Introduction

Yttrium monoantimonide has mainly been used as a non-magnetic reference or as a ‘solvent’ in monoantimonides of f-electron-elements solid solutions with anomalous physical properties such as dense Kondo behavior and complex magnetic ground-states1. Within that context it has been characterized as a metal by low-temperature specific heat measurements2. Later Hayashi et al. have shown that the first-order phase transition from the NaCl-type to a CsCl-type crystal structure occurs in YSb at 26 GPa3. That discovery induced numerous calculations of electronic structure of the compound, among them those by Tütüncü, Bagci and Srivastava, who directly compared electronic structure of YSb with that of LaSb4. Results of those calculations were very similar for both compounds, revealing low densities of states at Fermi level and characteristic anti-crossings leading to band inversion at X-points of the Brillouin zone. LaSb has a simple NaCl-type structure without broken inversion symmetry, perfect linear band crossing or perfect electron–hole symmetry, yet it exhibits the exotic magnetotransport properties of complex-structure semimetals like TaAs, NbP (Weyl semimetals)5,6, Cd3As2 (Dirac semimetal)7 and WTe2 (resonant compensated semimetal)8,9. Recently Tafti et al. discovered in LaSb field-induced resistivity plateau at low temperatures up to ≈15 K, ultrahigh mobility of carriers in the plateau region, quantum oscillations, and magnetoresistance (MR) of nearly one million percent at 9 T10. Their calculations, including spin-orbit coupling (SOC) effect, suggested that LaSb is a topological insulator with a 10 meV gap open near the X-point of the Brillouin zone. They also observed specific angular dependence of frequencies of quantum oscillations and ascribed them to two-dimensional Fermi surface (FS) possibly formed of topologically nontrivial states, and thus proposed LaSb as a model system for understanding the consequences of breaking time-reversal symmetry in topological semimetals10.

However, such angular dependence has already been observed in LaSb by de Haas-Van Alphen measurements and well explained by the presence of elongated pockets of 3D-Fermi surface11,12.

Tafti et al. also invoked the opening of insulating gap (i.e. metal-insulator transition) as a source of the field-induced resistivity plateau in LaSb10 but it should be noted that in the case of WTe2 the existence of a magnetic-field-driven metal-insulator transition has been excluded by means of Kohler scaling analysis of magnetoresistance13.

Motivated by these ambiguities in the interpretation of LaSb properties we decided to carry out a comprehensive characterization of magnetotransport properties of a sister compound YSb. We found that YSb displays physical properties in many aspects very similar to those of LaSb. Our results are in accord with those of other groups that appeared during preparation of our article14,15. The interpretation proposed by Ghimire et al. and Yu et al. follows that presented by Tafti et al. for LaSb10, implying the role of field-induced metal-insulator transition in YSb.

However, our analysis of magnetoresistance and Shubnikov–de Haas (SdH) effect provides strong support for 3D-Fermi surface scenario of magnetotransport, without necessity of invoking topologically non-trivial 2D states or metal-insulator transition in YSb.

Results

Magnetoresistance and the origin of its plateau

Electrical resistivity (ρ) was measured on two samples (denoted as #1 and #2) and its dependence on temperature in zero field is plotted in Supplementary Figure 1(a). Shape of ρ(T) curves is typical for a metal. When measured in different applied fields ρ(T) exhibits universal plateau at temperatures 2–15 K, as shown for sample #1 in Supplementary Figure 1(b). Temperature range of this plateau is very similar to that reported for LaSb10.

Magnetoresistance, MR ≡ [ρ(B) − ρ(B = 0)]/ρ(B = 0), is plotted versus magnetic field, B, in Fig. 1(a) and (b), for samples #1 and #2, respectively. Following the approach of Wang et al.13 we performed Kohler scaling analysis of magnetoresistance to test for the existence of a magnetic-field-driven metal-insulator transition in YSb. Figure 1(c) and (d) show very good Kohler scaling of our data, MR ∝ (B/ρ0)m, yielding at 2.5 K exponents m = 1.64 and 1.74, respectively, very close for both samples, despite significant difference of their MR values. Efficiency of this scaling indicates that resistivity plateau is due purely to the magnetoresistance, but not to a field-induced metal-insulator transition.

Figure 1
figure 1

(a) and (b) Magnetoresistance of samples #1 and #2, respectively, versus strength of applied magnetic field at different temperatures. (c) and (d) Kohler scaling of magnetoresistance, MR ∝ (B/ρ0)m fitted to 2.5 K data yields m = 1.64 and 1.74, for samples #1 and #2, respectively.

