The transition metal dichalcogenide Ir1−xPtxTe2 displays both superconductivity and a topological band structure. Using angle-resolved photoemission spectroscopy, we obtain a comprehensive understanding of the three-dimensional electronic structure in the normal state of Ir1−xPtxTe2 for doping levels from x = 0.1 to 0.4, which spans the composition range of a superconducting state to a non-superconducting state. Many features of the electronic structure can be attributed to strong Te–Te interactions between the layers of the layered crystal structure and can be resolved by photon energy dependent measurements. We demonstrate that the type-II Dirac fermions can be successfully tuned via Pt doping, where the Dirac point lies close to the Fermi level for x = 0.1. The band evolution vs doping provides a clearer understanding of the relationship between the superconductivity and electronic structure. In addition, the β band in the superconducting samples locates the system close to a type-II van Hove singularity, where spin triplet paring symmetry has been predicted. Our results provide a comprehensive understanding of the band structure of Ir1−xPtxTe2, and we discuss the possibilities of the existence of topological superconductivity in this system.

Transition metal dichalcogenides (TMDs) have attracted much attention due to their layered two-dimensional structure with strong in-plane bonding and weak out-of-plane van der Waals interactions. The layered structure of TMDs results in interesting physical phenomena, such as charge density wave (CDW) in TaSe2,1–3 superconductivity in NdSe2,4,5 and topological Dirac/Weyl semimetal (TDS) phases in PtTe(Se)2 and MoTe2.6–8 Recently, the class of TMDs with the 1T structure has been predicted to display a systematically controlled range of band structures including type-I and Lorentz-violating type-II 3D bulk Dirac states, as well as ladders of topological surface states.7–11 The topology of these band structures is determined by the p-orbital manifold of the chalcogenide, rather than the transition metal. This observation motivates a search for superconductivity within this class that may allow tuning the band structure to achieve topological superconductivity, which has been predicted to host exotic Majorana bound states that obey non-Abelian statistics and can be exploited in topological quantum computing.12 

The TMD IrTe2 is a member of this class, which also hosts superconductivity. However, it has a competing CDW ground state that reduces the 1T structure to a monoclinic structure at low temperature, which competes with superconductivity.13 The CDW ground state is associated with a type-II van Hove singularity (vHs) (namely, a vHs associated at momenta away from time-reversal invariant momenta) near the Fermi level (EF) at room temperature, which drives a reconstruction of the electronic structure below the CDW phase transition.14 To fully suppress the CDW transition and stabilize the 1T structure, IrTe2 can be doped with Pt, where a Dirac dispersion has been found near EF without affecting the superconducting properties.15–17 The coexistence of the topological band structure near EF and superconductivity in Ir1−xPtxTe2 suggests possible intrinsic topological superconductivity. Although the nature of the superconductivity in these TDSs and its link to a topological band structure is still unclear, they remain a promising class of materials to realize topological superconductivity. Our goal here is to characterize the 3D band structure with high resolution to identify the properties of the band structure that can induce superconductivity and look for connections between superconductivity and the topological band structure.

Here, we investigate the evolution of the 3D electronic band structures for a range of Pt doping, Ir1−xPtxTe2 (x = 0.1, 0.2, 0.3, and 0.4), using ARPES. The doping levels we study are above the CDW region in the phase diagram, and the samples maintain the 1T structure at low temperature. Strong 3D α (Te pz), β (Te px + py), and γ (Te pz) Fermi surface (FS) sheets have been observed, where the touching point of the β and γ bands defines a type-II Dirac point, namely, one with a tilted Dirac cone in momentum space. Doping with Pt successfully tunes the Fermi level (EF) to be at the Dirac point for x = 0.1 meV and 200 meV below EF for x = 0.4. We also find that the β band dispersion preserves a saddle point (SP) below EF in the superconducting samples, while it disappears in the more highly doped samples. This indicates that the superconducting samples are very close to a type-II van Hove singularity, which theoretically may lead to spin triplet pairing and suggests the possible existence of topological superconductivity in Ir1−xPtxTe2.18,19

