Introduction

The kagome lattice is a 2D network of corner-sharing triangles that has been intensively investigated the last years. Due to its unusual geometry, it offers a playground to study interesting physics, including frustrated, correlated1,2, exotic topological quantum1,2,3,4,5,6,7,8,9,10,11,12,13,14,15,16, topological Chern17, insulating and Weyl semimetal13,18 phases, originating from the interplay between magnetism and electronic topology. In fact, the kagome lattice has been realized in several materials including metal stannides, germannides19,20 as well as TmXn compounds with T = Mn, Fe, Co, X = Sn, Ge (m: n = 3:1, 3:2, 1:1)21. Recent studies demonstrated that Fe–Sn-based kagome compounds exhibiting interesting properties, such as large magnetic tunability1. Furthermore, they can host Dirac fermions and flat bands, as found in Fe3Sn222,23 and FeSn21,24. The existence of spin degenerate band touching points was linked to the generation of several interesting phenomena. Specifically, the anomalous Hall effect (AHE) results in a transverse spin-polarized charge current (charge current and spin current due to the imbalance of spin up and spin down electrons in ferromagnets) in response to a longitudinal charge current, in the absence of an external magnetic field25,26,27,28,29. This applies also to its thermal counterpart, the anomalous Nernst effect (ANE), in which the external stimuli is replaced by a thermal gradient30 as well as the Seebeck effect31.

Interestingly, the Fe–Sn-based intermetallics compounds not only exhibit attractive topological transport properties, but also show rich magnetic properties. In our previous studies32,33, a DFT screening of the Fe–Sn phase diagram was used to identify Fe–Sn-based phases with the potential to be stabilized upon alloying, and their magnetization and magnetocrystalline anisotropy were evaluated. The results revealed that a strong anisotropy as observed in Fe3Sn may also be found in other Fe–Sn-based phases, having high potential to be used as hard magnetic materials. Meanwhile, we applied the reactive crucible melting (RCM) approach to the Fe–Sn binary system, and observed three metastable intermetallic compounds, namely Fe3Sn, Fe5Sn3, Fe3Sn2, which are ferromagnetic and exist between 873 K and 1173 K. We found that such metastable phases can be synthesized using the RCM method at specific temperature ranges. What’s more, phase diagram of the Fe–Sn system reported in the literature19,34,35 has mentioned that the Fe3Sn was considered to be a metastable phase, and presented the relevant so-called metastable composition range and phase relations. According to Fayyazi’s32 work, the reactive crucible reproduced the corresponding phase relations as in the bulk samples at 998 K (α-Fe, Fe3Sn2, FeSn, and Sn) and 1023 K (Fe3Sn2, Fe3Sn, and FeSn), of which Fe3Sn can only be stabilized between 1023–1098 K during a non-equilibrium process as a metastable phase but disappears at 1123 K due to the presence of Fe5Sn3 phase. Accordingly, adding more details to the phase diagram of the metastable Fe3Sn phase, with the discovered temperature range based on the reported phase diagram is of great significance. Therefore, to further explore the interesting properties of metastable Fe–Sn phases, it is important to understand the phase diagram and thermodynamical properties of the Fe–Sn system.

In this work, we adopted our new measurements32,33 on the equilibria states of Fe3Sn, Fe5Sn3, Fe3Sn2, combined with the thermodynamic properties of such intermetallic phases obtained based on first-principles calculations. A consistent thermodynamic assessment of the Fe–Sn system was then developed based on all available experimental and first-principles results. Furthermore, the AHC and ANC of Fe3Sn were calculated and its dependence on the magnetization direction and doping were evaluated. We observed that there exist significant changes in AHC and ANC by tuning the Fermi energy via Mn-doping. Therefore, Fe3Sn renders itself a promising candidate for new transverse thermoelectric devices with potential applications.

