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

Electrical transport and magnetic properties of the triangular-lattice compound Zr2NiP2

Zongtang Wan, Yuqian Zhao, Jiabin Liu Yuesheng Li [email protected] Wuhan National High Magnetic Field Center and School of Physics, Huazhong University of Science and Technology, 430074 Wuhan, China
Abstract

We report the first investigation of the electrical and magnetic properties of the triangular-lattice compound Zr2NiP2 (space group PP63/mmcmmc). The temperature evolution of electrical resistivity follows the Bloch-Grüneisen-Mott law, and exhibits a typically metallic behavior. No transition is visible by both electrical and magnetic property measurements, and nearly no magnetization is detected (M0M_{0} << 0.002μB\mu_{\mathrm{B}}/Ni) down to 1.8 K up to 7 T. The metallic and nonmagnetic characters are well understood by the first-principles calculations for Zr2NiP2.

keywords:
Zr2NiP2, triangular lattice, metallic behavior, non-magnetism
journal: Chemical Physics Letters

1 Introduction

Quasi-two-dimensional materials with exotic properties have fascinated condensed matter physicists for decades. Among others, the prototype of a resonating-valence-bond quantum spin liquid state had been proposed by P. W. Anderson on the SS = 1/2 triangular-lattice Heisenberg antiferromagnetic model in 1973 [1]. The experimental realizations of SS = 1/2 triangular-lattice spin-liquid states had been extensively reported, e.g., in κ\kappa-(BEDT-TTF)2Cu2(CN)3 [2, 3], EtMe3Sb[Pd(dmit)2]2 [4], Ba3CuSb2O9 [5], YbMgGaO4 [6, 7], and NaYbCh2 (Ch = O, S, Se) [8, 9]. Recently, SS = 1 triangular-lattice spin-liquid ground states were also proposed [10, 11, 12, 13], and might be materialized by the Ni-based compounds, including NiGa2S4 [14, 15] and Ba3NiSb2O9 [16]. Besides this, unconventional high-temperature superconductors had been discovered in many quasi-two-dimensional compounds, including Cu-based [17, 18] and Fe-based [19] systems. It has been proposed for a long time that the high-temperature superconductivity can be considered as evolving from a doped spin-liquid state [20, 21]. The “smoking gun” evidence for the intimate relationship between the high-temperature superconductivity and spin liquid remain absent, and most of the existing studies on spin liquid are restricted in the quasi-two-dimensional geometrically frustrated Mott insulators. Therefore, it is intriguing to explore the physical properties of new Cu-, Ni- or Fe-based metals with quasi-two-dimensional geometrically frustrated lattices, e.g., the triangular lattice.

The Ni-based triangular-lattice compound Zr2NiP2 belongs to the family of the ternary pnictides Ln2NiX2 (Ln = Tb, Dy, Ho, Er and Zr; X = P, As), with the hexagonal modification (PP63/mmcmmc space group) [22]. Although these compounds have been synthesized and structurally characterized long time ago [22], other physical properties, e.g., electrical conductivity and magnetism, remain completely unknown. In this work, we conducted the electrical transport and magnetic property characterizations of Zr2NiP2. As shown in Figure 1, Ni atoms form a geometrically perfect triangular lattice (intralayer Ni-Ni distance of \sim 3.77 Å), and are spatially separated by the double layers of ZrP6 octahedra (interlayer Ni-Ni distance of \sim 6.44 Å). The temperature dependence of electrical resistivity is reasonably described by the Bloch-Grüneisen-Mott (BGM) law, and indicates the intrinsic metallic property of Zr2NiP2. Neither structural nor magnetic transition is observed by electrical resistivity and magnetic susceptibility down to 1.8 K. The low-TT magnetization is measured to be tiny, M0M_{0} \leq 0.002μB\mu_{\mathrm{B}}/Ni, up to 7 T. Ni should be in the nonmagnetic state, which accounts for the weak magnetism observed in Zr2NiP2.

