Synthesis and structural analysis. Figure 1a shows that the Gd3SnC crystallizes in the antiperovskite structure (space group: Pm\(\stackrel{-}{3}\)m) with face-centred Gd and body-centred C atoms in a Sn cubic lattice, as confirmed by Rietveld analysis of powder X-ray diffraction (XRD) pattern (Supplementary Table 1). Single-phase bulk Gd3SnC was synthesized via the melt-solidification process at one atmosphere. This indicates the thermodynamic stability of Gd3SnC antiperovskite, which gets additional support from the calculated phonon dispersions having no imaginary frequency (Supplementary Fig. 1). In contrast to the ABX3 perovskite, which has cation-centred BX6 octahedra with ionic bonds, Gd3SnC has anion-centred CGd6 octahedra with strong covalent Gd–C and metallic Gd–Gd bonds. This mixed bonding character of Gd bonds in CGd6 octahedra results in different Gd orbital levels from those of the Gd element and its ionic compounds. As shown in Fig. 1b, while Gd metal and Gd2O3 have ordinary Gd orbital levels of Gd–Gd metallic and Gd–O ionic bonds, respectively, Gd3SnC has unusual energy levels for Gd orbitals due to the coexistence of metallic and covalent bonds. Considering the shorter Gd–Gd bond length (3.48 Å) in CGd6 octahedra compared to that of Gd metal (~ 3.6 Å), one expects an enhanced metallicity and thus additional Fermi level (EF) crossings for Gd bands in Gd3SnC. On the other hand, the electronegativity difference between Gd and C is much smaller than that between Gd and O, resulting in the formation of covalent bonds between Gd and C. The energy difference between bonding and antibonding states of the covalent bonds is expected to be significantly smaller than that of ionic bonds between Gd and O. These mixed bonding characters of Gd valence orbitals in CGd6 octahedra can lead to a peculiar molecular orbital diagram with both metallic and covalent bonds as shown in Fig. 1b. Such observation indicates that Gd3SnC antiperovskite may exhibit unusual physical properties that are hardly found in elemental Gd and conventional Gd-related compounds.
Tricritical ferromagnetic behaviour. Figure 2a shows the FM phase transition of Gd3SnC with a Curie temperature (TC) of 100 K under a low magnetic field (H) of 0.1 kOe. In comparison to the itinerant FM in Gd metal with a TC of 297 K, different aspects of the FM order are observed12. A sharp increase in the magnetization (M) near the TC indicates a first-order transition, which is further evidenced by the hysteresis in the temperature(T) dependence of electrical resistivity (ρ) (Fig. 2e) and the negative slope in the M2 dependence of H/M (Supplementary Fig. 2e). In addition, no structural phase transition is observed around the TC, indicating that the FM ordering occurs while preserving the antiperovskite structure (Supplementary Fig. 3). We confirm the saturation magnetic moment of ∼6.8 µB per Gd obtained from the M–H curve (Fig. 2b), indicating this FM state comes mostly from the seven Gd 4f-electrons. A notable aspect of M (T) is a slight decrease in the ZFC curve below ~ 20 K (inset of Fig. 2a), which suggests a possible formation of singlets such as Kondo singlet13,14, as will be discussed in the following sections. This unusual behaviour in M indicates that the near EF states of Gd, which do not have contributions from fully localized f-orbitals, may be different from those of the familiar FM Gd metal and Gd-based compounds12,15.
The most peculiar aspect of the FM state in Gd3SnC is the tricritical behaviour that appears near the boundary between the first- and the second-order phase transitions16. Figure 2c shows a modified Arrott plot17 of the tricritical mean-field model for the FM order with βArrott = 0.25 and γArrott = 1 in (H/M)1/γArrott and M1/βArrott, respectively. Compared to other magnetic ordering models, it is apparent from the relative slopes (dashed red lines in Fig. 2c and Supplementary Fig. 2a–d) that the FM transition belongs to the universality class of the tricritical mean-field model (Fig. 2d). A clear crossover from a first-order to second-order transition is seen in the T dependent ρ and M curves under different H as shown in Fig. 2e,f, respectively. Three representative features in the data verify the crossover between FM orders: (i) disappearance of the peak in ρ at TC, (ii) suppression of the hysteresis in the ρ–T curve during a thermal cycle (insets of Fig. 2e), and (iii) smooth increase of M upon cooling (Fig. 2f). Such tricritical behaviour often emerges in heavy-fermion systems such as UGe218 and YbRh2Si219 due to the competition between Kondo coupling (TK) and Ruderman–Kittel–Kasuya–Yosida interaction20.
Heavy-fermions with non-Fermi liquid behaviour. More unusual properties of Gd3SnC may be found in the heat capacity (C) measurement results (Fig. 3). As a way to determine the unexpected behaviour, we compare the C of Gd3SnC to those of FM Gd metal and non-magnetic La3SnC antiperovskite, as shown in Fig. 3a. In contrast to the small C values of Gd metal and La3SnC antiperovskite in the low-T region, an extremely large C value with a Sommerfeld coefficient (γ) of ~ 1114 mJ mol− 1 K− 2 at 90 mK is observed for the Gd3SnC. The dramatic low-T rise of C is reminiscent of a nuclear Schottky anomaly, most likely from the Sn nucleus (the natural abundance for non-zero nuclear spin 13C is below 1%). However, our modelling, which considers 16 % natural abundance of 117Sn and 119Sn, rules out the Schottky anomaly by showing that the required internal field is about − 1287 kOe, a value much larger than any reported values21–23 (Supplementary Fig. 4d). In addition, the Gd nuclear spin cannot be the culprit as we do not see a similar rise in C for Gd metal. Therefore, we attribute the logarithmic rise of C for Gd3SnC below ~ 0.13 K (Supplementary Fig. 4a) to a non-Fermi-liquid (NFL) behaviour of itinerant conduction electrons24–26 (inset of Fig. 3a). A T-linear dependence of ρ in the low-T region also supports the NFL behaviour of itinerant conduction electrons in Gd3SnC (left panel in Fig. 3b). The transition from T-linear to T2 behaviour of ρ upon applying 140 kOe (right panel in Fig. 3b and Supplementary Fig. 5) is suggestive of a magnetic origin of the NFL behaviour27,28. Thus, we assume that an exotic quantum phenomenon coupled with a correlation between itinerant and localized electrons in the Gd3SnC antiperovskite appears as heavy fermions.
