Äîêóìåíò âçÿò èç êýøà ïîèñêîâîé ìàøèíû. Àäðåñ îðèãèíàëüíîãî äîêóìåíòà : http://www.issp.ac.ru/lhpp/PapersBarkalov/25.pdf
Äàòà èçìåíåíèÿ: Thu Apr 7 20:46:17 2016
Äàòà èíäåêñèðîâàíèÿ: Sun Apr 10 03:14:37 2016
Êîäèðîâêà:

Ïîèñêîâûå ñëîâà: m 8
AR TICLE
Received 10 Sep 2015 | Accepted 15 Feb 2016 | Published 14 Mar 2016
DOI: 10.1038/ncomms11038

OPEN

Superconductivity in Weyl semimetal candidate MoTe2
Yanpeng Qi1, Pavel G. Naumov1, Mazhar N. Ali2, Catherine R. Rajamathi1, Walter Schnelle1, Oleg Barkalov1, ¨ Michael Hanfland3, Shu-Chun Wu1, Chandra Shekhar1, Yan Sun1, Vicky Su1, Marcus Schmidt1, Ulrich Schwarz1, 4, Peter Werner4, Reinald Hillebrand4, Tobias Forster5, Erik Kampert5, Stuart Parkin4, R.J. Cava2, ¨ Eckhard Pippel 1, Binghai Yan1,6 & Sergey A. Medvedev1 Claudia Felser

Transition metal dichalcogenides have attracted research interest over the last few decades due to their interesting structural chemistry, unusual electronic properties, rich intercalation chemistry and wide spectrum of potential applications. Despite the fact that the majority of related research focuses on semiconducting transition-metal dichalcogenides (for example, MoS2), recently discovered unexpected properties of WTe2 are provoking strong interest in semimetallic transition metal dichalcogenides featuring large magnetoresistance, pressure-driven superconductivity and Weyl semimetal states. We investigate the sister compound of WTe2, MoTe2, predicted to be a Weyl semimetal and a quantum spin Hall insulator in bulk and monolayer form, respectively. We find that bulk MoTe2 exhibits superconductivity with a transition temperature of 0.10 K. Application of external pressure dramatically enhances the transition temperature up to maximum value of 8.2 K at 11.7 GPa. The observed dome-shaped superconductivity phase diagram provides insights into the interplay between superconductivity and topological physics.

1 Max Planck Institute for Chemical Physics of Solids, Nothnitzer Strae 40, 01187 Dresden, Germany. 2 Department of Chemistry, Princeton University, ¨ Princeton, New Jersey 08544, USA. 3 European Synchrotron Radiation Facility, BP 220, 38043 Grenoble, France. 4 Max Planck Institute of Microstructure Physics, 06120 Halle, Germany. 5 Dresden High Magnetic Field Laboratory (HLD-EMFL), Helmholtz-Zentrum Dresden-Rossendorf, 01328 Dresden, Germany. 6 Max Planck Institute for the Physics of Complex Systems, 01187 Dresden, Germany. Correspondence and requests for materials should be addressed to B.Y. (email: Yan@cpfs.mpg.de) or to S.A.M. (email: Sergiy.Medvediev@cpfs.mpg.de).

NATURE COMMUNICATIONS | 7:11038 | DOI: 10.1038/ncomms11038 | www.nature.com/naturecommunications

1


AR TICLE
ransition metal dichalcogenides (TMDs) have attracted tremendous attention due to their rich physics and promising potential applications1­11. TMDs share the same formula, MX2, where M is a transition metal (for example, Mo or W) and X is a chalcogenide atom (S, Se and Te). These compounds typically crystallize in many structures, including 2H-, 1T-, 1T0 - and Td-type lattices. The most common structure is the 2H phase, where M atoms are trigonal-prismatically coordinated by the chalcogenide atoms. These planes then stack on one other with van der Waals gaps inbetween. In contrast, the 1T structure corresponds to octahedral coordination of M. The 1T0 phase is a monoclinic lattice that can be interpreted as a distortion of the 1T phase by the formation of in-plane M­M bonds, resulting in a pseudohexagonal layer with zigzag metal chains. Finally, the Td phase is very similar to the 1T0 phase, but the layers stack in a direct fashion, resulting in a higher-symmetry orthorhombic structure. Depending on the synthesis technique, the same composition of MX2 can crystallize in a variety of structures with very different electronic properties. For example, MoTe2 exists in 2H, 1T0 and Td structures12­14, while WTe2 has commonly been observed in the Td structure15. The 2H and 1T compounds are primarily semiconducting, whereas the 1T0 and Td compounds are typically semimetallic. Very recently, semimetallic TMDs have attracted considerable attention because of the discovery of salient quantum phenomena. For instance, Td þ WTe2 has been found to exhibit an extremely large magnetoresistance16,17, pressure (P)-driven superconductivity (highest resistive transition temperature TcE7 K at 16.8 GPa) (refs 18,19), and a large and linear Nernst effect20. Further, this material has been theorized to constitute the first example of a type-II Weyl semimetal21. Moreover, the 1T0 -MX2 monolayer has been predicted to be a two-dimensional topological insulator6.