High quality Ir1−xPtxTe2 single crystals are synthesized by the flux method.16 (Sample characterization is shown in Fig. S1.) We study samples with different doping levels varying from x = 0.1 to x = 0.4. For x = 0.1, the superconducting transition temperature is ∼1.7 K, x = 0.2 (∼0.65 K), x = 0.3 (∼0.15 K), and no superconductivity is observed for x = 0.4 down to 400 mK. ARPES measurements are performed at the beamline 21-ID-1 of the National Synchrotron Light Source II (NSLS II) equipped with a Scienta DA30 electron analyzer. The photon energies used for mapping out the 3D electronic structures range from 60 eV to 120 eV, with an overall energy resolution better than 25 meV for all photon energies. The samples are cleaved in situ to achieve a fresh surface for the measurement in a vacuum chamber at a pressure below 2 × 10−11 Torr. All measurements are performed at 11 K.

To determine the 3D electronic structure of Ir1−xPtxTe2, we conduct ARPES measurements with various photon energies. This approach enables us to reconstruct the 3D electronic structures across the ΓMK plane (120 eV), AHL plane (96 eV), and TDS plane (90 eV). We describe how kz and the work function of the spectrometer are determined in the supplementary material. Figure 1 shows the 2D Fermi surfaces (FS’s) of Ir0.9Pt0.1Te2 measured at different photon energies that cover from the BZ center (ΓMK plane) to the BZ boundary (ALH plane), as illustrated in Fig. 1(a). For simplicity in terminology, we use Γ¯, M¯ for the high symmetry points in other kz planes. The 2D FS’s change dramatically for different kz, indicating the 3D band structure nature of Ir1−xPtxTe2. The strong 3D nature in such a layered 2D material originates from the strong Te–Te interaction between different layers, as reported previously in Ref. 20. There are mainly two bands α and β contributing to the FS’s, which originate from the Te pz orbital and px + py orbital, respectively. From the layout of the FS’s in Figs. 1(c)–1(h), the α and β bands are well resolved and quite separated from each other near the BZ center [Figs. 1(f)–1(h)], then start touching each other, and finally change dramatically when approaching the BZ boundary [Figs. 1(c)–1(f)]. Near the BZ boundary when the two FS sheets are close to each other, one can also see substantial spectral weight beside the two FS’s, making the bands less well resolved than when they are near the BZ center; this is due to the strong 3D behavior of the bands and the limited kz resolution of the measurements, as reported in other material systems.21 We extract the shape of the α and β FS’s from our experimental data and plot them in Figs. 1(c-ii)–1(h-ii). Our results are consistent with a previous calculation of Ir0.85Pt0.05Te2.20 A threefold symmetry of the band structure is clearly visible for all doping levels, especially for the kz plane closest to the BZ boundary (see the supplementary material, Figs. S3 and S4).

FIG. 1.

(a) Schematic of 3D BZ with high symmetry points indicated. (b) Stacking plot of different 2D Fermi surfaces along the kz direction. [(c)–(h)] (i) Layout of 2D Fermi surfaces measured at various photon energies from 96 eV to 120 eV, respectively, and (ii) schematic of the 2D Fermi surfaces extracted from the experimental data from 96 eV to 120 eV, respectively. [(c)–(h)] Two FSs α and β are resolved. To derive the sketches of the FS shown to the right of the measurements, the general shape of the experimental FS is matched to the calculated FS.20 A blurring of the FS is observed for planes [(c)–(f)] where the FS changes rapidly with kz.

FIG. 1.

(a) Schematic of 3D BZ with high symmetry points indicated. (b) Stacking plot of different 2D Fermi surfaces along the kz direction. [(c)–(h)] (i) Layout of 2D Fermi surfaces measured at various photon energies from 96 eV to 120 eV, respectively, and (ii) schematic of the 2D Fermi surfaces extracted from the experimental data from 96 eV to 120 eV, respectively. [(c)–(h)] Two FSs α and β are resolved. To derive the sketches of the FS shown to the right of the measurements, the general shape of the experimental FS is matched to the calculated FS.20 A blurring of the FS is observed for planes [(c)–(f)] where the FS changes rapidly with kz.

Close modal

To obtain a comprehensive picture of the 3D electronic structure of Ir1−xPtxTe2, we plot the band dispersions for all the samples on the high symmetry planes, ΓMK and AHL (Figs. 2 and 3). As we show, these planes are important to establish the 3D character and detect changes in dispersion with doping. To better describe the threefold symmetry characteristic of the 1T space group, we use two sets of high symmetry points M1K1, M2K2 in the ΓMK plane, H1L1, H2L2 in the AHL plane, and T1S1, T2S2 in the TDS plane. The electronic structure along the ΓMK plane is presented in Fig. 2, where the measured plane in momentum space is highlighted in the 3D BZ in Fig. 2(a–i). On this high symmetry plane, α and β are well separated in momentum space. Data for x = 0.1 and x = 0.3 are presented in Fig. 2 (data of all the samples are shown in Fig. S3). The 3D volume plot in Fig. 2(a–ii) clearly shows the sharp FS and the band dispersion of the x = 0.1 sample. One can see a flower-shaped FS α and a hexagonal-shaped FS β, as indicated in Fig. 2(b). The dispersion of the α and β bands along K2ΓK1 for x = 0.1 and x = 0.3 is plotted in Figs. 2(c) and 2(d), respectively, showing that the two FS’s are both hole-like. Along the ΓK2 direction, the α band crosses EF at 0.77 Å−1 for x = 0.1 and at 0.71 Å−1 for x = 0.3, while the β band crosses EF at 0.25 Å−1 in x = 0.1 and at 0.24 Å−1 in x = 0.3 [top inset in Figs. 2(c) and 2(d)]. The evolution of the α and β bands along the ΓK2 direction can be seen in the momentum distribution curves (MDCs) plotted in Fig. 2(e). The MDCs at the FS along the M2ΓM1 direction are presented in Figs. 2(f)–2(h). The α band crosses EF at 0.47 Å−1 for x = 0.1 and 0.53 Å−1 for x = 0.3, while the β band crosses EF at 0.27 Å−1 for x = 0.1 and 0.25 Å−1 for x = 0.3 along the ΓM2 direction. Comparing the spectra along both ΓK2 and ΓM2 directions for different x, doping does not result in a simple rigid band shift. In addition, the α and β pockets shrink upon doping up to x = 0.3 and then expand a little bit for x = 0.4. This might arise from a change in lattice parameters with doping. While Pt occupies the Ir sites, there is a change in the lattice parameter from 3.95 Å (x = 0.1) to 3.99 Å (x = 0.4), which is a ∼1% change.13 In addition to the electron doping effect, the bandwidth may also change and result in such a non-rigid band shift.

FIG. 2.

(a) (i) 3D BZ plot with the measured ΓMK plane highlighted in pink (left) and 2D BZ plot with the two high symmetry cuts M2ΓM1 and K2ΓK1 indicated (right). (ii) 3D volume plot of the band structure of Ir0.9Pt0.1Te2. (b) (i) The 2D FS at the ΓMK plane of Ir0.9Pt0.1Te2. (ii) The 2D FS at the ΓMK plane of Ir0.7Pt0.3Te2. [(c) and (d)] The band dispersion along the K2ΓK1 direction for x = 0.1 and x = 0.3, respectively. (e) The MDC’s of different samples along the K2ΓK1 direction at EF. [(f) and (g)] The band dispersion along the M2ΓM1 direction for x = 0.1 and x = 0.3, respectively. (h) The MDC’s of different samples along the M2ΓM1 direction at EF. The dispersions of the α band and β band are indicated by the black dashed lines; the red lines show the MDC’s at EF in (c), (d), (f), and (g).

FIG. 2.

(a) (i) 3D BZ plot with the measured ΓMK plane highlighted in pink (left) and 2D BZ plot with the two high symmetry cuts M2ΓM1 and K2ΓK1 indicated (right). (ii) 3D volume plot of the band structure of Ir0.9Pt0.1Te2. (b) (i) The 2D FS at the ΓMK plane of Ir0.9Pt0.1Te2. (ii) The 2D FS at the ΓMK plane of Ir0.7Pt0.3Te2. [(c) and (d)] The band dispersion along the K2ΓK1 direction for x = 0.1 and x = 0.3, respectively. (e) The MDC’s of different samples along the K2ΓK1 direction at EF. [(f) and (g)] The band dispersion along the M2ΓM1 direction for x = 0.1 and x = 0.3, respectively. (h) The MDC’s of different samples along the M2ΓM1 direction at EF. The dispersions of the α band and β band are indicated by the black dashed lines; the red lines show the MDC’s at EF in (c), (d), (f), and (g).

Close modal
FIG. 3.

(a) (i) 3D BZ plot with the measured ALH plane highlighted in green (left), and 2D BZ plot with the two high symmetry cuts L2AL1 and H2AH1 indicated (right). (ii) 3D volume plot of the band structure of Ir0.9Pt0.1Te2. (b) (i) The 2D FS at the ALH plane of Ir0.9Pt0.1Te2. (ii) The 2D FS at the ALH plane of Ir0.7Pt0.3Te2. [(c) and (d)] The band dispersion along the H2AH1 direction for x = 0.1 and x = 0.3, respectively. (e) The MDC’s of different samples along the H2AH1 direction at EF. [(f) and (g)] The band dispersion along the L2AL1 direction for x = 0.1 and x = 0.3, respectively. (h) The MDC’s of different samples along the L2AL1 direction at EF. The dispersions of the α band and γ band are indicated by the black dashed lines; the red lines show the MDC’s at EF in (c), (d), (f), and (g).

FIG. 3.

(a) (i) 3D BZ plot with the measured ALH plane highlighted in green (left), and 2D BZ plot with the two high symmetry cuts L2AL1 and H2AH1 indicated (right). (ii) 3D volume plot of the band structure of Ir0.9Pt0.1Te2. (b) (i) The 2D FS at the ALH plane of Ir0.9Pt0.1Te2. (ii) The 2D FS at the ALH plane of Ir0.7Pt0.3Te2. [(c) and (d)] The band dispersion along the H2AH1 direction for x = 0.1 and x = 0.3, respectively. (e) The MDC’s of different samples along the H2AH1 direction at EF. [(f) and (g)] The band dispersion along the L2AL1 direction for x = 0.1 and x = 0.3, respectively. (h) The MDC’s of different samples along the L2AL1 direction at EF. The dispersions of the α band and γ band are indicated by the black dashed lines; the red lines show the MDC’s at EF in (c), (d), (f), and (g).