Results and discussion

Metastable phase diagram

Most end-members in the sublattice models are not stable and their thermodynamic data are impossible to be determined by experiments. First principles are hence performed to estimate the Gibbs energies of the compounds and end-members at finite temperatures. In order to benchmark the current DFT calculations, the calculated crystallographic information of phases in the binary Fe–Sn system are listed in Table 1, in comparison with the available experimental data. The calculated lattice parameters of the solid phases at 0 K are in good agreement with the experimental results at room temperature. As one can see, the differences between the theoretical and experimental lattice constants are within 0.5% for all the phases. Note that, in our earlier study32,33, we showed, that the crystal structure of "Fe5Sn3” synthesized by the equilibrated alloy method, is not of the typically assumed hexagonal Laves structure (as shown in Table 1). We rather observed superstructure reflections in the powder XRD spectra that could not be explained by the hexagonal structure and we assigned to a modulated orthorhombic unit cell with lattice parameters of a = 4.221 Å, b = 7.322 Å, c = 5.252 Å. More details and explanations can be found in the refs. 32,33. Hence, we used this structure to do phonon calculations. Furthermore, the calculated phonon bands of such phases are shown in Fig. 1. To prove the validity of the calculations, as shown in Fig. 1, the phonon dispersion of BCC-Fe is compared with the experimental data36, presenting good agreement. Therefore, it is expected that the thermodynamical properties of the Fe–Sn intermetallic phases can also be accurately obtained based on DFT calculations. As shown in Fig. 1, no imaginary phonon modes exist for all the compounds, indicating that all the intermetallics are dynamically stable. And the quasi-harmonic approximation (QHA) can be used to calculate the thermodynamic properties.

Table 1 Lattice parameters of intermetallics from first-principles calculations compared with experimental values.
Fig. 1: Phonon dispersions of the pure elements and intermetallic phases in the Fe–Sn system.
figure 1

The black solid points represent the experimental data from ref. 37.

The thermodynamic properties at finite temperatures are evaluated based on the Gibbs free energies specified in Eq. (5). And from the thermodynamical point of view, we can derive the Gibbs free energies from the heat capacity. To obtain the accurate heat capacity of the intermetallics, we firstly compare the calculated heat capacity of the BCC-Fe with the available experimental data37, as shown in Fig. 2. Among that, the magnetic contribution to the heat capacity is analyzed following the theory of IHJ model38,39 and further improved version by Xiong40:

$${C}_{{p}_{{{{\rm{mag}}}}}}=R{{{\rm{\ln }}}}({\beta }^{* }+1)c(\tau ).$$
(1)
Fig. 2: Heat capacity of pure Fe and Sn from DFT calculations in comparison with the experiment data37.
figure 2

Those for all the intermetallics are also shown, experimental data of Fe3Sn is obtained from our previous studies32,33.

Figure 2 shows isobaric heat capacity obtained from our DFT calculations. It can be observed that the lattice vibrations dominate other contributions to the heat capacity. Interestingly, the correction made by adding electronic and magnetic heat capacities shifted the result toward bigger values and after that calculations show an excellent agreement with the experimental data37. More interestingly, the magnetic contribution to the heat capacity presents at the magnetic phase transition of BCC-Fe. These results prove the accuracy of the current methods and justify the following calculations for intermetallics. Using the same strategy, we calculate heat capacities of Fe5Sn3, Fe3Sn2, Fe3Sn, FeSn2, and FeSn at finite temperatures, as shown in Fig. 2, with the magnetic heat capacity evaluated using Inden model39. The heat capacity of Fe3Sn shows a good consistency between our calculations and experiments at low temperature, which also confirms the accuracy of current theoretical results. We note that such good agreements are supported by considering the magnetic contributions in the magnetic system.

After getting the thermodynamical properties of intermetallics, we used CALPHAD method41 to evaluate the thermodynamic model parameters of the Fe–Sn system, and the phase diagram and thermodynamic properties are calculated by Thermo-Calc42. Combining DFT and CALPHAD methods has already been successfully applied in different systems43,44. Supplementary Table 1 lists the modeled thermodynamic parameters of the Fe–Sn system. The calculated Fe–Sn phase diagram is presented in Fig. 3 along with the experimental data32,45,46,47,48,49,50,51,52,53,54,55,56. The comparison of the calculated temperatures and compositions of invariant reactions with experimental data45,46,47,48,49,50,51 as well as results from previous thermodynamic assessments35,57 are listed in Table 2.

Fig. 3: The optimized Fe–Sn phase diagram based on our thermodynamic modeling.
figure 3

The points correspond to the different experimental data32,45,46,47,48,49,50,51,52,53,54,55,56. The red lines indicate the reaction temperatures between the different phases.

Table 2 Summary of the invariant reactions in the Fe–Sn system.