In order to elucidate the dimensionality of FS we measured magnetoresistance of the sample #2 in fields applied at different angles to its surface, Fig. 2(a). Here θ = 0° denotes the field perpendicular to the sample surface and the current direction, whereas θ = 90° means that the field is parallel to the current. As shown in the inset to Fig. 2(a), ρ at strongest field of 9 T follows . This is typical behavior for materials without magnetic anisotropy, but the change of ρ expressed as anisotropic magnetoresistance, AMR ≡ [ρ(90°) − ρ(0°)]/ρ(0°), has an outstanding −80% value.

Figure 2
figure 2

(a) Resistivity of YSb (sample #2) at 2.5 K versus strength of magnetic field applied at different angles θ. Inset: resistivity at 2.5 K and in 9 T versus field rotation angle; blue line represents dependence. (b) Data of (a) replotted with B scaled by angle-dependent factor εθ. Inset: angle dependence of εθ; red line represents fit with εθ = (cos2θ + γ−2sin2θ)1/2 function yielding mass anisotropy γ = 3.4.

Moreover, when field strength is scaled by a factor εθ dependent on mass anisotropy and θ-angle, all ρ(T) data of Fig. 2(a) collapse on single curve, as shown in Fig. 2(b). Inset of Fig. 2(b) shows that values of εθ plotted against field angle θ can be perfectly fitted with εθ = (cos2θ + γ−2sin2θ)1/2 function, shown with red line. Such scaling has initially been proposed for anisotropic superconductors16, and recently used to interpret MR behavior of WTe2 based on its 3D electronic nature17. Parameter γ represents effective mass anisotropy of carriers mostly contributing to the magnetoresistance. We ascribe this behavior to a strongly anisotropic sheet of 3D-FS revealed by SdH effect data, as shown below.

Hall effect

Hall resistivity of sample #1h (cut from the same single crystal as #1) measured at several temperatures between 2.2 K and 300 K is shown in Fig. 3(a). Nonlinear ρxy(B) indicates that at least two types of charge carriers are responsible for the Hall effect observed in YSb. The ρxy(B) curves for temperatures from 2.2–15 K range are almost identical, which points to nearly constant carrier concentrations and mobilities, and coincides with the plateau of ρxx(T) observed in the same range of T. Changes of sign of ρxy(B) observed at T ≤ 100 K, indicate conducting bands of both electrons and holes, at higher temperatures ρxy(B) is positive in the whole range of magnetic field (0 < B ≤ 9T). Thus, the Hall contributions of holes and electrons nearly compensate, but both depend on temperature in different manner. Clear quantum oscillations are observed in ρxy in temperature range 2.2–15 K (cf. Fig. 3(a) and (b)).

Figure 3
figure 3

(a) Hall resistivity of YSb (sample #1h) versus magnetic field recorded at several temperatures. (b) Magnetic field dependent Hall conductivity at T = 2.2 K (left axis) and 300 K (right axis). Lines represent the fits with multiple-band model (Eq. 1).

Since ρxy ≪ ρxx, the off-diagonal component of conductivity tensor should be used for multiple-band analysis of Hall data. In this case simple Drude model can be used:

summing up conductivities of individual bands, with ni and μi denoting respectively concentration and mobility of carriers from the i-th band18. As Fig. 3(b) shows, for data collected at 300 K accounting for two bands yielded a good fit. On the other hand, fitting with two bands was insufficient for 2.2 K data, but addition of a contribution of another band with small concentration of more mobile holes brought a very satisfactory fit.

For LaSb Tafti et al.10 estimated uncompensated carrier concentration, n, and the Hall mobility mH, using the relations n = 1/eRH(0) (with RH(0) being the zero-temperature limit of RH(T)) and μH = RH(0)/ρ0. They obtained n ≈ 1020 cm−3 and μH ≈ 105 cm2/(Vs).

For our YSb sample the fit with with multiple-band model (Eq. 1), shown in Fig. 3(b), yielded at T = 2.2 K the concentrations: ne = 1.52 × 1020 cm−3, nh = 1.16 × 1020 cm−3, and mobilities μe = 2.7 × 103 cm2/(Vs), μh = 1.9 × 103 cm2/(Vs). Third band necessary for that fit consists of more mobile holes [n = 3.4 × 1019 cm−3, μ = 7.9 × 103 cm2/(Vs)]. These results are consistent with band calculations presented below.

