially suppressed [14]. Similarly, hole-like “lenses” are also present in EuMnBi 2 , but due to stronger hybridization and a greater energy gap (∼0.2 eV) related to SOC, Weyl states are absent in this case [14]. Very similar FS pockets were revealed in ARPES for CaMnBi 2 (isostructural with LaZn1−xBi2) and partially for SrMnBi 2 [58], i.e., square-like contours with hole-like “lenses” (described by the authors as “crescent moon-like” shapes) and electron-like thin pockets near the X point. Furthermore, the FS of SrMnBi 2 (isostructural with EuMnBi 2 ) was found to be similar to that of EuMnBi 2 , i.e., the hole-like “lenses” are very pronounced but there are unoccupied states close to E F around the X point [58]. The comparison of SrMnBi 2 and CaMnBi 2 crystal structures and their influence on the electronic structure properties are described in detail in [47]. Moreover, in both CaMnBi 2 and SrMnBi 2 , there are also hole-like FS pockets centered around the Γ point [58]. These hole-like shapes seem to be similar to those originating from bands Nos. 133 and 134, calculated here for LaZn0.5Bi2, as shown in Figure 5d. The small Dirac-like electron pockets of Nos. 137 and 138 located around the X point are a common feature in this family of compounds and can be found in compounds such as LaAgBi 2 [53], LaCuSb 2 [37,41], and LaAgSb 2 [52].
Based on the volume of each FS sheet in the BZ (Figure 5), we numerically estimated the electron/hole compensation ratios. Interestingly, almost perfect carrier compensation ( n e / n h 0.98 ) was found for LaZnBi 2 , and a lack of carrier compensation was found for LaZn0.5Bi2 ( n e / n h 1.23 ). These results were negligibly influenced by SOC. Higher n e / n h ratios were reported for LaSb 2 ( n e / n h 1.4 ) and LaAgSb 2 ( n e / n h 1.6 ), layered Dirac-like compounds [40]. The Fermi velocity maps calculated for LaZnBi 2 and LaZn0.5Bi2 indicate that the maximum carrier velocity values (reaching up to 1.2 × 10 6 m/s) occurred around the X point in LaZn0.5Bi2 and in the Γ –M direction in LaZnBi 2 . Similar results have been previously obtained for other similar compounds, e.g., (1.5–8) × 10 5 m/s for the Dirac state in LaAgSb 2 [12,13], (3–6.5) × 10 5 m/s for ZrSiS [30,45], and ∼1.6 × 10 6 m/s for one Dirac branch along Γ –M in SrMnBi 2 and CaMnBi 2 [17,58].

4. Conclusions

In this paper, we have carefully studied the electronic structures of selected LaZn1−xBi2 compounds for a model crystal structure corresponding to a real experimental composition of x = 0.5 and for a hypothetical x = 0 . The reduced Zn content ( x = 0.5 ) led to a downshift of the Fermi level by ∼0.8–1.0 eV, which promoted the contribution of the Bi 6p square-net states to the total DOS at E F compared to the hypothetical LaZnBi 2 . The value of the tolerance factor t, obtained for LaZn1−xBi2 materials, was estimated to be 0.95 , suggesting certain topological features, particularly expected in the more realistic composition with x = 0.5 . Indeed, almost linear Dirac-like bands were found in both systems. The conic-like shapes along the Γ –M and Z–A directions were more pronounced in LaZn0.5Bi2, and a Dirac point at E F in the middle of the Z–R line in LaZnBi 2 was present. Two bands at the X point, slightly below E F , formed another Dirac point (fourfold degenerate), which was protected by the nonsymmorphic and time-reversal symmetries along the X–R line in LaZnBi 2 , and another one caused by the vanishing of the asymmetric SOC at the time-reversal invariant X point in LaZn0.5Bi2. These features are quite similar to those in the reference square-net Dirac semimetals like LaAgSb 2 and ZrSiS. The overall anisotropic shapes of the Fermi surfaces and Fermi velocity values are also in good agreement with those reported for other square-net materials. One may expect that the transverse magnetoresistance in LaZn1−xBi2 is limited by the disorder associated with the Zn vacancies. However, for the composition x = 0.5 , the lack of carrier compensation suggests that in the case of linear field-dependent magnetoresistance, it can originate mainly from the Dirac-like states of the Bi square net. Further magnetotransport and ARPES studies are needed to examine the existence of Dirac states in the modeled LaZn0.5Bi2 material. In addition, a potential substitution of Zn with Ag ions may be used to shift E F and induce the topologically non-trivial states in these 112-type bismuthides.

