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

Pressure effects on the electronic structure, phonons, and superconductivity of noncentrosymmetric ThCoC2

Gabriel Kuderowicz    Paweł Wójcik    Bartlomiej Wiendlocha Faculty of Physics and Applied Computer Science, AGH University of Science and Technology, Aleja Mickiewicza 30, 30-059 Krakow, Poland [email protected]
Abstract

The electronic structure, phonons, electron-phonon coupling, and superconductivity are theoretically studied in noncentrosymmetric superconductor ThCoC2 as a function of pressure, in the pressure range 0 - 20 GPa. We found that the electronic band splitting induced by the spin-orbit coupling is enhanced under pressure. In spite of the overall stiffening of the crystal lattice, the electron-phonon coupling constant λ\lambda increases with pressure, from 0.583 at 0 GPa to 0.652 at 20 GPa. If the isotropic Eliashberg electron-phonon coupling theory is used to simulate the effect on the critical temperature TcT_{c}, such an increase in λ\lambda results in a substantial increase of TcT_{c} from 2.5 K at 0 GPa to 4 K at 20 GPa. This shows that examining the effect of pressure offers a chance to resolve the pairing mechanism in ThCoC2.

I Introduction

ThCoC2 is a low-temperature noncentrosymmetric superconductor with the critical temperature Tc2.5T_{c}\simeq 2.5 K Grant et al. (2014) and a puzzling nature of the superconducting state. The pairing mechanism has not been resolved yet, and the thermodynamic properties of the superconducting phase strongly deviate from the predictions of the Bardeen-Cooper-Shrieffer theory (BCS) Bardeen et al. (1957). Among them, we may list: (i) a non-exponential temperature dependence of the electronic specific heat Ce(T)C_{e}(T), with the reduced specific heat jump of ΔCe/γTc=0.86\Delta C_{e}/\gamma T_{c}=0.86, much lower than the BCS value of 1.43 Grant et al. (2014); (ii) the positive curvature of the temperature dependence of the upper critical magnetic field Hc2(T)H_{c2}(T); (iii) a non-s-wave temperature dependence of the magnetic field penetration depth λL(T)\lambda_{L}(T) Bhattacharyya et al. (2019). The measured λL(T)\lambda_{L}(T) was fitted Bhattacharyya et al. (2019) assuming a dd-wave nodal superconducting gap. Moreover, spin fluctuations were suggested to be the pairing mechanism in ThCoC2 Bhattacharyya et al. (2019).

Very interesting results were also obtained when the material was doped with nickel Grant et al. (2017). Critical temperature in ThCo1-xNixC2 increased with xx reaching Tc=12T_{c}=12 K for x=0.4x=0.4, with no signs of magnetism and a gradual suppressing of the non-BCS features of the superconducting phase towards a more conventional fully gapped superconductor Grant et al. (2017).

To shed more light on the nature of superconductivity in ThCoC2, we have recently presented two theoretical works Kuderowicz et al. (2021); Kuderowicz and Wiendlocha (2022). The electronic structure, phonons and the electron-phonon coupling were computed in Ref. Kuderowicz et al. (2021) using the density functional perturbation theory Giannozzi et al. (2009, 2017); Baroni et al. (2001). We found that strong spin-orbit coupling combined with the lack of inversion symmetry leads to a large band splitting with an average value of ΔE¯SOC\overline{\Delta E}_{SOC}\approx 150 meV, comparable to that observed in the triplet superconductors CePt3Si and Li2Pt3B (see Smidman et al. (2017) and references therein). That indicates that the conventional ss-wave gap symmetry will be strongly perturbed, as the pairing inside the spin-split bands in the strong spin-orbit coupling limit requires the odd parity of the gap with respect to the 𝐤𝐤\mathbf{k}\rightarrow-\mathbf{k} Samokhin et al. (2004). As far as the pairing mechanism is concerned, the computed electron-phonon coupling constant λ=0.59\lambda=0.59 was large enough to propose the electron-phonon interaction to be responsible for superconductivity in ThCoC2 Kuderowicz et al. (2021). However, our calculations within the isotropic Eliashberg formalism Eliashberg (1960) did not explain the experimental non-BCS features of ThCoC2, supporting the non-ss-wave picture of its superconductivity, but with the possibility of the electron-phonon coupling mechanism Kuderowicz et al. (2021).

Electron-phonon coupling hypothesis was further strengthened by calculations of the effect of Ni doping presented in Ref. Kuderowicz and Wiendlocha (2022). Using the simplified rigid muffin-tin approximation (RMTA), the evolution of the electron-phonon coupling constant in ThCo1-xNixC2 was analyzed. Although due to the usage of the RMTA approach, the calculated values of λ(x)\lambda(x) were systematically underestimated when compared to the experimental estimates, the strong increase of λ(x)\lambda(x) was found, which remained in a qualitative agreement with the experimental observation of the increase in Tc(x)T_{c}(x) under the assumption of the electron-phonon coupling mechanism of superconductivity in ThCo1-xNixC2.

In this work, we study the effects of pressure on the electronic structure, phonons, electron-phonon interaction, and superconductivity in ThCoC2, which should help to determine the pairing interaction if the experimental studies are undertaken. In this context, it is important to recall the interesting pressure dependence of superconductivity in the sister isostructural superconductor LaNiC2\mathrm{LaNiC_{2}}. The external pressure initially increases the critical temperature from 2.8 K at 0 GPa to 3.8 at 4 GPa and later decreases TcT_{c} to suppress superconductivity above 7 GPa Katano et al. (2014). The calculated electron-phonon coupling parameter λ\lambda and critical temperature TcT_{c}, on the other hand, were found to monotonously increase with pressure Wiendlocha et al. (2016), in agreement with the initial increase in Tc(p)T_{c}(p) but showing the non-phononic origin of the suppression of superconductivity at larger pressures.

II Computational details

Calculations were done for p = 5, 10, 15, and 20 GPa, and to be consistent with our earlier work for ambient pressure, all computational details were kept unchanged. Some of the results for p = 0 GPa are recalled here for a convenient comparison. The Quantum Espresso package Giannozzi et al. (2009, 2017) was used, with the Perdew-Burke-Ernzerhof (PBE) generalized gradient approximation for the exchange-correlation functional Perdew et al. (1996) and the Rappe-Rabe-Kaxiras-Joannopoulos (RRKJ) ultrasoft pseudopotentials ThC ; Dal Corso (2014). Electronic structure was calculated in the scalar-relativistic and fully relativistic way (i.e., including the spin-orbit coupling). Phonons and electron-phonon coupling functions were calculated in the scalar-relativistic way, as the inclusion of SOC had no important effect on the phonon structure nor on the electron-phonon interaction Kuderowicz et al. (2021).