The fit to data collected at 300 K yielded: the concentrations: ne = 1.34 × 1018 cm−3, nh = 1.94 × 1019 cm−3, and mobilities μe = 1.8 × 103 cm2/(Vs), μh = 8.2 × 102 cm2/(Vs), so YSb has similar concentration of carriers but with significantly lower mobility than LaSb. This seems to be the main reason for its significantly smaller magnetoresistance, as it was well demonstrated for WTe29.

Overall, Hall effect results are in perfect agreement with characteristics of Fermi surface presented in the next section.

Fermi surface analysis: electronic structure calculations and Shubnikov-de Haas effect

We performed the electronic structure calculations for YSb using a full potential all-electron local orbital code (FPLO) within GGA approximation. Figure 4 shows obtained energy band structure for YSb, with a few bands crossing Fermi level. Near X-point there is an anti-crossing present with a gap of ≈0.8 meV, similar to those reported for lanthanum monopnictides, which led to the proposal of 2D topologically non-trivial states at the origin of extraordinary behavior of their magnetoresistance10,19. Fermi surface resulting from our calculations is presented in Fig. 5: two electron sheets centered at X-points and three nested hole sheets centered at Γ-point. Our calculations are in good agreement with those of ref. 14.

Figure 4
figure 4

Relativistic energy band structure for YSb calculated using FPLO approach with GGA approximation.

Bands crossing Fermi level are marked in colors: electron band α in red, hole bands: β and δ in blue, ζ (not observed in SdH oscillations) in green. Vicinity of X-point with an anti-crossing and opened gap is shown as blow-up in inset.

Figure 5
figure 5

Fermi surface of YSb from electronic band calculations.

Green lines indicate cyclotron orbits (extreme cross-sections of FS-sheets α, β, δ and α1), for which we observed SdH oscillations in fields applied at θ = 0° (cf. Fig. 7 and Table 1).

Shubnikov–de Haas oscillations of resistivity are discernible for YSb at temperatures up to 15 K and in fields above 6 T, as seen in Figs 1(a), 1(b) and 2(a). Since ρxy ≪ ρxx, we may safely assume that the conductivity and analyze directly oscillations of ρxx. Points in Fig. 6 represent resistivity of sample #2 measured at 2.5 K, after subtraction of smooth background, plotted versus inverse field 1/B (for 7 < B < 9T). Complex shape of Δρxx(1/B) dependence indicates that observed SdH oscillation has several components. Indeed fast Fourier transform (FFT) analysis reveals clearly six well separated frequencies (Fig. 7(a)).

Figure 6
figure 6

Shubnikov-de Haas oscillations of resistivity at T = 2.5 K.

Blue solid line represents the fit with multiple Lifshitz-Kosevich function (Eq. 2).

Figure 7
figure 7

(a) FFT-frequency spectrum of ρxx(1/B) oscillations of sample #2 measured at different temperatures. Diamond symbols indicate normalized amplitudes obtained from fitting with multiple Lifshitz-Kosevich function shown in Fig. 6. Black solid line (for the sake of clarity offset by 5 units) represents the spectrum of ρxy(1/B) oscillations of sample #1h at 2.5 K. (b) Frequencies of two strongest FFT components of SdH oscillations shown in Fig. 2(a): α (circles) and 2α (squares), as well as 3α (triangles), versus sample rotation angle θ. Dashed lines represent , solid line corresponds to f = f0(sin2θ + r2cos2θ)−1/2 with r = 3.6. This line is redrawn for θ + π/2, reflecting the symmetry of cubic YSb lattice.

We chose to fit Δρxx(1/B) with the multi-frequency Lifshitz-Kosevich function20,21,22 (Eq. 2) because, as the maxima of total Δρxx do not correspond to the maxima of particular components with different frequencies, it is inadequate to determine phases in a multicomponent SdH oscillation using the so-called Landau-level fan diagram (plot of the values of 1/BN corresponding to the N-th maximum in Δρxx versus N).

where for each i-th component: fi is the frequency, represents the Dingle factor RD and comprises the temperature reduction factor RT and the spin factor RS. Detailed form of Eq. 2, taking into account harmonic components, and description of RD, RT and RS is presented in Supplementary Material.

