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

Suppression of both superconductivity and structural transition in hole-doped MoTe2 induced by Ta substitution

Siu Tung Lam    K. Y. Yip    Swee K. Goh [email protected] Department of Physics, The Chinese University of Hong Kong, Shatin, Hong Kong, China    Kwing To Lai [email protected] Department of Physics, The Chinese University of Hong Kong, Shatin, Hong Kong, China Shenzhen Research Institute, The Chinese University of Hong Kong, Shatin, Hong Kong, China Faculty of Science, The University of Hong Kong, Pokfulam Road, Hong Kong, China
Abstract

Type-II Weyl semimetal MoTe2 exhibits a first-order structural transition at TsT_{s} \sim250 K and superconducts at TcT_{c} \sim0.1 K at ambient pressure. Both TsT_{s} and TcT_{c} can be manipulated by several tuning parameters, such as hydrostatic pressure and chemical substitution. It is often reported that suppressing TsT_{s} enhances TcT_{c}, but our study shows a different behaviour when MoTe2 is hole-doped by Ta. When TsT_{s} is suppressed by Ta doping, TcT_{c} is also suppressed. Our findings suggest that the suppression of TsT_{s} does not necessarily enhance superconductivity in MoTe2. By connecting with the findings of electron-doped MoTe2, we argue that varying electron carrier concentration can effectively tune TcT_{c}. In addition, the Hall coefficient is enhanced around the doping region, where TsT_{s} is completely suppressed, suggesting that the critical scattering around the structural transition may also play a role in suppressing TcT_{c}.

I Introduction

Superconductivity is found, often by tuning the electronic properties via the application of hydrostatic pressure, in many topological semimetals, such as Cd3As2 [1], ZrTe5 [2], YPtBi [3, 4, 5, 6], WTe2 [7, 8, 9, 10] and MoTe2 [11, 12, 13, 14]. The exotic combination of topological bands and superconductivity offers a unique platform to search for topological superconductivity, where Majorana fermions can be used to develop topological quantum computation [15, 16].

Type-II Weyl semimetal MoTe2 [17, 18, 19, 20] is one of the promising candidates for hosting topological superconductivity, especially after the discovery of an edge supercurrent [21]. At ambient pressure, MoTe2 undergoes a first-order structural transition at TsT_{s}\sim 250 K, changing from a centrosymmetric (nonpolar) monoclinic 1T1T^{\prime} phase (space group: P21/mP2_{1}/m) to a noncentrosymmetric (polar) orthorhombic TdT_{d} phase (space group: Pmn21Pmn2_{1}) upon cooling. At TcT_{c}\sim 0.1 K, an additional superconducting phase transition occurs.

Owing to its low TcT_{c}, it is challenging to experimentally study the superconductivity of MoTe2. Finding a suitable way to control its TcT_{c} becomes an outstanding issue. Meanwhile, the competition between structural and superconducting transitions in MoTe2 has been reported in previous studies using a variety of tuning parameters. Through the application of pressure [11, 22, 23, 24, 25, 13], TsT_{s} is suppressed to 0 K at \sim10 kbar, resulting in a complete removal of the TdT_{d} phase at high pressures. Meanwhile, TcT_{c} is enhanced by 30-fold (\sim4 K) at \sim15 kbar. These behaviours demonstrate the anticorrelation between TsT_{s} and TcT_{c}. Similar anticorrelation can also be observed via isovalent chemical substitutions (S/Se substituting Te [22, 26]) and electron doping (Te deficiency [27] and Re substituting Mo [28]). Note that via the substitution of Mo by W, TsT_{s} is enhanced at ambient pressure, and the pressure-induced TcT_{c} is lower than that observed in the pristine MoTe2, demonstrating the anticorrelation between TsT_{s} and TcT_{c} again [29].

The superconductivity of hole-doped MoTe2 has not been studied to the same extent as the electron-doped counterpart. The introduction of hole carriers in monolayer MoTe2 through gating has been shown to reduce its TcT_{c} [30]. On the other hand, the effect of hole doping on bulk MoTe2 has been explored in depth through the substitution of Nb for Mo [31, 32]. Although no evidence of superconductivity with Tc>T_{c}> 2 K has been found up to the highest studied doping level xx = 0.22 in Mo1-xNbxTe2, indicating a lack of significant enhancement in TcT_{c} through hole doping, the hole-doping phase diagram of Mo1-xNbxTe2 in the normal state was extensively investigated by Sakai et al. [32]. They revealed that the suppression of TsT_{s} upon Nb doping is associated with a huge enhancement of thermopower at low temperatures, which they attributed to the critical scattering arising from the boundary of the nonpolar-to-polar transition around TsT_{s}.