NATURE COMMUNICATIONS | DOI: 10.1038/ncomms11038

T

The discovery of superconductivity in WTe2 is apparently contradictory to previous theoretical predictions22, which claim that 2H TMDs may become superconducting at high P, but the 1T0 phases will not. Thus the investigation of other TMDs for the appearance of superconductivity under pressure is of big interest. Molybdenum ditelluride (MoTe2) is unique among the TMDs since it is the only material that can be grown in both 2H and 1T0 forms, allowing for direct examination of this theory. If superconductivity exists in 1T0 -MoTe2, it may allow the topological edge states to also become superconducting because of the proximity effect in a bulk superconductor. This would open up a new platform for the study of topological superconductivity, which has potential application in quantum computation23. Regarding the recently anticipated Weyl semimetal phase in MoTe2 (ref. 24), discovery of superconductivity may introduce a new pathway for the exploration of topological superconductivity25­27 along with emergent space-time supersymmetry28. Here, we report on the transport properties of the 2H, 1T0 and Td polytypes of MoTe2 under various applied P. We find that Td-MoTe2 exhibits superconductivity with Tc ¼ 0.10 K, according to electrical resistivity (r) measurements. Application of relatively low pressures below 1 GPa dramatically enhances the Tc, and a dome-shaped Tc­P phase diagram is observed with maximum Tc ¼ 8.2 K at 11.7 GPa; this is B80 times larger than the ambient pressure value. In contrast, we do not observe any traces of superconductivity in the 2H phase, even when it becomes metallic under P. We assume that the extreme sensitivity of the superconductivity to P is a consequence of the unique electronic structure. Thus, MoTe2 presents the opportunity to study the interaction of topological physics and superconductivity in a bulk material. Results Structure and transport properties at ambient pressure. Prior physical properties measurements, synthesized 1T0 -MoTe2

a

b

(000) (011) (002) (004) [100] (022)

d
12
(meV)
Td 1T'

c

1T'

Td

8 4 0

E ­E

T

d

302

304 306 308 Volume (å3)

310

312

Figure 1 | MoTe2 crystal structure. (a) HAADF-STEM image of 1T0-MoTe2 along the [100] zone (scale bar, 0.5 nm). The red rectangle shows HAADF simulated image, and the red and blue spheres in the yellow rectangle represent Te and Mo atoms, respectively. (b) Corresponding electron diffraction images. (c) 1T0 and Td-MoTe2 crystal structures. (d) Energy-volume dependence for 1T0 and Td phases from DFT calculations.
2
NATURE COMMUNICATIONS | 7:11038 | DOI: 10.1038/ncomms11038 | www.nature.com/naturecommunications