Author Contributions

P.R.: conceptualization, methodology, investigation, visualization, and writing; M.J.W.: writing and supervision; M.S.-C.: conceptualization, resources, writing, and supervision. All authors have read and agreed to the published version of the manuscript.

Funding

The calculations for this work were performed at the Wroclaw Center for Networking and Supercomputing (Project No. 477).

Data Availability Statement

The datasets generated during and/or analyzed during the current study are available from the corresponding author upon reasonable request.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Armitage, N.P.; Mele, E.J.; Vishwanath, A. Weyl and Dirac semimetals in three-dimensional solids. Rev. Mod. Phys. 2018, 90, 015001. [Google Scholar] [CrossRef]
  2. Mizoguchi, H.; Matsuishi, S.; Hirano, M.; Tochibana, M.; Muromachi, E.T.; Kawaji, H.; Hosono, H. Coexistence of Light and Heavy Carriers Associated with Superconductivity and Antiferromagnetism in CeNi0.8Bi2 with a Bi Square Net. Phys. Rev. Lett. 2011, 106, 057002. [Google Scholar] [CrossRef]
  3. Han, F.; Malliakas, C.D.; Stoumpos, C.C.; Sturza, M.; Claus, H.; Chung, D.Y.; Kanatzidis, M.G. Superconductivity and strong intrinsic defects in LaPd1−xBi2. Phys. Rev. B 2013, 88, 144511. [Google Scholar] [CrossRef]
  4. Muro, Y.; Takeda, N.; Ishikawa, M. Magnetic and transport properties of dense Kondo systems, CeTSb2 (T = Ni, Cu, Pd and Ag). J. Alloys Compd. 1997, 257, 23–29. [Google Scholar] [CrossRef]
  5. Koyama, T.; Matsumoto, M.; Wada, S.; Muro, Y.; Ishikawa, M. Combination of Dense Kondo Effect and RKKY Interactions in CeCuSb2, Investigated by NMR. J. Phys. Soc. Jpn. 2001, 70, 3667–3672. [Google Scholar] [CrossRef]
  6. Inada, Y.; Tamizabel, A.; Sawai, Y.; Ikeda, S.; Shishido, H.; Okubo, T.; Yamada, M.; Onuki, Y.; Ebihara, T. Quasi-2D Fermi Surfaces of the Magnetic Compound CeAgSb2. Acta Phys. Pol. B 2003, 34, 1141–1144. Available online: http://www.actaphys.uj.edu.pl/R/34/2/1141/pdf (accessed on 31 January 2024).
  7. Zou, Y.; Logg, P.; Barber, M.; Feng, Z.; Ebihara, T.; Grosche, F.M. Thermodynamics of field-tuned quantum critical point in CeAgSb2. Phys. Status Solidi B 2013, 250, 529–532. [Google Scholar] [CrossRef]
  8. Chen, R.Y.; Liu, J.L.; Shi, L.Y.; Wang, C.N.; Zhang, S.J.; Wang, N.L. Possible orbital crossover in the ferromagnetic Kondo lattice compound CeAgSb2. Phys. Rev. B 2019, 99, 205107. [Google Scholar] [CrossRef]
  9. Adriano, C.; Rosa, P.F.S.; Jesus, C.B.R.; Mardegan, J.R.L.; Garitezi, T.M.; Grant, T.; Fisk, Z.; Garcia, D.J.; Reyes, A.P.; Kuhns, P.L.; et al. Physical properties and magnetic structure of the intermetallic CeCuBi2 compound. Phys. Rev. B 2014, 90, 235120. [Google Scholar] [CrossRef]
  10. Masuda, H.; Sakai, H.; Tokunaga, M.; Yamasaki, Y.; Miyake, A.; Shiogai, J.; Nakamura, S.; Awaji, S. Quantum Hall effect in a bulk antiferromagnet EuMnBi2 with magnetically confined two-dimensional Dirac fermions. Sci. Adv. 2016, 2, e1501117. [Google Scholar] [CrossRef] [PubMed]
  11. Wang, K.; Graf, D.; Petrovic, C. Quasi-two-dimensional Dirac fermions and quantum magnetoresistance in LaAgBi2. Phys. Rev. B 2013, 87, 235101. [Google Scholar] [CrossRef]