Table 1: Evolution of cell parameters with applied pressure. Atomic positions (in crystal coordinates) are in a form: Th (0,0,u), Co (0.5,0,v), C (0.5,±\pmy,z).
p (GPa) a (Å) b (Å) c (Å) u v y z
0 3.8214 4.5376 6.0708 -0.0014 0.6041 0.1561 0.3007
5 3.7781 4.5120 6.0219 -0.0011 0.6037 0.1569 0.3007
10 3.7392 4.4885 5.9786 -0.0009 0.6034 0.1575 0.3007
15 3.7031 4.4670 5.9408 -0.0006 0.6032 0.1581 0.3007
20 3.6705 4.4467 5.9063 -0.0004 0.6030 0.1586 0.3007
Refer to caption
Figure 1: The noncentrosymmetric unit cell of ThCoC2\mathrm{ThCoC_{2}} visualized with xcrysden Kokalj (2003). Numbers are used in Table 2 to define interatomic distances. In the right panel we may distinguish the layers of Th separated by the layers of Co-C rings.

Crystal unit cell parameters and atomic positions were theoretically optimized, starting from the p = 0 GPa experimental values Gerss and Jeitschko (1986). Electronic structure was then calculated using a Monkhorst-Pack grid of 12312^{3} k-points, whereas the Fermi surface was calculated on a 18318^{3} mesh. Wavefunction energy and charge density cutoffs were set to 130130 Ry and 13001300 Ry, respectively. Dynamical matrices were calculated on a 4x4x4 q-point mesh and the interatomic force constants (IFC) and phonon dispersion relations were computed using the Fourier interpolation technique. Finally, the Eliashberg electron-phonon interaction functions α2F(ω)\alpha^{2}F(\omega) were computed using the self-consistent first order variation of the crystal potential Baroni et al. (2001). The spectral functions α2F(ω)\alpha^{2}F(\omega) were then used to determine the pressure dependence of the electron-phonon coupling (EPC) parameter λ\lambda. The thermodynamic properties of the superconducting phase were determined using the isotropic Eliashberg formalism to see how they evolve under pressure, but one has to be aware that the superconducting state in ThCoC2 at p = 0 GPa could not be accurately described within this formalism. The isotropic theory is used here to simulate the trend in Tc(p)T_{c}(p), as well as to prepare the grounds for a discussion of future experimental results.

Refer to caption
Figure 2: (a) Birch-Murnaghan equation of state fitted to ThCoC2\mathrm{ThCoC_{2}} unit cell volumes under pressure. (b) Evolution of cell parameters (lines are guides for the eye).
Table 2: Distances between neighboring atoms (Å) in ThCoC2\mathrm{ThCoC_{2}} under pressure p (GPa). Atoms are numbered as shown in Fig. 1. Δd=d20GPad0GPa\mathrm{\Delta d=d_{20\ GPa}-d_{0\ GPa}}.
p C1 - C2 - C2 - Co1 - Co1 - Th1 - Th2 -
C2 Th1 Th2 C2 C3 Co1 Co1
0 1.4165 2.7442 2.7412 1.9646 1.9734 3.0641 3.0345
5 1.4156 2.7184 2.7152 1.9504 1.9571 3.0384 3.0094
10 1.4142 2.6954 2.6920 1.9377 1.9428 3.0154 2.9868
15 1.4128 2.6746 2.6705 1.9260 1.9306 2.9949 2.9659
20 1.4108 2.6555 2.6513 1.9153 1.9194 2.9759 2.9469
Δd\mathrm{\Delta d} -0.0057 -0.0887 -0.0899 -0.0493 -0.0540 -0.0882 -0.0876
ΔdGPa\mathrm{\frac{\Delta d}{GPa}} -0.0003 -0.0044 -0.0045 -0.0025 -0.0027 -0.0044 -0.0044

III Results and discussion

Refer to caption
Figure 3: Electronic dispersion relation of ThCoC2\mathrm{ThCoC_{2}} under selected pressures. In panels (a-c) the scalar-relativistic and fully relativistic results are plotted for 0, 10 and 20 GPa, while in (d) the relativistic results for all pressures are gathered to visualize their relative differences.

III.1 Crystal structure

ThCoC2 crystallizes in the base-centered orthorhombic structure Amm2, spacegroup no. 38, shown in Fig. 1. In the structure we may distinguish the layers of Th separated by the layers of Co-C rings, stacked along aa axis. Evolution of the unit cell parameters is presented in Table 1 and Fig. 2. The strongest compression of the unit cell under pressure is seen in the aa direction, i.e. perpendicular to the aforementioned Th and Co-C layers, whereas the most resistant to pressure is bb direction, along the covalent Co-C and C-C bonds. As a consequence, b/ab/a and c/ac/a ratios increase with pressure. Atomic positions change only slightly, and to illustrate how the crystal structure evolved, the distances between neighboring atoms are shown in Table 2. The carbon-carbon bond is the stiffest as its bond length shortens an order of magnitude less than in the other pairs. Distances between cobalt and carbon decrease roughly two times less than for the carbon - thorium and thorium - cobalt pairs.

The change in the unit cell volume with pressure is shown in Fig. 2(a) where it is fitted using the Birch-Murnaghan equation of state Birch (1947). The computed bulk modulus is equal to B=188B=188 GPa.

Refer to caption
Figure 4: The Fermi surface (FS) of ThCoC2\mathrm{ThCoC_{2}} calculated with SOC. Top row shows FS sheets for 0 GPa (a-d), middle for 10 GPa (e-h) and bottom for 20 GPa (i-l). FS visualized using fermisurfer Kawamura (2019). Color marks the Fermi velocity.

III.2 Electronic structure

Evolution of the electronic dispersion relations in ThCoC2 with pressure is presented in Fig. 3. Panels (a-c) show dispersion relations for 0, 10 and 20 GPa, while panel (d) gathers all results in one graph to visualize their relative differences. In a similar way we will present changes in electronic densities of states, phonon dispersions, phonon densities of states and Eliashberg functions. As discussed in Kuderowicz et al. (2021), at ambient pressure in the scalar-relativistic case, two bands cross the Fermi level forming two sheets of the Fermi surface (FS), with one large, dominating FS sheet and the other consisting of two small electron pockets. When the spin-orbit coupling is taken into account, due to the lack of the inversion center, the band structure is split, forming four FS sheets, shown for p = 0, 10, and 20 GPa in Fig. 4. The largest change in E(𝐤)E({\bf k}) relation due to pressure is observed near Γ\mathrm{\Gamma} and T points. The electron band at Γ\mathrm{\Gamma}, which is just above EFE_{F} for p = 0 GPa, is pushed down and eventually crosses the Fermi level, contributing to FS by forming an additional tube along kzk_{z} direction, seen in Fig. 4 for p = 10 GPa.

Orbital characters of the main Fermi surface sheet for pressures of 0, 10 and 20 GPa are shown in Supplemental Material sup . Contribution to Fermi surface is mostly from Th-6d, Co-3d and C-2p orbitals, with some contribution from Th-f states, in agreement with contributions to total DOS at EFE_{F} discussed below (see Fig. 5). The anisotropy of orbital character of FS is moderate and is enhanced under pressure – the tube which is formed in the central part of FS above 10 GPa has mostly C-p and Th-f character, which were more uniformly distributed at ambient pressure. That may enhance the anisotropic properties of the superconducting phase under pressure in ThCoC2.