NATURE COMMUNICATIONS | DOI: 10.1038/ncomms11038

AR TICLE
a
12 10 MoTe2
Cooling Warming

samples were structurally characterized (Fig. 1) using singlecrystal x-ray diffraction (SXRD) and high-angle annular dark-field scanning transmission electron microscopy (HAADFSTEM). The atomic arrangement of the 1T0 structure was determined using high-resolution HAADF-STEM images and diffraction patterns, as shown in Fig. 1a,b and Supplementary Fig. 1a,b. The crystal structures of 1T0 and Td-MoTe2 are sketched in Fig. 1c. At room temperature, the crystals exhibit the expected monoclinic 1T0 -MoTe2 structure, while the SXRD measurements at 120 K indicate a transition into the orthorhombic Td structure. The 1T0 -MoTe2 structure crystallizes in the P21/m space group with lattice parameters of a ¼ 6.320 å, b ¼ 3.469 å, c ¼ 13.86 å and b ¼ 93.917°; these results are consistent with the previously reported structure12. The Raman spectra at ambient P contain two characteristic peaks (Supplementary Fig. 1c), which are due to the Ag and Bg vibrational modes of the 1T0 -MoTe2 structure; this is also in agreement with a previous report29. A full structural solution was obtained for the orthorhombic Td phase at 120 K, the refined parameters are given in Supplementary Tables 1 and 2. Temperature dependence of electrical resistivity of MoTe2 down to a minimum temperature of Tmin ¼ 0.08 K at ambient pressure is presented in Fig. 2. In contrast to the 2H phase, which displays semiconducting behaviour, 1T0 -MoTe2 is semimetallic in nature. At zero field, the room-temperature resistivity is r ¼ 1.0 á 10 þ 5 O m, which decreases to 2.8 á 10 þ 7 O m at 0.25 K, yielding a residual resistance ratio (RRR) E36. At TE250 K an anomaly with thermal hysteresis (Fig. 2a, inset) is observed, which is associated with the first-order structural phase transition from the 1T0 to the Td polytype14,30. A range of magnetotransport properties has been measured at zero pressure on our MoTe2 crystals (Supplementary Figs 2­4 and Supplementary Note 1). From Hall effect measurements, MoTe2 shows dominant electron-type transport. Within a single-band model the electron concentration ne is estimated to 5 á 1019 cm þ 3 at 2 K and 8 á 1020 cm þ 3 at 300 K (Supplementary Fig. 2), which is close to reported values29. In addition, Td-MoTe2 gradually becomes superconducting below TB0.3 K (the onset of transition), while zero resistance is observed at Tc ¼ 0.10 K (Fig. 2b). Note that, although potential superconductivity atB0.25 K in MoTe2 has been briefly mentioned in the literature31, no related data have been published. 1T0 ­Td structural transition under pressure. It is well known that high pressure can effectively modify lattice structures and the corresponding electronic states in a systematic fashion. Hence, we measured r(T) for the same 1T0 -MoTe2 single crystal at various pressure values P (Fig. 3). Figure 3a shows the typical r(T) curves for P up to 34.9 GPa. For increasing P, the metallic characteristic becomes stronger and r decreases over the entire temperature range. At low pressures, resistance curves exhibit an anomaly at a temperature Ts, associated with the monoclinic 1T0 ­orthorhombic Td structural phase transition similarly to the ambient pressure data. With pressure increase, the resistivity anomaly becomes less pronounced whereas the temperature of anomaly Ts is significantly shifted to lower T and disappears completely above 4 GPa. Thus, the application of P tends to stabilize the monoclinic phase. In addition, the Raman spectra recorded at room temperature under different pressures (Fig. 4a) contain only two characteristic peaks for the 1T0 -structure Ag and Bg modes29. The frequencies of both vibrational modes increase gradually with no discontinuities as P increases (Fig. 4b) indicating the absence of major structural phase transition in the whole studied pressure range at room temperature. The SXRD data (Fig. 4c and Supplementary Fig. 5) also indicate that application of pressure

Resistivity (10­6 m)

8
1.2

6 4 2 0

Resistivity (10 m)

­6

1.0

0.8 200

220

240

260

280

Temperature (K)

0

50

100 150 200 Temperature (K)

250

300

b 0.4

Resistivity (10­6 m)

0.3

0.2

0.1

0.0 0.0

0.2

0.4 0.6 0.8 Temperature (K)

1.0

1.2

Figure 2 | Resistivity of 1T0-MoTe2 at ambient pressure. (a) Temperature-dependent resistivity at near zero pressure. Inset: anomaly with hysteresis observed at B250 K. This hysteresis is associated with the structural phase transition from 1T0-MoTe2 to Td-MoTe2. (b) Resistivity detail from 0.08 to 1.2 K. Superconductivity is observed with onset at E0.25 K and zero resistance at Tc ¼ 0.10 K.

stabilizes the monoclinic 1T0 structure. Increase of P at room temperature results in enhancement of monoclinic distortion (increase of the monoclinic angle b). In an isothermal run at 135 K the reversible orthorhombic Td to monoclinic 1T0 transition is observed at E0.8 GPa (E0.4 GPa) at pressure increase (decrease) (Fig. 4c). Thus, application of P well below 1 GPa decreases the temperature of structural transition to below 135 K. Furthermore, at PE1.5 GPa, the 1T0 structure remains stable down to at least 80 K. The quantitative discrepancy in the Ts values derived from structural and resistivity data is most likely due to nonhydrostatic pressure conditions in the resistivity measurements, and the thermal hysteresis since the resistivity curves are recorded with increasing temperatures. The stability of MoTe2 in different phases can be explained using total energy calculations within density-functional theory (DFT). The optimized lattice constants are very close to experimental values for both phases, as shown in Supplementary Fig. 6 and Supplementary Table 3. After evaluating the total energies of the two phases at ambient pressure, we found that the Td phase exhibits slightly lower energy (0.5 meV per formula unit) than the 1T0 phase. This is consistent with the fact that the low- and high-T phases are Td and 1T0 , respectively, without external pressure. As the 1T0 phase can be obtained by sliding between layers of the Td phase, the former exhibits a slightly
3

NATURE COMMUNICATIONS | 7:11038 | DOI: 10.1038/ncomms11038 | www.nature.com/naturecommunications


AR TICLE
a
5 MoTe 4

NATURE COMMUNICATIONS | DOI: 10.1038/ncomms11038

0.8 GPa
2

Resistivity (10­6 m)

3

2

1

34.9 GPa

0

0

50

100

150 Temperature (K)

200

250

300

b 0.4
m)

c 0.20
Resistivity (10­6 m)
34.9 GPa

0.3
4.4 GPa

0.15

Resistivity (10

­6

0.2
0.8 GPa

0.10
11.7 GPa

0.1

7.2 GPa 11.7 GPa

0.05
27.1 GPa 21.2 GPa

0.0

4

6 8 Temperature (K)

10

0.00

0

2

4 6 8 Temperature (K)

10

d
0.3
3.0 T

e
4
11.7 GPa 1.1 GPa Fitting

m)

3 0.2
0T 1.6 T
0Hc2(T)

Resistivity (10

­6

2

0.1
0.8 T

0.4 T

1

0.0 0 2 4 6 8 Temperature (K) 10

0 0 2 4 6 Temperature (K) 8

Figure 3 | Transport properties of 1T0-MoTe2 as a function of pressure. (a) Electrical resistivity as a function of temperature for pressures of 0.76 þ 34.9 GPa. The anomaly associated with the structural transition is completely suppressed with increasing pressure. (b,c) Electrical resistivity as a function of temperature for pressures of 0.7 þ 11.7 and 11.7 þ 34.9 GPa, respectively. Clear electrical resistivity drops and zero-resistance behaviour are apparent. Tc increases under increasing pressure and a dome-shaped superconducting phase in pressure­temperature space is observed for the maximum superconducting transition temperature corresponding to Tc ¼ 8.2 K at 11.7 GPa. (d) Temperature dependence of resistivity under different magnetic fields of up to 3 T at 11.2 GPa. (e) Temperature dependence of MoTe2 upper critical field Hc2. Tc is defined as temperature at which resistivity drops to 90% of its residual value in normal state. The red curve is the best least squares fit of the equation Hc2(T) ¼ Hc2*(1--T/Tc)1 × a to the experimental data.

4

NATURE COMMUNICATIONS | 7:11038 | DOI: 10.1038/ncomms11038 | www.nature.com/naturecommunications


NATURE COMMUNICATIONS | DOI: 10.1038/ncomms11038

AR TICLE
b
Raman shift (cm­1)
350 300 250 200 150
B A

a
MoTe2

0.8 GPa

g g

Intensity (arb. u.)

3.5 GPa

0

10 20 Pressure (GPa)

30

8.8 GPa

(degrees)

c

96

94

Monoclinic angle

14.8 GPa

92
295 K 135 K 135 K

21.2 GPa

150

200

250

300

350

90 0.0

Raman shift (cm­1)

0.5 1.0 1.5 Pressure (GPa)

2.0

Figure 4 | High-pressure Raman spectroscopy and structural studies of 1T0-MoTe2. (a) Pressure-dependent Raman signals for 1T0-MoTe2 at room temperature. The Raman spectra contain two characteristic peaks due to the Ag and Bg vibrational modes of the 1T0-MoTe2 structure. (b) Frequencies of Ag and Bg modes as function of pressure. The frequencies of both vibrational modes increase gradually and continuously as the pressure increases. (c) Pressure dependence of the monoclinic angle b obtained from SXRD studies. Isothermal compression at room temperature (red filled squares) shows increase of the monoclinic distortion with pressure, whereas reversible orthorhombic Td­monoclinic 1T0 transition is observed in isothermal compression (filled blue circles)/decompression (open blue circles) run at 135 K. The values of Raman frequencies in b and monoclinic angle in c at each pressure are average values obtained from several Raman spectra (XRD patterns) collected from different areas across the sample. The error bars for Raman frequencies in b and monoclinic angle in c due to s.d. are smaller than the symbols size.

Temperature (K)

smaller equilibrium volume than the latter, as also revealed from the lattice parameters measured via SXRD. As illustrated by the energy-volume profile in Fig. 1d, external pressure will stabilize the 1T0 phase with the smaller volume (and correspondingly higher density) by increasing the shift between neighbouring layers. The dome-shaped superconductivity behaviour. Our pressure studies have revealed that the Tc is very sensitive to pressure. That is, Tc increases dramatically to 5 K at relatively low pressures below 1 GPa, before beginning a slower increase to a maximum Tc of 8.2 K at 11.7 GPa (Figs 3b and 5). Beyond this pressure, Tc decreases and no superconductivity with Tc41.5 K is found at P434.9 GPa (Fig. 3c). Remarkably, the drastic increase of Tc at low pressures is associated with a sharp decrease of the 1T0 ­Td structural phase transition temperature Ts. Subsequently at higher pressures, Tc still increases to its maximum value with increasing P but with significantly lower rate. Our findings demonstrate that the strong enhancement of Tc at relatively low P is associated with suppression of the 1T0 ­Td structural phase transition. All the characteristic temperatures in the above experimental results are summarized in the T­P phase diagram in Fig. 5. A dome-shaped superconducting phase boundary is obtained for MoTe2, with a sharp slope towards the zero-P end of the diagram. The bulk character of the superconductivity is confirmed by observations of the magnetic shielding effect in the low pressure range and at 7.5 GPa (Supplementary Fig. 7). The onset temperatures of the diamagnetism are consistent with that of the resistivity drop and confirm the drastic increase of Tc in the

300 MoTe2 250 200 150
Ts (resistivity) Ts Tc Tc Tc Tc (XRD) (run 1) (run 2) (decompression) (shielding effect)

25 20
Temperature (K)

15 10

100 50 0 5 Superconductivity 0 5 10 15 20 25 Pressure (GPa) 30 0 35

Figure 5 | MoTe2 electronic phase diagram. The black and green squares represent the structural phase transition temperature Ts obtained from resistivity and single-crystal synchrotron x-ray diffraction data. The red, blue and olive circles represent the Tc extracted from various electrical resistance measurements, and the magenta triangles represent the Tc determined from the magnetization measurements. The error bars deduced from resistivity measurements values of Tc (red, olive and blue solid circles) due to s.d. of resistivity values (Methods section) are smaller than the symbols size.

low pressure range (Fig. 5). Further, we conducted resistivity measurements in the vicinity of Tc for various external magnetic fields. As can be seen in Fig. 3d, the zero-resistance-point Tc under P ¼ 11.2 GPa is gradually suppressed with increasing field.
5

NATURE COMMUNICATIONS | 7:11038 | DOI: 10.1038/ncomms11038 | www.nature.com/naturecommunications


AR TICLE
Deviating from the Werthamer­Helfand­Hohenberg theory based on the single-band model, the upper critical field, Hc2(T), of MoTe2 has a positive curvature close to Tc (H ¼ 0), as shown in Fig. 3e. This is similar to the behaviours of both WTe2 (ref. 18) and NbSe2 (ref. 32). The experimental Hc2(T) data can be described within the entire T/Tc range by the expression Hc2(T) ¼ Hc2*(1--T/Tc)1 × a (refs 18,33). The fitting parameter Hc2* ¼ 4.0 T can be considered as the upper limit for the upper critical field Hc2(0), which yields a Ginzburg­Landau coherence length xGL(0) of B9 nm. The corresponding data obtained at P ¼ 1.1 GPa is also shown in Fig. 3e. It is also worth noting that our estimated value of Hc2(0) is well below the Pauli-Clogston limit. We repeated the high-pressure experiments using different crystal flakes. Similar superconducting behaviour with almost identical Tc was observed. For comparison with 1T0 -MoTe2, we also measured r(T) for the 2H-MoTe2 single crystal at various pressure values. We found a pressure-induced metallization at 15 GPa (Supplementary Fig. 8), which is consistent with previous theoretical predictions22. However, in contrast, we did not detect any signature of superconductivity in the 2H phase for pressures up to 40 GPa. Discussion For MoTe2, the superconducting behaviour in the low-P region clearly differs from that in the high-P region. Under quite low P, the sharp increase in Tc is concomitant with a strong suppression of the structural transition, which is reminiscent of observations for other superconductors with various kinds of competing phase transitions. The drastic increase of the Tc occurs within the Td phase, which is shown by DFT calculations to be a Weyl semimetal (Supplementary Fig. 9a and Supplementary Note 2) with a band structure around the Fermi level, which is extremely sensitive to changes in the lattice constants24,34. Thus, one can expect that dramatic structural and electronic instabilities emerge in the low-P region, which may account for the strong enhancement of Tc. At higher pressures, the topologically trivial (due to inversion and time reversal symmetry) 1T0 phase (Supplementary Fig. 9b and Supplementary Note 2) remains stable in the whole temperature range. Although within this phase Tc still continues to increase up to its maximum value, the rate of the increase is significantly lower and this growth is naturally explained by the increase of the electronic density of states at the Fermi level in the 1T0 phase (Supplementary Fig. 9c). Thorough exploration of superconductivity in MoTe2 from both experimental and theoretical perspectives is required. Methods
Single-crystal growth. 1T0 -MoTe2 crystals were grown via chemical vapour transport using polycrystalline MoTe2 powder and TeCl4 as a transport additive35. Molar quantities of Mo (Sigma Aldrich 99.99%) were ground in combination with purified Te pieces (Alfa Aesar 99.99%), pressed into pellets and heated in an evacuated quartz tube at 800 °C for 7 days. Crystals were obtained by sealing 1 g of this powder and TeCl4 (3 mg ml þ 1) in a quartz ampoule, which was then flushed with Ar, evacuated, sealed and heated in a two-zone furnace. Crystallization was conducted from (T2) 1,000 to (T1) 900 °C. The quartz ampoule was then quenched in ice water to yield the high-temperature monoclinic phase. The obtained crystals were silver-gray and rectangular in shape. 2H-MoTe2 crystals were grown using a similar method, but without quenching. Structural and transport measurements at ambient pressure. The structures of the MoTe2 crystals were investigated using SXRD with Mo Ka radiation. To analyse the atomic structure of the material, HAADF-STEM was performed. The dependence of the electrical resistivity r on temperature T was measured using a conventional four-probe method (low-frequency alternating current, Physical Property Measurement System (PPMS), Quantum Design). Temperatures down to 0.08 K were achieved using a home-built adiabatic demagnetization stage. The pulsed magnetic field experiments were conducted at the Dresden High Magnetic Field Laboratory (Helmholtz-Zentrum Dresden-Rossendorf, HLD-HZDR).
6

NATURE COMMUNICATIONS | DOI: 10.1038/ncomms11038

Experimental details of high-pressure measurements. A non-magnetic diamond anvil cell was used for r measurements under P values of up to 40 GPa. A cubic BN/epoxy mixture was used for the insulating gaskets and Pt foil was employed in the electrical leads. The diameters of the flat working surface of the diamond anvil and the sample chamber were 500 and 200 mm, respectively. The initial sample thickness was E40 mm. Electrical resistivity at zero magnetic field was measured using the dc current in van der Pauw technique in a customary cryogenic setup (lowest achievable temperature 1.5 K). The resistivity values were defined as an average of five successive measurements at constant temperature. Resistivity measurements in magnetic field were performed on PPMS. Pressure was measured using the ruby scale36 by measuring the luminescence from small chips of ruby placed in contact with the sample. Magnetization was measured on MoTe2 (m ¼ 3.1 mg) in a pressure cell (m ¼ 170 mg) for Pr0.7 GPa and TZ0.5 K (Quantum Design Magnetic Property Measurement System (MPMS), iQuantum 3He insert). Shielding (after zero-field cooling) and Meiner effect curves (in field-cooling) were recorded. The high-P Raman spectra were recorded using a customary micro-Raman spectrometer with a HeNe laser as the excitation source and a single-grating spectrograph with 1 cm þ 1 resolution. Raman scattering was calibrated using Ne lines with an uncertainty of ±1cm þ 1. High-pressure diffraction experiments have been performed at ID09A synchrotron beamline using monochromatic x-ray beam (E ¼ 30 keV, l ¼ 0.413 å) focused to 15 á 10 mm2 on the sample37. We used a membrane-driven highpressure cell equipped with Boehler-Almax seats and diamond anvil design, allowing an opening cone of 64°. The culet size was 600 mm and the sample was loaded together with He as pressure transmission medium into a hole in a stainless steel gasket preindented to B80 mm with an initial diameter of 300 mm. Low temperature data were collected in a He-flow cryostat. Single-crystal data have been collected by a vertical-acting o-axis rotation, with an integrated step scan of 0.5° and a counting time of 1 s per frame. Diffraction intensities have been recorded with a Mar555 flat-panel detector. Diffraction data have been processed and analysed with CrysAlisPro-171.37.35 and Jana2006 software. Pressures were measured with the ruby fluorescence method36. DFT calculations. DFT calculations were performed using the Vienna Ab-initio Simulation Package with projected augmented wave potential38,39. The exchange and correlation energy was considered at the generalized gradient approximation level for the geometry optimization40, and the electronic structure was calculated using the hybrid functional (HSE06)41. Spin­orbital coupling was included in all calculations. Van der Waals corrections were included via a pair-wise force field of the Grimme method42. In the lattice relaxation, the volumes were fixed while lattice constants and atomic positions were optimized. The pressure was derived by fitting the total energy dependence on the volume with the Murnaghan equation43. After checking the k convergence, the 24 á 12 á 8 and 7 á 5 á 3 k-meshes with Gaussian-type smearing were used for the generalized gradient approximation (Supplementary Fig. 10) and HSE06 calculations, respectively. The band structures, density of states and Fermi surfaces were interpolated in a dense k-mesh of 200 á 200 á 200 using the maximally localized Wannier functions44 extracted from HSE06 calculations.

References
1. Wilson, J. A. & Yoffe, A. D. The Transition Metal Dichalcogenides. Discussion and interpretation of the observed optical, electrical and structural properties. Adv. Phys. 18, 193­335 (1969). 2. Klemm, R. A. Pristine and intercalated transition metal dichalcogenide superconductors. Phys. C 514, 86­94 (2015). 3. Morris, R. C., Coleman, R. V. & Bhandari, R. Superconductivity and magnetoresistance in NbSe2. Phys. Rev. B 5, 895­901 (1972). 4. Morosan, E. et al. Superconductivity in CuxTiSe2. Nat. Phys. 2, 544­550 (2006). 5. Moncton, D. E., Axe, J. D. & DiSalvo, F. J. Neutron scattering study of the charge-density wave transitions in 2H-TaSe2 and 2H-NbSe2. Phys. Rev. B 16, 801­819 (1977). 6. Qian, X., Liu, J., Fu, L. & Li, J. Quantum spin Hall effect in two-dimensional transition metal dichalcogenides. Science 346, 1344­1347 (2014). 7. Xu, X., Yao, W., Xiao, D. & Heinz, T. F. Spin and pseudospins in layered transition metal dichalcogenides. Nat. Phys. 10, 343­350 (2014). 8. Bates, J. B., Gruzalski, G. R., Dudney, N. J., Luck, C. F. & Yu, X. Rechargeable thin-film lithium batteries. Solid State Ion. 70/71, 619­628 (1994). 9. Li, Y. et al. MoS2 nanoparticles grown on graphene: an advanced catalyst for the hydrogen evolution reaction. J. Am. Chem. Soc. 133, 7296­7299 (2011). 10. Zhang, Y. J., Oka, T., Suzuki, R., Ye, J. T. & Iwasa, Y. Electrically switchable chiral light-emitting transistor. Science 344, 725­728 (2014). 11. Lin, Y.-C., Dumcenco, D. O., Huang, Y.-S. & Suenaga, K. Atomic mechanism of the semiconducting-to-metallic phase transition in single-layered MoS2. Nat. Nanotechnol. 9, 391­396 (2014). 12. Clarke, R., Marseglia, E. & Hughes, H. P. A low-temperature structural phase transition in b-MoTe2. Philos. Mag. B 38, 121­126 (1978). 13. Puotinen, D. & Newnhan, R. E. The crystal structure of MoTe2. Acta Crystallogr. 14, 691­692 (1961).

NATURE COMMUNICATIONS | 7:11038 | DOI: 10.1038/ncomms11038 | www.nature.com/naturecommunications


NATURE COMMUNICATIONS | DOI: 10.1038/ncomms11038

AR TICLE
39. Kresse, G. & Furthmuller, J. Efficiency of ab-initio total energy calculations for ¨ metals and semiconductors using a plane-wave basis set. Comput. Mater. Sci. 6, 15­50 (1996). 40. Perdew, J. P., Burke, K. & Ernzerhof, M. Generalized gradient approximation made simple. Phys. Rev. Lett. 77, 3865­3868 (1996). 41. Heyd, J., Scuseria, G. E. & Ernzerhof, M. Hybrid functionals based on a screened Coulomb potential. J. Chem. Phys. 118, 8207­8215 (2003). 42. Grimme, S. Semiempirical GGA-type density functional constructed with a long-range dispersion correction. J. Comput. Chem. 27, 1787­1799 (2006). 43. Murnaghan, F. D. The compressibility of media under extreme pressures. Proc. Natl Acad. Sci. USA 30, 244­247 (1944). 44. Marzari, N. & Vanderbilt, D. Maximally localized generalized Wannier functions for composite energy bands. Phys. Rev. B 56, 12847­12865 (1997).