As Gd f-electron levels locate deep in the binding energy, the heavy-fermion with the NFL behaviour should originate from the electrons of other orbitals, which is revealed from the analysis of the C of FM Gd3SnC. The magnetic contribution (Cmag) to the C of FM Gd3SnC is obtained by subtracting the C of non-magnetic La3SnC without 4f-electrons (La metal with the configuration of [Xe]4f05d16s2) from those of Gd3SnC with seven 4f-electrons (Gd metal with the configuration of [Xe]4f75d16s2). We find two distinct features in Cmag: a large Cmag in the high-T region (Cmag,high, Fig. 3c) and a small Cmag in the low-T region (Cmag, low, inset of Fig. 3c). While the Cmag,high is attributed to the FM order of 4f7-electrons, which agrees well with the prediction of the renormalization-group approach29 in the scheme of tricritical ferromagnetism (Supplementary Fig. 2f), the origin of Cmag,low is ambiguous. From the T-dependent magnetic entropy (ΔS) obtained by integrating Cmag (T) (Supplementary Fig. 4b), we can assign which electrons are responsible for Cmag,low. Figure 3d shows a large difference in ΔS between the high- and low-T regions: 0.5Rln8 and 0.05Rln2, respectively. Since 0.5Rln8 reflects the contribution from seven 4f-electrons (4f7) involved in FM ordering, one may notice that the much smaller value of 0.05Rln2 has no correlation with 4f-electrons but is ascribed to 5d-electrons, which are responsible for the heavy-fermion with the NFL behaviour.
Dual character of Gd 5d electrons. This unconventional and anomalous d-electron heavy-fermion state in f-block Gd-based Gd3SnC is, if true, the first case of FM-assisted heavy-fermion behaviour in 5d-electron systems. Density functional theory (DFT) calculations along with angle-resolved photoemission spectroscopy (ARPES) measurements provide a clear picture for the formation of the heavy-fermion state, which is induced not by 4f-electrons but by 5d-electrons of Gd element in the context of mixed bonding characters of the antiperovskite structure. Figure 4a shows ARPES data along the Γ–X–Γ direction taken above (125 K) and below (90 and 13 K) the TC. The 125 K data shows a fast-dispersing band centred at the X point. As the T is lowered to 90 K, the band shifts down due to the exchange splitting of both 4f- and 5d-electrons (Supplementary Fig. 6d). In addition to the fast-dispersing band (green curve), there is a weakly dispersing band near EF (red curve in Fig. 4a, Supplementary Fig. 6a–c). Upon cooling to 13 K, the fast-dispersing band shifts further down, and the weakly dispersing band near EF disappears. These changes in the electronic structure might be related to the hybridization of fast and weakly dispersing bands below 20 K. The T-dependent electronic structure change is also seen in the kx–ky plane of Fermi surface maps in Fig. 4b; the Fermi surface pocket becomes larger as the T decreases. This enlarged Fermi surface is a typical phenomenon as found in various heavy-fermion systems27,30−32. The calculated partial density of states (DOS) clearly demonstrates that the Gd 5d-orbitals are almost solely responsible for the states near EF, while Sn 5p and C 2p derived states are located away from EF (Fig. 4c). Meanwhile, we confirm that the contribution of Gd 4f-orbitals into the band structure near EF is negligible but responsible for the FM ordering (Supplementary Fig. 7).
We note that the calculated band structure shows an anomalous flat band feature near EF, which is also a signature of heavy-fermion behaviour (Fig. 4d,e). The experimentally observed flatness of the fast-dispersing band at the X point (13 K in Fig. 4a) is indeed well-reproduced by band calculations (highlighted as yellow colour in Fig. 4e). The fat band analysis reveals that the band hybridization triggers the appearance of a flat band feature near EF. Importantly, the two hybridized bands originate from the only Gd 5d-orbitals: t2g bands from the Gd–Gd bonds (red curves) and eg bands from the mostly Gd–C bonds (green curves). Thus, we believe that the flat t2g band observed at 90 K (Fig. 4a) is lifted above EF after the band hybridization process (Supplementary Fig. 6e). Finally, we emphasize the importance of antiperovskite structure with mixed bonding nature, which is the key to imparting the unusual heavy-fermion behaviour to Gd3SnC. The crystal orbital Hamiltonian population (COHP) analysis (Fig. 4f) explains how the chemical bonds of CGd6 octahedra in antiperovskite structure can induce the unusual band structure. The COHP analysis suggests two types of orbital overlap: a small orbital overlap in metallic Gd–Gd bond responsible for the t2g bands near EF (− ICOHP ~ 0.5 eV/bond, Supplementary Table 2) and large orbital overlap in covalent Gd–C bond for the eg bands below EF (− ICOHP ~ 1.5 eV/bond). This mixed bonding nature with the different degrees of orbital overlap leads to the unusual band dispersion of Gd 5d-electrons, as illustrated in Fig. 4g − i. Thus, the duality of Gd 5d-electrons participating in the mixed chemical bonds in CGd6 octahedra of the Gd3SnC antiperovskite structure is a key factor that determines the unusual physical properties.