  12. Shi, X.; Richard, P.; Wang, K.; Liu, M.; Matt, C.E.; Xu, N.; Dhaka, R.S.; Ristic, Z.; Qian, T.; Yang, Y.-F.; et al. Observation of Dirac-like band dispersion in LaAgSb2. Phys. Rev. B 2016, 93, 081105. [Google Scholar] [CrossRef]
  13. Wang, K.; Petrovic, C. Multiband effects and possible Dirac states in LaAgSb2. Phys. Rev. B 2012, 86, 155213. [Google Scholar] [CrossRef]
  14. Borisenko, S.; Evtushinsky, D.; Gibson, Q.; Yaresko, A.; Koepernik, K.; Kim, T.; van den Brink, M.A.J.; Hoesch, M.; Fedorov, A.; Haubold, E.; et al. Time-reversal symmetry breaking type-II Weyl state in YbMnBi2. Nat. Commun. 2019, 10, 3424. [Google Scholar] [CrossRef] [PubMed]
  15. Chaudhuri, D.; Cheng, B.; Yaresko, A.; Gibson, Q.D.; Cava, R.J.; Armitage, N.P. Optical investigation of the strong spin-orbit-coupled magnetic semimetal YbMnBi2. Phys. Rev. B 2017, 96, 075151. [Google Scholar] [CrossRef]
  16. Wang, Y.-Y.; Xu, S.; Sun, L.-L.; Xia, T.-L. Quantum oscillations and coherent interlayer transport in a new topological Dirac semimetal candidate YbMnSb2. Phys. Rev. Mater. 2018, 2, 021201. [Google Scholar] [CrossRef]
  17. Ray, S.J.; Alff, L. Superconductivity and Dirac fermions in 112-phase pnictides. Phys. Status Solidi B 2017, 254, 1600163. [Google Scholar] [CrossRef]
  18. Klemenz, S.; Lei, S.; Schoop, L.M. Topological Semimetals in Square-Net Materials. Annu. Rev. Mater. Res. 2019, 49, 185–206. [Google Scholar] [CrossRef]
  19. Tamura, S.; Katayama, N.; Yamada, Y.; Sugiyama, Y.; Sugawara, K.; Sawa, H. Various Arsenic Network Structures in 112-Type Ca1–xLaxFe1–yPdyAs2 Revealed by Synchrotron X-ray Diffraction Experiments. Inorg. Chem. 2017, 56, 3030–3035. [Google Scholar] [CrossRef] [PubMed]
  20. Klemenz, S.; Hay, A.K.; Teicher, S.M.L.; Topp, A.; Cano, J.; Schoop, L.M. The Role of Delocalized Chemical Bonding in Square-Net-Based Topological Semimetals. J. Am. Chem. Soc. 2020, 142, 6350–6359. [Google Scholar] [CrossRef] [PubMed]
  21. Zelinska, O.Y.; Mar, A. Structure and electrical resistivity of rare-earth zinc bismuthides REZn1−xBi2 (RE = La, Ce, Pr). J. Alloys Compd. 2008, 451, 606–609. [Google Scholar] [CrossRef]
  22. Lakshmi, K.V.; Menon, L.; Nigam, A.K.; Das, A.; Malik, S.K. Magneto-resistance studies on RTSb2 compounds (R = La, Ce and T = Ni, Cu). Phys. B Condens. Matter 1996, 223–224, 289–291. [Google Scholar] [CrossRef]
  23. Kodama, K.; Wakimoto, S.; Igawa, N.; Shamoto, S.; Mizoguchi, H.; Hosono, H. Crystal and magnetic structures of the superconductor CeNi0.8Bi2. Phys. Rev. B 2011, 83, 214512. [Google Scholar] [CrossRef]
  24. Retzlaff, R.; Buckow, A.; Komissinskiy, P.; Ray, S.; Schmidt, S.; Mühlig, H.; Schmidl, F.; Seidel, P.; Kurian, J.; Alff, L. Superconductivity and role of pnictogen and Fe substitution in 112-LaPdxPn2 (Pn = Sb, Bi). Phys. Rev. B 2015, 91, 104519. [Google Scholar] [CrossRef]
  25. Ramachandran, K.K.; Genet, C.; Mar, A. Quaternary rare-earth arsenides REAg1−xZnyAs2 (RE = La–Nd, Sm, Gd–Dy) with tetragonal SrZnBi2- and HfCuSi2-type structures. J. Solid State Chem. 2015, 231, 204–211. [Google Scholar] [CrossRef]
  26. Malick, S.; Ghosh, A.; Barman, C.K.; Alam, A.; Hossain, Z.; Mandal, P.; Nayak, J. Weak antilocalization effect and triply degenerate state in Cu-doped CaAuAs. Phys. Rev. B 2022, 105, 165105. [Google Scholar] [CrossRef]