The fit including six components was of very good quality, as shown by blue line in Fig. 6. Obtained parameters are collected in Table 1. All frequencies converged almost exactly to values obtained from FFT analysis. We ascribe oscillations with frequencies of 720 and 1072 T to second and third harmonic of the strongest one with fi = 360 T. The phases φi of fundamental oscillations resulting from the fit are close to Onsager phase factor of 1/2 expected for free electrons. Thus all components of SdH oscillation can be assigned to 3D-FS sheets predicted by band calculations (shown above in Fig. 5) and no Berry phase of π was observed, which could reveal topologically non-trivial charge carriers.

Table 1 Parameters obtained from fit of the multiple Lifshitz-Kosevich function (Eq. 2) to data-points shown in Fig. 6.

We performed FFT analysis for all data sets presented in Fig. 2(a), which allowed us to observe angular behavior of frequencies corresponding to all extreme cross-sections of Fermi sheets and compare them to those derived from our band structure calculations, as shown in Fig. 7(a). Three of observed frequencies were clearly changing upon rotation of the magnetic field: the principal, labeled as α (T at θ = 0°), and its harmonics, 2α and 3α. These frequencies are plotted versus θ in Fig. 7(b).

It became apparent from the shape of FS obtained from band calculations (cf. Fig. 5) that angular behavior of follows a cross-section area of a prolate ellipsoid (which well approximates the shape of electron sheet centered at X-point shown in Fig. 5) by the ((1 − cos θ) 0 1) plane passing by X-point. When magnetic field is applied along [0 0 1] direction the plane is perpendicular to Γ-X line. When field is tilted from [0 0 1] towards [1 0 0] by the angle θ the cross-section increases as . This holds for a two-axial ellipsoid described by the equation: (x/kx)2 + (y/kx)2 + (z/kz)2 = 1, with r = kz/kx. Band structure shown in Fig. 4 yields r ≈ 3.6 (with kx estimated as average size of α-sheet of FS along X-U and X-W lines, and kz as its size along Γ-X line).

After rotation by θ = 90° the α frequency meets the one denoted as α1, initially (i.e. at θ = 0°) corresponding to the largest cross-section of the same ellipsoidal FS sheet. Two other observed frequencies, β and δ do not change notably with θ, as expected for almost isotropic hole Fermi sheets centered at Γ-point.

We plotted S(θ) (for r = 3.6) in Fig. 7(b) as solid lines. The dependence, expected for two-dimensional FS, behaves similarly and is shown for comparison with dashed lines. The dependence was used by Tafti et al.10 as a hint of possible topologically-nontrivial states in LaSb. However, Fermi surface of that compound has already been well characterized by band calculations and angle-dependent de Haas–van Alphen measurements11,12, revealing FS very similar to the one we found in YSb, namely consisting of one ellipsoidal electron sheet centered at X-point and two isotropic hole pockets centered at Γ-point. Hasegawa11 and Settai et al.12 assigned angular dependence of the principal de Haas–van Alphen frequency (identical to the SdH frequency in ref. 10) to the cross-section of the ellipsoidal sheet S(θ) in accord with our interpretation of S(θ) behavior for YSb. This underscores the similarities between these two compounds and implies that there is no need to invoke topologically non-trivial states to explain exotic magnetotransport properties neither in YSb nor in LaSb (contrary to ref. 10).

Discussion and Conclusions

YSb is another material displaying giant magnetoresistance (1100% in 9 T), three orders of magnitude smaller than that of sister compound LaSb10, thus it cannot be termed ‘extreme magnetoresistance’ (XMR). This is due mainly to its lower carrier mobility and weaker electron-hole compensation revealed by our Hall effect measurement.

Kohler scaling analogous to that shown in Fig. 1(c) and (d) has recently been used to explain the remarkable up-turn behavior of MR in WTe2 without the field-induced metal-insulator transition or significant contribution of an electronic structure change13. The same authors have shown that perfect carrier compensation leads to exponent m = 2 in this scaling. Kohler scaling for YSb yielded for our samples the exponents m = 1.64 and 1.74, which seems related to weaker carrier compensation than nearly perfect one in WTe2, where m = 1.9213. Thus, analogously to WTe2, Kohler scaling indicates that the field-induced metal-insulator transition is unnecessary to explain up-turn and low-temperature plateau of resistivity in YSb. The origin of the up-turn is a combination of magnetoresistance with the low-temperature resistivity plateau present already at zero field. Given the similarity of YSb and LaSb the same may also be true for the latter compound.