Close modal

The band structures at the BZ boundary [AHL plane as indicated in Fig. 3(a)] of x = 0.1 and x = 0.3 samples are plotted in Fig. 3. The topology of the FS on the AHL plane is very different from the one on the ΓMK plane, indicating a strong 3D band structure. Figure 3(a–ii) shows the 3D volume plot of the band structure for x = 0.1, where the threefold symmetry is clearly visible. The outer α FS turns into a star-shape [Fig. 3(b)], while the β FS evolves into several broken pieces. The band dispersions of the α band along the H2AH1 direction show a sharp Fermi crossing at 0.52 Å−1, while the β band bends back at around 0.37 Å−1, which contributes to small hole pockets for x = 0.1 in Fig. 3(c).The β band also contributes to a saddle point along the L2AL1 direction, which becomes a van Hove singularity in the parent compound IrTe2, as will be discussed later.14 For x = 0.3, the α band crosses EF at 0.5 Å−1 along the AH2 direction in Fig. 3(d). The α band dispersion along the L2AL1 direction is broad close to the BZ boundary L, but we still observe a sharp crossing at ∼0.4 Å−1 for x = 0.1, as plotted in Fig. 3(f). This crossing is due to the effect of the highly 3D nature of the α band close to the A point and is a consequence of strong hopping across pz orbitals perpendicular to the TMD layers, as reported before. A weak Fermi crossing around 0.25 Å−1 (named β′) in Fig. 3(d) and 0.75 Å−1 (named α′) have also been resolved, which originate from the projection of the β and α bands from other kz planes due to the limited kz resolution. The β band bends back and sits slightly below EF at the A point, contributing to the lower branch of the type-II Dirac crossing bands, which is discussed below. Figure 3(g) presents the band structure of x = 0.3 along the L2AL1 direction. The dispersions of all the samples along both directions are listed in Fig. S4, and their MDCs at EF are presented in Figs. 3(e) and 3(h) to show the evolution of each band upon doping. The α band along AH2 shrinks in k and the β band sinks down to EF upon doping, which is qualitatively consistent with electron doping into the system. A tiny electron pocket γ starts to appear at x = 0.3, as shown in Figs. 3(d) and 3(g), it is non-degenerate with the β band at the A point. Since the type-II Dirac crossing along the kz direction is defined by the degenerate point of the β and γ bands, the Dirac point will be closer to the A point than to the Γ point.

To search for the Dirac point position, we scan kz with different photon energies and locate the Dirac point at 96 eV photon energy, which is close to the A point. The band structures in the TDS plane for x = 0.1 and x = 0.3 are plotted in Fig. 4. From the 3D volume plot in Fig. 4(a) and the FS plots in Fig. 4(b), a non-zero intensity is visible at the BZ center. In the 2D plots of the band dispersions in Figs. 4(c)–4(g), the finite density at the D point is from the lower cone in the x = 0.1 sample and the upper cone in the x = 0.3 sample, respectively. The β′ band along the T2DT1 direction is from the projections of the β band in other kz planes. From the dispersions of the β and γ bands, one sees that the β band for both x = 0.1 and x = 0.3 goes up compared to their dispersions in the AHL plane, while in the x = 0.3 sample, the γ band sinks down to touch the β band at 120 meV below EF. The evolution of the β and γ bands is consistent with type II Dirac fermion behavior. To see the Dirac dispersion more clearly, we measure the band dispersion at lower photon energy (50 eV) along the DAL plane with enhanced energy resolution. The band dispersions across the Dirac point in the TDS plane for all samples are plotted in Fig. 5, where their in-plane dispersion directions, taking into consideration the experimental geometry, are indicated with the red arrows on top of each figure. The shallow Dirac band dispersions add difficulty to the direct imaging of its dispersion along the kz direction due to the limited kz resolution. However, one sees that the top of the β band at the D point sinks upon Pt doping, and at the same time, the bottom of the γ band at the D point starts to appear at x = 0.2. Such evolution of β and γ bands gives a type II bulk Dirac point (BDP) around 40 meV, 120 meV, and 200 meV below the EF for x = 0.2, 0.3, and 0.4 samples [Fig. 5(e)]. This result demonstrates that the type-II Dirac points can be successfully tuned by the Pt doping effect in IrTe2.

FIG. 4.

(a) (i) 3D BZ plot with the measured DST plane highlighted in green (left) and 2D BZ plot with the two high symmetry cuts T2DT1 and S1DS2 indicated (right). (ii) 3D volume plot of the band structure of Ir0.9Pt0.1Te2. (b) The 2D FS at the ALH plane of Ir0.9Pt0.1Te2. (c) The band dispersion along the DT direction of Ir0.9Pt0.1Te2. (d) The band dispersion along the DS direction of Ir0.9Pt0.1Te2. [(e)–(g)] The 2D FS at the DST plane, dispersion along the T2DT1 direction, and dispersion along the S1DS2 direction of the Ir0.7Pt0.3Te2, respectively. The dispersions of the α, β, and γ bands are indicated by the black dashed lines; the red lines show the MDC’s at EF in (c), (d), (f), and (g).

FIG. 4.

(a) (i) 3D BZ plot with the measured DST plane highlighted in green (left) and 2D BZ plot with the two high symmetry cuts T2DT1 and S1DS2 indicated (right). (ii) 3D volume plot of the band structure of Ir0.9Pt0.1Te2. (b) The 2D FS at the ALH plane of Ir0.9Pt0.1Te2. (c) The band dispersion along the DT direction of Ir0.9Pt0.1Te2. (d) The band dispersion along the DS direction of Ir0.9Pt0.1Te2. [(e)–(g)] The 2D FS at the DST plane, dispersion along the T2DT1 direction, and dispersion along the S1DS2 direction of the Ir0.7Pt0.3Te2, respectively. The dispersions of the α, β, and γ bands are indicated by the black dashed lines; the red lines show the MDC’s at EF in (c), (d), (f), and (g).

Close modal
FIG. 5.

[(a)–(d)] High-resolution band dispersions across BDP for x = 0.1, x = 0.2, x = 0.3, and x = 0.4, respectively. The in-plane directions are indicated in each panel according to the measurement geometry. (e) The EDC’s across the Dirac point of different samples. The intensity dip in each EDC indicates the BDP position.

FIG. 5.

[(a)–(d)] High-resolution band dispersions across BDP for x = 0.1, x = 0.2, x = 0.3, and x = 0.4, respectively. The in-plane directions are indicated in each panel according to the measurement geometry. (e) The EDC’s across the Dirac point of different samples. The intensity dip in each EDC indicates the BDP position.

Close modal

The superconducting behavior of Ir1−xPtxTe2 shows a dome-like feature in the superconducting phase diagram: with increased doping, the structural and CDW transitions are suppressed, giving rise to superconductivity.13 This phase diagram mimics those of cuprate and iron pnictide superconductors, suggesting the possibility of unconventional superconductivity in this system. This behavior is in contrast to that in a related 1T superconductor, PdTe2, which is thought to be a conventional BCS superconductor.22–24 Although the thermal conductivity results demonstrate a nodeless gap structure in Ir1−xPtxTe2,25 suggesting conventional s-wave superconductivity, the nature of the superconducting pairing symmetry in this system remains inconclusive. A second notable feature of Ir1−xPtxTe2 is that the type-II Dirac points are very close to the Fermi level for the x = 0.1, 0.2 samples, suggesting that intrinsic topological superconductivity might be induced due to the finite density of states around the Dirac points.15 It has also been predicted that spin-triplet paring can happen in Ir1−xPtxTe2 when the band filling is near a type-II vHs away from the time-reversal invariant momenta.18,19 For the parent compound, IrTe2, the CDW transition is associated with a vHs originating from the β band, schematically shown in Fig. 6(a), which reconstructs strongly below the transition temperature.14 To characterize the evolution of this type-II vHs in the doped samples using ARPES, we plot the dispersion of the β band in the HAL plane at the BZ boundary of all the samples in Fig. 6(c). A saddle point (SP) can be observed along the AL1 direction from the 3D dispersion of the β band in Figs. 6(a) and 6(c)–6(f). The observation of a saddle point is apparent in Fig. 6(a) where the extrema of the parabolic dispersion of the β band vary from a maximum at the TP1/TP2 point to a minimum at the SP point, where the saddle point is located. As the sample is doped, the vHs persists and moves from around 150 meV below EF for x = 0.1 and 200 meV below EF for x = 0.2. For x = 0.3 and 0.4, the saddle point weakens considerably. Notably, the extremum at the TP1/TP2 point, where the DOS is large, touches the FS for x = 0.1, implying that the β band plays an important role in the superconductivity. However, how these states interact with the Dirac point derived from the β band and γ band is an open question needed to understand the pairing symmetry and existence of topological superconductivity.

FIG. 6.

(a) Contour plot (black lines) of the energy dispersion of the β band in the AH plane showing a generic set of saddle points. The maximum of the parabolic dispersion of the β band along AH1 and AH2 are indicated as TP1 and TP2; the minimum along the AL1 direction is indicated as SP. (b) FS of x = 0.1 along the AHL plane. The red circles indicate the tiny FS’s formed by the top of the β band. [(c)–(f)] Band dispersions along the H1AL1 direction for x = 0.1, 0.2, 0.3, and 0.4, respectively. The β band dispersions are indicated by the red dashed lines. The red arrows indicate the TP1 and SP positions.

FIG. 6.

(a) Contour plot (black lines) of the energy dispersion of the β band in the AH plane showing a generic set of saddle points. The maximum of the parabolic dispersion of the β band along AH1 and AH2 are indicated as TP1 and TP2; the minimum along the AL1 direction is indicated as SP. (b) FS of x = 0.1 along the AHL plane. The red circles indicate the tiny FS’s formed by the top of the β band. [(c)–(f)] Band dispersions along the H1AL1 direction for x = 0.1, 0.2, 0.3, and 0.4, respectively. The β band dispersions are indicated by the red dashed lines. The red arrows indicate the TP1 and SP positions.

Close modal

In summary, we obtain a detailed 3D band structure of Ir1−xPtxTe2 with doping levels from x = 0.1 to 0.4. With different photon energies, we are able to distinguish the band structure evolution across different kz planes. A large outer FS sheet α, an inner FS sheet β, and tiny γ FS sheets together form the 3D band structure. The type-II Dirac point represented by the crossing of the β and γ bands along the kz direction is clearly resolved. With Pt doping, we are able to tune the energetics of the Dirac point in Ir1−xPtxTe2. One might expect similar results by alkali metal dosing, which will be an informative experiment to carry out. More intriguing is the coexistence of superconductivity and Dirac points close to the Fermi level, together with the existence of a type-II van Hove singularity for x = 0.1 and 0.2. Together, these features leave open the possibility of hosting topological superconductivity, which can be exploited to future topological computing, calling for future theory to understand the relation between the superconductivity and the topological nontrivial bands. Experimentally, our comprehensive 3D band structure study in the normal state calls for further low-temperature studies of quantum oscillations in electronic transport measurements, such as scanning tunneling spectroscopy.

See the supplementary material for details of sample characterization, kz determination of the electronic structure, and electronic structures of all the samples along ΓMK and ZAR planes.

The data that support the findings of this study are available from the corresponding author upon reasonable request.

This work was supported by the U.S. Department of Energy (DOE), Offices of Science, Office of Basic Energy Sciences under Award No. DE-SC0019211, the National Natural Science Foundation of China (Grant Nos. U1732273, 11904165, and 11904166). Operation of ESM beamline at the National Synchrotron Light Source was supported by the U.S. DOE User Facility Program operated for the DOE Office of Science by Brookhaven National Laboratory under Contract No. DE-AC02-98CH10886.

1.
S. V.
Borisenko
 et al., “
Pseudogap and charge density waves in two dimensions
,”
Phys. Rev. Lett.
100
(
19
),
196402
(
2008
).
2.
K.
Rossnagel
, “
On the origin of charge-density waves in select layered transition-metal dichalcogenides
,”
J. Phys.: Condens. Matter
23
(
21
),
213001
(
2011
).
3.
Y. W.
Li
 et al., “
Folded superstructure and degeneracy-enhanced band gap in the weak-coupling charge density wave system 2H−TaSe2
,”
Phys. Rev. B
97
(
11
),
115118
(
2018
).
4.
T.
Yokoya
 et al., “
Fermi surface sheet-dependent superconductivity in 2H-NbSe2
,”
Science
294
(
5551
),
2518
2520
(
2001
).
5.
X.
Xi
 et al., “
Ising pairing in superconducting NbSe2 atomic layers
,”
Nat. Phys.
12
(
2
),
139
143
(
2015
).
6.
J.
Jiang
 et al., “
Signature of type-II Weyl semimetal phase in MoTe2
,”
Nat. Commun.
8
,
13973
(
2017
).
7.
M.
Yan
 et al., “
Lorentz-violating type-II Dirac fermions in transition metal dichalcogenide PtTe2
,”
Nat. Commun.
8
(
1
),
257
(
2017
).
8.
Y.
Li
 et al., “
Topological origin of the type-II Dirac fermions in PtSe2
,”
Phys. Rev. Mater.
1
(
7
),
074202
(
2017
).
9.
M. S.
Bahramy
 et al., “
Ubiquitous formation of bulk Dirac cones and topological surface states from a single orbital manifold in transition-metal dichalcogenides
,”
Nat. Mater.
17
(
1
),
21
28
(
2018
).
10.
H. J.
Noh
 et al., “
Experimental realization of type-II Dirac fermions in a PdTe2 superconductor
,”
Phys. Rev. Lett.
119
(
1
),
016401
(
2017
).
11.
F.
Fei
 et al., “
Nontrivial Berry phase and type-II Dirac transport in the layered material PdTe2
,”
Phys. Rev. B
96
(
4
),
041201
(
2017
).
12.
C.
Nayak
 et al., “
Non-Abelian anyons and topological quantum computation
,”
Rev. Mod. Phys.
80
(
3
),
1083
1159
(
2008
).
13.
S.
Pyon
,
K.
Kudo
, and
M.
Nohara
, “
Superconductivity induced by bond breaking in the triangular lattice of IrTe2
,”
J. Phys. Soc. Jpn.
81
(
5
),
053701
(
2012
).
14.
T.
Qian
 et al., “
Structural phase transition associated with van Hove singularity in 5d transition metal compound IrTe2
,”
New J. Phys.
16
(
12
),
123038
(
2014
).
15.
B.-B.
Fu
 et al., “
Realization of low-energy type-II Dirac fermions in (Ir1−xPtx)Te2 superconductors
,”
Chin. Phys. B
28
(
3
),
037103
(
2019
).
16.
F.
Fei
 et al., “
Band structure perfection and superconductivity in type-II Dirac semimetal Ir1−xPtxTe2
,”
Adv. Mater.
30
(
35
),
1801556
(
2018
).
17.
A. F.
Fang
 et al., “
Structural phase transition in IrTe2: A combined study of optical spectroscopy and band structure calculations
,”
Sci. Rep.
3
,
1153
(
2013
).
18.
Z. Y.
Meng
 et al., “
Evidence for spin-triplet odd-parity superconductivity close to type-II van Hove singularities
,”
Phys. Rev. B
91
(
18
),
184509
(
2015
).
19.
H.
Yao
and
F.
Yang
, “
Topological odd-parity superconductivity at type-II two-dimensional van Hove singularities
,”
Phys. Rev. B
92
,
035132
(
2015
).
20.
D.
Ootsuki
 et al., “
Te 5p orbitals bring three-dimensional electronic structure to two-dimensional Ir0.95Pt0.05Te2
,”
Phys. Rev. B
89
(
10
),
104506
(
2014
).
21.
Q.
Song
 et al., “
Electronic structure of the titanium-based oxypnictide superconductor Ba0.95Na0.05Ti2Sb2O and direct observation of its charge density wave order
,”
Phys. Rev. B
93
(
2
),
024508
(
2016
).
22.
H.
Leng
 et al., “
Type-I superconductivity in the Dirac semimetal PdTe2
,”
Phys. Rev. B
96
(
22
),
220506
(
2017
).
23.
Amit
and
Y.
Singh
, “
Heat capacity evidence for conventional superconductivity in the type-II Dirac semimetal PdTe2
,”
Phys. Rev. B
97
(
5
),
054515
(
2018
).
24.
O. J.
Clark
 et al., “
Fermiology and superconductivity of topological surface states in PdTe2
,”
Phys. Rev. Lett.
120
(
15
),
156401
(
2018
).
25.
S. Y.
Zhou
 et al., “
Nodeless superconductivity in Ir1−xPtxTe2 with strong spin-orbital coupling
,”
Europhys. Lett.
104
(
2
),
27010
(
2013
).

Supplementary Material