The average value of E(𝐤)E({\bf k}) splitting due to SOC for the states located around the Fermi level as a function of pressure is shown in Fig. 6(b). The ΔE¯SOC\overline{\Delta E}_{SOC} starts to increase above 5 GPa when the additional, Γ\Gamma-centered tube starts to contribute to FS. The growth is approximately linear with the a rate of 1.5 meV/GPa, reaching 185 meV at 20 GPa111The average splitting at p = 0 GPa reported here is 162 meV is slightly larger than the (rounded down) value of 150 meV reported in Ref.Kuderowicz et al. (2021) due to a denser kk-mesh used to calculate the average.. That suggests that the unconventional features of superconductivity in ThCoC2, likely driven by the strong antisymmetric spin-orbit coupling, should be even more pronounced under elevated pressure.

The electronic density of states (DOS) near the Fermi level is presented in Fig. 5 and the change of DOS(EFE_{F}) with pressure in Fig. 6(a) and Table 3. Under the pressure, the DOS curve around EFE_{F} is flattened, due to pressure-enhanced hybridization we can observe lowering of maxima and rising of minima of DOS(EE). The DOS(EFE_{F}) gradually decreases with pressure with a ratio of 1.1×102-1.1\times 10^{-2} eV/1{}^{-1}/GPa which compares to 2.7×103-2.7\times 10^{-3} eV/1{}^{-1}/GPa in LaNiC2 Wiendlocha et al. (2016), thus, the effect is much stronger in ThCoC2. The largest contributions to the total DOS at the Fermi level are from the 6d Th, 3d Co, and 2p C states, with a noticeable contribution from 5f states of Th. The major contributors to DOS(EF)(E_{F}) do not change with pressure, however the relative importance of Th-5f states is growing with increasing pressure, due to the above-mentioned increase in the Fermi surface area with the Th-5f orbital character.

Table 3: Density of states of ThCoC2\mathrm{ThCoC_{2}} at the Fermi energy.
N(EF)N(E_{F}) p 0 5 10 15 20
(eV1)\mathrm{(eV^{-1})} (GPa)
w/o SOC total 2.07 2.02 2.00 1.91 1.85
with SOC total 2.14 2.04 2.03 1.92 1.86
Th-6d 0.59 0.55 0.53 0.50 0.48
Th-5f 0.25 0.26 0.27 0.26 0.25
Co-3d 0.76 0.72 0.69 0.66 0.64
C2-2p 0.44 0.44 0.45 0.43 0.42
Refer to caption
Figure 5: Total and partial electronic density of states of ThCoC2\mathrm{ThCoC_{2}} under selected pressures. Magenta line marks the Fermi energy. In panels (a-c) fully relativistic total and partial DOS results are plotted for 0, 10 and 20 GPa and compared to the scalar-relativistic total DOS, while in (d) the relativistic results for all pressures are gathered to visualize their relative differences.

III.3 Phonons and electron-phonon coupling

As we found for the zero pressure case Kuderowicz et al. (2021), spin-orbit coupling makes no significant changes for the phonon spectrum of ThCoC2, as well as for the electron-phonon interactions. The computed scalar-relativistic value of the EPC constant for p = 0, λ=0.583\lambda=0.583, is only slightly smaller than the relativistic λ=0.590\lambda=0.590. Because of this, we restricted our calculations of the pressure evolution of the dynamic properties and EPC in ThCoC2 to the scalar-relativistic case. To make sure that SOC may be also neglected for phonons under pressure, giving the correct evolution of λ(p)\lambda(p), calculations including SOC were done for a single case of p = 10 GPa and also here no important differences were found (see Supplemental Material sup for further details). Similarly to the p = 0 case, at 10 GPa the scalar-relativistic λ=0.609\lambda=0.609 is slightly smaller than the relativistic λ=0.618\lambda=0.618. The relative effect of pressure on the EPC parameter between 10 GPa and 0 GPa, λ(10)/λ(0)\lambda(10)/\lambda(0), is 4.7% increase, when computed with SOC, and 4.5% increase, when SOC is neglected, thus the scalar-relativistic approximation is sufficient to drive conclusions for the evolution of λ\lambda with pressure.

Refer to caption
Figure 6: (a) Total and partial electronic density of states at the Fermi energy of ThCoC2\mathrm{ThCoC_{2}} under selected pressures. (b) Average SOC splitting of the large Fermi surface sheet. Lines are guides for the eye.

Figure 7 presents the phonon dispersion curves ω(𝐪)\omega(\mathbf{q}) and the phonon density of states F(ω)F(\omega) for selected pressures. The characteristic structure of the phonon spectrum, driven by the large difference in atomic masses between Th, Co, and C, is the separation into three regions, seen in the partial F(ω)F(\omega) plots. The lowest-frequency part has the majority of Th vibrations, next Co phonon branches dominate and the high frequency part is contributed by carbon atoms. The highest single mode near 3535 THz (p = 0 GPa) is contributed by the C-C bond-stretching mode.

Refer to caption
Figure 7: Phonon dispersion relations and phonon densities of states of ThCoC2\mathrm{ThCoC_{2}} under selected pressures. In panels (a-c) results are plotted for 0, 10 and 20 GPa and in the phonon dispersion plots blue shading shows the phonon linewidths (multiplied by 15). In panel (d) results for all pressures are gathered to visualize their relative evolution with pressure, here the phonon linewidths are not shown.

Under elevated pressure, most of the phonon branches move towards higher frequencies, which is especially visible for carbon modes. The average phonon frequency increases with an approximate rate of 0.057 THz/GPa, reaching 11.76 THz at 20 GPa, as shown in Fig. 8(a). On the other hand, near R, T, and Z points, the lowest acoustic mode is softened. Similar softening was found in LaNiC2 Wiendlocha et al. (2016). The real-space atomic displacements for these phonon modes are visualized in the Supplemental Material sup . In T point, where the effect is the strongest, Th vibrates along the cc axis, whereas Co and C move out of phase, along aa.

Before moving to the discussion of the electron-phonon coupling, it is worth to recall how the quantities important for the interaction strength depend on the phonon frequencies. The electron-phonon coupling matrix elements g𝐪ν(𝐤,i,j)g_{{\bf q}\nu}({\bf k},i,j) are defined as Wierzbowska et al. (2005); Heid et al. (2010); Giustino (2017)

g𝐪ν(𝐤,i,j)==s2Msω𝐪νψi,𝐤+𝐪|dVSCFdu^νsϵ^νs|ψj,𝐤,\begin{split}g_{{\bf q}\nu}({\bf k},i,j)&=\\ =\sum_{s}&\sqrt{{\hbar\over 2M_{s}\omega_{{\bf q}\nu}}}\langle\psi_{i,{\bf k+q}}|{dV_{\rm SCF}\over d{\hat{u}}_{\nu s}}\cdot\hat{\epsilon}_{\nu s}|\psi_{j,{\bf k}}\rangle,\end{split} (1)

where ss enumerates atoms in the unit cell, i,ji,j are band indexes, MsM_{s} is the atomic mass, dVSCFdu^νs{dV_{\rm SCF}\over d{\hat{u}}_{\nu s}} is a change of electronic potential calculated in self-consistent cycle due to the displacement u^νs{\hat{u}}_{\nu s} of an atom ss, ϵ^νs\hat{\epsilon}_{\nu s} is the ν\nu-th phonon polarization vector and ψi,𝐤\psi_{i,{\bf k}} is the electronic wave function. In eq.(1) phonon frequency appears under the square root in the denominator. The square of matrix elements are integrated over the Fermi surface and multiplied by frequency to contribute to the phonon linewidths γ𝐪ν\gamma_{{\bf q}\nu}: Wierzbowska et al. (2005); Grimvall (1981); Giustino (2017):