  27. Koepernik, K.; Eschrig, H. Full-potential nonorthogonal local-orbital minimum-basis band-structure scheme. Phys. Rev. B 1999, 59, 1743–1757. [Google Scholar] [CrossRef]
  28. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1996, 77, 3865–3868. [Google Scholar] [CrossRef]
  29. Eschrig, H.; Richter, M.; Opahle, I. Relativistic solid state calculations. In Relativistic Electronic Structure Theory, Part 2. Applications (Theoretical and Computational Chemistry); Schwerdtfeger, P., Ed.; Elsevier: Amsterdam, The Netherlands, 2004; Volume 14, pp. 723–776. [Google Scholar]
  30. Schoop, L.M.; Ali, M.N.; Straßer, C.; Topp, A.; Varykhalov, A.; Marchenko, D.; Duppel, V.; Parkin, S.S.P.; Lotsch, B.V.; Ast, C.R. Dirac cone protected by non-symmorphic symmetry and three-dimensional Dirac line node in ZrSiS. Nat. Commun. 2016, 7, 11696. [Google Scholar] [CrossRef]
  31. Brylak, M.; Möller, M.H.; Jeitschko, W. Ternary Arsenides ACuAs2 and Ternary Antimonides AAgSb2 (A = Rare-Earth Elements and Uranium) with HfCuSi2-Type Structure. J. Solid State Chem. 1995, 115, 305–308. [Google Scholar] [CrossRef]
  32. Stoyko, S.S.; Mar, A. Ternary rare-earth zinc arsenides REZn1–xAs2 (RE = La–Nd, Sm). J. Solid State Chem. 2011, 184, 2360–2367. [Google Scholar] [CrossRef]
  33. Salamakha, L.P.; Mudryi, S.I. Crystal structure of the RZn1−xSb2 compounds (R = La, Ce). J. Alloys Compd. 2003, 359, 139–142. [Google Scholar] [CrossRef]
  34. Zelinska, O.Y.; Mar, A. Structure and physical properties of rare-earth zinc antimonides REZn1–xSb2 (RE = La, Ce, Pr, Nd, Sm, Gd, Tb). J. Solid State Chem. 2006, 179, 3776–3783. [Google Scholar] [CrossRef]
  35. Lin, X.; Stoyko, S.S.; Mar, A. Ternary rare-earth zinc arsenides REZn2As3 (RE = La–Pr) containing defect fluorite-type slabs. J. Solid State Chem. 2013, 199, 189–195. [Google Scholar] [CrossRef]
  36. Seibel, E.M.; Schoop, L.M.; Xie, W.; Gibson, Q.D.; Webb, J.B.; Fuccillo, M.K.; Krizan, J.W.; Cava, R.J. Gold-Gold Bonding: The Key to Stabilizing the 19-Electron Ternary Phases LnAuSb (Ln = La–Nd and Sm). J. Am. Chem. Soc. 2015, 137, 1282–1289. [Google Scholar] [CrossRef] [PubMed]
  37. Ruszała, P.; Winiarski, M.J.; Samsel-Czekała, M. Dirac-like band structure of LaTESb2 (TE = Ni, Cu, and Pd) superconductors by DFT calculations. Comput. Mater. Sci. 2018, 154, 106–110. [Google Scholar] [CrossRef]
  38. Ren, W.; Wang, A.; Graf, D.; Liu, Y.; Zhang, Z.; Yin, W.-G.; Petrovic, C. Absence of Dirac states in BaZnBi2 induced by spin-orbit coupling. Phys. Rev. B 2018, 97, 035147. [Google Scholar] [CrossRef]
  39. Wang, Y.-Y.; Guo, P.-J.; Yu, Q.-H.; Xu, S.; Liu, K.; Xia, T.-L. Magneto-transport and electronic structures of BaZnBi2. New J. Phys. 2017, 19, 123044. [Google Scholar] [CrossRef]
  40. Ruszała, P.; Winiarski, M.J.; Samsel-Czekała, M. Dirac-Like Electronic-Band Dispersion of LaSb2 Superconductor and its Counterpart LaAgSb2. Acta Phys. Pol. A 2020, 138, 748–751. [Google Scholar] [CrossRef]
  41. Chamorro, J.R.; Topp, A.; Fang, Y.; Winiarski, M.J.; Ast, C.R.; Krivenkov, M.; Varykhalov, A.; Ramshaw, B.J.; Schoop, L.M.; McQueen, T.M. Dirac fermions and possible weak antilocalization in LaCuSb2. APL Mater. 2019, 7, 121108. [Google Scholar] [CrossRef]