2 Materials and methods

The polycrystalline samples of Zr2NiP2 were synthesized from stoichiometric mixtures of zirconium (99.5%, aladdin), nickel (99.8%,alfa), and red phosphorus (99.999%, aladdin). To prevent oxidation, the starting materials were mixed and pressed into pellets in an argon atmosphere glove box. The pellets wrapped in platinum foil were transferred into a quartz tube, which were sealed under vacuum (inner pressure of \sim 2×\times10-9 bar). The sealed tube was heated to 500 C at a rate of 1 C/min, and its temperature was maintained for three days. After that, the tube was further heated to 1000 C for seven days, and then cooled down to room temperature at a rate of \sim -1 C/min. Air-insensitive and polycrystalline black Zr2NiP2 samples were obtained, as shown in Figure 1(c).

The phase purity of the sample was checked by powder x-ray diffraction (XRD, Rigaku) measurements with Cu KαK_{\alpha} radiation (λ\lambda = 1.5418 Å), and a small weight fraction of nonmagnetic impurity ZrO2, 3.2±\pm0.1%, was found from Rietveld refinement (see Figure 1(c)), possibly due to the short exposure to air during the synthesis process. The reaction between the mixtures and the quartz tube was prevented by the platinum foil. The General Structure Analysis System (GSAS) program was used for the Rietveld refinements [23]. The ratio of Zr:Ni:P was semi-quantitatively determined by a scanning electron microscope equipped with a x-ray energy dispersive spectrometer (SEM–EDX, JEOL) to be \sim (2.0±\pm0.3):1:(1.9±\pm0.3), consistent with the ideal ratio of 2:1:2 within the resolution. The electrical resistivity down to 2 K was measured by using the standard four-probe ac technique on two cuboid-shaped as-grown samples of Zr2NiP2 (S1 and S2) in a physical property measurement system (PPMS, Quantum Design) at both 0 and 9 T. The dc magnetization measurements were carried out in a magnetic property measurement system (MPMS, Quantum Design) down to 1.8 K up to 7 T, by using the as-grown samples of \sim 40 mg. As the magnetization signal of the title compound is very weak, we carefully chose a quartz holder without evident background in the magnetic property measurements.

Refer to caption
Figure 1: (a) Crystal structure of Zr2NiP2. NiP3 triangles (red) and ZrP6 octahedra (olive) are shown, and the dashed black lines mark the unit cell. (b) The regular triangular lattice of Ni atoms. The interlayer distance of Ni-Ni is cc/2 \sim 6.443 Å, much larger than the intralayer one aa \sim 3.772 Å. (c) X-ray powder diffraction measured at 300 K and Rietveld refinement for Zr2NiP2. The inset shows a typical as-grown sample of Zr2NiP2.

We conducted first-principles calculations for Zr2NiP2 based on the density functional theory (DFT) with the generalized gradient approximation in the Vienna Ab initio Simulation Package (VASP) [24, 25, 26]. The crystal structure experimentally determined by XRD was used as the starting one, and the kinetic energy cutoff for the plane waves was specified to be 330 eV. The 6×\times6×\times6 kk-mesh was used, and the lattice parameters were fixed to their experimental values. The residual forces were below 0.003 eV/Å and the total energy was converged to 10-8 eV in the fully optimized structure with zz(Zr) = 0.10995. We performed both DFT simulations with Coulomb repulsions UU = 0 and 5 eV on Ni sites, but no essential difference was found. Furthermore, we also relaxed the lattice parameters simultaneously, and obtained aa = 3.775 Å, cc = 13.07 Å, and zz(Zr) = 0.10996, reasonably consistent with the experimental result (see Table 1).

3 Results and discussion

3.1 Crystal structure

Table 1: Crystal structure of the Zr2NiP2 phase determined by XRD at 300 K. The space group is PP63/mmcmmc (No.194) with aa = 3.7717(1) Å and cc = 12.8860(4) Å, and the final Rietveld refinement results in RpR_{\mathrm{p}} = 2.46% and RwpR_{\mathrm{wp}} = 3.99%. The weight fraction of the impurity phase ZrO2 (space group PP 21/cc) is refined to be, 3.2±\pm0.1%. Due to the low intensities of the reflections of the impurity phase (Figure 1(c)), we only refined the phase fraction and lattice parameters of ZrO2, but fixed other atom parameters as reported in reference [27]. The resulted aa(ZrO2) = 5.143(2) Å, bb(ZrO2) = 5.285(3) Å, cc(ZrO2) = 5.373(3) Å, and β\beta(ZrO2) = 99.17(2) are reasonably consistent with the reported result of ZrO2 [27].
Atoms Wyckoff position xx yy zz Occ. 100×Uiso\times U_{\mathrm{iso}}2)
Zr 4f 1/3 2/3 0.10995(4) 1 1.10(2)
Ni 2b 0 0 1/4 1 2.34(5)
P1 2a 0 0 0 1 1.22(7)
P2 2c 1/3 2/3 3/4 1 1.75(6)