Nevertheless, it remains uncertain how TcT_{c} evolves and what the correlation of TsT_{s} and TcT_{c} is upon hole doping. Understanding these issues can help us reveal the key factors that control TcT_{c} of MoTe2. In this article, we study the effect of hole doping on MoTe2 via the substitution of Mo by Ta. Transport measurements were conducted down to \sim30 mK to track the evolution of both TsT_{s} and TcT_{c}, and surprisingly, we found that both TsT_{s} and TcT_{c} are suppressed and eventually vanish with increasing hole doping, contrary to the anticorrelation between TsT_{s} and TcT_{c} established in MoTe2 controlled by other tuning parameters.

Refer to caption
Figure 1: (a) X-ray diffraction (XRD) spectra of single crystals of Mo1-xTaxTe2. The peaks of (00LL) are indexed in the figure. (b) Enlarged XRD spectra near the peak of (002). The (002) peak shifts progressively toward a higher diffraction angle 2θ2\theta when xx increases.

II Experiment

Single crystals of Mo1-xTaxTe2 were grown by the self-flux method. The mixture of Mo powder (99.999%\%, Alfa Aesar), Te (99.99999%\% lumps, Ultimate Material), and Ta powder (99.99%\%, Sigma Aldrich) were first placed into an alumina crucible, with a stoichiometric ratio of Mo:Ta:Te = 1x-x:xx:20. The alumina crucible was inserted into a quartz tube before the quartz tube was sealed under a vacuum. The sealed ampule was then heated to 1100 C within 24 hours and stayed for 24 hours, followed by slow cooling to 880 C for 400 hours. Finally, the ampule was taken out from the furnace at 880 C and centrifuged to remove the excess Te flux. X-ray diffraction (XRD) data were collected at room temperature by using a Rigaku X-ray diffractometer with CuKαK_{\alpha} radiation. The chemical compositions were characterized by a JEOL JSM-7800F scanning electron microscope equipped with an Oxford energy-dispersive X-ray (EDX) spectrometer. A standard four-probe method was used to measure temperature-dependent resistance in a Bluefors dilution refrigerator with a base temperature of 30 mK. A standard six-probe method was used to measure the Hall effect in a Quantum Design Physical Property Measurement System with a temperature range from 300 K to 2 K and a magnetic field of ±14\pm 14 T.

III Results and discussion

Figure 1(a) shows the XRD spectra for the Mo1-xTaxTe2 single crystals with x=x= 0, 0.021, 0.042, 0.046, 0.065, 0.078, 0.097, 0.118, and 0.173. The peaks shown in all spectra are well indexed by the (00LL) planes originating from the pattern of 1T1T^{\prime}-MoTe2, confirming that all crystals are single-crystalline 1T1T^{\prime}-MoTe2 at room temperature. Figure 1(b) focuses on the (002) peaks of all samples, which reveal a monotonic shift to a higher 2θ\theta when xx increases, indicating a shrinking crystal structure. As the covalent radius of Ta is smaller than that of Mo, this provides crystallographic evidence that Ta is systemically substituting Mo with increasing xx. These Mo1-xTaxTe2 crystals measured in XRD were also examined by EDX, from which we determined their elemental compositions and hence the values of xx in each sample. The EDX results are consistent with the findings in XRD spectra. (see Supplemental Material for more details [33].)

Refer to caption
Figure 2: (a) Temperature dependence of resistivity ρ(T)\rho(T) of Mo1-xTaxTe2 at zero magnetic fields. The warm-up (cool-down) data are plotted as solid (dashed) curves. (b) Low-temperature ρ(T)\rho(T) normalized to the value of ρ(1K)\rho(1~{}\mathrm{K}), displaying the superconducting transitions.

Figure 2(a) illustrates the temperature dependence of resistivity ρ(T)\rho(T) of Mo1-xTaxTe2 with x=00.173x=0-0.173 measured under zero magnetic field. All samples exhibit metallic behaviour. A thermal hysteresis can be observed in pristine MoTe2 (x=0x=0) around 150–250 K when the resistivity was measured upon increasing (solid curves) and decreasing temperature (dashed curves), indicating the appearance of the first-order structural transition [34, 11, 13, 14, 35]. This transition persists up to x=0.097x=0.097. With increasing xx, the transition shifts gradually toward lower temperatures, and the hysteresis loop becomes broader. When x0.118x\geq 0.118, no hysteresis is observed in the whole temperature range, suggesting that the structural transition vanishes at the high doping region. Figure 2(b) shows the resistivity data normalized to the value of ρ(T)\rho(T) at 1 K at the low-temperature region. A superconducting transition, where TcT_{c} is defined at which the resistivity drops to zero, is observed at x=0x=0 with TcT_{c}\sim 0.1 K, which is consistent with the previous studies [11, 12, 21, 13, 14, 22, 23, 24, 25, 35]. When xx increases, TcT_{c} generally reduces despite a small enhancement to \sim0.25 K at x=x= 0.042. At x=x= 0.065, a small drop of resistivity without reaching zero resistivity is observed near the base temperature, indicating that the bulk superconductivity is heavily suppressed and only trace superconductivity is detected. When xx further increases (0.078\geq 0.078), the resistivity data shows no signs of superconductivity.

To probe the evolution of the Fermi surface of Mo1-xTaxTe2, we conducted the Hall effect measurements. Figure 3 illustrates the magnetic field dependence of Hall resistivity ρxy(B)\rho_{xy}(B) of Mo1-xTaxTe2 with x=0x=0, 0.021, 0.065 and 0.173 at different temperatures measured during warm-up. ρxy(B)\rho_{xy}(B) data of samples with other doping can be found in Fig. S2 in Supplemental Material [33]. At x=0x=0 (Fig. 3(a)), ρxy(B)\rho_{xy}(B) has a negative slope at the whole temperature range. At low temperatures, ρxy(B)\rho_{xy}(B) shows a non-linear feature. These features are consistent with the semimetallic nature of MoTe2, which exhibits nearly perfect electron-hole compensation with a high electron mobility [36, 13]. After introducing Ta doping, the slope of ρxy(B)\rho_{xy}(B) at x=0.021x=0.021 (Fig. 3(b)) begins to turn positive at high temperatures. When xx further increases, the slope is always positive at all measured temperatures (see Figs. 3(c) and (d) as examples). This trend indicates that Ta doping introduces hole carriers to the samples and the hole carriers are dominant at x>0.021x>0.021. Moreover, the additional hole carriers destroy the nearly perfect electron-hole compensation, resulting in the linear positive slope of ρxy(B)\rho_{xy}(B) at x>0.021x>0.021.

Refer to caption
Figure 3: Magnetic field dependence of Hall resistivity ρxy(B)\rho_{xy}(B) of Mo1-xTaxTe2 with (a) x=0x=0, (b) x=0.021x=0.021, (c) x=0.065x=0.065, and (d) x=0.173x=0.173 collected during warm-up. The colour scale at the right indicates the measured temperature.

To further visualize the temperature evolution of the Hall effect of Mo1-xTaxTe2, we extract the Hall coefficient RHR_{H} from the slope of ρxy(B)\rho_{xy}(B) in the linear region, and the temperature evolution of RHR_{H} is plotted in Fig. 4. The RHR_{H} data measured at high temperatures during cool-down are also displayed. We find that a thermal hysteresis can also be observed in the RHR_{H} data of the samples from x=0x=0 to x=0.097x=0.097, while the hysteresis is absent in the sample with x0.118x\geq 0.118. These results are consistent with the observation of the first-order structural transition in the ρ(T)\rho(T) data in Fig. 2(a). At x=0x=0 (Fig. 4(b)), RHR_{H} shows a strong temperature dependence below 50 K, which is similar to the result reported in previous studies [36, 34]. Upon Ta doping, RHR_{H} shifts toward the positive side due to the introduction of additional hole carriers, while the temperature dependence is relatively mild compared to x=0x=0. The most prominent temperature profile of RHR_{H} in Ta-doped samples is x=0.065x=0.065, where the magnitude of RHR_{H} (|RH||R_{H}|) gradually increases with decreasing temperature and reaches the maximum value at 2 K. Interestingly, our results show that |RH||R_{H}| at 2 K is the largest around x=0.065x=0.065 (see also Fig. 3(c), where ρxy(B)\rho_{xy}(B) has a steeper slope at 2 K compared to that in Fig. 3(d)), which is different from the expectation that RHR_{H} would increase toward the positive side when xx increases. This issue will be further discussed in the later section.

Refer to caption
Figure 4: Temperature dependence of Hall coefficient RHR_{H} of (a) Mo1-xTaxTe2 with x0x\neq 0 and (b) pristine MoTe2 (x=0x=0). The closed (open) symbols represent the warm-up (cool-down) data. The cool-down data are only shown at high temperatures.

We summarize our results and construct a temperature-doping phase diagram of Mo1-xTaxTe2 in Fig. 5, which shows the Ta-doping dependence of TsT_{s} and TcT_{c}. The structural transition temperatures acquired during warm-up (Ts,warmT_{s,warm}) and cool-down (Ts,coolT_{s,cool}) are defined by the extrema of the first derivative of ρ(T)\rho(T) around the transition (see Fig. S3 in Supplemental Material [33]). Both Ts,warmT_{s,~{}warm} and Ts,coolT_{s,~{}cool} show a generally decreasing trend with increasing xx. Compared to Ts,warmT_{s,~{}warm}, Ts,coolT_{s,~{}cool} decreases more rapidly with increasing xx. When x0.118x\geq 0.118, both Ts,warmT_{s,~{}warm} and Ts,coolT_{s,~{}cool} are completely suppressed. On the other hand, after experiencing a local maximum at x=x= 0.042, TcT_{c} also decreases when xx increases, and drops to zero at x0.065x\geq 0.065 (before TsT_{s} vanishes). The disappearance of superconductivity is unique in our hole-doping phase diagram; in the previous phase diagram studies of MoTe2 upon pressure [11, 22, 23, 24, 25, 13], isovalent chemical substitution [22, 26], and electron doping [27, 28], they typically show the anticorrelation of TcT_{c} and TsT_{s} as well as a huge enhancement of TcT_{c}.

Refer to caption
Figure 5: Temperature-doping phase diagram of Mo1-xTaxTe2. The upward (downward) blue triangles represent TsT_{s} defined from the temperature-dependent resistivity data measured during warm-up (cool-down). The solid cyan circles represent TcT_{c}. The solid curves are guides for the eyes. The colour contour denotes the temperature dependence of Hall coefficient RHR_{H} at different doping levels.

To shed light on the issue of why the superconductivity of MoTe2 is suppressed upon hole doping, a contour plot of RHR_{H} is overlaid in Fig. 5. We reveal that |RH||R_{H}| is significantly enhanced around the region when TsT_{s} is suppressed to zero (x0.1x\sim 0.1), and TcT_{c} vanishes when the enhancement of |RH||R_{H}| emerges at x0.05x\sim 0.05. Compared to the previous studies with other tuning parameters, while TcT_{c} increases, low-temperature |RH||R_{H}| has either a weak electron-doping dependence [27, 28] or decreases with pressure [13]. Meanwhile, a similar enhancement of RHR_{H} has been observed in another hole doping study, Nb-doped MoTe2 [32]. Such enhancement is associated with the enhancement of thermopower divided by temperature S/TS/T, which is maximum around the region where TsT_{s} is completely suppressed; our RHR_{H} contour plot is reminiscent of the contour plot of S/TS/T reported in the phase diagram of Nb-doped MoTe2 (Fig. 1(b) in Ref. [32]). According to Sakai et al.’s argument, both enhancements of RHR_{H} and S/TS/T are attributed to the strong fluctuation or phase separation around the nonpolar-polar structural transition, giving rise to some critical scattering effects on the carriers [32]. Combining this statement with our phase diagram, the critical scattering may also hinder the formation of Cooper pairs, and therefore suppress superconductivity. Further investigations on the competition between superconductivity and critical scattering are highly desired to confirm this picture.

Another possible explanation for the suppression of superconductivity is related to the change in the Fermi surface topology upon hole doping. Cho et al. [27] have performed theoretical calculations on the impact of electron and hole doping on TcT_{c}. While they have attributed the increase in TcT_{c} upon electron doping (arising from Te vacancy in MoTe2-x) to the enhancement of the density of states at the Fermi level (N(EF)N(E_{F})) and the electron-phonon coupling constant (λ\lambda), they have also predicted that, upon hole doping, N(EF)N(E_{F}) and λ\lambda will be suppressed and therefore TcT_{c} will decrease, which is consistent with our experimental findings. Cho et al. further attributed the change in λ\lambda to phonon vectors connecting between electron Fermi pockets, which are enlarged upon electron doping according to their calculations. In contrast, upon hole doping, electron pockets shrink and only spherical-shaped hole pockets remain at the Γ\Gamma point [32, 27]. In the situation without phonon vectors linking between electron pockets, λ\lambda will be suppressed and hence TcT_{c} will be reduced. Therefore, our study has provided solid experimental evidence to showcase Cho et al.’s theoretical prediction.

Refer to caption
Figure 6: Temperature-doping phase diagram of hole-doped and electron-doped MoTe2. The hole-doping data (solid symbols) are our findings of Mo1-xTaxTe2, as shown in Fig. 5, while the electron-doping data (open symbols) are adapted from the findings of Te-deficient MoTe2-x from Cho et al. [27].

To further elaborate on this idea, we connect our hole-doping phase diagram with the electron-doping phase diagram (based on the result of Te-deficient MoTe2 from Cho et al. [27]) and plot the combined phase diagram in Fig. 6. It unambiguously shows the asymmetry between the hole-doping phase and electron-doping diagrams, which is reminiscent of different behaviours between hole-doped and electron-doped cuprate superconductors [37, 38]. While TsT_{s} shows a similar suppression upon both hole- and electron-doping, the doping dependence of TcT_{c} behaves differently. At the electron-doping region (the right-hand side of Fig. 6), TcT_{c} is largely enhanced. However, when we move to the hole-doping region (the left-hand side of Fig. 6), TcT_{c} is heavily suppressed. This demonstrates a clear trend that TcT_{c} can be induced and enhanced when the electron carrier concentration increases, no matter what the phase is.

Meanwhile, although the critical scattering around the structural transition may contribute to the suppression of superconductivity, our result shows that the tuning of the carrier concentration, which controls the phonon nesting vector(s), provides an effective means to vary the TcT_{c} of MoTe2, regardless of the suppression of TsT_{s}. These findings provide experimental evidence that enhancing the TcT_{c} of MoTe2 by solely increasing the electron carrier concentration while preserving the topologically nontrivial TdT_{d} phase is possible. Such property can potentially boost the progress of the search for topological superconductivity in MoTe2, which is currently hindered by its low TcT_{c}.

IV Conclusions

In summary, we have investigated the phase diagram of Ta-doped MoTe2, Mo1-xTaxTe2, with x=00.173x=0-0.173 through magnetotransport measurements. Single crystals of Mo1-xTaxTe2 were successfully grown by the self-flux method. X-ray diffraction and energy-dispersive X-ray spectroscopy have confirmed that Mo is partially substituted by Ta in the doped samples. By measuring the temperature dependence of resistivity and the Hall effect, we have revealed that the structural transition temperature TsT_{s} is completely suppressed at x0.11x\sim 0.11, while the superconducting transition TcT_{c} generally decreases upon Ta doping and finally vanishes at x0.08x\sim 0.08. This behaviour is in contrast to the previous phase diagrams constructed based on applying pressure, isovalent doping, or electron doping, which show the enhancement of TcT_{c} when TsT_{s} is suppressed. Moreover, the Hall coefficient is found to be enhanced at low temperatures around the region where TsT_{s} is suppressed to zero, suggesting that the critical scattering arising from the structural temperature may have some contributions to the suppression of TcT_{c}. By comparing our findings with the phase diagram of electron-doped MoTe2, we argue that the electron carrier concentration in MoTe2 is a key factor in controlling TcT_{c}, which offers a straightforward way to boost the TcT_{c} of MoTe2.

Notes added: After the first submission of this article, we noticed a recently published article [39] which reports an enhancement of TcT_{c} in Ta-doped MoTe2. Our results do not agree with those of Ref. [39]. The discrepancy may be attributed to methodological differences. First, Ref. [39] used a different crystal growth condition. Second, we determine our TcT_{c} values based on the observation of zero resistivity while Ref. [39] deduced their TcT_{c} values from the onset of the transition in resistivity. We note that zero resistivity has not been observed in the doped samples in Ref. [39].

Acknowledgements.
We acknowledge Xinyou Liu, Ying Kit Tsui, Wei Zhang, and Lingfei Wang for fruitful discussions, and financial support from the Research Grants Council of Hong Kong (GRF/14300419, GRF/14301020 and A-CUHK402/19), CUHK Direct Grant (4053463, 4053528, 4053408 and 4053461), and the National Natural Science Foundation of China (12104384).

References

  • He et al. [2016] L. He, Y. Jia, S. Zhang, X. Hong, C. Jin,  and S. Li, “Pressure-induced superconductivity in the three-dimensional topological Dirac semimetal Cd3As2,” npj Quantum Mater. 1, 16014 (2016).
  • Zhou et al. [2016a] Y. Zhou, J. Wu, W. Ning, N. Li, Y. Du, X. Chen, R. Zhang, Z. Chi, X. Wang, X. Zhu, P. Lu, C. Ji, X. Wan, Z. Yang, J. Sun, W. Yang, M. Tian, Y. Zhang,  and H.-k. Mao, “Pressure-induced superconductivity in a three-dimensional topological material ZrTe5,” Proc. Natl. Acad. Sci. U.S.A. 113, 2904 (2016a).
  • Meinert [2016] M. Meinert, “Unconventional Superconductivity in YPtBi and Related Topological Semimetals,” Phys. Rev. Lett. 116, 137001 (2016).
  • Butch et al. [2011] N. P. Butch, P. Syers, K. Kirshenbaum, A. P. Hope,  and J. Paglione, “Superconductivity in the topological semimetal YPtBi,” Phys. Rev. B 84, 220504 (2011).
  • Bay et al. [2012] T. V. Bay, T. Naka, Y. K. Huang,  and A. de Visser, “Superconductivity in noncentrosymmetric YPtBi under pressure,” Phys. Rev. B 86, 064515 (2012).
  • Kim et al. [2018] H. Kim, K. Wang, Y. Nakajima, R. Hu, S. Ziemak, P. Syers, L. Wang, H. Hodovanets, J. D. Denlinger, P. M. R. Brydon, D. F. Agterberg, M. A. Tanatar, R. Prozorov,  and J. Paglione, “Beyond triplet: Unconventional superconductivity in a spin-3/2 topological semimetal,” Sci. Adv. 4, eaao4513 (2018).
  • Pan et al. [2015] X.-C. Pan, X. Chen, H. Liu, Y. Feng, Z. Wei, Y. Zhou, Z. Chi, L. Pi, F. Yen, F. Song, X. Wan, Z. Yang, B. Wang, G. Wang,  and Y. Zhang, “Pressure-driven dome-shaped superconductivity and electronic structural evolution in tungsten ditelluride,” Nat. Commun. 6, 7805 (2015).
  • Kang et al. [2015] D. Kang, Y. Zhou, W. Yi, C. Yang, J. Guo, Y. Shi, S. Zhang, Z. Wang, C. Zhang, S. Jiang, A. Li, K. Yang, Q. Wu, G. Zhang, L. Sun,  and Z. Zhao, “Superconductivity emerging from a suppressed large magnetoresistant state in tungsten ditelluride,” Nat. Commun. 6, 7804 (2015).
  • Lu et al. [2016] P. Lu, J.-S. Kim, J. Yang, H. Gao, J. Wu, D. Shao, B. Li, D. Zhou, J. Sun, D. Akinwande, D. Xing,  and J.-F. Lin, “Origin of superconductivity in the Weyl semimetal WTe2\mathrm{WT}{\mathrm{e}}_{2} under pressure,” Phys. Rev. B 94, 224512 (2016).
  • Chan et al. [2017] Y. T. Chan, P. L. Alireza, K. Y. Yip, Q. Niu, K. T. Lai,  and S. K. Goh, “Nearly isotropic superconductivity in the layered Weyl semimetal WTe2{\mathrm{WTe}}_{2} at 98.5 kbar,” Phys. Rev. B 96, 180504 (2017).
  • Qi et al. [2016] Y. Qi, P. G. Naumov, M. N. Ali, C. R. Rajamathi, W. Schnelle, O. Barkalov, M. Hanfland, S.-C. Wu, C. Shekhar, Y. Sun, V. Süß, M. Schmidt, U. Schwarz, E. Pippel, P. Werner, R. Hillebrand, T. Förster, E. Kampert, S. Parkin, R. J. Cava, C. Felser, B. Yan,  and S. A. Medvedev, “Superconductivity in Weyl semimetal candidate MoTe2,” Nat. Commun. 7, 11038 (2016).
  • Rhodes et al. [2017] D. Rhodes, R. Schönemann, N. Aryal, Q. Zhou, Q. R. Zhang, E. Kampert, Y.-C. Chiu, Y. Lai, Y. Shimura, G. T. McCandless, J. Y. Chan, D. W. Paley, J. Lee, A. D. Finke, J. P. C. Ruff, S. Das, E. Manousakis,  and L. Balicas, “Bulk Fermi surface of the Weyl type-II semimetallic candidate γMoTe2\gamma-{\mathrm{MoTe}}_{2},” Phys. Rev. B 96, 165134 (2017).
  • Hu et al. [2019] Y. J. Hu, Y. T. Chan, K. T. Lai, K. O. Ho, X. Guo, H.-P. Sun, K. Y. Yip, D. H. L. Ng, H.-Z. Lu,  and S. K. Goh, “Angular dependence of the upper critical field in the high-pressure 1T1{T}^{{}^{\prime}} phase of MoTe2{\mathrm{MoTe}}_{2},” Phys. Rev. Mater. 3, 034201 (2019).
  • Hu et al. [2020] Y. J. Hu, W. C. Yu, K. T. Lai, D. Sun, F. F. Balakirev, W. Zhang, J. Y. Xie, K. Y. Yip, E. I. P. Aulestia, R. Jha, R. Higashinaka, T. D. Matsuda, Y. Yanase, Y. Aoki,  and S. K. Goh, “Detection of hole pockets in the candidate type-II Weyl semimetal MoTe2{\mathrm{MoTe}}_{2} from Shubnikov–de Haas quantum oscillations,” Phys. Rev. Lett. 124, 076402 (2020).
  • Sato and Ando [2017] M. Sato and Y. Ando, “Topological superconductors: a review,” Rep. Prog. Phys. 80, 076501 (2017).
  • Li and Xu [2019] Y. Li and Z.-A. Xu, “Exploring topological superconductivity in topological materials,” Adv. Quantum Technol. 2, 1800112 (2019).
  • Soluyanov et al. [2015] A. A. Soluyanov, D. Gresch, Z. Wang, Q. Wu, M. Troyer, X. Dai,  and B. A. Bernevig, “Type-II Weyl semimetals,” Nature 527, 495 (2015).
  • Deng et al. [2016] K. Deng, G. Wan, P. Deng, K. Zhang, S. Ding, E. Wang, M. Yan, H. Huang, H. Zhang, Z. Xu, J. Denlinger, A. Fedorov, H. Yang, W. Duan, H. Yao, Y. Wu, S. Fan, H. Zhang, X. Chen,  and S. Zhou, “Experimental observation of topological Fermi arcs in type-II Weyl semimetal MoTe2,” Nat. Phys. 12, 1105 (2016).
  • Huang et al. [2016] L. Huang, T. M. McCormick, M. Ochi, Z. Zhao, M.-T. Suzuki, R. Arita, Y. Wu, D. Mou, H. Cao, J. Yan, N. Trivedi,  and A. Kaminski, “Spectroscopic evidence for a type II Weyl semimetallic state in MoTe2,” Nat. Mater. 15, 1155 (2016).
  • Jiang et al. [2017] J. Jiang, Z. K. Liu, Y. Sun, H. F. Yang, C. R. Rajamathi, Y. P. Qi, L. X. Yang, C. Chen, H. Peng, C.-C. Hwang, S. Z. Sun, S.-K. Mo, I. Vobornik, J. Fujii, S. S. P. Parkin, C. Felser, B. H. Yan,  and Y. L. Chen, “Signature of type-II Weyl semimetal phase in MoTe2,” Nat. Commun. 8, 13973 (2017).
  • Wang et al. [2020] W. Wang, S. Kim, M. Liu, F. A. Cevallos, R. J. Cava,  and N. P. Ong, “Evidence for an edge supercurrent in the Weyl superconductor MoTe2,” Science 368, 534 (2020).
  • Takahashi et al. [2017] H. Takahashi, T. Akiba, K. Imura, T. Shiino, K. Deguchi, N. K. Sato, H. Sakai, M. S. Bahramy,  and S. Ishiwata, “Anticorrelation between polar lattice instability and superconductivity in the Weyl semimetal candidate MoTe2{\mathrm{MoTe}}_{2},” Phys. Rev. B 95, 100501 (2017).
  • Heikes et al. [2018] C. Heikes, I.-L. Liu, T. Metz, C. Eckberg, P. Neves, Y. Wu, L. Hung, P. Piccoli, H. Cao, J. Leao, J. Paglione, T. Yildirim, N. P. Butch,  and W. Ratcliff, “Mechanical control of crystal symmetry and superconductivity in Weyl semimetal MoTe2{\mathrm{MoTe}}_{2},” Phys. Rev. Mater. 2, 074202 (2018).
  • Lee et al. [2018] S. Lee, J. Jang, S.-I. Kim, S.-G. Jung, J. Kim, S. Cho, S. W. Kim, J. Y. Rhee, K.-S. Park,  and T. Park, “Origin of extremely large magnetoresistance in the candidate type-II Weyl semimetal MoTe2-x,” Sci. Rep. 8, 13937 (2018).
  • Guguchia et al. [2017] Z. Guguchia, F. von Rohr, Z. Shermadini, A. T. Lee, S. Banerjee, A. R. Wieteska, C. A. Marianetti, B. A. Frandsen, H. Luetkens, Z. Gong, S. C. Cheung, C. Baines, A. Shengelaya, G. Taniashvili, A. N. Pasupathy, E. Morenzoni, S. J. L. Billinge, A. Amato, R. J. Cava, R. Khasanov,  and Y. J. Uemura, “Signatures of the topological s+- superconducting order parameter in the type-II Weyl semimetal TdT_{d}-MoTe2,” Nat. Commun. 8, 1082 (2017).
  • Chen et al. [2016] F. C. Chen, X. Luo, R. C. Xiao, W. J. Lu, B. Zhang, H. X. Yang, J. Q. Li, Q. L. Pei, D. F. Shao, R. R. Zhang, L. S. Ling, C. Y. Xi, W. H. Song,  and Y. P. Sun, “Superconductivity enhancement in the S-doped Weyl semimetal candidate MoTe2,” Appl. Phys. Lett. 108, 162601 (2016).
  • Cho et al. [2017] S. Cho, S. H. Kang, H. S. Yu, H. W. Kim, W. Ko, S. W. Hwang, W. H. Han, D.-H. Choe, Y. H. Jung, K. J. Chang, Y. H. Lee, H. Yang,  and S. W. Kim, “Te vacancy-driven superconductivity in orthorhombic molybdenum ditelluride,” 2D Mater. 4, 021030 (2017).
  • Mandal et al. [2018] M. Mandal, S. Marik, K. P. Sajilesh, Arushi, D. Singh, J. Chakraborty, N. Ganguli,  and R. P. Singh, “Enhancement of the superconducting transition temperature by Re doping in Weyl semimetal MoTe2{\mathrm{MoTe}}_{2},” Phys. Rev. Mater. 2, 094201 (2018).
  • Dahal et al. [2020] R. Dahal, L. Z. Deng, N. Poudel, M. Gooch, Z. Wu, H. C. Wu, H. D. Yang, C. K. Chang,  and C. W. Chu, “Tunable structural phase transition and superconductivity in the Weyl semimetal Mo1xWxTe2{\mathrm{Mo}}_{1-x}{\mathrm{W}}_{x}{\mathrm{Te}}_{2},” Phys. Rev. B 101, 140505 (2020).
  • Rhodes et al. [2021] D. A. Rhodes, A. Jindal, N. F. Q. Yuan, Y. Jung, A. Antony, H. Wang, B. Kim, Y.-c. Chiu, T. Taniguchi, K. Watanabe, K. Barmak, L. Balicas, C. R. Dean, X. Qian, L. Fu, A. N. Pasupathy,  and J. Hone, “Enhanced superconductivity in monolayer TdT_{d}-MoTe2,” Nano Lett. 21, 2505 (2021).
  • Ikeura et al. [2015] K. Ikeura, H. Sakai, M. S. Bahramy,  and S. Ishiwata, “Rich structural phase diagram and thermoelectric properties of layered tellurides Mo1-xNbxTe2,” APL Mater. 3, 041514 (2015).
  • Sakai et al. [2016] H. Sakai, K. Ikeura, M. S. Bahramy, N. Ogawa, D. Hashizume, J. Fujioka, Y. Tokura,  and S. Ishiwata, “Critical enhancement of thermopower in a chemically tuned polar semimetal MoTe2,” Sci. Adv. 2, e1601378 (2016).
  • [33] See Supplemental Material for the representative EDX spectrum with all doped samples, the raw data of Hall resistivity ρxy\rho_{xy} of all doping levels, from which we extracted RHR_{H}, and the dρ(T)/dTd\rho(T)/dT data on cooling or warming for determination of TsT_{s} .
  • Zandt et al. [2007] T. Zandt, H. Dwelk, C. Janowitz,  and R. Manzke, “Quadratic temperature dependence up to 50 K of the resistivity of metallic MoTe2,” J. Alloys Compd. 442, 216 (2007).
  • Yip et al. [2023] K. Y. Yip, S. T. Lam, K. H. Yu, W. S. Chow, J. Zeng, K. T. Lai,  and S. K. Goh, “Drastic enhancement of the superconducting temperature in type-II Weyl semimetal candidate MoTe2 via biaxial strain,” APL Mater. 11, 021111 (2023).
  • Zhou et al. [2016b] Q. Zhou, D. Rhodes, Q. R. Zhang, S. Tang, R. Schönemann,  and L. Balicas, “Hall effect within the colossal magnetoresistive semimetallic state of MoTe2,” Phys. Rev. B 94, 121101 (2016b).
  • Armitage et al. [2010] N. P. Armitage, P. Fournier,  and R. L. Greene, “Progress and perspectives on electron-doped cuprates,” Rev. Mod. Phys. 82, 2421 (2010).
  • Keimer et al. [2015] B. Keimer, S. A. Kivelson, M. R. Norman, S. Uchida,  and J. Zaanen, “From quantum matter to high-temperature superconductivity in copper oxides,” Nature 518, 179 (2015).
  • Zhang et al. [2023] Y. Zhang, F. Fei, R. Liu, T. Zhu, B. Chen, T. Qiu, Z. Zuo, J. Guo, W. Tang, L. Zhou, et al., “Enhanced Superconductivity and Upper Critical Field in Ta-Doped Weyl Semimetal TdT_{d}-MoTe2,” Advanced Materials 35, 2207841 (2023).