Comparing results of SdH measurements with those of electronic structure calculations we obtained comprehensive description of the Fermi surface of YSb. Presence of both electron and hole sheets of similar volumes provides partial charge compensation responsible for its strong magnetoresistance. Band structures of YSb and LaSb are very similar. All Fermi sheets in YSb but the smallest one centered at Γ-point have their counterparts in LaSb10,11,12. Our analysis of angular behavior of SdH frequencies in YSb indicates it is related to the three-dimensional FS, in line with Hasegawa11 and Settai et al.12 findings for LaSb, but not connected to possible non-trivial topology of electronic structure analogous to that suggested by Tafti et al.10

Angular behavior of MR can also be perfectly explained by anisotropy of 3D-FS. When field strength is scaled by the angle-dependent factor εθ, all data of Fig. 2(a) collapse on single curve. The effective mass anisotropy factor γ = 3.4, obtained from the fit of εθ(θ) with the expression , is in excellent agreement with kz/kx = 3.6 we estimated for α-sheet of FS. This is not surprising, since the mass anisotropy directly reflects the shape of FS, but it shows that angular behavior of MR in YSb is mainly governed by anisotropic form of α-sheet of FS. That sheet corresponds to the electron band, all other FS-sheets contain holes and are nearly isotropic. The effective mass and mobility of α-sheet electrons change significantly with field angle, which strongly modifies the magnetoresistance.

It has been proposed that the magnetic field induces the reconstruction of the FS in a Dirac semimetal by breaking the time reversal invariance23,24,25,26. Assisted by the high mobility of carriers such reconstruction has been suggested to induce very large MR observed in Cd3As2 and NbSb227,28. We also observe features in the electronic structure of YSb, buried under the Fermi level, which may possibly allow the magnetic field to transform this compound into Dirac semimetal. A small gap between inverted bands near the X-point (cf. inset to Fig. 4) might result in topologically non-trivial states. The effect of FS reconstruction could be similar to temperature-induced Lifshitz transition in WTe229, whereas its mechanism might be related, for example, to that of Lifshitz transition driven by magnetic field in CeIrIn530. Very recently Dirac states have been observed by angle-resolved-photoemission spectroscopy in NbSb, a compound with bulk electronic structure very similar to that of YSb31, however topologically protected states were not detected in YSb by this method32.

Although a small contribution of topologically non-trivial 2D states cannot be completely excluded our analysis of magnetoresistance and Shubnikov–de Haas effect provides strong support for 3D-Fermi surface scenario of magnetotransport properties in YSb. Analogous field-induced properties of LaSb can most probably be also described in the framework of 3D multiband model.

Methods

Measurements were performed using a Physical Property Measurement System (Quantum Design) on two samples cut from one single crystal and labeled as #1 and #1 h, and a sample cut from another single crystal and labeled #2. All samples had shapes of rectangular cuboid with all edges along 〈1 0 0〉 crystallographic directions. Their sizes were: 0.56 × 0.25 × 0.12 mm3, 0.4 × 0.47 × 0.13 mm3 and 0.41 × 0.32 × 0.09 mm3, for samples #1, #1 h and #2, respectively. The electric current was always flowing along [1 0 0] crystallographic direction. Single crystals were grown from Sb flux and their NaCl-type crystal structure was confirmed by powder X-ray diffraction carried out using an X’pert Pro (PANanalytical) diffractometer with Cu-Kα radiation. No other phases were detected and lattice parameter of 6.163 Å was determined, reasonably close to literature value 6.155 Å 3. Electronic structure calculations were carried out using FPLO-9.00-34 code within generalized gradient approximation (GGA) method33. The full-relativistic Dirac equation was solved self-consistently, treating exactly all relativistic effects, including the spin-orbit interaction without any approximations. The Perdew–Burke–Ernzerhof exchange-correlation potential34 was applied and the energies were converged on a dense k mesh with 243 points. The convergence was set to both the density (10−6 in code specific units) and the total energy (10−8 Hartree). For the Fermi surface a 643 mesh was used to ensure accurate determination of the Fermi level.

Additional Information

How to cite this article: Pavlosiuk, O. et al. Giant magnetoresistance, three-dimensional Fermi surface and origin of resistivity plateau in YSb semimetal. Sci. Rep. 6, 38691; doi: 10.1038/srep38691 (2016).

Publisher's note: Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.