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

Enhanced Superconductivity in Bilayer PtTe2 by Alkali-Metal Intercalations

Danhong Wu Siyuan Laboratory, Guangzhou Key Laboratory of Vacuum Coating Technologies and New Energy Materials, Department of Physics, Jinan University, Guangzhou 510632, China    Yiping Lin Siyuan Laboratory, Guangzhou Key Laboratory of Vacuum Coating Technologies and New Energy Materials, Department of Physics, Jinan University, Guangzhou 510632, China    Lingxiao Xiong Siyuan Laboratory, Guangzhou Key Laboratory of Vacuum Coating Technologies and New Energy Materials, Department of Physics, Jinan University, Guangzhou 510632, China    Junjie Li Siyuan Laboratory, Guangzhou Key Laboratory of Vacuum Coating Technologies and New Energy Materials, Department of Physics, Jinan University, Guangzhou 510632, China    Tiantian Luo Siyuan Laboratory, Guangzhou Key Laboratory of Vacuum Coating Technologies and New Energy Materials, Department of Physics, Jinan University, Guangzhou 510632, China    Deyi Chen Siyuan Laboratory, Guangzhou Key Laboratory of Vacuum Coating Technologies and New Energy Materials, Department of Physics, Jinan University, Guangzhou 510632, China    Feipeng Zheng [email protected] Siyuan Laboratory, Guangzhou Key Laboratory of Vacuum Coating Technologies and New Energy Materials, Department of Physics, Jinan University, Guangzhou 510632, China
Abstract

Layered platinum tellurium (PtTe2) were recently synthesized with controllable layer numbers down to monolayer limit. Using abinitioab~{}initio calculation based on anisotropic Midgal-Eliashberg formalism, we show that by rubidium (Rb) intercalation, weak superconductivity in bilayer PtTe2 can be significantly boosted with superconducting TcT_{\text{c}} = 8 K in the presence of spin-orbit coupling (SOC). The intercalant on one hand mediates the interlayer coupling and serve as an electron donor, leading to large density of states at Fermi energy. On the other hand, it increases the mass-enhancement parameter with electron-phonon coupling strength comparable to that of Pt. The potassium intercalated bilayer PtTe2 has a comparable TcT_{\text{c}} to the case of Rb intercalation. The relatively high TcT_{\text{c}} with SOC combined with experimental accessible crystal structures suggest that these superconductors are promising platforms to study the novel quantum physics associated with two-dimensional superconductivity, such as the recently proposed type-II Ising superconductivity.

I Introduction

Transition metal dichalcogenides (TMDCs) exhibit intriguing physical properties including superconductivity, charge-density wave, Dirac semimetals, among which the two-dimensional (2D) superconductivity has received growing attention, where the reduced dimensionality leads to unique behaviors compared to their bulk counterparts. One notable example is monolayer NbSe2: (1) the enhanced charge-density wave and reduced superconductivity in going from bulk to the monolayer [1]; (2) the transition from a two-gap superconductor (bulk) to a single-gap one (monolayer) [2, 3], owning to the competition between charge-density wave and superconductivity [4]; (3) suppressed magnetic instability by charge-density wave [5]; (4) Ising superconductivity, where the inversion-symmetry-broken crystal leads to large Zeeman type spin-orbit coupling (SOC), resulting in the extremely large in-plane upper critical field [6, 7]. Besides the inversion-symmetry-broken Ising superconductivity, the large critical fields were also observed in 2D centrosymmetric systems known as type-II Ising superconductivity, arising from multiple degenerate orbitals with spin-orbital locking [8, 9, 10]. Clearly, realizing 2D superconductivity in layered TMDCs is desirable to offer potential platforms for studying the novel quantum physics and the interplay among different orders in 2D limit.

The group VIII TMDCs, MTe2, where M=Ni, Pd, Pt, recently attach increasing attentions for the experimentally verified type-II Dirac semimetal [11, 12, 13], pressure induced superconductivity in their bulk phases [14, 15], and novel physics in their ultrathin films [9, 16, 17, 18, 19, 20, 21]. In particular, monolayer NiTe2 was predicted to be an intrinsic superconductor, and lithium intercalation can boost the superconducting transition temperature (henceforth TcT_{\text{c}}) of bilayer NiTe2 up to 11.3 K [16]. Few-layer PdTe2 were experimentally verified to be type-II Ising superconductors though with TcT_{\text{c}}<< 1 K [21]. Very recently, their homologues PtTe2 crystals were reported to be synthesized with controllable thickness down to monolayer limit [18]. Different from NiTe2, monolayer PtTe2 is an intrinsic semiconductor with band gap about 0.8 eV. Interlayer coupling in PtTe2 is expected to be stronger, due to the smaller interlayer spacing in bulk PtTe2. Furthermore, SOC is supposed to be stronger in PtTe2, as Pt element is much heavier than Pd and Ni. The above differences, combined with the longing for the 2D superconductor motivate us to study the possibility of the emergence of superconductivity in 2D PtTe2 crystals.

This Letter reports an abinitioab~{}initio study on electron-phonon coupling (EPC) and superconducting properties of PtTe2. We show that the weak superconductivity in bilayer PtTe2(Pt2Te4) can be significantly boosted by alkali-metal intercalations. In particular, rubidium (Rb) intercalation leads to the formation of a thermodynamically stable crystal with the stoichiometry of RbPt2Te4, where the Rb occupy all the octahedral sites. Based on anisotropic Midgal-Eliashberg formalism, the TcT_{\text{c}} of the RbPt2Te4 is computed to be 8 K with SOC, which is very high among TMDCs. The homologues KPt2Te4 is shown to have comparable TcT_{\text{c}} to RbPt2Te4. The mechanism of the remarkable boosted superconductivity and the effect of SOC are systematically analyzed.