14. Zandt, T., Dwelk, H., Janowitz, C. & Manzke, R. Quadratic temperature dependence up to 50 K of the resistivity of metallic MoTe2. J. Alloys Compd. 442, 216­218 (2007). 15. Brown, B. E. The crystal structures of WTe2 and high-temperature MoTe2. Acta Crystallogr. 20, 268­274 (1966). 16. Ali, M. N. et al. Large, non-saturating magnetoresistance in WTe2. Nature 514, 205­208 (2014). 17. Ali, M. N. et al. Correlation of crystal quality and extreme magnetoresistance of WTe2. Europhys. Lett. 110, 67002 (2015). 18. Pan, X.-C. et al. Pressure-driven dome-shaped superconductivity and electronic structural evolution in tungsten ditelluride. Nat. Commun. 6, 7805 (2015). 19. Kang, D. et al. Superconductivity emerging from suppressed large magnetoresistant state in WTe2. Nat. Commun. 6, 7804 (2015). 20. Zhu, Z. et al. Quantum oscillations, thermoelectric coefficients, and the fermi surface of semimetallic WTe2. Phys. Rev. Lett. 114, 176601 (2015). 21. Soluyanov, A. et al.Type II Weyl Semimetals. Preprint at http://arxiv.org/abs/ 1507.01603 (2015). ´ ´ 22. Riflikova, M., Martonak, R. & Tosatti, E. Pressure-induced gap closing and metallization of MoSe2 and MoTe2. Phys. Rev. B 90, 035108 (2014). 23. Fu, L. & Kane, C. L. Superconducting proximity effect and Majorana fermions at the surface of a topological insulator. Phys. Rev. Lett. 100, 096407 (2008). 24. Sun, Y., Wu, S.-C., Ali, M. N., Felser, C. & Yan, B. Prediction of the Weyl semimetal in the orthorhombic MoTe2. Phys. Rev. B 92, 161107 (2015). 25. Cho, G. Y., Bardarson, J. H., Lu, Y.-M. & Moore, J. E. Superconductivity of doped Weyl semimetals: Finite-momentum pairing and electronic analog of the 3He-A phase. Phys. Rev. B 86, 214514 (2012). 26. Wei, H., Chao, S.-P. & Aji, V. Odd-parity superconductivity in Weyl semimetals. Phys. Rev. B 89, 014506 (2014). 27. Hosur, P., Dai, X., Fang, Z. & Qi, X.-L. Time-reversal-invariant topological superconductivity in doped Weyl semimetals. Phys. Rev. B 90, 045130 (2014). 28. Jian, S.-K., Jiang, Y.-F. & Yao, H. Emergent Spacetime Supersymmetry in 3D Weyl Semimetals and 2D Dirac Semimetals. Phys. Rev. Lett. 114, 237001 (2015). 29. Keum, D. H. et al. Bandgap opening in few-layered monoclinic MoTe2. Nat. Phys. 11, 482­487 (2015). 30. Hughes, H. P. & Friend, R. H. Electrical resistivity anomaly in b-MoTe2. J. Phys. C Solid State Phys. 11, L103­L105 (1978). 31. Hulliger, F. Crystal Chemistry of Chalcogenides and Pnictides of the Transition Elements. in Structure and Bonding, Vol. 4, 83­229 (Springer-Verlag, 1968). ´ 32. Suderow, H., Tissen, V. G., Brison, J. P., Martinez, J. L. & Vieira, S. Pressure induced effects on the Fermi surface of superconducting 2H-NbSe2. Phys. Rev. Lett. 95, 117006 (2005). 33. Muller, K. H. et al. The upper critical field in superconducting MgB2. J. Alloys ¨ Compd. 322, L10­L13 (2001). 34. Wang, Z. et al. MoTe2: Weyl and Line Node Topological Metal. Preprint at http://arxiv.org/abs/1511.07440 (2015). 35. Fourcaudot, G., Gourmala, M. & Mercier, J. Vapor phase transport and crystal growth of molybdenum trioxide and molybdenum ditelluride. J. Cryst. Growth 46, 132­135 (1979). 36. Mao, H. K., Xu, J. & Bell, P. M. Calibration of the ruby pressure gauge to 800 kbar under quasi-hydrostatic conditions. J. Geophys. Res. 91, 4673­4676 (1986). 37. Merlini, M. & Hanfland, M. Single-crystal diffraction at megabar conditions by synchrotron radiation. High Press. Res. 33, 511­522 (2013). 38. Kresse, G. & Hafner, J. Ab initio molecular dynamics for open-shell transition metals. Phys. Rev. B 48, 13115­13118 (1993).

Acknowledgements
Y.Q. acknowledges financial support from the Alexander von Humboldt Foundation. We would like to thank C. Klausnitzer, M. Nicklas and R. Koban for their help with highpressure magnetic measurements. This work was financially supported by the Deutsche Forschungsgemeinschaft (DFG, Project No. EB 518/1-1 of DFG-SPP 1666 `Topological Insulators') and by a European Research Council (ERC) Advanced Grant, No. (291472) `Idea Heusler'.

Author contributions
B.Y. and C.F. conceived the project. M.N.A., C.R.R. and V.S. prepared the samples and performed XRD structural characterization. E.P., P.W. and R.H. performed TEM studies. W.S. performed ambient pressure transport measurements and Meiner effect measurements at low pressures. C.S., T.F. and E.K. performed magneto-transport measurements at ambient pressure. Y.Q., P.G.N., O.B. and S.A.M. performed high-pressure electrical resistivity, Raman spectroscopy and magnetic susceptibility measurements. M.H. performed high-pressure SXRD studies. S.C.W., Y.S. and B.Y. carried out the theoretical calculations. All authors discussed the results of the studies. Y.Q., B.Y., W.S. and S.A.M. co-wrote the paper. All authors commented on the manuscript.

Additional information
Supplementary Information accompanies this paper at http://www.nature.com/ naturecommunications Competing financial interests: The authors declare no competing financial interests. Reprints and permission information is available online at http://npg.nature.com/ reprintsandpermissions/ How to cite this article: Qi, Y. et al. Superconductivity in Weyl semimetal candidate MoTe2. Nat. Commun. 7:10038 doi: 10.1038/ncomms11038 (2016). This work is licensed under a Creative Commons Attribution 4.0 International License. The images or other third party material in this article are included in the article's Creative Commons license, unless indicated otherwise in the credit line; if the material is not included under the Creative Commons license, users will need to obtain permission from the license holder to reproduce the material. To view a copy of this license, visit http://creativecommons.org/licenses/by/4.0/

NATURE COMMUNICATIONS | 7:11038 | DOI: 10.1038/ncomms11038 | www.nature.com/naturecommunications

7