  42. Seibel, E.M.; Xie, W.; Gibson, Q.D.; Cava, R.J. Structure and magnetic properties of the REAuBi2 (RE = La–Nd, Sm) phases. J. Solid State Chem. 2015, 230, 318–324. [Google Scholar] [CrossRef]
  43. Petrovic, C.; Bud’ko, S.L.; Strand, J.D.; Canfield, P.C. Anisotropic properties of rare earth silver dibismites. J. Magn. Magn. Mater. 2003, 261, 210–221. [Google Scholar] [CrossRef]
  44. Akiba, K.; Umeshita, N.; Kobayashi, T.C. Observation of superconductivity and its enhancement at the charge density wave critical point in LaAgSb2. Phys. Rev. B 2022, 106, L161113. [Google Scholar] [CrossRef]
  45. Uykur, E.; Maulana, L.Z.; Schoop, L.M.; Lotsch, B.V.; Dressel, M.; Pronin, A.V. Magneto-optical probe of the fully gapped Dirac band in ZrSiS. Phys. Rev. Res. 2019, 1, 032015. [Google Scholar] [CrossRef]
  46. Topp, A.; Queiroz, R.; Grüneis, A.; Müchler, L.; Rost, A.W.; Varykhalov, A.; Marchenko, D.; Krivenkov, M.; Rodolakis, F.; McChesney, J.L.; et al. Surface Floating 2D Bands in Layered Nonsymmorphic Semimetals: ZrSiS and Related Compounds. Phys. Rev. X 2017, 7, 041073. [Google Scholar] [CrossRef]
  47. Lee, G.; Farhan, M.A.; Kim, J.S.; Shim, J.H. Anisotropic Dirac electronic structures of AMnBi2 (A = Sr,Ca). Phys. Rev. B 2013, 87, 245104. [Google Scholar] [CrossRef]
  48. Klemenz, S.; Schoop, L.; Cano, J. Systematic study of stacked square nets: From Dirac fermions to material realizations. Phys. Rev. B 2020, 101, 165121. [Google Scholar] [CrossRef]
  49. Young, S.M.; Kane, C.L. Dirac Semimetals in Two Dimensions. Phys. Rev. Lett. 2015, 115, 126803. [Google Scholar] [CrossRef]
  50. Shao, D.-F.; Zhang, S.-H.; Dang, X.; Tsymbal, E.Y. Tunable two-dimensional Dirac nodal nets. Phys. Rev. B 2018, 98, 161104. [Google Scholar] [CrossRef]
  51. Fu, B.-B.; Yi, C.-J.; Zhang, T.-T.; Caputo, M.; Ma, J.-Z.; Gao, X.; Lv, B.Q.; Kong, L.-Y.; Huang, Y.-B.; Richard, P.; et al. Dirac nodal surfaces and nodal lines in ZrSiS. Sci. Adv. 2019, 5, eaau6459. [Google Scholar] [CrossRef]
  52. Rosmus, M.; Olszowska, N.; Bukowski, Z.; Starowicz, P.; Piekarz, P.; Ptok, A. Electronic Band Structure and Surface States in Dirac Semimetal LaAgSb2. Materials 2022, 15, 7168. [Google Scholar] [CrossRef]
  53. Murase, M.; Sasagawa, T. Large Magnetoresistance and Dirac Line Node in LaAgBi2. J. Phys. Soc. Jpn. 2020, 89, 055003. [Google Scholar] [CrossRef]
  54. Ekahana, S.A.; Wu, S.-C.; Jiang, J.; Okawa, K.; Prabhakaran, D.; Hwang, C.-C.; Mo, S.-K.; Sasagawa, T.; Felser, C.; Yan, B.; et al. Observation of nodal line in non-symmorphic topological semimetal InBi. New J. Phys. 2017, 19, 065007. [Google Scholar] [CrossRef]
  55. Regmi, S.; Dhakal, G.; Kabeer, F.C.; Harrison, N.; Kabir, F.; Sakhya, A.P.; Gofryk, K.; Kaczorowski, D.; Oppeneer, P.M.; Neupane, M. Observation of multiple nodal lines in SmSbTe. Phys. Rev. Mater. 2022, 6, L031201. [Google Scholar] [CrossRef]
  56. Feng, Y.; Jiang, Q.; Feng, B.; Yang, M.; Xu, T.; Liu, W.; Yang, X.; Arita, M.; Schwier, E.F.; Shimada, K.; et al. Rashba-like spin splitting along three momentum directions in trigonal layered PtBi2. Nat. Commun. 2019, 10, 4765. [Google Scholar] [CrossRef] [PubMed]