II Computational methods

Density-functional theory and density-functional perturbation theory calculations were performed with the exchange-correlation functional of PBE [22] to study the crystal structures, electronic structures, EPC of bulk, and few-layer  PtTe2 before and after alkali-metal intercalations [23, 24]. The norm-conserving pseudopotentials of FHI98 [25] and ONCV  [26] were used to describe the interaction between valance and core electrons. The Kohn-shame valance states were expanded as plane waves below 80 Rydberg. A 18×\times18 (18×\times18×\times12) 𝐤\mathbf{k}-mesh and a 6×\times6 (6×\times6×\times4) 𝐪\mathbf{q}-mesh were adopted to calculate the ground states of charge density and phonons for few-layer (bulk) systems, whereupon the electron-phonon coupling (EPC) matrix element gmn,ν(𝐤,𝐪)g_{mn,\nu}(\mathbf{k},\mathbf{q}) are calculated, which quantifies the scattering amplitude between the electronic states with wavevector 𝐤\mathbf{k}, band index mm (𝐤\mathbf{k}, mm), and (𝐤\mathbf{k}+𝐪\mathbf{q}, nn) via a phonon with branch ν\nu and wavevector 𝐪\mathbf{q}. Then above quantities are interpolated to the 𝐤{\mathbf{k}}-grid of 120×\times120 (60×\times60×\times36) and 𝐪{\mathbf{q}}-grid of 60×6060\times 60 (30×\times30×\times20) [28, 27], based on which the mass-enhancement parameter [α2F(ω)/ω\alpha^{2}F(\omega)/\omega] are computed, where α2F(ω)\alpha^{2}F(\omega) is the Eliashberg spectrum, defined as:

α2F(ω)=12vBZd𝐪ΩBZω𝐪νλ𝐪νδ(ωω𝐪ν),\alpha^{2}F(\omega)=\frac{1}{2}\sum_{v}\int_{\mathrm{BZ}}\frac{d\mathbf{q}}{\Omega_{\mathrm{BZ}}}\omega_{\mathbf{q}\nu}\lambda_{\mathbf{q}\nu}\delta\left(\omega-\omega_{\mathbf{q}\nu}\right), (1)

where λ𝐪ν\lambda_{\mathbf{q}\nu} are phonon-momentum-resolved EPC constant [27], ΩBZ\Omega_{\mathrm{BZ}} is the volume of the 1st Brillouin zone, and δ(ωω𝐪ν)\delta\left(\omega-\omega_{\mathbf{q}\nu}\right) is replaced with a gaussian function with a broadening of 0.5 meV. Using the same 𝐤{\mathbf{k}}- and 𝐪{\mathbf{q}}-grids, the temperature-dependent superconducting gaps [Δ(𝐤,T)\Delta(\mathbf{k},T)] are obtained by solving the anisotropic Midgal-Eliashberg equations with the Matsubara frequencies below 0.23 eV on imaginary axis [27, 29], followed by performing analytic continuation to the real axis with Padé functions.

III results and discussions

Monolayer PtTe2 is composed of Te-Pt-Te triatomic layers, where each Pt is octahedral-coordinated by six Te atoms. Its bulk counterpart, 1TT-PtTe2  is formed by the AA stacking of such monolayers along zz direction, separated by an interlayer spacing d=2.57d={\color[rgb]{0,0,0}2.57} Å [30], with the hexagonal lattice constants aa = 4.01 and c=5.24c={\color[rgb]{0,0,0}5.24} Å [18]. The optimized lattice constants are aa =4.08, and cc = 5.27 Å using the FHI98 pseudopotential [25], in nice agreement with the experiment. To study the SOC effect on EPC and superconducting properties, we also adopt the ONCV pseudopotential [26], which yields the optimized a=4.10a=4.10, and c=5.38c={\color[rgb]{0,0,0}5.38} Å, slightly larger the experimental one. Nonetheless, we found that the above two pseudopotentials yield consistent computational results (see Sec. S1 [31] for details). In addition to the different pseudopotentials, the van der Waals (vdW) correlations [32, 33, 34] are also found to have little influence on the electronic structures (see Sec. S1[31]). Therefore, we mainly report in the main text, the results of the ONCV without vdW correction. It should also be mentioned that we involve the projector augmented wave type of pseudopotentials for more efficient calculations of abinitioab~{}initio molecular dynamic (AIMD), and the determination of convex hull for the Rb intercalated bilayer PtTe2, which will be discussed soon. More details of the motivation of using the above methods can be found [35]. The interlayer distance d=2.57d={\color[rgb]{0,0,0}2.57} Å of the bulk PtTe2 is much smaller than that of other layered TMDCs (e.g., 2.90 Å for 2HH-NbSe2 [36], 2.89 Å for TdT_{d}-WTe2 [37]) and even smaller than its homologues NiTe2 (2.63 Å) [38], indicating the potential strong interlayer coupling. Indeed, it was reported that the dispersive bandstructure along cc^{*} suppresses its superconductivity in bulk PtTe2 [39]. Besides, our computational results show that the electron-doped monolayer PtTe2 can exhibit overall large EPC strength (λ\lambda) and isotropic superconducting TcT_{\text{c}} as shown in Sec. S2 [31]. Therefore, it is expected that the superconductivity can be emerged in PtTe2 by electron doping and weakening the interlayer coupling. Thus, to enhance the superconductivity in layered PtTe2, the intercalation of alkali-metal atoms should be a reasonable way, as the intercalants can act as electron donors and relieve the interlayer coupling by expanding the interlayer spacing. To realize 2D superconductivity in PtTe2, we then study the alkali-metal intercalated bilayer PtTe2.

Refer to caption
Figure 1: (a) Side and top views of a slab model of rubidium intercalated bilayer PtTe2 (schematic diagram). The black rhombus indicates a two-dimensional hexagonal unit cell. (b) Formation energies as a function of Rb concentration (xx) calculated in a 3×\times3 supercell via density-functional theory (DFT; see main text). The dashed pink line illustrates the convex hull between bilayer PtTe2 (x=0x=0) and the body-centered cubic rubidium crystal (x=1x=1).

We begin by studying the energy favorable geometry structures of Rb-intercalated bilayer PtTe2, whose stoichiometry can be generated as Rbm(Pt2Te4)n, where mm and nn are integers, counting the number of Rb and Pt2Te4 units, respectively. We use a 3×\times3 supercell to calculate the formation energies ΔE\Delta E with respect to the Rb concentration fraction x=m/(m+n)x=m/(m+n), defined as ΔE(x)\Delta E(x)= EE[Rbx(Pt2Te4)1-x] - xx E [Rb] - (1-xx) EE[Pt2Te4], where EE[Rb] and EE[Pt2Te4] are the energies of a body-center cubic Rb crystal per atom and a bilayer PtTe2 per Pt2Te4 unit, respectively. Fig. 1(b) displays the corresponding results, where ΔE(x)\Delta E(x) decrease as the increase of xx and reach the minimum at x=0.5x=0.5, corresponding to the crystal with the stoichiometry RbPt2Te4. Further increasing the number of Rb, the ΔE\Delta E begins to increase. The resultant convex hull shown in Fig. 1(b) suggests that the RbPt2Te4 crystal is thermodynamically stable with respect to any other stoichiometry. Fig. 1(a) displays the crystal structure of the RbPt2Te4, where the Rb atoms occupy all the octahedral sites at the midpoints of the nearest pairs of Pt atoms in the adjacent monolayers, which is similar to the case of lithium intercalated bilayer NiTe2 [16]. AIMD simulations [24, 40] further suggest that the RbPt2Te4 crystal is stable at room temperature without structure distortion in 10 picoseconds (see Sec. S3 [31]).

Refer to caption
Figure 2: (a) Electronic bandstructure of bilayer PtTe2. (b) Electronic bandstructures of RbPt2Te4 with/without SOC. The corresponding projected DOS (without SOC) onto pp-like orbitals of Te and dd-like orbitals of Pt atoms are shown in the right panel. The projected weights (without SOC) onto the dd-like orbitals (c) and pp-like orbitals (d) for the electronic states of RbPt2Te4  within an energy window of 100 meV near ϵF\epsilon_{\text{F}}. The colors of the dots represent the projection weights (see the colorbar). The black solid lines indicate the Fermi surface and the hexagons are the first Brillouin zone.

Figs. 2(a) and S4(a) display the electronic bandstructures of bilayer (henceforth Pt2Te4) and monolayer PtTe2, respectively. The monolayer is computed to have a band gap of 0.8 eV, which agrees well with the experiment [18]. On going from the monolayer to the bilayer, a semiconductor-to-metal transition is seen with the strong modification of the electronic structure around Fermi energy (ϵF\epsilon_{\text{F}}). The mostly remarkable change is a band derived mainly from the pzp_{z}-like orbitals of Te drops down and crosses ϵF\epsilon_{\text{F}} (see Sec. S4 for the projected bandstructures [31]). This can be assigned to the overlap between the pzp_{z}-like orbitals of the Te atoms in the bilayer, suggesting the strong interlayer coupling. The calculated bandstructure of the Pt2Te4 is again consistent with the ARPES measurements [18]. After Rb intercalation, the interlayer spacing expands and the interlayer coupling weakens, accompanied by the electrons transfer from Rb to its adjacent Te layers. The synergy effects of Rb-mediated interlayer coupling and electron doping make the bandstructure of the RbPt2Te4 resemble that of the electron doped monolayer PtTe2, but with slight band splitting [compare Figs. 2(b) and S4(a)]. One can see from Fig. 2(b) that without SOC, there are two conduction bands crossing the ϵF\epsilon_{\text{F}}, leading to the formation of several electron pockets depicted by the solid black circles in Figs. 2(c)– 2(d). The two bands intersect at the Brillouin zone corners, K and K’, at the energy \sim12 meV above ϵF\epsilon_{\text{F}}, which is responsible for the emergent tiny cirles centered at K and K’. When the SOC is included, though the Kramers degeneracy is preserved due to the inversion symmetry at the site of Rb, the double degenerated bands at K and K’ are split, leading to the avoided crossing near ϵF\epsilon_{\text{F}}  and a \sim57 meV gap opening at K and K’, giving rise to a slight reduced density of states at ϵF\epsilon_{\text{F}} [N(0)N(0)]. In addition to the small Fermi circles, the clover-shaped pockets centered at K and K’ are also noted, each of them consists of three separate petals. Centered at M and Γ\Gamma points, there are also two electron pockets. The calculated projected electronic density of states (DOS) [right panel of Fig. 2(b)] and the momentum-resolved DOS for the states near ϵF\epsilon_{\text{F}} [Figs. 2(c)–2(d)] suggest that these states near K, K’ are significantly contributed by the pp-like orbitals, and the remaining states are assigned to the hybridization of the dd- and pp-like orbitals.

Refer to caption
Figure 3: (a) Phonon dispersion (ω𝒒ν\omega_{\boldsymbol{q}\nu}) of RbPt2Te4 without SOC, and its corresponding EPC constants λ𝒒ν\lambda_{\boldsymbol{q}\nu}. (b) The mass-enhancement parameters α2F(ω)/ω\alpha^{2}F(\omega)/\omega of Pt2Te4 and RbPt2Te4 with/without SOC. (c) Partial mass-enhancement parameters α2F(ω,j,n^)/ω\alpha^{2}F(\omega,j,\hat{n})/\omega, where jj= Pt, Te, Rb, and n^\hat{n} is the unit projection direction vector. (d) Bandstructures of RbPt2Te4 (without SOC) with/without atomic displacements of the phonon modes shown in the inset (see main text). The displacements are tabulated in Table S2 [31], where one can see the Te atoms have the largest displacements of about 0.05 Å. (e) Partial EPC constants arising from various atomic vibrations: λ(j,n^)=20𝑑ωα2F(ω,j,n^)/ω\lambda(j,\hat{n})=2\int_{0}^{\infty}d\omega\alpha^{2}F(\omega,j,\hat{n})/\omega.

The computed phonon dispersions (ω𝒒ν\omega_{\boldsymbol{q}\nu}) of the RbPt2Te4 [Figs. 3(a) and  S5(b)] and Pt2Te4 [Fig. S5(a)] show that all the ω𝐪ν0\omega_{\mathbf{q}\nu}\geq 0, suggesting the dynamically stabilities of the crystals. The SOC  is found to have little influence on ω𝒒ν\omega_{\boldsymbol{q}\nu} of RbPt2Te4 [Fig. S5(b)], and slightly decrease the intensity of the mass-enhancement parameter [α2F(ω)/ω\alpha^{2}F(\omega)/\omega; see Fig. 3(b)], but the main features of the α2F(ω)/ω\alpha^{2}F(\omega)/\omega is similar to the case without SOC. For easier analysis, we first focus on the case without SOC. Fig. 3(b) displays the comparison of the α2F(ω)/ω\alpha^{2}F(\omega)/\omega and the corresponding accumulated EPC strength λ(ω)=20ωα2F(ω)/ω𝑑ω\lambda(\omega)=2\int_{0}^{\omega}\alpha^{2}F(\omega)/\omega d\omega (dashed lines) between RbPt2Te4 and Pt2Te4. It is found that after the intercalation of Rb, the λ(ω)\lambda(\omega) are remarkably enhanced in the energy region between 5 – 13 meV. Particularly, three intensive peaks are found located at 6.6, 9.6 and 12.3 meV, respectively. To further study the mechanism of the enhanced EPC, we decomposed the α2F(ω)\alpha^{2}F(\omega)  into the contributions from in-plane (xyxy) and out-of-plane (zz) vibrations of each atom by computing

α2F(ω,j,n^)=12vBZd𝐪ΩBZω𝐪vλ𝐪vδ(ωω𝐪ν)|n^e𝐪,νj|2,\alpha^{2}F(\omega,j,\hat{n})=\frac{1}{2}\sum_{v}\int_{\mathrm{BZ}}\frac{d\mathbf{q}}{\Omega_{\mathrm{BZ}}}\omega_{\mathbf{q}v}\lambda_{\mathbf{q}v}\delta\left(\omega-\omega_{\mathbf{q}\nu}\right)|\hat{n}\cdot e^{j}_{\mathbf{q},\nu}|^{2}, (2)

where e𝐪,νje^{j}_{\mathbf{q},\nu} is the component of atom jj (Pt, Rb, Te) in the eigenvector of the dynamic matrix with phonon momentum 𝐪{\mathbf{q}} and modes ν\nu, and n^\hat{n} is the unit projection direction vector, which is chosen along in-plane (xyxy) and out-of-plane (zz) directions. The computed α2F(ω,j,n^)/ω\alpha^{2}F(\omega,j,\hat{n})/\omega and the corresponding accumulated EPC constants λ(j,n^)=20𝑑ωα2F(ω,j,n^)/ω\lambda(j,\hat{n})=2\int_{0}^{\infty}d\omega\alpha^{2}F(\omega,j,\hat{n})/\omega are shown in Figs. 3(c) and  3(e), respectively. The most remarkable change is the doubled λ(Texy)\lambda(\text{Te}_{xy}) to 0.56 after the Rb intercalation [Fig. 3(e)]. This is consistent with the computed α2F(ω,Texy)/ω\alpha^{2}F(\omega,\text{Te}_{xy})/\omega [Fig. 3(c)], where α2F(ω,Texy)/ω\alpha^{2}F(\omega,\text{Te}_{xy})/\omega dominates the total spectrum of α2F(ω)/ω\alpha^{2}F(\omega)/\omega in 5 – 13 meV, suggesting the strongly coupling of the in-plane vibration of Te atoms with the electrons. In particular, the two pairs of double degenerated modes Eg and Eu can be seen at Γ\Gamma with energy of 12.08 and 12.11 meV, respectively [Fig. 3(a)], which are derived from Texy vibrations [see Fig. S6(a) [31]]. Among them, the two Eg modes exhibit large λ𝒒ν\lambda_{\boldsymbol{q}\nu}. The vibrational pattern for one of the Eg is shown in the inset of Fig. 3(d). By comparing the bandstructures with and without such phonon displacements imposed on the equilibrant RbPt2Te4 crystal, one can see that the Texy displacements have strong influences on the bands near ϵF\epsilon_{\text{F}}, especially around K, where a Lifshitz transition is noted [Fig. 3(d)]. The secondary contributions to λ\lambda arise from the Tez, Ptxy and Rbxy vibrations as shown in Fig. 3(e). By further examining the corresponding spectra in Fig. 3(c), one can see that the contribution of Tez vibration is mainly associated with the peaks at 6.6 and 12.3 meV, while Ptxy exhibits a relatively uniform contribution in the whole energy region. Interestingly, the Rb also exhibits moderate contribution to λ\lambda, as the computed λ(Rb)=0.27\lambda(\text{Rb})=0.27 is comparable to λ(Pt)\lambda(\text{Pt}) = 0.31 [see Fig. 3(e)]. The Rb vibrations promote the α2F(ω)/ω\alpha^{2}F(\omega)/\omega peak at 9.6 meV [Fig. 3(c)], being on par with the contribution of Texy at the same energy.

Refer to caption
Figure 4: Electronic momentum-resolved EPC constants λ𝐤\lambda_{\mathbf{k}} (see main text) for Pt2Te4 (a), RbPt2Te4 without (b) and with (c) spin-orbit coupling. The corresponding Fermi surfaces are indicated by the solid lines. (d) Histograms of temperature-dependent superconducting gaps Δ𝐤(T)\Delta_{\mathbf{k}}(T) for those electronic states 𝐤\mathbf{k}, whose Kohn-Sham energies are 0.1 eV around ϵF\epsilon_{\text{F}}. (e) The distribution of the Δ𝐤\Delta_{\mathbf{k}} at T=2T=2~{}K in an extensive Brillouin zone. SOC is taken into account.

We turn to study the mechanism of the enhanced EPC in RbPt2Te4 by analyzing the electronic states involved in EPC. We computed the 𝐤\mathbf{k}-resolved EPC constant λ𝐤\lambda_{\mathbf{k}}, defined as follow in this work:

λ𝐤=1N(0)m,n,νdωωd𝐪ΩBZ|gmn,ν(𝐤,𝐪)|2δ(ϵn𝐤)\displaystyle\lambda_{\mathbf{k}}=\frac{1}{N(0)}\sum_{m,n,\nu}\int\frac{d\omega}{\omega}\int\frac{d\mathbf{q}}{\Omega_{\text{BZ}}}\left|g_{mn,\nu}(\mathbf{k},\mathbf{q})\right|^{2}\delta\left(\epsilon_{n\mathbf{k}}\right) (3)
×δ(ϵm𝐤+𝐪)δ(ωω𝐪ν),\displaystyle\times\delta\left(\epsilon_{m\mathbf{k}+\mathbf{q}}\right)\delta\left(\omega-\omega_{\mathbf{q}\nu}\right),

where ΩBZ\Omega_{\text{BZ}} is the volume of the first Brillouin zone, gmn,ν(𝐤,𝐪)g_{mn,\nu}(\mathbf{k},\mathbf{q}) is the EPC matrix element, and the δ\delta functions related to electron and phonon energies are replaced by gaussian functions with broadening of 10 and 0.5 meV, respectively. The λ𝐤\lambda_{\mathbf{k}} is related to total EPC constant by λ=d𝐤ΩBZλ𝐤\lambda=\int\frac{d\mathbf{k}}{\Omega_{\text{BZ}}}\lambda_{\mathbf{k}}. By comparing λ𝐤\lambda_{\mathbf{k}} of Pt2Te4 and RbPt2Te4 shown in Figs. 4(a) and  4(b), one can see the clover-shaped circles at K and K’ are expanded, and the additional Fermi circles emerge after Rb intercalation, promoting the N(0)N(0) from 1.54 to 4.09 states/eV/Pt. Consequently, the λ\lambda in RbPt2Te4  is remarkably boosted to 1.4, as more pairs of electronic states can be involved in the EPC. Furthermore, the contribution of the λ\lambda of RbPt2Te4 are almost arising from the 𝐤\mathbf{k} near K and K’ as shown in Fig. 4(b), which are mainly derived from the electronic states of Te atoms [Fig. 2(d)]. By similar analysis, the states in the remaining sections of the Fermi surface, derived from the pp-dd hybridization of Te and Pt atoms, also contribute to λ\lambda. The SOC is found to reduce the λ𝐤\lambda_{\mathbf{k}} near K, K’ [Fig. 4(c)], attributable to the reduced N(0)N(0) arising from band splitting at K, K’[Fig. 2(b)]. Consequently, λ\lambda decreases to 1.3 with SOC. Except that, the main feature of the computed λ𝐤\lambda_{\mathbf{k}} is unchanged. Therefore, it is evident that the large λ𝐤\lambda_{\mathbf{k}} in RbPt2Te4 is mainly contributed by the pairing of electronic states of Te atoms near K, K’ Fermi pockets.

The significantly enhanced λ\lambda indicates the boost of the superconducting TcT_{\text{c}}. Indeed, using McMillan-Allen-Dynes approach [41, 42, 43] based on the calculated λ\lambda = 1.30, the logarithmic average of the phonon frequencies ωlog=104.5\omega_{\text{log}}=104.5~{}K, and μ=0.205\mu^{*}=0.205 evaluated by μ[0.26N(0)/[1+N(0)]\mu^{*}\approx[0.26N(0)/[1+N(0)] [44], the isotropic superconducting TcT_{\text{c}} of RbPt2Te4 is calculated to be 6.6 K with SOC. In contrast, the computed TcT_{\text{c}} for Pt2Te4 are only 1.1 K. The reliability of the computed isotropic TcT_{\text{c}} is also examined by computing bulk PtTe2 as discussed in Sec. S7, where one can see the computed λ\lambda for bulk PtTe2 is 0.35 and the corresponding TcT_{\text{c}} is 0 K, in nice agreement with previous study that the bulk PtTe2 is non-superconducting with λ=0.33\lambda=0.33 [39]. To obtain more reliable superconducting TcT_{\text{c}} for the systems with reduced dimensionality and anisotropic Fermi surface [29, 45, 46], we further evaluate the TcT_{\text{c}} of RbPt2Te4 by solving the full Midgal-Eliashberg gap equation [29, 45]. Compared to the slight decrease of the aforementioned isotropic λ\lambda, the SOC has relatively noticeable influences on the anisotropic superconducting properties. Fig. 4(d) shows the evolution of the energy distribution of superconducting gaps (Δ𝐤\Delta_{\mathbf{k}}) with respect to temperatures. At TT = 2 K without SOC, the Δ𝐤\Delta_{\mathbf{k}} are distributed in the energy range between 0.9 – 2.4 meV with an average value of 1.6 meV, showing strong anisotropy. When SOC is involved, the overall decreased Δ𝐤\Delta_{\mathbf{k}}  and the suppressed anisotropic distribution of Δ𝐤\Delta_{\mathbf{k}}  are seen, as the average value of Δ𝐤\Delta_{\mathbf{k}} decreases to \sim1.2 meV, and the energy distribution range is reduced to 0.6 – 1.6 meV. The corresponding distribution of Δ𝐤\Delta_{\mathbf{k}} in an extended Brillouin zone is shown in Fig. 4(e), where one can see that the  Δ𝐤\Delta_{\mathbf{k}}  exhibit anisotropic distribution to some extents, in which the relatively large  Δ𝐤\Delta_{\mathbf{k}}  comes from the Fermi pockets around K and K’, dominated by the electronic states of Te atoms as analyzed before. As the temperature increases, the values of  Δ𝐤\Delta_{\mathbf{k}} gradually reduce and finally vanish at T=8KT=8~{}K, suggesting the TcT_{\text{c}} of the RbPt2Te4 is \sim8 K, which is higher than the isotropic one (6.6 K) due to the anisotropic Δ𝐤\Delta_{\mathbf{k}}. We also study the intercalation of other alkali metal elements including lithium (Li), sodium(Na), potassium(K) and caesium(Cs), in which the KPt2Te4 is expected to have comparable TcT_{\text{c}} to the RbPt2Te4, followed by the TcT_{\text{c}} of CsPt2Te4, whereas the LiPt2Te4 and NaPt2Te4 are with relatively low TcT_{\text{c}} [see Sec. S8 for details].

IV conclusion

In summary, we have predicted the enhanced EPC and superconducting TcT_{\text{c}}  in RbPt2Te4 and understood the corresponding mechanisms from abinitioab~{}initio calculations. Firstly, according to the computed convex hull, phonon dispersions and abinitioab~{}initio molecular dynamic simulations, the intercalated Rb are energetically favorable to occupy all the octahedral sites in the interlayer gallery, forming the thermodynamically stable RbPt2Te4 crystal. The TcT_{\text{c}} of RbPt2Te4  is computed to be 8 K with the anisotropic superconducting gaps based on the anisotropic Midgal-Eliashberg formalism in the presence of SOC, though the pristine Pt2Te4  have a very low TcT_{\text{c}}. Such a remarkable enhancement is assigned to the effects of the Rb intercalations from two sides. On one hand, the synergy effect of Rb-mediated interlayer coupling and electron doping lead to the significant promotion of N(0)N(0), accompanied by the markedly enhanced EPC of Te phonons. On the other hand, the Rb directly contributes to the EPC by significantly increasing the intensity of α2F(ω)/ω\alpha^{2}F(\omega)/\omega peak at 9.6 meV. The SOC reduces the TcT_{\text{c}} and the anisotropic superconducting gaps by splitting the band degeneracy near ϵF\epsilon_{\text{F}}. The KPt2Te4 is proposed to have comparable TcT_{\text{c}} to the RbPt2Te4. The reliability of our predictions is supported by the consistent computational results of electronic structures of few-layer PtTe2, and the EPC and superconductivity of bulk PtTe2 with previous studies.

Considering bilayer PtTe2 has been experimentally synthesized  [18], and the intercalation of alkali-metal atoms therein can be experimentally realized [47, 48, 49, 50], our predictions can be straightforwardly probed. Therefore, the experimental accessibility, combined with the relatively high superconducting TcT_{\text{c}} with SOC will make these superconductors promising platforms to investigate intriguing novel quantum physics associated with 2D superconductivity. For example, as we have shown before, the RbPt2Te4 crystal has three-fold rotational symmetries, which are preserved at high symmetry momenta of K and K’, at which the double degenerated electronic bands very close to ϵF\epsilon_{\text{F}} are noted without SOC, and they will be further split by SOC [Fig. 2(b)]. The KPt2Te4 also exhibits similar results (Fig. S10). The above results suggest that the alkali-metal intercalated bilayer PtTe2 are potential candidates for realizing the type-II Ising superconductivity with relatively high superconducting TcT_{\text{c}} [8], which calls for further studies.

Acknowledgements.
This work is supported by National Natural Science Foundation of China (Grants No. 11804118), Guangdong Basic and Applied Basic Research Foundation (Grants No. 2021A1515010041), open project funding of Guangzhou Key Laboratory of Vacuum Coating Technologies and New Energy Materials (KFVEKFVE20200001). The Calculations were performed on high-performance computation cluster of Jinan University, and Tianhe Supercomputer System. D.H. Wu, Y.P. Lin and L.X. Xiong contribute equally to this work.

References

  • Xi et al. [2015] X. Xi, L. Zhao, Z. Wang, H. Berger, L. Forró, J. Shan, and K. F. Mak, Nat. Nanotechnol. 10, 765 (2015).
  • Noat et al. [2015] Y. Noat, J. A. Silva-Guillén, T. Cren, V. Cherkez, C. Brun, S. Pons, F. Debontridder, D. Roditchev, W. Sacks, L. Cario, P. Ordejón, A. García, and E. Canadell, Phys. Rev. B 92, 134510 (2015).
  • Khestanova et al. [2018] E. Khestanova, J. Birkbeck, M. Zhu, Y. Cao, G. L. Yu, D. Ghazaryan, J. Yin, H. Berger, L. Forró, T. Taniguchi, K. Watanabe, R. V. Gorbachev, A. Mishchenko, A. K. Geim, and I. V. Grigorieva, Nano Lett. 18, 2623 (2018),  .
  • Zheng and Feng [2019] F. Zheng and J. Feng, Phys. Rev. B 99, 161119 (2019).
  • Zheng et al. [2018] F. Zheng, Z. Zhou, X. Liu, and J. Feng, Phys. Rev. B 97, 081101(R) (2018).
  • Xi et al. [2016] X. Xi, Z. Wang, W. Zhao, J.-H. Park, K. T. Law, H. Berger, L. Forró, J. Shan, and K. F. Mak, Nat. Phys. 12, 139 (2016),  .
  • Xing et al. [2017] Y. Xing, K. Zhao, P. Shan, F. Zheng, Y. Zhang, H. Fu, Y. Liu, M. Tian, C. Xi, H. Liu, J. Feng, X. Lin, S. Ji, X. Chen, Q. K. Xue, and J. Wang, Nano Lett. 17, 6802 (2017), .
  • Wang et al. [2019] C. Wang, B. Lian, X. Guo, J. Mao, Z. Zhang, D. Zhang, B. L. Gu, Y. Xu, and W. Duan, Phys. Rev. Lett. 123, 126402 (2019), .
  • Liu et al. [2020a] M. Liu, C. Wu, Z. Liu, Z. Wang, D. X. Yao, and D. Zhong, Nano Res. 13, 1733 (2020a),  .
  • Falson et al. [2020] J. Falson, Y. Xu, M. Liao, Y. Zang, K. Zhu, C. Wang, Z. Zhang, H. Liu, W. Duan, K. He, H. Liu, J. H. Smet, D. Zhang, and Q. K. Xue, Science 367, 1454 (2020).
  • Zhang et al. [2017] K. Zhang, M. Yan, H. Zhang, H. Huang, M. Arita, Z. Sun, W. Duan, Y. Wu, and S. Zhou, Phys. Rev. B 96, 125102 (2017).
  • Yan et al. [2017] M. Yan, H. Huang, K. Zhang, E. Wang, W. Yao, K. Deng, G. Wan, H. Zhang, M. Arita, and H. Yang, Nat. Commun. 8, 257 (2017).
  • Huang et al. [2016] H. Huang, S. Zhou, and W. Duan, Phys. Rev. B 94, 121117 (2016).
  • Qi et al. [2020] M. Qi, C. An, Y. Zhou, H. Wu, B. Zhang, C. Chen, Y. Yuan, S. Wang, Y. Zhou, X. Chen, R. Zhang, and Z. Yang, Phys. Rev. B 101, 1 (2020).
  • Xiao et al. [2017] R. C. Xiao, P. L. Gong, Q. S. Wu, W. J. Lu, M. J. Wei, J. Y. Li, H. Y. Lv, X. Luo, P. Tong, X. B. Zhu, and Y. P. Sun, Phys. Rev. B 96, 075101 (2017).
  • Zheng et al. [2020] F. Zheng, X.-B. Li, Y. Lin, L. Xiong, and J. Feng, Phys. Rev. B 101, 100505(R) (2020), .
  • Zhao et al. [2018] B. Zhao, W. Dang, Y. Liu, B. Li, J. Li, J. Luo, Z. Zhang, R. Wu, H. Ma, and G. Sun, J. Am. Chem. Soc. 140, 14217 (2018).
  • Lin et al. [2020] M. K. Lin, R. A. B. Villaos, J. A. Hlevyack, P. Chen, R. Y. Liu, C. H. Hsu, J. Avila, S. K. Mo, F. C. Chuang, and T. C. Chiang, Phys. Rev. Lett. 124, 036402 (2020).
  • Deng et al. [2019] K. Deng, M. Yan, C. P. Yu, J. Li, X. Zhou, K. Zhang, Y. Zhao, K. Miyamoto, T. Okuda, W. Duan, Y. Wu, X. Zhong, and S. Zhou, Sci. Bull. 64, 1044 (2019).
  • Liu et al. [2018] C. Liu, C. S. Lian, M. H. Liao, Y. Wang, Y. Zhong, C. Ding, W. Li, C. L. Song, K. He, X. C. Ma, W. Duan, D. Zhang, Y. Xu, L. Wang, and Q. K. Xue, Phys. Rev. Mater. 2, 094001 (2018).
  • Liu et al. [2020b] Y. Liu, Y. Xu, J. Sun, C. Liu, Y. Liu, C. Wang, Z. Zhang, K. Gu, Y. Tang, C. Ding, H. Liu, H. Yao, X. Lin, L. Wang, Q. K. Xue, and J. Wang, Nano Lett. 20, 5728 (2020b).
  • Perdew et al. [1996] J. P. Perdew, K. Burke, and M. Ernzerhof, Phys. Rev. Lett. 77, 3865 (1996), .
  • Giannozzi et al. [2009] P. Giannozzi, S. Baroni, N. Bonini, M. Calandra, R. Car, C. Cavazzoni, D. Ceresoli, G. L. Chiarotti, M. Cococcioni, I. Dabo, A. Dal Corso, S. de Gironcoli, S. Fabris, G. Fratesi, R. Gebauer, U. Gerstmann, C. Gougoussis, A. Kokalj, M. Lazzeri, L. Martin-Samos, N. Marzari, F. Mauri, R. Mazzarello, S. Paolini, A. Pasquarello, L. Paulatto, C. Sbraccia, S. Scandolo, G. Sclauzero, A. P. Seitsonen, A. Smogunov, P. Umari, and R. M. Wentzcovitch, J. Phys. Condens. Matter 21, 395502 (2009).
  • Kresse and Furthmüller [1996] G. Kresse and J. Furthmüller, Phys. Rev. B 54, 11169 (1996).
  • Fuchs and Scheffler [1999] M. Fuchs and M. Scheffler, Comput. Phys. Commun. 119, 67 (1999).
  • Hamann [2013] D. R. Hamann, Phys. Rev. B 88, 85117 (2013).
  • Poncé et al. [2016] S. Poncé, E. R. Margine, C. Verdi, and F. Giustino, Comput. Phys. Commun. 209, 116 (2016), .
  • Mostofi et al. [2008] A. A. Mostofi, J. R. Yates, Y.-S. Lee, I. Souza, D. Vanderbilt, and N. Marzari, Comput. Phys. Commun. 178, 685 (2008).
  • Margine and Giustino [2013] E. R. Margine and F. Giustino, Phys. Rev. B 87, 024505 (2013), .
  • [30] Furuseth, S., K. Selte, and A. Kjekshus, Acta Chemica Scandinavica 19, 257 (1965).
  • [31] Supplemental Materials.
  • Thonhauser et al. [2007] T. Thonhauser, V. R. Cooper, S. Li, A. Puzder, P. Hyldgaard, and D. C. Langreth, Phys. Rev. B 76, 125112 (2007).
  • Sabatini et al. [2012] R. Sabatini, E. Küçükbenli, B. Kolb, T. Thonhauser, and S. De Gironcoli, J. Phys. Condens. Matter 24, 424209 (2012).
  • Hamada [2014] I. Hamada, Phys. Rev. B 89, 121103 (2014).
  • [35] The optimized structure parameters of bulk PtTe2 using the FHI98 (without vdW), which is a type of norm-conserving pseudopotential, are in nice agreement with experiment (see Table S1 [31]). Therefore, it is the best choice to using the FHI98 for all the calculations. However, we did not find the fully relativistic version of the FHI98, so that the SOC effect cannot be taken into account. Therefore, we turn to the ONCV pseudopotential, another type of norm-conserving pseudopotential, which has both the versions, so that the calculations with and without SOC can be performed, and the comparison of the results with and without SOC can be made. Therefore, we report the results of the ONCV (Figs. 2, 3 and 4) in the main text. Regarding the vdW correction, we find that the structure parameters optimized using ONCV with and without vdW are both reasonable and the former ones are closer to experiment (Table S1 [31]). On the other hand, we find that the density functional perturbation theory calculations are hard to converge when the vdW is included in the phonon calculations for the Pt2Te4 and RbPt2Te4. Therefore, the calculations of phonons, electron-phonon coupling and superconducting properties are using ONCV without vdW in the main text. We also examined the electronic structures computed using the ONCV with and without vdW and found that the vdW correlation has little influence on the electronic structure of RbPt2Te4, especially for the DOS at Fermi energy (Sec. 1 [31]). Futhermore, we also find the consistent results of superconducting TcT_{\text{c}} of RbPt2Te4 using the ONCV (11 K, Fig. S13[31]) and FHI98 (11.5 K, Fig. S11[31]) without vdW and SOC. Therefore, we believe that the results obtained using ONCV without vdW correction, which we report in the main text, are reliable. It should also be mentioned that for more efficient calculations of the systems containing large numbers of atoms and electrons, we adopt projector augmented wave pseudopotentials with an energy truncation of 29.4 Rydberg, in cooperation with vdW correction, for the abinitioab~{}initio molecular dynamic simulations of RbPt2Te4 using a 5×\times5×\times1 supercell (see Sec. S3 [31]), and for determining the convex hull of Rbx(Pt2Te4)1-x in a 3×\times3×\times1 supercell [see Fig. 1(b)]. This type of pseudopotential also yields reasonable structure parameters as shown in Table S1 [31].
  • [36] K. Selte and A. Kjekshus, Acta chem. scand. 18, 697 (1964).
  • Zheng et al. [2016] F. Zheng, C. Cai, S. Ge, X. Zhang, X. Liu, H. Lu, Y. Zhang, J. Qiu, T. Taniguchi, K. Watanabe, S. Jia, J. Qi, J.-H. Chen, D. Sun, and J. Feng, Adv. Mater. 28, 4845 (2016).
  • [38] Peacock, M. A., Thompson, R. M., American Mineralogist, 31, 204 (1946).
  • Kim et al. [2018] K. Kim, S. Kim, J. S. Kim, H. Kim, J. H. Park, and B. I. Min, Phys. Rev. B 97, 165102 (2018).
  • Allen and Tildesley [1989] M. P. Allen and D. J. Tildesley, Computer simulation of liquids (Oxford university press, 1989).
  • Giustino [2017] F. Giustino, Rev. Mod. Phys. 89, 015003 (2017), .
  • McMillan [1968] W. L. McMillan, Phys. Rev. 167, 331 (1968).
  • Allen and Dynes [1975] P. B. Allen and R. C. Dynes, Phys. Rev. B 12, 905 (1975).
  • Bennemann and Garland [1972] K. H. Bennemann and J. W. Garland, in AIP Conf. Proc., Vol. 4 (AIP, 1972) pp. 103–137.
  • Choi et al. [2002] H. J. Choi, D. Roundy, H. Sun, M. L. Cohen, and S. G. Louie, Nature 418, 758 (2002).
  • Sanna et al. [2012] A. Sanna, S. Pittalis, J. K. Dewhurst, M. Monni, S. Sharma, G. Ummarino, S. Massidda, and E. K. Gross, Phys. Rev. B 85, 184514 (2012).
  • Sajadi et al. [2018] E. Sajadi, T. Palomaki, Z. Fei, W. Zhao, P. Bement, C. Olsen, S. Luescher, X. Xu, J. A. Folk, and D. H. Cobden, Science 362, 922 (2018).
  • Fan et al. [2019] X. Fan, H. Chen, J. Deng, X. Sun, L. Zhao, L. Chen, S. Jin, G. Wang, and X. Chen, Inorg. Chem. 58, 7564 (2019).
  • Nakata et al. [2019] Y. Nakata, K. Sugawara, A. Chainani, K. Yamauchi, K. Nakayama, S. Souma, P.-Y. Chuang, C.-M. Cheng, T. Oguchi, and K. Ueno, Phys. Rev. Mater. 3, 71001 (2019).
  • Bianchi et al. [2012] M. Bianchi, R. C. Hatch, Z. Li, P. Hofmann, F. Song, J. Mi, B. B. Iversen, Z. M. Abd El-Fattah, P. Löptien, L. Zhou, A. A. Khajetoorians, J. Wiebe, R. Wiesendanger, and J. W. Wells, ACS Nano 6, 7009 (2012).