Figure 1(c) shows the powder XRD pattern measured at 300 K and the final Rietveld refinement. Except the possible nonmagnetic ZrO2 (weight fraction \sim 3%), no other impurity phases are visible in the samples of Zr2NiP2. The grain size is estimated to be larger than 100 nm, according to the narrow full width at half maximum \sim 0.09 111The instrumental resolution is about 0.05. observed at 2θ\theta = 34.5 (see Figure 1(c)) [23]. The detailed crystallographic parameters and atomic positions are listed in Table 1. The refinement result suggests that the title compound has a hexagonal structure and the PP63/mmcmmc space group with two formulae per unit cell, as previously reported in reference [22]. Moreover, our refined crystal structure shows quantitative agreement with the previous result [22]. Ni atoms occupy the 2b Wyckoff position and form a geometrically perfect triangular lattice (see Figure 1(b)). Along the cc axis, the triangular layers of Ni atoms are stacked in an “AA”-type stacking fashion, and well separated by the double layers of ZrP6 octahedra. The interlayer Ni-Ni distance is cc/2 = 6.4430(2) Å, much larger than the intralayer Ni-Ni distance of aa = 3.7717(1) Å. The Ni-Zr distances \sim 2.8 Å are reasonably short, which may have considerable influence on the magnetism of Zr2NiP2. All of these indicate a quasi-two-dimensional nature of the system of Zr2NiP2. Moreover, no antisite mixing was reported [22] due to the large chemical difference among Zr, Ni, and P atoms, and thus the structural disorder should have marginal effects in Zr2NiP2.

3.2 Electrical transport

Refer to caption
Figure 2: Temperature dependence of the electrical resistivity (ρ\rho) measured at (a) 0 and (b) 9 T on two as-grown samples of Zr2NiP2 (S1 and S2). The colored lines present the BGM fits (see main text) from 2 to 300 K. (c) The magnetoresistance of Zr2NiP2 measured at 9 T, MR(9 T) = [ρ\rho(9 T)ρ-\rho(0 T)]/ρ\rho(0 T). The green lines show 1/[ρ\rho(0 T)]2 measured on S1 and S2, respectively.
Table 2: The BGM fitting parameters for Zr2NiP2. Here, σLS\sigma_{\mathrm{LS}} presents the least-squares standard deviation.
ρ0\rho_{0} (μΩ\mu\Omega\cdotm) ΘD\Theta_{\mathrm{D}} (K) RR (μΩ\mu\Omega\cdotm/K) KK (μΩ\mu\Omega\cdotm/K3) σLS\sigma_{\mathrm{LS}} (μΩ\mu\Omega\cdotm)
S1, 0 T 0.93 342 0.0052 4×\times10-10 0.0088
S2, 0 T 0.98 349 0.0057 1×\times10-9 0.0108
S1, 9 T 1.05 363 0.0051 8×\times10-10 0.0109
S2, 9 T 1.10 364 0.0054 8×\times10-10 0.0122

Figure 2(a) and (b) show the electrical resistivity (ρ\rho) measured on two independent as-grown samples of Zr2NiP2 at μ0H\mu_{0}H = 0 and 9 T down to 2 K, and no significant sample dependence is observed. ρ\rho decreases with the decrease of temperature, and gradually reaches a minimum below \sim 25 K. The measured temperature dependence can be well described by the BGM law [28, 29],

ρ(T)=ρ0+4RΘD(TΘD)50ΘD/Tx5dx(ex1)(1ex)KT3,\rho(T)=\rho_{0}+4R\Theta_{\mathrm{D}}(\frac{T}{\Theta_{\mathrm{D}}})^{5}\int_{0}^{\Theta_{\mathrm{D}}/T}\frac{x^{5}dx}{(e^{x}-1)(1-e^{-x})}-KT^{3}, (1)

where ρ0\rho_{0} is the temperature-independent residual resistivity, the second term accounts for the electron-phonon scattering, and the T3T^{3} term is due to the Mott-type ss-dd electron scattering. All the least-squares fitting parameters are listed in Table 2. The resulted ρ0\rho_{0}, Debye temperature ΘD\Theta_{\mathrm{D}}, and electron-phonon coefficient RR of Zr2NiP2 show typically metallic features [30, 31]. In comparison, the coefficient KK that is directly related to the strength of the ss-dd interaction is about one order of magnitude smaller than those of many other metals [30, 31].

Figure 2(c) shows the magnetoresistance measured on Zr2NiP2 at 9 T, MR(9 T) = [ρ\rho(9 T)ρ-\rho(0 T)]/ρ\rho(0 T), which also exhibits the common positive values of metals, MR \sim several percent [32, 33]. As presented in Figure 2(c), the temperature evolution of MR(9 T) roughly follows that of 1/[ρ\rho(0 T)]2 in the case of Zr2NiP2. To understand this observation, we consider both holes and electrons (n1n_{1}, q1q_{1}, μ1\mu_{1}, and n2n_{2}, q2q_{2}, μ2\mu_{2}, respectively),

𝐣=(n1q1μ11+B2μ12+n2q2μ21+B2μ22)𝐄+(n1q1μ121+B2μ12+n2q2μ221+B2μ22)𝐄×𝐁,\mathbf{j}=(\frac{n_{1}q_{1}\mu_{1}}{1+B^{2}\mu_{1}^{2}}+\frac{n_{2}q_{2}\mu_{2}}{1+B^{2}\mu_{2}^{2}})\mathbf{E}+(\frac{n_{1}q_{1}\mu_{1}^{2}}{1+B^{2}\mu_{1}^{2}}+\frac{n_{2}q_{2}\mu_{2}^{2}}{1+B^{2}\mu_{2}^{2}})\mathbf{E}\times\mathbf{B}, (2)

where the electric field 𝐄\mathbf{E} and magnetic field 𝐁\mathbf{B} are applied perpendicular and parallel to the zz axis, respectively. For simplicity, we assume the carrier density n1n_{1} = n2n_{2} = nn, zero-field carrier mobility μ1\mu_{1} = μ2-\mu_{2} = μ\mu, and charge q1q_{1} = q2-q_{2} = ee, so that the Hall term is negligible and current density 𝐣\mathbf{j} is parallel to 𝐄\mathbf{E}. Thereby, the magnetoresistance is given as, MR(BB) = B2μ2B^{2}\mu^{2} = B2B^{2}/[2neρne\rho(0 T)]2, in agreement with the experimental observation.

The electrical transport characterizations (Figure 2) consistently suggest that Zr2NiP2 is a typical metal without any evident transition down to 2 K. Furthermore, both ρ\rho and low MR imply a nonmagnetic state of Ni in Zr2NiP2. To confirm this, we conducted the magnetic property measurements (see below).

3.3 Magnetic properties

Refer to caption
Figure 3: (a) Magnetic susceptibilities measured under zero-field cooling (ZFC) and field cooling (FC) at 0.1 T down to 1.8 K for Zr2NiP2. The scatters show the contributions of the bulk diamagnetism χbulk\chi_{\mathrm{bulk}}. (b) Field dependence of isothermal magnetization measured at various temperatures. HpH_{\mathrm{p}} at 1.8 K is marked. The Brillouin fits (lines) to the experimental magnetization (scatters) by fixing (c) and tuning (d) the gg factor.

Neither clear anomaly/kink nor splitting between the ZFC and FC susceptibilities is observed (see Figure 3(a)), which also precludes the existence of magnetic transition in Zr2NiP2 down to 1.8 K. Above \sim 10 K, the magnetic susceptibility is negative, χ\chi << 0, thus suggesting a diamagnetism. In contrast, the susceptibility exhibits a profound Curie-like tail at low temperatures, which is most likely to originate from a paramagnetic impurity. To confirm this, we fit the experimental magnetization (MM) by a combination of a Brillouin function for free ss = 1 Ni spins (impurity) and a linear magnetization for the bulk system,

M(H,T)=fgμBexp(2μ0HgμBkBT)1exp(2μ0HgμBkBT)+exp(μ0HgμBkBT)+1+χbulk(T)H,M(H,T)=fg\mu_{\mathrm{B}}\frac{\exp(\frac{2\mu_{0}Hg\mu_{\mathrm{B}}}{k_{\mathrm{B}}T})-1}{\exp(\frac{2\mu_{0}Hg\mu_{\mathrm{B}}}{k_{\mathrm{B}}T})+\exp(\frac{\mu_{0}Hg\mu_{\mathrm{B}}}{k_{\mathrm{B}}T})+1}+\chi_{\mathrm{bulk}}(T)H, (3)

where ff is the fraction of free spins. When the gg factor is fixed to the expected value gg = 2 with negligible spin-orbit coupling, we obtain ff = 0.25% and χbulk\chi_{\mathrm{bulk}} = -0.0052, -0.0043, -0.0037, -0.0034, -0.0037 cm3/mol-Ni at 1.8, 3, 5 10, 50 K, respectively, with the least-squares standard deviation σLS\sigma_{\mathrm{LS}} = 0.0002μB\mu_{\mathrm{B}}/Ni (see Figure 3(c)). Moreover, we also treat gg as an adjustable parameter, the goodness of fit gets better (σLS\sigma_{\mathrm{LS}} = 0.00004μB\mu_{\mathrm{B}}/Ni), and we find gg = 3.3, ff = 0.09%, and χbulk\chi_{\mathrm{bulk}} = -0.0026, -0.0027, -0.0029, -0.0032, -0.0037 cm3/mol-Ni at 1.8, 3, 5 10, 50 K, respectively (see Figure 3(d)). Both combined fits suggest that the bulk system of Zr2NiP2 remains diamagnetic/nonmagnetic down to 1.8 K, i.e., χbulk\chi_{\mathrm{bulk}} << 0. The fitted χbulk\chi_{\mathrm{bulk}}(TT) is roughly consistent with the total susceptibility measured at high temperatures, but doesn’t exhibit a profound Curie-like tail at low temperatures (see Figure 3(a)).

No obvious magnetic hysteresis is observed, thus precluding ferromagnetic freezing down to 1.8 K (Figure 3(b)). The diamagnetism is confirmed by MM measured above \sim 10 K, as dMM/dHH << 0. Below \sim 10 K, the paramagnetism of tiny impurities (i.e. dMM/dHH >> 0) dominates at HH << HpH_{\mathrm{p}}, whereas the bulk diamagnetism dominates at HH >> HpH_{\mathrm{p}} after the polarization of the impurity spins. At μ0Hp\mu_{0}H_{\mathrm{p}} = 2-3 T and 1.8 \leq TT \leq 10 K, MM exhibits a broad hump with the maximum magnetization, M0M_{0} \leq 0.002μB\mu_{\mathrm{B}}/Ni, suggesting that most of Ni atoms are in the nonmagnetic state in our samples of Zr2NiP2.

Refer to caption
Figure 4: The total, Zr-, Ni-, and P-DOS of Zr2NiP2 calculated by DFT with UU = (a) 0 eV and (c) 5 eV, respectively. (b,d) Partial enlarged views of DOS near the Fermi energies (EFE_{\mathrm{F}} = 0 eV). The spin-up and spin-down states are distinguished by positive and negative DOS, respectively, for clarity. (e) The band structure. (f) The projection to the Ni-3dd states.

3.4 First-principles calculations

To understand the observed metallic and nonmagnetic characters, we calculated the density of states (DOS) by DFT+UU for Zr2NiP2. Figure 4 shows the total and partial DOS of Zr2NiP2 calculated with UU = 0 and 5 eV, respectively. Several dispersive bands passing over the Fermi energy (EFE_{\mathrm{F}} = 0 eV) are clearly observed with significant DOS (see Figure 4(b), (d), and (e)), suggesting a typically metallic character of Zr2NiP2. The DOS at EE \geq EFE_{\mathrm{F}} is dominantly contributed by Zr atoms, which accounts for the metallic nature.

The core diamagnetic contribution to the magnetic susceptibility can be estimated as χD\chi_{\mathrm{D}} \sim -MW×\times2π×\pi\times10-6 cm3/mol (SI unit) [34] \sim -2×\times10-3 cm3/mol-Ni that is close to the experimental value χbulk\chi_{\mathrm{bulk}} \sim -3×\times10-3 cm3/mol-Ni (see Figure 3(a)), where MW = 303 is the molecular weight of Zr2NiP2. Therefore, it is reasonable to estimate that the Pauli susceptibility (χPauli\chi_{\mathrm{Pauli}}) should be less than \sim 3×\times10-3 cm3/mol-Ni. We can obtain the DOS at the Fermi level DD(EFE_{\mathrm{F}}) = χPauli\chi_{\mathrm{Pauli}}/(NAμ0μB2N_{\mathrm{A}}\mu_{0}\mu_{\mathrm{B}}^{2}) << 7 States/eV, which does not contradict with our DFT results. Alternately, the Pauli susceptibility can be estimated as χPauli\chi_{\mathrm{Pauli}} = NAμ0μB2N_{\mathrm{A}}\mu_{0}\mu_{\mathrm{B}}^{2}DD(EFE_{\mathrm{F}}) = 2×\times10-4 cm3/mol-Ni, where DD(EFE_{\mathrm{F}}) \sim 0.5 States/eV is calculated by DFT (see Figure 4(b) and (d)). Usually, it is very difficult to precisely measure such a small Pauli susceptibility.

In the DFT+UU simulations, we considered both the initial antiferromagnetic and ferromagnetic structures for Ni atoms, ||\uparrow\downarrow\rangle and ||\uparrow\uparrow\rangle with a large magnetic moment mm = 2μB\mu_{\mathrm{B}}, but finally got the same nonmagnetic state with mm << 0.01μB\mu_{\mathrm{B}} after optimizations. Moreover, the spin-up and spin-down DOS are almost symmetric (see Figure 4), which naturally explains the non-magnetism observed in Zr2NiP2 (see above) [35, 36]. The number of states (occupancy) integrated the Ni DOS (Figure 4(a)) over -15 \leq EE \leq 0 eV (Fermi energy EFE_{\mathrm{F}} = 0 eV) is calculated to be 9.5, and thus the Ni valence may be \sim +0.5 in Zr2NiP2. The band structure projected to the Ni-3dd states is shown in Figure 4(f), most of the weight is below the Fermi level, suggesting Ni tends to have the nonmagnetic electronic configuration d10d^{10}.

There exists no essential difference between the DFT results calculated with UU = 0 and 5 eV. Furthermore, we also performed the exactly similar DFT+UU simulations for the other members of the family, Ln2NiX2 with Ln = Y, In, Lu, Zr and X = P, As, and obtained similar electrical and magnetic properties, i.e., metallic nature and nonmagnetic state of Ni. These results imply that the metallic and nonmagnetic ground state of Zr2NiP2 is fairly stable.

4 Conclusions

Polycrystalline and nearly pure-phase Zr2NiP2 has been successfully synthesized by the solid-state method. Structural characterization shows that the Ni atoms are arranged on a geometrically perfect triangular lattice and spatially separated by the double layers of ZrP6 octahedra. The first electrical transport and magnetic property measurements consistently demonstrate that the Ni atoms are in the nonmagnetic state. The main actors of metallicity of Zr atoms are responsible for the observed metallic behavior, as revealed by the first-principles calculations. Our work should pave the way to exploring the intriguing properties of the Ni-based triangular-lattice family Ln2NiX2, as well as other quasi-two-dimensional materials.

Acknowledgement

We thank Prof. Kan Zhao, Ying Li, and Bin Xi for helpful discussion. This work was supported by the Fundamental Research Funds for the Central Universities, HUST: 2020kfyXJJS054.

References

  • [1] P. W. Anderson, Resonating valence bonds: A new kind of insulator?, Mater. Res. Bull. 8 (2) (1973) 153–160.
  • [2] Y. Shimizu, K. Miyagawa, K. Kanoda, M. Maesato, G. Saito, Spin liquid state in an organic Mott insulator with a triangular lattice, Phys. Rev. Lett. 91 (2003) 107001.
  • [3] S. Yamashita, Y. Nakazawa, M. Oguni, Y. Oshima, H. Nojiri, Y. Shimizu, K. Miyagawa, K. Kanoda, Thermodynamic properties of a spin-1/2 spin-liquid state in a κ\kappa-type organic salt, Nat. Phys. 4 (6) (2008) 459–462.
  • [4] T. Itou, A. Oyamada, S. Maegawa, M. Tamura, R. Kato, Quantum spin liquid in the spin-1/2 triangular antiferromagnet EtMe3Sb[Pd(dmit)2]2, Phys. Rev. B 77 (2008) 104413.
  • [5] H. D. Zhou, E. S. Choi, G. Li, L. Balicas, C. R. Wiebe, Y. Qiu, J. R. D. Copley, J. S. Gardner, Spin liquid state in the SS = 1/2 triangular lattice Ba3CuSb2O9, Phys. Rev. Lett. 106 (2011) 147204.
  • [6] Y. Li, H. Liao, Z. Zhang, S. Li, F. Jin, L. Ling, L. Zhang, Y. Zou, L. Pi, Z. Yang, J. Wang, Z. Wu, Q. Zhang, Gapless quantum spin liquid ground state in the two-dimensional spin-1/2 triangular antiferromagnet YbMgGaO4, Sci. Rep. 5 (2015) 16419.
  • [7] Y. Li, G. Chen, W. Tong, L. Pi, J. Liu, Z. Yang, X. Wang, Q. Zhang, Rare-earth triangular lattice spin liquid: a single-crystal study of YbMgGaO4, Phys. Rev. Lett. 115 (16) (2015) 167203.
  • [8] W. Liu, Z. Zhang, J. Ji, Y. Liu, J. Li, X. Wang, H. Lei, G. Chen, Q. Zhang, Rare-earth chalcogenides: A large family of triangular lattice spin liquid candidates, Chin. Phys. Lett. 35 (2018) 117501.
  • [9] M. Baenitz, P. Schlender, J. Sichelschmidt, Y. A. Onykiienko, Z. Zangeneh, K. M. Ranjith, R. Sarkar, L. Hozoi, H. C. Walker, J.-C. Orain, H. Yasuoka, J. van den Brink, H. H. Klauss, D. S. Inosov, T. Doert, NaYbS2: A planar spin-12\frac{1}{2} triangular-lattice magnet and putative spin liquid, Phys. Rev. B 98 (2018) 220409(R).
  • [10] Z.-X. Liu, Y. Zhou, T.-K. Ng, Possibility of SS = 1 spin liquids with fermionic spinons on triangular lattices, Phys. Rev. B 81 (2010) 224417.
  • [11] M. Serbyn, T. Senthil, P. A. Lee, Exotic SS = 1 spin-liquid state with fermionic excitations on the triangular lattice, Phys. Rev. B 84 (2011) 180403(R).
  • [12] C. Xu, F. Wang, Y. Qi, L. Balents, M. P. A. Fisher, Spin liquid phases for spin-1 systems on the triangular lattice, Phys. Rev. Lett. 108 (2012) 087204.
  • [13] S. Bieri, M. Serbyn, T. Senthil, P. A. Lee, Paired chiral spin liquid with a Fermi surface in SS = 1 model on the triangular lattice, Phys. Rev. B 86 (2012) 224409.
  • [14] S. Nakatsuji, Y. Nambu, H. Tonomura, O. Sakai, S. Jonas, C. Broholm, H. Tsunetsugu, Y. Qiu, Y. Maeno, Spin disorder on a triangular lattice, Science 309 (5741) (2005) 1697–1700.
  • [15] S. Nakatsuji, Y. Nambu, S. Onoda, Novel geometrical frustration effects in the two-dimensional triangular-lattice antiferromagnet NiGa2S4 and related compounds, J. Phys. Soc. Jpn. 79 (2010) 011003.
  • [16] J. G. Cheng, G. Li, L. Balicas, J. S. Zhou, J. B. Goodenough, C. Xu, H. D. Zhou, High-pressure sequence of Ba3NiSb2O9 structural phases: New SS = 1 quantum spin liquids based on Ni2+, Phys. Rev. Lett. 107 (2011) 197204.
  • [17] J. G. Bednorz, K. A. Müeller, Possible high TCT_{\mathrm{C}} superconductivity in the Ba-La-Cu-O system, Zeitschrift für Physik B 64 (1986) 189–193.
  • [18] M. K. Wu, J. R. Ashburn, C. J. Torng, P. H. Hor, R. L. Meng, L. Gao, Z. J. Huang, Y. Q. Wang, C. W. Chu, Superconductivity at 93 K in a new mixed-phase Y-Ba-Cu-O compound system at ambient pressure, Phys. Rev. Lett. 58 (1987) 908–910.
  • [19] Y. Kamihara, T. Watanabe, M. Hirano, H. Hosono, Iron-based layered superconductor La[O1-xFx]FeAs (xx = 0.05-0.12) with TCT_{\mathrm{C}} = 26 K, J. Am. Chem. Soc. 130 (2008) 3296–3297.
  • [20] P. W. Anderson, The resonating valence bond state in La2CuO4 and superconductivity, Science 235 (4793) (1987) 1196–1198.
  • [21] P. A. Lee, N. Nagaosa, X.-G. Wen, Doping a Mott insulator: Physics of high-temperature superconductivity, Rev. Mod. Phys. 78 (2006) 17–85.
  • [22] E. H. E. Ghadraoui, J. Y. Pivan, R. Guérin, M. Sergent, New ternary pnictides Ln2NiX2 (X = P, As) with a filled tip-type structure, Mat. Res. Bull. 23 (6) (1988) 891–898.
  • [23] A. C. Larson, R. B. Von Dreele, General Structure Analysis System (GSAS), Los Alamos National Laboratory Report LAUR 86–748, 2004.
  • [24] G. Kresse, J. Furthmüller, Efficiency of ab-initio total energy calculations for metals and semiconductors using a plane-wave basis set, Comput. Mater. Sci. 6 (1996) 15–50.
  • [25] G. Kresse, J. Furthmüller, Efficient iterative schemes for ab initio total-energy calculations using a plane-wave basis set, Phys. Rev. B 54 (1996) 11169–11186.
  • [26] J. P. Perdew, K. Burke, M. Ernzerhof, Generalized gradient approximation made simple, Phys. Rev. Lett. 77 (1996) 3865–3868.
  • [27] D. K. Smith, W. Newkirk, The crystal structure of baddeleyite (monoclinic ZrO2) and its relation to the polymorphism of ZrO2, Acta Cryst. 18 (6) (1965) 983–991.
  • [28] N. F. Mott, H. Jones, The theory of the properties of metals and alloys, Oxford University Press, London, 1958.
  • [29] G. Grimvall, The Electron-Phonon Interaction in Metals, North-Holland, Amsterdam, 1981.
  • [30] D. Kaczorowski, P. Rogl, K. Hiebl, Magnetic behavior in a series of cerium ternary intermetallics: Ce2{\mathrm{Ce}}_{2}T2In (T = Ni, Cu, Rh, Pd, Pt, and Au), Phys. Rev. B 54 (1996) 9891–9902.
  • [31] A. P. Pikul, D. Kaczorowski, P. Rogl, Y. Grin, Kondo effect in the presence of crystal-field in Ce-Ni-Ge compounds, phys. stat. sol. (b) 236 (2) (2003) 364–367.
  • [32] J. Nickel, Magnetoresistance overview, Hewlett-Packard Laboratories, Technical Publications Dept, 1995.
  • [33] A. B. Pippard, Magnetoresistance in metals, Cambridge University Press, London, 1989.
  • [34] O. Kahn, Molecular magnetism, Wiley-VCH: New York, 1993.
  • [35] A. Ali, J.-M. Zhang, I. Muhammad, X.-M. Wei, I. Ahmad, M. U. Rehman, Changing the electronic and magnetic properties of monolayer HfS by doping and vacancy defects: Insight from first-principles calculations, Phys. Status Solidi B 257 (6) (2020) 1900768.
  • [36] Y. C. Cheng, Z. Y. Zhu, W. B. Mi, Z. B. Guo, U. Schwingenschlögl, Prediction of two-dimensional diluted magnetic semiconductors, Phys. Rev. B 87 (10) (2013) 100401.