  57. Acosta, C.M.; Fazzio, A.; Dalpian, G.M. Zeeman-type spin splitting in nonmagnetic three-dimensional compounds. NPJ Quantum Mater. 2019, 4, 41. [Google Scholar] [CrossRef]
  58. Feng, Y.; Wang, Z.; Chen, C.; Shi, Y.; Xie, Z.; Yi, H.; Liang, A.; He, S.; He, J.; Peng, Y.; et al. Strong Anisotropy of Dirac Cones in SrMnBi2 and CaMnBi2 Revealed by Angle-Resolved Photoemission Spectroscopy. Sci. Rep. 2014, 4, 5385. [Google Scholar] [CrossRef] [PubMed]
  59. Bosak, A.; Souliou, S.-M.; Faugeras, C.; Heid, R.; Molas, M.R.; Chen, R.-Y.; Wang, N.-L.; Potemski, M.; Tacon, M.L. Evidence for nesting-driven charge density wave instabilities in the quasi-two-dimensional material LaAgSb2. Phys. Rev. Res. 2021, 3, 033020. [Google Scholar] [CrossRef]
  60. Thirupathaiah, S.; Efremov, D.; Kushnirenko, Y.; Haubold, E.; Kim, T.K.; Pienning, B.R.; Morozov, I.; Aswartham, S.; Buchner, B.; Borisenko, S.V. Absence of Dirac fermions in layered BaZnBi2. Phys. Rev. Mater. 2019, 3, 024202. [Google Scholar] [CrossRef]
  61. Ryu, H.; Park, S.Y.; Li, L.; Ren, W.; Neaton, J.B.; Petrovic, C.; Hwang, C.; Mo, S.-K. Anisotropic Dirac Fermions in BaMnBi2 and BaZnBi2. Sci. Rep. 2018, 8, 15322. [Google Scholar] [CrossRef]
Figure 1. (a) Unit cells used for calculations. Total and partial densities of states calculated for (b) LaZnBi 2 and (c) LaZn0.5Bi2 in the fully relativistic (with SOC) GGA approach.
Figure 1. (a) Unit cells used for calculations. Total and partial densities of states calculated for (b) LaZnBi 2 and (c) LaZn0.5Bi2 in the fully relativistic (with SOC) GGA approach.
Crystals 14 00209 g001
Figure 2. The BZ schemes and electronic band structures obtained (GGA) with SOC. (a) The BZ of the considered systems, with high-symmetry lines denoted in blue. (b) The same BZ but with different lines and paths highlighted (see main text for a detailed explanation). (c,d) Electronic band structures of LaZn1−xBi2 compounds for x = 0.5 and x = 0 , respectively. (e,f) Electronic band structures of template square-net Dirac semimetals ZrSiS and LaAgSb 2 , respectively. In all electronic structure plots, the first (lowest) band that crosses E F is indicated in blue; the subsequent bands are indicated in red, green, and orange (highest band), respectively.
Figure 2. The BZ schemes and electronic band structures obtained (GGA) with SOC. (a) The BZ of the considered systems, with high-symmetry lines denoted in blue. (b) The same BZ but with different lines and paths highlighted (see main text for a detailed explanation). (c,d) Electronic band structures of LaZn1−xBi2 compounds for x = 0.5 and x = 0 , respectively. (e,f) Electronic band structures of template square-net Dirac semimetals ZrSiS and LaAgSb 2 , respectively. In all electronic structure plots, the first (lowest) band that crosses E F is indicated in blue; the subsequent bands are indicated in red, green, and orange (highest band), respectively.
Crystals 14 00209 g002