γ𝐪ν=2πω𝐪νijd3kΩBZ|g𝐪ν(𝐤,i,j)|2×δ(E𝐤,iEF)δ(E𝐤+𝐪,jEF).\begin{split}\gamma_{{\bf q}\nu}=&2\pi\omega_{{\bf q}\nu}\sum_{ij}\int{d^{3}k\over\Omega_{\rm BZ}}|g_{{\bf q}\nu}({\bf k},i,j)|^{2}\\ &\times\delta(E_{{\bf k},i}-E_{F})\delta(E_{{\bf k+q},j}-E_{F}).\end{split} (2)

This multiplication cancels the direct frequency dependence, thus the linewidths do not directly depend on ω𝐪ν\omega_{{\bf q}\nu}, rather than on the wavefunction matrix elements, which enter eq.(1).

The summation of the phonon linewidths divided by frequency gives the Eliashberg function α2F(ω)\alpha^{2}F(\omega):

α2F(ω)=12πN(EF)𝐪νδ(ωω𝐪ν)γ𝐪νω𝐪ν,\alpha^{2}F(\omega)={1\over 2\pi N(E_{F})}\sum_{{\bf q}\nu}\delta(\omega-\omega_{{\bf q}\nu}){\gamma_{{\bf q}\nu}\over\hbar\omega_{{\bf q}\nu}}, (3)

which is then inversely proportional to ω𝐪ν\omega_{{\bf q}\nu}. Finally, the EPC constant λ\lambda is calculated as the integral of the Eliashberg function, again divided by frequency Grimvall (1981),

λ=20ωmaxα2F(ω)ωdω,\lambda=2\int_{0}^{\omega_{\rm max}}\frac{\alpha^{2}F(\omega)}{\omega}\text{d}\omega, (4)

which results in the proportionality of λγ𝐪νω𝐪ν2\lambda\propto\frac{\gamma_{{\bf q}\nu}}{\omega_{{\bf q}\nu}^{2}}, thus for majority of the electron-phonon superconductors pressure-induced lattice stiffening (increase in ω𝐪ν\omega_{{\bf q}\nu}) dominates, reducing λ\lambda.

Table 4: (Top part) Phonon modes’ contributions λν\lambda_{\nu} to the electron-phonon coupling constant λ=νλν\lambda=\sum_{\nu}\lambda_{\nu} of ThCoC2\mathrm{ThCoC_{2}} for 0 and 20 GPa. Δλ\Delta\lambda is the difference between 20 GPa and 0 GPa results. (Bottom part) Mean phonon frequencies (in THz): global ω\left<\omega\right> and for each of the phonon modes ων\left<\omega\right>_{\nu}.
p (GPa) λ\lambda λ1\lambda_{1} λ2\lambda_{2} λ3\lambda_{3} λ4\lambda_{4} λ5\lambda_{5} λ6\lambda_{6} λ7\lambda_{7} λ8\lambda_{8} λ9\lambda_{9} λ10\lambda_{10} λ11\lambda_{11} λ12\lambda_{12}
0 0.5835 0.0755 0.0396 0.0420 0.0856 0.0658 0.0398 0.0814 0.1020 0.0104 0.0166 0.0184 0.0064
20 0.6521 0.1391 0.0412 0.0457 0.0974 0.0607 0.0395 0.1017 0.0806 0.0079 0.0152 0.0167 0.0064
Δλ\Delta\lambda 0.0686 0.0636 0.0016 0.0037 0.0118 -0.0051 -0.0003 0.0203 -0.0214 -0.0025 -0.0014 -0.0017 0.00
p (GPa) ω\left<\omega\right> ω1\left<\omega\right>_{1} ω2\left<\omega\right>_{2} ω3\left<\omega\right>_{3} ω4\left<\omega\right>_{4} ω5\left<\omega\right>_{5} ω6\left<\omega\right>_{6} ω7\left<\omega\right>_{7} ω8\left<\omega\right>_{8} ω9\left<\omega\right>_{9} ω10\left<\omega\right>_{10} ω11\left<\omega\right>_{11} ω12\left<\omega\right>_{12}
0 10.6188 2.3185 2.6588 3.3310 4.6398 5.4076 7.2053 8.9207 10.1717 14.5968 16.0684 17.3748 34.7618
20 11.7611 2.4635 2.9604 3.8626 5.0101 6.1745 8.3831 10.1379 10.6632 16.1909 18.7990 20.1257 36.4349
Refer to caption
Figure 8: (a) Average phonon frequency, (b) logarithmic average frequency ωln\omega_{\rm ln} [eq.(8, (c) integral II [eq.(5 and (d) electron-phonon coupling constant λ\lambda of ThCoC2\mathrm{ThCoC_{2}}.

To have a convenient single-number parameter which shows how the electronic contribution to the electron-phonon coupling changes with pressure, cutting off the direct dependence on the phonon frequency, one might calculate the sum of all phonon linewidths. The more straightforward way to do it is to calculate the integral of the Eliashberg function multiplied by frequency:

I=0ωmaxωα2F(ω)dω,I=\int_{0}^{\omega_{\rm max}}\omega\cdot\alpha^{2}F(\omega){\rm d}\omega, (5)

as it carries similar physical information. Substituting eq.(1-3) to the above formula, one can see that II does not explicitly depend on phonon frequency Gutowska et al. (2021):

I=12πN(EF)0ωmax𝑑ω𝐪νδ(ωω𝐪ν)γqν=1N(EF)0ωmax𝑑ω𝐪νδ(ωω𝐪ν)×ijd3kΩBZ|s12Msψi,𝐤+𝐪|dVSCFdu^νsϵ^νs|ψj,𝐤|2×δ(E𝐤,iEF)δ(E𝐤+𝐪,jEF).\begin{split}I=&\frac{1}{2\pi\hbar N(E_{F})}\int_{0}^{\omega_{max}}d\omega\sum_{{\bf q}\nu}\delta(\omega-\omega_{{\bf q}\nu})\gamma_{\textbf{q}\nu}\\ =&\frac{1}{N(E_{F})}\int_{0}^{\omega_{max}}d\omega\sum_{{\bf q}\nu}\delta(\omega-\omega_{{\bf q}\nu})\\ &\times\sum_{ij}\int\frac{d^{3}k}{\Omega_{BZ}}\left|\sum_{s}\frac{1}{\sqrt{2M_{s}}}\langle\psi_{i,{\bf k+q}}|{dV_{\rm SCF}\over d{\hat{u}}_{\nu s}}\cdot\hat{\epsilon}_{\nu s}|\psi_{j,{\bf k}}\rangle\right|^{2}\\ &\times\delta(E_{{\bf k},i}-E_{F})\delta(E_{{\bf k+q},j}-E_{F}).\\ \end{split} (6)