Using the reactive crucible melting (RCM) approach, it is found that 3 metastable intermetallic compounds, i.e., Fe3Sn, Fe5Sn3, and Fe3Sn2, can be stabilized between 873 K and 1173 K. Furthermore, we are convinced that the phase diagram reported in the literature is inaccurate in the temperature interval 1023–1038 K and Fe3Sn can exist at 1023 K. Thus, the metastable phase Fe3Sn is introduced by considering the current accurate experimental results. Obviously, good agreement between the optimized and experimental (see Supplementary Figs. 1 and 2) shows the calculated thermodynamic properties of the compounds in current CALPHAD modeling, first-principles calculations, previous CALPHAD modeling and the experimental data. The calculated thermodynamic properties in this work are consistent with experimental data.

Topological transport properties

Our calculations demonstrate that Fe3Sn exhibits the largest AHC among the Fe–Sn family. The calculated x-component of AHC (σx) for the equilibrium lattice parameters, with the magnetization direction along [100]-axis, reaches 757 S cm−1 at Fermi energy, as shown in Fig. 4a. This value is surprisingly large compared to the more than 3 times smaller value of 200 S cm−1 for Fe3Sn258. In addition, Li et al. reported an experimentally measured value of 613 S cm−1 value as well as a calculated of 507 S cm−1 for Fe5Sn329, both being smaller than our value for Fe3Sn. In addition, compared to other ferromagnetic kagome materials, it ranks among the largest reported, being larger than 380 S cm−1 of LiMn6Sn659 and 223 S cm−1 for GdMn6Sn660, but lower than the largest reported value of 1130 S cm−1 for Co3Sn2S261. Correspondingly, the ANC of Fe3Sn, evaluated at T = 300 K, is also the largest among the Fe–Sn family. Specifically, it reaches −2.71 A m−1 K−1 (see Fig. 5b) being more than two times larger than the reported value of 1 A m−1 K−1 for Fe3Sn230. Compared to the other kagome materials, Fe3Sn exhibits a reasonably large ANC, being larger than 1.29 A m−1 K−1 and 0.20 A m−1 K−1 reported for ZrMn6Sn6 and MgMn6Sn662, but smaller than the largest reported value of 10 A m−1 K−1 for Co3Sn2S263. Compounds with large AHC and ANC values are promising candidates for transverse thermoelectric devices64,65,66,67.

Fig. 4: Anomalous Hall/Nernst conductivities and Band gap of Fe3Sn with M[100].
figure 4

a AHC components as a function of energy. b ANC components as a function of energy. c The band gap (eV) for a BZ-slice at kz = −0.131 with its energies within the shaded energy range of part (a). d The distribution of the x-component (σx = σyz) of the full AHC at each part of the BZ.

Fig. 5: Tunability of anomalous Hall and Nernst conductivities in respect to strain and magnetization direction for Fe3Sn.
figure 5

a The x-component (σx = σyz) of the full AHC as a function of energy for different doping concentrations x of (Fe1−xMnx)3Sn. b The x-component (αx = αyz) of the full ANC as a function of energy for different doping concentrations x of (Fe1−xMnx)3Sn. c AHC as a function of energy for different magnetization directions and doping concentrations of (Fe1−xMnx)3Sn. d ANC as a function of energy for different magnetization directions and concentrations of (Fe1−xMnx)3Sn.

Symmetry plays a crucial role in determining the shape of the AHC and ANC tensors. AHC and ANC strongly depend on the Berry curvature which behaves as a pseudovector under the application of any symmetry operation68,69 and transforms according to the formula

$${{{\bf{s}}}}{{{\mathbf{\Omega }}}}\left({{{\bf{r}}}}\right)=\pm \!\det \left({{{\bf{D}}}}\left({{{\bf{R}}}}\right)\right){{{\bf{D}}}}\left({{{\bf{R}}}}\right){{{\mathbf{\Omega }}}}\left({{{{\bf{s}}}}}^{-1}{{{\bf{r}}}}\right),$$
(2)