Figure 3. Detailed studies of band dispersion along the X–R line for LaZn1−xBi2. (a,b) Electronic band structure plotted along a closed path on the BZ boundary for LaZn0.5Bi2 and LaZnBi 2 , respectively. (c) Characteristic lines marked in blue on the BZ scheme, corresponding to the locus of respective Dirac-like band crossings. (d,e) Electronic bands that cross E F along the X–R line, denoted in (a) by a red dashed ellipse. D X ( D R ) denotes the point that lies closer to the X (R) point in the BZ. (f,g) Quasi-2D Dirac-like band dispersion around the D XR point, plotted along two perpendicular directions in the BZ for LaZnBi 2 , denoted by a red dashed ellipse in (b).
Figure 3. Detailed studies of band dispersion along the X–R line for LaZn1−xBi2. (a,b) Electronic band structure plotted along a closed path on the BZ boundary for LaZn0.5Bi2 and LaZnBi 2 , respectively. (c) Characteristic lines marked in blue on the BZ scheme, corresponding to the locus of respective Dirac-like band crossings. (d,e) Electronic bands that cross E F along the X–R line, denoted in (a) by a red dashed ellipse. D X ( D R ) denotes the point that lies closer to the X (R) point in the BZ. (f,g) Quasi-2D Dirac-like band dispersion around the D XR point, plotted along two perpendicular directions in the BZ for LaZnBi 2 , denoted by a red dashed ellipse in (b).
Crystals 14 00209 g003
Figure 4. Detailed studies of band dispersion near the D ZR Dirac point along the Z–R line for (a,c,d) three different perpendicular directions for LaZnBi 2 . (b) The BZ, marked with the respective points and lines we used to plot the graphs in (a,c,d).
Figure 4. Detailed studies of band dispersion near the D ZR Dirac point along the Z–R line for (a,c,d) three different perpendicular directions for LaZnBi 2 . (b) The BZ, marked with the respective points and lines we used to plot the graphs in (a,c,d).
Crystals 14 00209 g004
Figure 5. Fermi surface pieces calculated in the GGA mode for (a) LaZnBi 2 without SOC, (b) LaZnBi 2 with SOC, (c) LaZn0.5Bi2 without SOC, and (d) LaZn0.5Bi2 with SOC. It is clearly visible that a lack of inversion symmetry eliminates the Kramers degeneracy in (d), compared to the centrosymmetric case (b). For a detailed description of FSs, please see the main text.
Figure 5. Fermi surface pieces calculated in the GGA mode for (a) LaZnBi 2 without SOC, (b) LaZnBi 2 with SOC, (c) LaZn0.5Bi2 without SOC, and (d) LaZn0.5Bi2 with SOC. It is clearly visible that a lack of inversion symmetry eliminates the Kramers degeneracy in (d), compared to the centrosymmetric case (b). For a detailed description of FSs, please see the main text.
Crystals 14 00209 g005
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Ruszała, P.; Winiarski, M.J.; Samsel-Czekała, M. LaZn1−xBi2 as a Candidate for Dirac Nodal-Line Intermetallic Systems. Crystals 2024, 14, 209. https://doi.org/10.3390/cryst14030209

AMA Style

Ruszała P, Winiarski MJ, Samsel-Czekała M. LaZn1−xBi2 as a Candidate for Dirac Nodal-Line Intermetallic Systems. Crystals. 2024; 14(3):209. https://doi.org/10.3390/cryst14030209

Chicago/Turabian Style

Ruszała, Piotr, Maciej J. Winiarski, and Małgorzata Samsel-Czekała. 2024. "LaZn1−xBi2 as a Candidate for Dirac Nodal-Line Intermetallic Systems" Crystals 14, no. 3: 209. https://doi.org/10.3390/cryst14030209

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop