2024 Arxiv Ni3In2S2
2024 Arxiv Ni3In2S2
2024 Arxiv Ni3In2S2
Metal Ni3In2S2
1
School of Physical Sciences, Jawaharlal Nehru University, New Delhi-110067, India
Kagome metals gain attention as they manifest a spectrum of quantum phenomena, including
superconductivity, charge order, frustrated magnetism, and intertwined correlated states of
condensed matter. With regard to electronic band structure, several of the them exhibit non-trivial
topological characteristics. Here, we present a thorough investigation on the growth and the
physical properties of single crystals of Ni3In2S2 which is established to be a Dirac nodal line
Kagome metal. Extensive characterization is attained through temperature and field-dependent
resistivity, angle-dependent magnetoresistance and specific heat measurements. In most metals,
the Fermi liquid behaviour is mostly restricted to a narrow range of temperature. In Ni3In2S2, this
characteristic feature has been observed for an extensive temperature range of 82 K. This is
attributed to the strong electron-electron correlation in the material. Specific heat measurements
reveal a high Kadowaki-Woods ratio which is in good agreement with strongly correlated systems.
Almost linear positive magnetoresistance follows the conventional Kohler scaling which depicts
the applicability of semi-classical theories. The angle-dependent magneto-resistance been
explained using the Voigt-Thomson formula. Furthermore, de-Haas van Alphen oscillations are
observed in magnetization vs. magnetic field measurement which sheds light on the topological
features in the Shandite Ni3In2S2.
1. Introduction
In general, Kagome compounds with their corner-shared triangle networks for electron
transport, generate intricate quantum interference processes giving rise to an admixture of flat
bands, saddle points, and van Hove singularities (VHSs) [13-15]. Of particular interest is Ni3In2S2
which belongs to the family of Shandite structured Kagome-metal, which requires in-depth study
for clarification because the material that is currently available has inconsistent and contradicting
information. From density-functional theory (DFT) calculations and symmetry-based theories,
Ni3In2S2 is determined to be a nonmagnetic topological semimetal with six endless Dirac-nodal
lines near the Fermi level [16-17]. Earlier reports have claimed that this non-magnetic Kagome
metal exhibits a high magnetoresistance of 2000% which is attributed to the high mobility and
small effective mass of the conduction electrons along with de Haas van Alphen (dHvA)
oscillation and Wilson ratio (Rw) greater than unity, suggesting may have strong correlation effects
in Ni3In2S2 [16]. Another study suggests that Ni3In2S2 is a completely compensated semimetal
with record-high mobility, and extremely high magnetoresistance [17]. However, the physical
mechanism underpinning the electrical conduction in Ni3In2S2 needs more clarity vis a vis its band
structure. In this work, we provide additional evidence for strong electron correlation by means of
corroborative heat capacity measurements, supporting the notion that correlation effects are
prominent in Ni3In2S2. The role played by electron correlation, which is the basis for charge density
wave [18-20] and superconductivity [21-23] in analogous systems, also needs to be integrated for
the explanation of transport data. It is well known that the electron-electron interaction strongly
influences the transport and thermodynamic characteristics of a metallic system. For instance, a
low temperature T2 behavior can be observed in the temperature-dependent resistivity in good
metals which is ascribed to electron-electron interaction. The typical behaviour is often called the
Baber law [24-26]. Generally, in metallic systems, this T2 nature exists only till low temperatures
(below 20K) [27]. At high temperatures, electron-phonon scattering dominates over the electron-
electron scattering, which ultimately leads to a linear T or T5 dependence. However, recent studies
on dichalcogenides like TiTe2 [28] and TiS2 [29] have revealed the presence of electron-electron
interaction at temperatures as high as 400K. But, the reason behind this unconventional behaviour
still requires in-depth analysis. Furthermore, it is obvious to inquire whether such unusual
behaviour exists in other materials such as the Kagome dichalogenides.
In this paper, we report on the synthesis and extensive magneto-transport and thermodynamic
characterization of single crystals of Ni3In2S2. We observe a quadratic temperature dependence of
resistivity (T2) in a broad temperature window. In the high-temperature region, on the other hand,
a linear dependence on temperature is observed. Furthermore, the specific heat study reconfirms
the existence of strong electron-electron interaction in a wide range of temperatures. The
Kadowaki-Woods ratio, which is a measure of strength of electronic correlation, is found to be
comparable to strongly correlated systems such as MoOCl2, Na0.7CoO2 and many other heavy
fermionic systems [30-32]. The linear positive magnetoresistance of Ni3In2S2 is 286% at 2K,
following the standard Kohler equation, shedding light on the semi-classical origin elevated
magnetoresistance. This type of non-saturating MR could be associated with high carrier
concentration [33], carrier spin polarization [34], or extremely high mobility [35-36]. We further
validate the semi-classical nature of Magnetoresistance in our studied material by elucidating the
angle-dependent Magnetoresistance using the Voigt-Thomson formula. Furthermore, in the
magnetic susceptibility data, we found faint but noticeable oscillations i.e., de Haas van Alphen
oscillations (dHvA). This result adds more evidence to support the sample's intrinsic topological
nature.
2. Experimental Techniques
2.1. Sample Synthesis
Single crystals of Ni3In2S2 were synthesized using a modified Bridgemann furnace. Nickel
(Powder, Aldrich, 99.99%), Indium (Powder, Thermo scientific, 99.999% pure) and Sulfur
(Powder, Aldrich, 99.98% pure) were taken in stoichiometric ratio and then ground using
mortar and pestle. The powder mixture was sealed in a quartz tube under the pressure of 10-2
mbar and then fired at 1000°C for 24hrs followed by cooling down to 800°C over 100 hrs.
Finally, the ampule was water quenched at 800°C. Shiny silver crystal flakes were obtained.
About four crystals were chosen for transport measurements, the reported data in this
manuscript were taken on a 2.1 × 1.8 × 0.4 mm3 sized single crystal. The electrical contacts
between the sample and copper wires were made using conductive silver epoxy, which was
dried using heat treatment. For the specific heat measurements, a 35mg crystal of size
1.24 × 0.96 × 0.38 mm3 was used. Apiezon N grease was used to attach to the sample to the
platform.
The crystal phase and structure of the sample were identified using powder x-ray diffraction
(XRD) in Rigaku Miniflex D/MAX 25000PC (Cu-Kα radiation, λ=1.5460Å) at room
temperature. Figure 1(a) shows the XRD data of single crystal of Ni3In2S2 which was cleaved
along (0 0 l) plane. This Shandite Kagome-metal belongs to space group 𝑅3̅𝑚 (group no. 166).
The lattice parameters from the refined XRD data are: a=b=5.37Å and c=13.53Å, which is in
agreement with previous reports [16-17]. Evidently, the space group and lattice parameters of
Ni3In2S2 and Co3Sn2S2 are much close to one another [37]. The conventional unit cell of
Ni3In2S2 is shown in the figure (1b). Inset (b) shows the single-XRD pattern i.e., Laue’s pattern,
which reconfirm the phase purity of the crystal with the powder-XRD data.
The temperature (T) dependence of electrical resistivity at zero external magnetic fields is shown
in Figure (2). The increasing resistivity with evolving temperature indicates the typical metallic
behaviour of the sample in the broad temperature range of 3.6 -300K. The residual resistivity ratio
RRR is estimated to be 42.4. No abrupt change in slope is observed in RT data which implies the
absence of any magnetic phase transition in this compound. The curve is fitted with the resistivity
equation ρ(T) = ρres + ρe−e (T) + ρe−ph (T); where the residual resistivity ρres =
0.67013 µΩ. cm, ρe−e (T)~T 2 and ρe−ph (T)~T. According to the fitting analysis in Fig. (2),
resistivity exhibits a T2 dependence at low temperatures, that is sustained to a remarkably high
temperature of 86K (light orange region). This points towards the Fermi liquid behaviour of the
sample. It is to be noted that this Fermi liquid behaviour also known as Baber law [24], can be
confirmed through the specific heat study as will be explained in the next section. The resistivity
varies at low temperature region as ρ(T) = ρres + 𝐴𝑇 2 [17], where ρres is the residual resistivity
and coefficient A denotes the electron-electron scattering effects and is typically proportional to
the square inverse of Fermi energy i.e., 𝐴 ∝ 1/𝜀𝐹2 [38]. The value of A obtained from the fitting of
resistivity data is 0.726𝑛Ω. 𝑐𝑚. 𝐾 −2 . This finding suggests that for transport in Ni3In2S2, electron-
electron scattering dominates over a broad temperature range. The presence of Fermi liquid
behaviour in such a broad temperature range is has not been observed in the previously reported
Shandite compounds. Above this temperature (82K) , the temperature dependence of resistivity
switches to a linear T dependency (light blue zone). In essence, it is observed that in Ni3In2S2, the
resistivity temperature scaling switches from T2 to 𝑇 with a small crossover region (light green).
The temperature dependence in the small cross over region is found to be ~T3/2. Similar small
crossover sections have also been observed in layered compounds [30], where the Baber law holds
at low temperatures followed by linear behaviour at higher temperatures. The linear temperature
dependence of resistivity above 108K is reflective of dominance of electron-phonon scattering
similar to what is observed in Shandite family Co3In2S2 [39].
3.2. Specific Heat
Additionally, heat capacity measurements have been carried out to verify the linear temperature
dependence of specific heat at low temperatures which is another determining feature of Fermi
liquids. The specific heat of a metal at low temperature is given by 𝐶 = 𝛾𝑇 + 𝛽𝑇 3 [40] where 𝛾 is
the Sommerfeld coefficient which indicates the electronic contribution, and 𝛽 is the phononic
contribution. In order to probe the electronic contributions to specific heat, one can study the
Kadawaki-Woods scaling in fermi liquid systems [34]. In Figure (3) we show the measured
specific heat data for single crystalline Ni3In2S2. From the previously mentioned specific heat
equation, the value of 𝛾 and 𝛽 obtained are as follows: 𝛾 = 4.81𝑚𝐽. 𝑚𝑜𝑙 −1 . 𝐾 −2 and 𝛽 =
1.7𝑚𝐽. 𝑚𝑜𝑙 −1 . 𝐾 −4 . In single band free electron model, 𝛾 can be related with the effective mass of
2
1
𝑘 2 𝜋
3
electrons with the relation: 𝛾 = 𝑚∗ 𝑛3 ( ℏ𝐵 ) (3 ) , from which the value of effective mass can be
obtained as 𝑚∗ = 1.17𝑚𝑒 . It is worth noting that the effective mass obtained for this sample is
much higher than the others compounds in the same family [37,*]. Previously T. T Zhang et al
[15] had reported a high Wilson ratio () for the same compound. This high value indicates towards
the strong electronic correlation and is in resonance with our results of specific heat measurement.
The phononic contribution parameter β, is connected with the Debye temperature ƟD. The Debye
temperature for Ni3In2S2 is estimated to be about 92.09K, which is quite high as compared to other
shandite materials.
Now from Figure (3) we have demonstrated that in Ni3In2S2, the resistivity temperature scaling
directly transitions from T2 to T. The above-mentioned predicted Debye temperature aids us in
comprehending this unusual trait. Below the cross-over temperature, the transport is dominated by
electron-electron scattering. We note that in Ni3In2S2, crossover temperature and Debye
temperature are comparable to each other.
The A coefficient and 𝛾 coefficient obtained from the scaling relation for resistivity curve and
specific heat, both contain electron-electron scattering effects. The Kadawaki-Woods ratio 𝛼 =
𝐴/𝛾 2 is a significant metric that investigates the relation of the electron-electron interaction rate
and the effective mass [41]. Usually, the value of the ratio is approximately same for materials
belonging to the same class. While for many transition metals, 𝛼 𝑇𝑀 ~0.4 𝜇Ω. 𝑐𝑚. 𝑚𝑜𝑙 2 𝐾 2 𝐽−2 ,
whereas for many heavy fermionic metals 𝛼𝐻𝐹 ~10 𝜇Ω. 𝑐𝑚. 𝑚𝑜𝑙 2 𝐾 2 𝐽−2 . Here, based on the
outcomes of our experiments, we obtained 𝛼𝑁𝑖3𝐼𝑛2𝑆2 ~31.37 𝜇Ω. 𝑐𝑚. 𝑚𝑜𝑙 2 𝐾 2 𝐽−2 which is three
times higher than the value for heavy fermionic compounds. Such high Kadawaki-Woods ratio has
been previously obtained many oxide perovskites [31, 42-43]. This high value of α attributes to
the strong electron correlation and is consistent with the broad T2 scaling in the temperature
dependent resistivity behaviour.
3.3. Magnetoresistance
where ρ(B) and ρ(B = 0) being the resistivity under the applied fixed magnetic field and zero
magnetic field respectively. The direction of applied current and the magnetic field is
perpendicular to each other (I ⊥ B).
The sample exhibited a notable peak in positive Magnetoresistance (MR), reaching a maximum
value of 286% at 2K under a 6T magnetic field. It is noteworthy that our experimental MR value
is comparatively lower than the reported results on Ni3In2S2, wherein the MR reaches an
exceptionally high value of 15518% at 2K and 13T magnetic field [17]. This discrepancy in MR
values is anticipated, attributable to lower carrier mobilities and the absence of effective charge
compensation within our sample. According to our findings, the mobility in the Ni3In2S2 single
crystal at 2K is ~30.63 𝑐𝑚2 . 𝑉. 𝑠, and the carrier density is ~1.72 × 1019 𝑐𝑚−3 . The MR%
remarkably decreases with increasing temperature as shown in inset of fig. 3(a). As the MR varies
with the direction of applied magnetic field (anisotropy in MR), the exixtance of open orbit in
Fermi surface can be a reasonable explanation for the non-saturating behaviour of MR [45]. In
Dirac semimetals, this kind of linear MR originated due to coupling of electrons with the spin-
orbit interaction of the material [46-48].
Futhermore, the application of the Kohler rule properly captures the analysis of transverse
magnetoresistance (MR) and illuminates the complex mechanisms governing the scattering
dynamics of charge carriers. The Kohler rule is defined by a basic equation (1) using the principles
of semi-classical band theory, providing a more sophisticated explanation of the intricate processes
underlying the observed MR behavior [49]. The Kohlr scaling relation is formulated as follows:
∆𝜌 𝐵 𝑚
𝑀𝑅 = = 𝛼 (𝜌 ) (2)
𝜌0 0
Here, in equation (2), the parameters (α, m) are predefined constants, and ρ 0 represents the zero-
field resistivity. Adhering to this Kohler rule, the MR data is systematically represented as MR~
𝐵
curve at different temperatures and is shown in figure (b). In accordance with Kohler’s rule, the
𝜌0
MR curve convergence into a single line denotes the homogeneity in the charge carrier scattering
process throughout the system's temperature range. The parameter 'm' derived from fitting equation
(1), exhibits a value of 1.05 at 2K, indicative of an almost linear dependence of Magnetoresistance
(MR) on the applied magnetic field. Consequently, the value of 'm' increases noticeably as the
temperature rises. The quantum linear MR (QLMR) is pronounced in Dirac semimetals due to
their unique band structure, and the presence of Dirac points such QLMR of Dirac semimetals has
been studied extensively in recent past years and is an important tool for studying the electronic
and transport properties of these materials. However, Fang, Hongwei, et al found the parabolic
dependence of MR on applied magnetic field which suggest this compound as fully compensated
semimetal [17].
Additionally, the angular dependence of MR has been studied. Figure (4) shows the angle-
dependent magnetotransport characteristics, where Ɵ is the angle between direction of the
magnetic field and the current, as shown in the inset of Figure 4(a). Figure 4(b) shows the Rxx as
a function of angle Ɵ at different temperatures at fixed magnetic field B=4T. It is observed that at
Ɵ=90° and 270°, the resistance value reach at its highest point, and at Ɵ=0°, 180° and 360° the
resisistance get minimum. The anisotropic orbital magnetoresistance of a system is defined by the
Voigt-Thomson formula [44-45].
where 𝑅⊥ = 𝑅𝑥𝑥 (𝜗 = 90°) and 𝑅∥ = 𝑅𝑥𝑥 (𝜗 = 0°). The red solid curve in Figure (4) shows that
the Voigt-Thomson formula that provides a good fit with the measured values of resisistance
𝑅𝑥𝑥 (𝜃). This outcome further confirms the semi-classical MR origin in our sample, that relate to
the extent of Lorentz force in this measurement configuration.
3.4. Magnetization
Quantum oscillations at low-temperature and high magnetic fields is a very useful tool to probe
the electronic properties of materials such as determining the Fermi surface of material, effective
mass of charge carriers. Figure (5) shows the isothermal magnetization measured with an applied
external magnetic field up to 10T H || c for our sample. The de Haas-van Alphen (dHvA)
oscillations were present when magnetic field exceeding 7T at 2K, indicating the fermi pockets
associated with dHvA oscillations. Moderate oscillations with amplitude 0.046 emu/g at 2K were
seen. The major oscillation frequencies Fα=12.02T, Fβ=19.24T and Fγ=28.85T were derived from
the fast Fourier transform (FFT) analysis of this oscillatory magnetization as shown in figure 5.
The cross-sectional area (AF) of Fermi surface can be calculated by assuming the circular cross-
2πe
section of surface along (0 0 l) plane using Onsager relation AF = F, where F is the frequency
ℏ
of the oscillation. The cross-sectional area of Fermi surfaces is 1.15×10-3Å-2, 1.18×10-3Å-2 and
2.76×10-3Å-2 corresponding to the frequencies Fα, Fβ and Fγ respectively. The obtained surface area
values are used further to calculate the Fermi vectors are as follows Kα =1.91×10-2Å-1, Kβ=1.94×10-
2 A
Å-1 and Kγ=2.96×10-2Å-1 respectively using the formula K F = √ πF .
This dHvA oscillations can be described using Lifshitz-Kosevich (LK) formula [53] in
which more detailed information about the Berry phase can be extracted by the fitting of LK
formula as given bellow:
F
∆m ∝ R T R D R S sin [2π( − γ − δ)]
μ0 B
2π2 kB m∗ T⁄
ℏeμ0 H 2π2 k B m∗ T⁄
Where R T = 2π2 kB m∗ T⁄
due to temperature effect, R D = exp ( ℏeμ0 H) due
sin( ℏeμ0 H)
to impurity scattering, and R S = cos(πgm∗ ⁄2me ) due to spin-splitting effect. The oscillation of
magnetization is due to the sine term as mentioned in the equation. The phase term −γ − δ, in
1 ΦB
which γ = 2 − is called Onsager phase directly linked with Berry phase ΦB . The other term δ
2π
is additional phase shift, can be determined by the dimensionality of the fermi surface. δ = 0 or ±
1⁄ for 2D and 3D cases where + is for the electronlike cases and – for the holelike cases.
8
4. Conclusion
In summary, we have successfully grown high-quality single crystals of Ni3In2S2, which is a Dirac
nodal line Kagome metal. Fermi-liquid behavior up to 86 K, which is the highest among the
existing Shandite Kagome semimetals, is observed. This is corroborated by specific heat studies
within the corresponding temperature range. The Kadowaki-Woods ratio of Ni3In2S2 is in good
agreement with the values of strongly correlated metals. Almost linear, unsaturated
magnetoresistance as well as the anisotropy in MR are attributed to the existence of open orbits in
the Fermi surface of Ni3In2S2. Employing the Kohler scaling to magneto-resistance data we
checked the applicability of semi-classical theories to magneto-transport in Ni3In2S2. This is
further complemented by the implementation of the Voigt-Thomson formula that is well-fitted in
the angle-dependent magnetoresistance. The observation of de Haas van Alphen oscillations at 2K
substantiates the quantum behavior of Ni3In2S2. Our findings offer insight into the interplay
between electron correlation and topological band structure in Kagome metals, setting the stage
for deeper exploration and comprehension of these intriguing materials.
5. Acknowledgement
P. Das and P. Saha acknowledge UGC-NET JRF for financial support. M. Singh and P. Kumar
thank CSIR for providing JRF. We are grateful to the FIST program of the Department of Science
and Technology, Government of India for the use of the low-temperature high magnetic field
measurement facility at JNU. We acknowledge funding support from DST towards the
procurement of chemicals and consumables from the project (DST/NM/TUE/QM-10/2109(G)/6).
We acknowledge Advanced Instrumentation Research Facility (AIRF), JNU for use of the PPMS
measurement facility.
References
Figure 1: Structural characterization of Ni3In2S2 single crystal: Figure (a) shows the XRD pattern
of Ni3In2S2 single crystal cleaved along (0 0 l) plane. Inset (i) shows the typical image of single
crystal. Inset (ii) shows the Lau’s pattern of single crystal. (b) Schematic view of unit cell of
Ni3In2S2 with 𝑅3̅𝑚 symmetry. (c) SEM image of the single crystal shows the layered structure of
Ni3In2S2.
Figure 3: The plot of Cp/T vs. T2 with a linear fitting at a low temperature of 12K by Cp/T = βT2 +
γ in solid line (red). The inset of figure 3 shows the temperature dependence of specific heat of the
single crystal Ni3In2S2 in the temperature range 2-160K.
Figure 4: (a) The magnetic field dependence of transverse resistance at different temperature.
Transverse magnetoresistance 𝑀𝑅(%) = [𝑅(𝐻) − 𝑅(0)]/𝑅(0) × 100 measured at the indicated
temperatures. Inset of 4(a) shows the variation of MR with respect of temperature. (b) The
MR~B/ρ0 plot data was systematically fitted using Kohler’s rule in the relevant equation (1) at
different temperatures.
Figure 5: Magneto-transport behavior of Ni3In2S2 single crystal. (a) The angular dependence of
Rxx at fixed temperature 5K and magnetic field 4T. The solid curve (red) is the fitting using Voigt-
Thomson Formula. The inset of (a) clarifies the angle Ɵ. (b) The angular dependency of ρxx at
different temperatures at fixed magnetic field of 4T. Inset (i) shows the variation of 𝑅⊥ = 𝑅𝑥𝑥 (𝜗 =
90°) and 𝑅∥ = 𝑅𝑥𝑥 (𝜗 = 0°) with temperatures. Inset (ii) shows the variation of anisotropy with
temperatures.
Figure (6): The dHvA oscillations in Ni3In2S2. (a) Isothermal magnetization data under 𝐻 ∥ 𝑐 at
2K is shown in inset of (a). Zoom in view of magnetization data from 8-8.4T. (b) The
magnetization oscillation after background substruction is shown in inset of (b). The corresponding
Fourier transforming of the oscillating component.
Figure 1
Figure 2
Figure 3
Figure 4
Figure 5
Figure 6