where \({{{\mathbf{\Omega }}}}\left({{{\bf{r}}}}\right)\) denotes the pseudovector Berry curvature, \({{{\bf{D}}}}\left({{{\bf{R}}}}\right)\) the three-dimensional representation of a symmetry operation without the translation part and s an arbitrary symmetry operation. That is, the symmetry operations of the magnetic point group will govern the shape of the tensors. Particularly, the ferromagnetic Fe3Sn belongs to the magnetic space groups \(Cmc^{\prime} m^{\prime}\) (BNS: 63.463), \(Cm^{\prime} cm^{\prime}\) (BNS: 63.464) and \(P{6}_{3}/mm^{\prime} c^{\prime}\) (BNS: 194.270) for the magnetic moments of Fe atoms pointing along the [100]-, [010]- and [001]-axis, respectively. Hence the presence of the 2x, 2y and 2z rotation axes for each of magnetic space groups transform the Berry curvature according to:

$$\begin{array}{l}{{{\rm{For}}}}\,{2}_{x}\,{{{\rm{with}}}}\,M| | [100]\\ {{{{\mathbf{\Omega }}}}}_{x}\left({k}_{x},-{k}_{y},-{k}_{z}\right)={{{{\mathbf{\Omega }}}}}_{x}\left({k}_{x},{k}_{y},{k}_{z}\right)\\ {{{{\mathbf{\Omega }}}}}_{y}\left({k}_{x},-{k}_{y},-{k}_{z}\right)=-{{{{\mathbf{\Omega }}}}}_{y}\left({k}_{x},{k}_{y},{k}_{z}\right)\\ {{{{\mathbf{\Omega }}}}}_{z}\left({k}_{x},-{k}_{y},-{k}_{z}\right)=-{{{{\mathbf{\Omega }}}}}_{z}\left({k}_{x},{k}_{y},{k}_{z}\right).\\ {{{\rm{For}}}}\,{2}_{y}\,{{{\rm{with}}}}\,M| | [010]\\ {{{{\mathbf{\Omega }}}}}_{x}\left(-{k}_{x},{k}_{y},-{k}_{z}\right)=-{{{{\mathbf{\Omega }}}}}_{x}\left({k}_{x},{k}_{y},{k}_{z}\right)\\ {{{{\mathbf{\Omega }}}}}_{y}\left(-{k}_{x},{k}_{y},-{k}_{z}\right)={{{{\mathbf{\Omega }}}}}_{y}\left({k}_{x},{k}_{y},{k}_{z}\right)\\ {{{{\mathbf{\Omega }}}}}_{z}\left(-{k}_{x},{k}_{y},-{k}_{z}\right)=-{{{{\mathbf{\Omega }}}}}_{z}\left({k}_{x},{k}_{y},{k}_{z}\right).\\ {{{\rm{For}}}}\,{2}_{z}\,{{{\rm{with}}}}\,M| | [001]\\ {{{{\mathbf{\Omega }}}}}_{x}\left(-{k}_{x},-{k}_{y},{k}_{z}\right)=-{{{{\mathbf{\Omega }}}}}_{x}\left({k}_{x},{k}_{y},{k}_{z}\right)\\ {{{{\mathbf{\Omega }}}}}_{y}\left(-{k}_{x},-{k}_{y},{k}_{z}\right)=-{{{{\mathbf{\Omega }}}}}_{y}\left({k}_{x},{k}_{y},{k}_{z}\right)\\ {{{{\mathbf{\Omega }}}}}_{z}\left(-{k}_{x},-{k}_{y},{k}_{z}\right)={{{{\mathbf{\Omega }}}}}_{z}\left({k}_{x},{k}_{y},{k}_{z}\right).\end{array}$$
(3)

The summation over the whole Brillouin zone forces σy and σz for the magnetization direction along the [100]-axis to vanish, and equivalently σx and σz and σx and σy for the magnetization along [010] and [001], respectively. However, there is no such condition for σx, σy, and σz for magnetization direction along [100], [010], and [001] axes, respectively, and therefore they are allowed to have finite values.

AHC and ANC are proportional to the sum of the Berry curvature of the occupied bands, evaluated in the whole Brilouin zone (BZ), as defined in Eq (11). Since the Berry curvature depends on the energy difference between two adjacent bands, therefore it is expected that Weyl nodes as well as nodal lines, located close to the reference energy, contribute significantly to the total value, as shown in refs. 70,71 and confirmed for MnZn62 and Mn3PdN72, respectively. Explicit band structure search reveals the presence of numerous Weyl nodes and nodal lines within the shaded energy range [−0.118, −0.018] eV of Fig. 4a that are expected to contribute to the total AHC value. In order to identify the origin of the AHC contribution, we split the BZ into 216 cubes, within which the AHC is evaluated (see Fig. 4d). Since the major contribution originates from the diagonals, located within \({k}_{z}\in \left(-0.166,0.000\right)\) (and \({k}_{z}\in \left(0.000,0.166\right)\)), as illustrated in Fig. 4d, it is fruitful to investigate the band gap within this kz range. Taking as an example the kz = −0.131 plane, we plot the difference of the two involved bands as a black and white plot where the black areas correspond to small gap regions whereas white areas to large gap regions (see Fig. 4c). The shape of the gap plot is in complete agreement with the distribution of the AHC within the specified area, demonstrating that small gap regions similar to those within the square \({k}_{x},{k}_{y}\in \left(0.333,0.500\right)\) of Fig. 4c, contribute dominantly to the total AHC value.

Interesting topological transport properties can arise away from the charge-neutral point. One important observation is that the AHC curve of Fe3Sn exhibits a sharp peak of 1308 S cm−1 located at 60 meV below the Fermi level, as shown in Fig. 4. Therefore, an interesting question is whether tuning the Fermi level to match the position of the peak is doable by means of doping. In order to investigate this possibility, we consider (Fe1−xMnx)3Sn for various values of x, with \(x\in \left[0,0.2\right]\), indicating the percentage of Mn doping to the system. By using virtual crystal approximation (VCA) calculations, we compute the AHC curve for different x, as illustrated in Fig. 5a. It is noted that the position of the peak approaches the Fermi level while the Mn dopand concentration is increased and it hits the Fermi energy at approximately x = 0.15 (black curve). The existence of the AHC peak and its location affects the calculated ANC. While the energy of the peak is lower than the Fermi energy (x < 0.15), the ANC is gradually decreased from −2.71 A m−1 K−1 for x = 0 to −1.58 A m−1 K−1 for x = 0.15. Once the energy of the peak gets larger than the Fermi energy (x > 0.15), ANC changes sign and jumps to 3.63 A m−1 K−1 for x = 0.2. Figure 5b shows the calculated ANC curves for various x, demonstrating that Fe3Sn offers an interesting playground of controlling the ANC by doping even with a sign change.

Since (Fe0.85Mn0.15)3Sn exhibits the closest peak to the Fermi energy, its dynamical stability was checked by calculating its phonon dispersion (see Supplementary Fig. 3). To mimic the disorder structure, (Fe0.85Mn0.15)3Sn, a supercell, containing 80 atoms in special quasi-random structure (SQS)73, was generated by the mcsqs code of the ATAT package74. As illustrated in Supplementary Fig. 3, the absence of imaginary modes indicates (Fe0.85Mn0.15)3Sn is dynamically stable.

Tuning the magnetization direction allows easier ANC modifications. In an attempt to tune the AHC and ANC of Fe3Sn, we considered different magnetization directions i.e., along [100], [010], and [001] axes. Our results show no impact of the magnetization direction to the AHC and ANC values along [100] and [010] axis, where the values remain practically unchanged at 757 S cm−1 and −2.58 A m−1 K−1 due to the underlying hexagonal symmetry. On the other hand, a small change is noticed for direction along [001], where the AHC (ANC) is tuned to 676 S cm−1 (−2.06 A m−1 K−1), see Fig. 5c and d (solid curves). Despite the minor changes in the AHC and ANC values at Fermi energy, a larger impact of the altering of the magnetization direction is observed away from the charge-neutral point for both magnetization directions and doping concentrations (solid and dashed curves, respectively). Specifically, the AHC peak of 1308 S cm−1 at 60 meV below the Fermi energy is moved closer to the Fermi energy, at 35 meV below the Fermi energy, and further reduces its maximum value to 886 S cm−1 when the magnetization direction is along the [001]-axis. The outcome of this change is more obvious in the ANC, where the zero value of the [001] direction is located closer to the Fermi energy, being useful for future applications. It is finally noted that for M[001], doping results in ANC sign change (Fig. 5d, red curves).

Based on DFT calculations, the thermodynamical properties of the Fe–Sn system and the topological transport properties of Fe3Sn are studied. Thermodynamic modeling of the Fe–Sn phase diagram has been re-established. The problems concerning invariant reactions of intermetallics are remedied under our newly measured temperature ranges. First-principles phonon calculations with the QHA approach were performed to calculate the thermodynamic properties at finite temperatures. Thermodynamic properties, phonon dispersions of pure elements, and intermetallics were predicted to make up the shortage of experimental data. A set of self-consistent thermodynamic parameters are obtained by the CALPHAD approach. Further, we evaluated the AHC and ANC of Fe3Sn with magnetization direction and doping being perturbations. The calculated AHC of 757 S cm−1 is the largest among all reported members of the Fe–Sn family. It is noted that the nodal lines combined with the extended small gap areas constitute the main contribution to the total AHC and they can further be tuned by doping Mn at the Fe sites, allowing the manipulation of the AHC and ANC values and offering good candidate materials for promising transverse thermoelectric devices. In addition, promising high-throughput calculations75,76 can be performed to search for more intriguing magnetic intermetallic compounds with singular topological transport properties, assisted by automated Wannier function construction77 for transport property calculations.

Methods

First-principles calculations

Our calculations were performed using the generalized gradient approximation (GGA) for the exchange-correlation functional, in the parameterization of Perdew–Burke–Ernzerhof78 for the Vienna ab initio Simulation Package (VASP)79,80. The energy cutoff is set at 600 eV and at least 5000 k-points in the first Brillouin zone with Γ-centered k-mesh were used for the hexagonal lattices (Fe3Sn, FeSn, and Fe5Sn3), while for all the other structures, Monkhorst–Pack grids were used. The energy convergence criterion was set as 10−6 eV, and 10−5 eV Å−1 was set as the tolerance of forces during the structure relaxation. The enthalpy of formation, \({{{\Delta }}}_{f}H({{{{\rm{Fe}}}}{}_{{{{\rm{xSn}}}}}}_{{{{\rm{y}}}}})\), for the FexSny intermetallic compounds was obtained following

$${{{\Delta }}}_{{{{\rm{f}}}}}H({{{{\rm{Fe}}}}{}_{\rm{x}}{\rm{Sn}}}_{{{{\rm{y}}}}})={E}_{{{{{\rm{Fe}}}}{}_{{{{\rm{xSn}}}}}}_{{{{\rm{y}}}}}}-\frac{x}{x+y}{E}_{{{{\rm{Fe}}}}}-\frac{y}{x+y}{E}_{{{{\rm{Sn}}}}},$$
(4)

where all the total energies for the equilibrium phases in their corresponding stable structures were obtained after structural relaxation.

For the phonon calculations, the frozen phonon approach was applied using the PHONOPY package80. The temperature-dependent thermodynamical properties were calculated by using the quasi-harmonic approximation81. The Gibbs free energy G(T, P) at temperature T and pressure P can be obtained from the Helmholtz free energy F(T, V) as follows82:

$$G(P,T)-PV=F(T,V)={E}_{0}(V)+{F}_{{{{\rm{vib}}}}}(V,T)+{F}_{{{{\rm{el}}}}}(V,T)+{F}_{{{{\rm{magn}}}}}(V,T),$$
(5)

where E0(V) is the total energy at zero Kelvin without the zero-point energy contribution, which were determined by fitting of the energies with respect to the volume data using the Birch–Murnaghan equation of state (EOS)83. Fvib corresponds to the lattice vibration contribution to the Helmholtz energy, which can be derived from the phonon density of states (PhDOS), g(ω, V), by using the following equation82:

$${F}_{{{{\rm{vib}}}}}(V,T)={k}_{B}T\int\nolimits_{0}^{\infty }{{{\rm{\ln }}}}\,\left[2\sinh \frac{\hslash \omega }{2{k}_{B}T}\right]g(\omega ,V){{{\rm{d}}}}\omega ,$$
(6)

where kB and are the Boltzmann constant and reduced Planck constant, respectively, and ω denotes the phonon frequency for a given wave vector q. The PhDOS g(ω, V) can be obtained by integrating the phonon dispersion in the Brillouin zone. The third term Fel represents the electronic contribution to the Helmholtz free energy, obtained by84:

$${F}_{{{{\rm{el}}}}}(V,T)={E}_{{{{\rm{el}}}}}(V,T)-T\cdot {S}_{{{{\rm{el}}}}}(V,T)$$
(7)

where Eel(V, T) and Sel(V, T) indicate the electronic energy and electronic entropy, respectively. With the electronic DOS, both terms can be formulated as84:

$${E}_{{{{\rm{el}}}}}(V,T)=\int n\left(\epsilon \right)f\epsilon {{{\rm{d}}}}\epsilon -\int\nolimits_{-\infty }^{{\epsilon }_{{{{\rm{F}}}}}}n(\epsilon ,V){{{\rm{d}}}}\epsilon ,$$
(8)
$${S}_{{{{\rm{el}}}}}(V,T)=-{k}_{B}\int n\epsilon [f{{{\rm{ln }}}}f+(1-f){{{\rm{ln }}}}(1-f)]{{{\rm{d}}}}\epsilon ,$$
(9)

where n(ϵ) is the electronic DOS, f represents the Fermi-Dirac distribution function and ϵF is the Fermi energy.

Finally, based on the original Inden–Hillert–Jarl (IHJ) model38,39 and further improved expression by Xiong40, the magnetic Gibbs energy can be formulated as:

$${G}_{{{{\rm{magn}}}}}=RT{{{\rm{\ln }}}}({\beta }^{* }+1)f(\tau ),$$
(10)

where τ is T/T*, T* is the critical temperature (the Cutie temperature TC for ferromagnetic materials or the Neel temperature TN for antiferromagnetic materials). β* is the effective magnetic moment per atom40. And the relative parameters are summarized in Supplementary Table 2. Note that, we adopted the experimental critical temperatures and calculated magnetic moments.

In order to evaluate AHC, we projected the Bloch wave functions onto maximally localized wannier functions (MLWF) using Wannier90, following ref. 85. A total number of 124 MLWFs, originating from the s, p and d orbitals of Fe atoms and the s and p orbitals of Sn atoms, are used. AHC is obtained by integrating the Berry curvature according to the formula:

$$\begin{array}{l}{\sigma }_{\alpha \beta }=-\frac{{{{{\rm{e}}}}}^{2}}{\hslash }\int \frac{{{{\rm{d}}}}{{{\bf{k}}}}}{{\left(2\pi \right)}^{3}}\sum f\left[\epsilon \left({{{\bf{k}}}}\right)-\mu \right]{{{\Omega }}}_{n,\alpha \beta }\left({{{\bf{k}}}}\right),\\ {{{\Omega }}}_{n,\alpha \beta }\left({{{\bf{k}}}}\right)=-2{{{\rm{Im}}}}\mathop{\sum}\limits_{m\ne n}\frac{\left\langle {{{\bf{k}}}}n| {v}_{\alpha }| {{{\bf{k}}}}m\right\rangle \left\langle {{{\bf{k}}}}m| {v}_{\beta }| {{{\bf{k}}}}n\right\rangle }{{\left[{\epsilon }_{m}\left({{{\bf{k}}}}\right)-{\epsilon }_{n}\left({{{\bf{k}}}}\right)\right]}^{2}},\end{array}$$
(11)

with μ, f, n, m, \({\epsilon }_{{{{\rm{n}}}}}\left({{{\bf{k}}}}\right)\), \({\epsilon }_{{{{\rm{m}}}}}\left({{{\bf{k}}}}\right)\), and vα being the Fermi level, the Fermi-Dirac distribution function, the occupied Bloch band, the empty Bloch band, their corresponding energy eigenvalues and the Cartesian component of the velocity operator. The integration is performed on a 270 × 270 × 350 mesh using Wanniertools86. ANC is evaluated using an in-house developed Python script, following the formula:

$${\alpha }_{\alpha \beta }=-\frac{1}{{{{\rm{e}}}}}\int {{{\rm{d}}}}\epsilon \frac{\partial f}{\partial \mu }{\sigma }_{\alpha \beta }\left(\epsilon \right)\frac{\epsilon -\mu }{T}$$
(12)

where T, e, and ϵ are the temperature, the electronic charge and the energy point within the integration energy window, respectively. An energy grid of 1000 points within the window \(\left[-0.5,0.5\right]{{{\rm{eV}}}}\) with respect to the Fermi level was chosen.

Mn doping at Fe sites is performed by using the virtual crystal approximation (VCA)87 as implemented in VASP88,89. In this approximation, virtual, fictitious atoms that behave in between the parent atoms are inserted. VCA techniques have been used to describe prototypical doped γ − FeMn systems90 and magnetic anisotropy energy of L10FePt and Fe1−xMnxPt91 and additionally topological transport properties of Co3−xNixSn2S292 and Fe3Co alloys93.