In Fig. 7 we see that some of the phonon linewidths are increasing with pressure, especially for the softened acoustic part of the spectrum between T, Z, and A points, which shows that the softening is associated with the enhanced electron-phonon coupling. On the other hand, for the optic Co and C modes, e.g., in the X-Γ\Gamma direction, γ𝒒ν\gamma_{\boldsymbol{q}\nu} are decreasing. Eliashberg functions for selected pressures are presented in Fig. 9 and an increase in the lower-frequency part can be seen. The overall effect on the electron-phonon coupling is the resultant of competing factors: the increase of the ”electronic” II integral, plotted in Fig. 8(c), and increase in the average phonon frequency, shown in Fig. 8(a). The tendency towards stronger coupling takes over and the EPC constant λ\lambda is increasing with pressure with a rate of about 3.5×1033.5\times 10^{-3}/GPa, as shown in Fig. 8(d). This is 20% slower than found for LaNiC2, where it was equal to 4.4×1034.4\times 10^{-3}/GPa Wiendlocha et al. (2016).

Analyzing the increase in λ\lambda in more details, Fig.  10 additionally compares the p = 0 and 20 GPa Eliashberg functions decoupled over all the phonon modes [panels (a,b)], as well as α2F(ω)/ω\alpha^{2}F(\omega)/\omega in panel (c). The substantial increase of the Eliashberg function for the first acoustic mode is well visible. For other modes generally the increase in phonon frequencies is accompanied by the increase in α2F(ω)\alpha^{2}F(\omega) values due to the larger phonon linewidths. This gives a larger α2F(ω)/ω\alpha^{2}F(\omega)/\omega function, whose integral determines the value of λ\lambda via eq.(4). Electron-phonon coupling constants for each of the 12 phonon modes λi\lambda_{i}, computed based on the data presented in Fig. 10(a,b) are collected in Table 4. The pressure changes in the cumulative electron-phonon coupling constant λ(ω)\lambda(\omega) are shown in Fig. 11. What we can conclude is that the increase in λ\lambda is mainly driven by the increase in λ1\lambda_{1} which is contributed by the lowest acoustic phonon mode, the one which softens under pressure in some regions of the Brillouin zone. As a consequence, the average frequency of this mode (see ω1\langle\omega\rangle_{1} in Table 4) increases only a little, allowing for the increase in λ1\lambda_{1}. Changes in λν\lambda_{\nu} of the other modes generally compensate each other when summed over, thus the change in λ1\lambda_{1} is a key factor for the λ\lambda enhancement. The increase in λ\lambda will tend to increase the critical temperature TcT_{c}.

Refer to caption
Figure 9: Panels (a-c): Eliashberg function of ThCoC2\mathrm{ThCoC_{2}} under pressures of 0, 10 and 20 GPa. Orange shading is the phonon density of states renormalized to have the same area as the corresponding Eliashberg function. In panel (d) Eliashberg functions for all pressures are gathered to visualize their relative evolution with pressure.

The second important parameter in terms of TcT_{c} is the characteristic phonon frequency, which sets the energy window for the electron-phonon interaction, and to which TcT_{c} is proportional to. That is the Debye frequency in the BCS theory and in the McMillan TcT_{c} formula McMillan (1968). The pressure-induced increase in the average phonon frequency suggests that this parameter should increase as well. However, in the more accurate Allen-Dynes formula:

kBTc=ωln1.2exp(1.04(1+λ)λμ(1+0.62λ)),k_{B}T_{c}=\frac{\hbar\omega_{\rm ln}}{1.2}\exp\left(-\frac{1.04(1+\lambda)}{\lambda-\mu^{*}(1+0.62\lambda)}\right), (7)

which is an analytic approximated solution of the isotropic Eliashberg gap equations Allen and Dynes (1975), the characteristic frequency is the ”logarithmic average”, defined on the basis of the Eliashberg function:

Refer to caption
Figure 10: Partial Eliashberg functions and normalized partial phonon density of states of ThCoC2\mathrm{ThCoC_{2}} for (a) 0 GPa and (b) 20 GPa. (c) Eliashberg functions devided by frequency for 0 and 20 GPa.
Refer to caption
Figure 11: Frequency distribution of the electron-phonon coupling constant of ThCoC2\mathrm{ThCoC_{2}} calculated as λ(ω)=20ωα2F(ω)ωdω,\lambda(\omega)=2\int_{0}^{\omega}\frac{\alpha^{2}F(\omega^{\prime})}{\omega^{\prime}}\text{d}\omega^{\prime}, .
ωln=exp(2λ0ωmaxα2F(ω)lnωdωω).\omega_{\rm ln}=\exp\left(\frac{2}{\lambda}\int_{0}^{\omega_{max}}\alpha^{2}F(\omega)\ln\omega\frac{d\omega}{\omega}\right). (8)

Increases in both the average phonon frequency and the EPC constant λ\lambda results in a non-monotonic evolution of ωln\omega_{\rm ln}, shown in Fig. 8(b). Logarithmic average increases at low pressures, reaches maximum at 5 GPa and then decreases.

III.4 Superconductivity

The results obtained in the preceding sections allow to simulate the pressure effects on superconductivity in ThCoC2 for the electron-phonon coupling scenario. As we have discussed in Ref. Kuderowicz et al. (2021), the experimental value of the critical temperature TcT_{c} at zero pressure can be obtained from the calculated Eliashberg function and the standard Allen-Dynes eq.(7), if the Coulomb pseudopotential parameter is set to μ0.16\mu^{*}\simeq 0.16. The need of using this higher than typical (0.10 - 0.13) value is most likely related to the more complex nature of the superconducting phase than covered by the Allen-Dynes formula, derived based on the isotropic Eliashberg formalism. Similarly, by direct numerical solving the isotropic Eliashberg equations, we have reproduced Kuderowicz et al. (2021) the experimental value of TcT_{c} for μ=0.268\mu^{*}=0.268 when α2F(ω)\alpha^{2}F(\omega) computed with SOC was used.222The value of μ\mu^{*} used for solving the Eliashberg gap equations, if is to be compared to the parameter used with the Allen-Dynes formula, has to be re-scaled due to the dependence of μ\mu^{*} on the numerical cutoff frequency: 1μ~=1μ+ln(ωcωmax)\frac{1}{\widetilde{\mu^{*}}}=\frac{1}{\mu^{*}}+\ln\left(\frac{\omega_{c}}{\omega_{\rm max}}\right). That gives μ~=0.195\widetilde{\mu^{*}}=0.195, see Refs. Allen and Dynes (1975); Kuderowicz et al. (2021) for further details. This and the deviation of the calculated thermodynamic properties of the superconducting phase (temperature dependence of the electronic specific heat, magnetic field penetration depth, and critical field) pointed to the conclusion that the superconductivity in ThCoC2 goes beyond the isotropic Eliashberg picture, nevertheless with a high probability of the electron-phonon coupling mechanism Kuderowicz et al. (2021). As we are, however, able to reproduce the magnitude of TcT_{c} at p = 0 accepting the enhanced μ\mu^{*} values, we may simulate the pressure effect on TcT_{c} by keeping the zero-pressure μ\mu^{*} that reproduces the initial TcT_{c}.

Starting with TcT_{c} computed using the Allen-Dynes formula, TcT_{c} increases almost linearly with pressure, as shown in Fig. 12 (blue squares). μ=0.161\mu^{*}=0.161 is used here, as it reproduces the experimental Tc=2.55T_{c}=2.55 K at 0 GPa when SOC is neglected. In spite of the decline of the density of state N(EF)N(E_{F}), in general detrimental to superconductivity, we observe increase of the critical temperature with a pressure, caused by the enhancement of EPC. The increase in λ\lambda takes over the decrease in ωln\omega_{\rm ln} observed above 5 GPa, thus the critical temperature increases monotonically with an approximately constant ratio of 0.073 K/GPa, reaching 3.11 K at 10 GPa and 4.01 K at 20 GPa. That is quite a strong pressure effect, experimentally accessible for verification even under moderate pressures of several GPa. Following the trend in λ(p)\lambda(p), the magnitude of the pressure-induced change in TcT_{c} in ThCoC2 is lower comparing to that computed for LaNiC2 Wiendlocha et al. (2016), where the ratio was 0.129 K/GPa.

Further on, we simulate how the TcT_{c} and other thermodynamic properties would change with pressure when computed using the Eliashberg isotropic gap equations. That will allow to conclude on the deviations of superconductivity in ThCoC2 under pressure from the isotropic state once the experimental works are reported. A full description of the applied theoretical model with mathematical formulas can be found in Ref. Kuderowicz et al. (2021). Here we keep the same numerical parameters used in solving the Eliashberg equations, i.e., the cut-off frequency ωc=4ωmax\omega_{c}=4\omega_{max} Carbotte (1990) and the number of Matsubara frequencies M=6500M=6500. As mentioned, the Coulomb pseudopotential has been chosen in a way to reproduce the critical temperature Tc=2.55T_{c}=2.55 K for p = 0. In the model without the spin-orbit coupling, this is μ=0.257\mu^{*}=0.257, which corresponds to re-scaled Allen-Dynes formula value  Allen and Dynes (1975); Kuderowicz et al. (2021) μ~=0.189\widetilde{\mu^{*}}=0.189, still above the commonly used approximations of 0.100.130.10-0.13 Morel and Anderson (1962). The changes of the critical temperature under pressure are presented in Fig. 12 (red dots), and are slightly smaller, comparing to predictions from the Allen-Dynes formula (rate of 0.64 K/GPa), nevertheless the substantial increase in TcT_{c} is predicted as well.

Refer to caption
Figure 12: Critical temperature TcT_{c} as a function of the pressure p, determined from the Allen-Dynes formula (blue squares) and from the numerical solving of Eliashberg equations (red dots). In both cases μ\mu^{*} was adjusted to reproduce the experimental zero-pressure TcT_{c} value. Inset: Temperature dependence of the superconducting order parameter Δn=1\Delta_{n=1} for different values of p.

The temperature dependence of Δn=1(T)\Delta_{n=1}(T) for different pressures are presented in the inset of Fig. 12 and undergo the following formula

Δ(T)=Δ(0)1(TTc)Γ,\Delta(T)=\Delta(0)\sqrt{1-\left(\frac{T}{T_{c}}\right)^{\Gamma}}, (9)

with Γ\Gamma which only slightly depends on p, Γ=3.28±0.01\Gamma=3.28\pm 0.01, remaining close to that predicted from the BCS theory, ΓBCS3.0\Gamma_{BCS}\approx 3.0. When increasing the pressure, the dimensionless ratio RΔ=2Δ(0)/kBTcR_{\Delta}=2\Delta(0)/k_{B}T_{c} changes in the range 3.553.583.55-3.58, being close to the BCS value of 3.533.53.

The self-consistent solution of the Eliashberg equations can then be used to calculate the electronic specific heat in the superconducting state CeSC_{e}^{S} Kuderowicz et al. (2021). Above TcT_{c}, the specific heat of the normal state is given by CeN(T)=π23kB2N(EF)(1+λ)TC_{e}^{N}(T)=\frac{\pi^{2}}{3}k_{B}^{2}N(E_{F})(1+\lambda)T. The temperature dependences of the specific heat for different pressure values are presented in Fig. 13.

Refer to caption
Figure 13: Temperature dependence of the electronic specific heat Ce(T)C_{e}(T) for different pressure values. Dashed lines represent the BCS predictions. Results for μ=0.257\mu^{*}=0.257.

For comparison, by dashed lines, we have also marked the BCS curves, which predict the exponential behavior of the specific heat at low temperatures. In this regime, as shown in Fig. 13, the Eliashberg solutions slightly deviate from the BCS curves and do not exhibit the Ceexp[Δ(0)/kBT]C_{e}\propto\exp[-\Delta(0)/k_{B}T] dependence. At high temperatures, CeS(T)C_{e}^{S}(T) is close to BCS relation, reaching the jump of reduced specific heat at TcT_{c}, ΔCe/γTc\Delta C_{e}/\gamma T_{c} from 1.371.37 for p = 0 to 1.471.47 for p=20p=20 GPa, almost equal to the weak-coupling BCS limit of 1.431.43.

Although electronic specific heat exhibits a temperature dependence close to ordinary BCS theory, the London penetration depth λL\lambda_{L} behaves differently. In Fig. 14 we present λL/λL(T=0,p=0)\lambda_{L}/\lambda_{L}(T=0,p=0) calculated from the Eliashberg formalism Kuderowicz et al. (2021); Carbotte (1990). Regardless of pressure, λL\lambda_{L} significantly deviates from the BCS prediction, which is an effect of the retardation processes included in the Eliashberg theory.

Refer to caption
Figure 14: The normalized London penetration depth λL/λL(T=0,p=0)\lambda_{L}/\lambda_{L}(T=0,p=0) as a function of the normalized temperature T/TcT/T_{c}.

With increasing pressure, the London penetration depth decreases slightly in the low temperature regime.

Refer to caption
Figure 15: Normalized upper critical field Hc2/Hc2(T=0,p=0)H_{c2}/H_{c2}(T=0,p=0) as a function of normalized temperature T/TcT/T_{c}.

Finally, we have also analyzed how the upper critical field, Hc2H_{c2}, changes within the pressure, if evaluated within the Eliashberg model Kuderowicz et al. (2021); Carbotte (1990). Our calculations for zero pressure Kuderowicz et al. (2021) showed that ThCoC2 was close to the clean limit, thus in present calculations we use the clean limit, assuming zero scattering rate.

The obtained temperature dependence of the upper critical field normalized to Hc2H_{c2} at T=0T=0 and p = 0 is shown in Fig. 15. In opposite to the penetration depth which does not change significantly with the pressure, the upper critical field increases more than twice when the pressure changes from 0 to 2020 GPa. Similarly as in the case of the critical temperature, it results from the increase of the EPC induced by pressure.

IV Summary

We have analyzed the pressure evolution of the electronic structure, phonons and the electron-phonon coupling in ThCoC2 in the range p = 0 - 20 GPa. Calculations showed that the density of states at the Fermi level slowly decreases with p. The band splitting induced by the spin-orbit coupling increases with p, reaching the value of 185 meV at 20 GPa. This suggests that under elevated pressure the non-BCS features in the thermodynamic properties of the superconducting phase in ThCoC2 may be even more pronounced than at ambient pressure. The electron-phonon coupling is enhanced under pressure due to the increase in the phonon linewidths. The overall EPC parameter λ\lambda increases from 0.583 at 0 GPa to 0.652 at 20 GPa, which is strongly related to the softening of the lowest acoustic phonon mode. If the superconducting pairing in ThCoC2 is based on the electron-phonon coupling, this should raise the superconducting critical temperature and the upper magnetic critical field. When the isotropic Eliashberg formalism is used to calculate TcT_{c}, enhancement of the electron-phonon coupling strength results in a significant increase of TcT_{c}, with a rate of 0.064 - 0.073 K/GPa.

Acknowledgements

This work was supported by the National Science Centre (Poland), project no. 2017/26/E/ST3/00119 and in part by the PL-Grid infrastructure.

References

  • Grant et al. (2014) T Grant, A J S Machado, D J Kim,  and Z Fisk, “Superconductivity in non-centrosymmetric ThCoC2,” Superconductor Science and Technology 27, 035004 (2014).
  • Bardeen et al. (1957) J. Bardeen, L. N. Cooper,  and J. R. Schrieffer, “Theory of Superconductivity,” Phys. Rev. 108, 1175–1204 (1957).
  • Bhattacharyya et al. (2019) A. Bhattacharyya, D. T. Adroja, K. Panda, Surabhi Saha, Tanmoy Das, A. J. S. Machado, O. V. Cigarroa, T. W. Grant, Z. Fisk, A. D. Hillier,  and P. Manfrinetti, “Evidence of a Nodal Line in the Superconducting Gap Symmetry of Noncentrosymmetric ThCoC2{\mathrm{ThCoC}}_{2},” Phys. Rev. Lett. 122, 147001 (2019).
  • Grant et al. (2017) T. W. Grant, O. V. Cigarroa, P. F. S. Rosa, A. J. S. Machado,  and Z. Fisk, “Tuning of superconductivity by Ni substitution into noncentrosymmetric ThCo1xNixC2\mathrm{ThC}{\mathrm{o}}_{1-x}\mathrm{N}{\mathrm{i}}_{x}{\mathrm{C}}_{2},” Phys. Rev. B 96, 014507 (2017).
  • Kuderowicz et al. (2021) Gabriel Kuderowicz, Paweł Wójcik,  and Bartlomiej Wiendlocha, “Electronic structure, electron-phonon coupling, and superconductivity in noncentrosymmetric ThCoC2{\mathrm{ThCoC}}_{2} from ab initio calculations,” Phys. Rev. B 104, 094502 (2021).
  • Kuderowicz and Wiendlocha (2022) Gabriel Kuderowicz and Bartlomiej Wiendlocha, “The effect of Ni doping on the electronic structure and superconductivity in the noncentrosymmetric ThCoC2,” Journal of Magnetism and Magnetic Materials 546, 168688 (2022).
  • Giannozzi et al. (2009) Paolo Giannozzi, Stefano Baroni, Nicola Bonini, Matteo Calandra, Roberto Car, Carlo Cavazzoni, Davide Ceresoli, Guido L Chiarotti, Matteo Cococcioni, Ismaila Dabo, Andrea Dal Corso, Stefano de Gironcoli, Stefano Fabris, Guido Fratesi, Ralph Gebauer, Uwe Gerstmann, Christos Gougoussis, Anton Kokalj, Michele Lazzeri, Layla Martin-Samos, Nicola Marzari, Francesco Mauri, Riccardo Mazzarello, Stefano Paolini, Alfredo Pasquarello, Lorenzo Paulatto, Carlo Sbraccia, Sandro Scandolo, Gabriele Sclauzero, Ari P Seitsonen, Alexander Smogunov, Paolo Umari,  and Renata M Wentzcovitch, “QUANTUM ESPRESSO: a modular and open-source software project for quantum simulations of materials,” Journal of Physics: Condensed Matter 21, 395502 (2009).
  • Giannozzi et al. (2017) P Giannozzi, O Andreussi, T Brumme, O Bunau, M Buongiorno Nardelli, M Calandra, R Car, C Cavazzoni, D Ceresoli, M Cococcioni, N Colonna, I Carnimeo, A Dal Corso, S de Gironcoli, P Delugas, R A DiStasio Jr, A Ferretti, A Floris, G Fratesi, G Fugallo, R Gebauer, U Gerstmann, F Giustino, T Gorni, J Jia, M Kawamura, H-Y Ko, A Kokalj, E Kucukbenli, M Lazzeri, M Marsili, N Marzari, F Mauri, N L Nguyen, H-V Nguyen, A Otero de-la Roza, L Paulatto, S Ponce, D Rocca, R Sabatini, B Santra, M Schlipf, A P Seitsonen, A Smogunov, I Timrov, T Thonhauser, P Umari, N Vast, X Wu,  and S Baroni, “Advanced capabilities for materials modelling with QUANTUM ESPRESSO,” Journal of Physics: Condensed Matter 29, 465901 (2017).
  • Baroni et al. (2001) Stefano Baroni, Stefano de Gironcoli, Andrea Dal Corso,  and Paolo Giannozzi, “Phonons and related crystal properties from density-functional perturbation theory,” Rev. Mod. Phys. 73, 515–562 (2001).
  • Smidman et al. (2017) M Smidman, M B Salamon, H Q Yuan,  and D F Agterberg, “Superconductivity and spin-orbit coupling in non-centrosymmetric materials: a review,” Reports on Progress in Physics 80, 036501 (2017).
  • Samokhin et al. (2004) K. V. Samokhin, E. S. Zijlstra,  and S. K. Bose, “CePt3Si:{\mathrm{CePt}}_{3}\mathrm{Si}: An unconventional superconductor without inversion center,” Phys. Rev. B 69, 094514 (2004).
  • Eliashberg (1960) G. M. Eliashberg, “Interactions between electrons and lattice vibrations in a superconductor,” Soviet Physics-JETP 11, 696 (1960).
  • Katano et al. (2014) Susumu Katano, Hideya Nakagawa, Kazuyuki Matsubayashi, Yoshiya Uwatoko, Hideto Soeda, Takahiro Tomita,  and Hiroki Takahashi, “Anomalous pressure dependence of the superconductivity in noncentrosymmetric LaNiC2\mathrm{LaNi}{\mathrm{C}}_{2}: Evidence of strong electronic correlations,” Phys. Rev. B 90, 220508(R) (2014).
  • Wiendlocha et al. (2016) B. Wiendlocha, R. Szczȩśniak, A. P. Durajski,  and M. Muras, “Pressure effects on the unconventional superconductivity of noncentrosymmetric LaNiC2{\mathrm{LaNiC}}_{2},” Phys. Rev. B 94, 134517 (2016).
  • Perdew et al. (1996) John P. Perdew, Kieron Burke,  and Matthias Ernzerhof, “Generalized Gradient Approximation Made Simple,” Phys. Rev. Lett. 77, 3865–3868 (1996).
  • (16) Pseudopotential were accessed from https://www.quantum-espresso.org/pseudopotentials. The following files were used: Th.pbe-spfn-rrkjus_psl.1.0.0.UPF, Th.rel-pbe-spfn-rrkjus_psl.1.0.0.UPF, Co.pbe-spn-rrkjus_psl.0.3.1.UPF, Co.rel-pbe-spn-rrkjus_psl.0.3.1.UPF and C.pbe-n-rrkjus_psl.1.0.0.UPF.
  • Dal Corso (2014) Andrea Dal Corso, “Pseudopotentials periodic table: From H to Pu,” Computational Materials Science 95, 337 – 350 (2014).
  • Kokalj (2003) Anton Kokalj, “Computer graphics and graphical user interfaces as tools in simulations of matter at the atomic scale,” Computational Materials Science 28, 155 – 168 (2003).
  • Gerss and Jeitschko (1986) M.H. Gerss and W. Jeitschko, “The crystal structures of ternary actinoid iron (cobalt, nickel) carbides with composition 1:1:2,” Materials Research Bulletin 21, 209 – 216 (1986).
  • Birch (1947) Francis Birch, “Finite elastic strain of cubic crystals,” Phys. Rev. 71, 809–824 (1947).
  • Kawamura (2019) Mitsuaki Kawamura, “Fermisurfer: Fermi-surface viewer providing multiple representation schemes,” Computer Physics Communications 239, 197 – 203 (2019).
  • (22) See Supplemental Materials for Fig. S1 with the Brillouin zone, Figs. S2-S3 where the full-relativistic and scalar-relativistic phonons are compared, Fig. S4 with the real-space atomic displacements for the first acoustic phonon mode at R, T and Z q-points, and Fig. S5 with the orbital character of Fermi surface.
  • Wierzbowska et al. (2005) Malgorzata Wierzbowska, Stefano de Gironcoli,  and Paolo Giannozzi, “Origins of low-and high-pressure discontinuities of Tc in niobium,” arXiv preprint cond-mat/0504077  (2005).
  • Heid et al. (2010) R. Heid, K.-P. Bohnen, I. Yu. Sklyadneva,  and E. V. Chulkov, “Effect of spin-orbit coupling on the electron-phonon interaction of the superconductors Pb and Tl,” Phys. Rev. B 81, 174527 (2010).
  • Giustino (2017) Feliciano Giustino, “Electron-phonon interactions from first principles,” Rev. Mod. Phys. 89, 015003 (2017).
  • Grimvall (1981) G. Grimvall, The electron-phonon interaction in metals (North-Holland, Amsterdam, 1981).
  • Gutowska et al. (2021) Sylwia Gutowska, Karolina Górnicka, Paweł Wójcik, Tomasz Klimczuk,  and Bartlomiej Wiendlocha, “Strong-coupling superconductivity of SrIr2{\mathrm{SrIr}}_{2} and SrRh2{\mathrm{SrRh}}_{2}: Phonon engineering of metallic Ir and Rh,” Phys. Rev. B 104, 054505 (2021).
  • McMillan (1968) W. L. McMillan, “Transition Temperature of Strong-Coupled Superconductors,” Phys. Rev. 167, 331–344 (1968).
  • Allen and Dynes (1975) P. B. Allen and R. C. Dynes, “Transition temperature of strong-coupled superconductors reanalyzed,” Phys. Rev. B 12, 905–922 (1975).
  • Carbotte (1990) J. P. Carbotte, “Properties of boson-exchange superconductors,” Rev. Mod. Phys. 62, 1027–1157 (1990).
  • Morel and Anderson (1962) P Morel and P W Anderson, “Calculation of the superconducting state parameters with retarded electronphonon interaction,” Phys. Rev. 125, 1263 (1962).

V Supplemental Material

Supplemental Material contains:

Fig. 16 of the Brillouin zone with the high symmetry points.
Fig. 17 which compares phonon dispersion relations, phonon density of states and Eliashberg function of ThCoC2 computed for p = 0 GPa with and without spin-orbit coupling (SOC).
Fig. 18 which compares phonon dispersion relations, phonon density of states and Eliashberg function of ThCoC2 computed for p = 10 GPa with and without spin-orbit coupling (SOC).
Fig. 19 with the displacement patterns of atoms vibrating in the first acoustic phonon mode at R, T anz Z 𝐪\mathbf{q} points at p = 0 GPa. Amplitude of vibrations is exaggerated. Atomic displacements remain similar for larger pressures.
Fig. 20

Figures 17 and 18 show that the spin-orbit coupling have a minor effect on the phonon dispersion relations and Eliashberg functions of ThCoC2 also under elevated pressure.

Orbital characters of the main Fermi surface sheet for pressures of 0, 10 and 20 GPa are shown in Fig. 20. Contribution to Fermi surface is mostly from Th-6d, Co-3d and C-2p orbitals, with some contribution from Th-f states, in agreement with contributions to total DOS at EFE_{F} discussed in the main text. The anisotropy of orbital character of FS is moderate and is enhanced under pressure – the tube which is formed in the central part of FS above 10 GPa has mostly C-p and Th-f character, which were more uniformly distributed at ambient pressure. Also the relative contribution to FS from 5f states of Th is growing under pressure. That may enhance anisotropy of the superconducting properties of ThCoC2.

Refer to caption
Figure 16: High symmetry points in the first Brillouin Zone of a base centered orthorhombic cell.
Refer to caption
Figure 17: The effect of SOC on phonon dispersion relation, density of states and Eliashberg function of ThCoC2\mathrm{ThCoC_{2}} under 0 GPa.
Refer to caption
Figure 18: The effect of SOC on phonon dispersion relation, density of states and Eliashberg function of ThCoC2\mathrm{ThCoC_{2}} under 10 GPa.
Refer to caption
Figure 19: Visualization of ThCoC2\mathrm{ThCoC_{2}} atoms’ vibrations in the first acoustic mode at (a) R, (b) T and (c) Z 𝐪\mathbf{q}-points. Gray balls are thorium atoms, blue are cobalt and yellow are carbon atoms.
Refer to caption
Figure 20: Orbital character of states forming the Fermi surface of ThCoC2\mathrm{ThCoC_{2}}. Color scale is from 0.0 (blue) to 0.45 (red), and all orbital contributions at each kk-point sum to 1.0.