CALPHAD modeling

Pure elements

The Gibbs free energies for pure Fe and Sn were taken from the Scientific Group Thermodata Europe (SGTE) pure element database94, which was described by:

$${}^{\circ }{G}_{i}^{\phi }(T)={G}_{i}^{\phi }(T)-{H}_{i,{{{\rm{SER}}}}}(298.15\,K)=a+bT+cT{{{\rm{\ln }}}}(T)+d{T}^{2}+e{T}^{3}+f{T}^{-1}+g{T}^{7}+h{T}^{-9}\,,$$
(13)

where i represents the pure elements Fe or Sn, Hi,SER(298.15K) is the molar enthalpy of element i at 298.15 K in its standard element reference (SER) state, and a to h are known coefficients.

Solution phases

The solution phases, Liquid, BCC_A2, FCC_A1 and BCT_A5 phases are described using the substitutional solution model, with the corresponding molar Gibbs free energy formulated as:

$${G}_{m}^{\varphi }={x}_{{{{\rm{Fe}}}}}{G}_{{{{\rm{Fe}}}}}^{\varphi }(T)+{x}_{{{{\rm{Sn}}}}}{G}_{{{{\rm{Sn}}}}}^{\varphi }(T)+RT({x}_{{{{\rm{Fe}}}}}{{{\rm{\ln }}}}{x}_{{{{\rm{Fe}}}}}+{x}_{Sn}{{{\rm{\ln }}}}{x}_{{{{\rm{Sn}}}}})+{G}^{ex}+{G}^{{{{\rm{magn}}}}},$$
(14)

where xFe and \({x}_{{{{\rm{Sn}}}}}\) are the mole fraction of Fe and Sn in the solution, respectively. Taken from SGTE94, \({G}_{i}^{\varphi }\) denotes the molar Gibbs free energy of pure Fe and Y in the structure φ at the given temperature. Gex denotes the excess Gibbs energy of mixing, which measures the deviation of the actual solution from the ideal solution behavior, modeled using a Redlich–Kister polynomial95:

$${G}^{ex}={x}_{{{{\rm{Fe}}}}}{x}_{{{{\rm{Sn}}}}}{\mathop{\sum }\limits_{j = 0}^{n}}^{(j)}{L}_{{{{\rm{Fe,Sn}}}}}^{\varphi }{({x}_{{{{\rm{Fe}}}}}-{x}_{{{{\rm{Sn}}}}})}^{j}.$$
(15)

The jth interaction parameter between Fe and Sn is described by \({}_{}^{(j)}{{{{\rm{L}}}}}_{{{{\rm{Fe,Sn}}}}}^{\varphi }\), which is modeled in terms of a*+b*T.

Stoichiometric intermetallic compounds

Fe5Sn3, Fe3Sn2, Fe3Sn, FeSn, and FeSn2 were considered as stoichiometric phases. The Gibbs free energies per mole atom of these phases were thus expressed as follows:

$${G}_{m}^{{{{{\rm{Fe}}}}{}_{{{{\rm{xSn}}}}}}_{{{{\rm{y}}}}}}=\frac{x}{x+y}{G}_{{{{\rm{Fe,SER}}}}}+\frac{y}{x+y}{G}_{{{{\rm{Sn,SER}}}}}+{{\Delta }}{G}_{f}^{{{{{\rm{Fe}}}}{}_{{{{\rm{xSn}}}}}}_{{{{\rm{y}}}}}}(T)\,,$$
(16)

where \({{\Delta }}{G}_{f}^{{{{{\rm{Fe}}}}{}_{{{{\rm{xSn}}}}}}_{{{{\rm{y}}}}}}(T)\) is the Gibbs free energy of formation of the stoichiometric compound FexSny which can be expressed as:

$${{\Delta }}{G}_{f}^{{{{{\rm{Fe}}}}{}_{{{{\rm{xSn}}}}}}_{{{{\rm{y}}}}}}(T)={A}_{3}+{B}_{3}T\,,$$
(17)

where the coefficients A3, B3 are the parameters to be optimized. Since there is no experimental data of the thermodynamic properties for such intermetallic phases, the calculated enthalpies of formation for these phases from DFT calculations were treated as initial values of the coefficient A3 in Eq. (17) in the present optimization.