Skip to main content

Main menu

  • Home
  • Articles
    • Current
    • Special Feature Articles - Most Recent
    • Special Features
    • Colloquia
    • Collected Articles
    • PNAS Classics
    • List of Issues
  • Front Matter
    • Front Matter Portal
    • Journal Club
  • News
    • For the Press
    • This Week In PNAS
    • PNAS in the News
  • Podcasts
  • Authors
    • Information for Authors
    • Editorial and Journal Policies
    • Submission Procedures
    • Fees and Licenses
  • Submit
  • Submit
  • About
    • Editorial Board
    • PNAS Staff
    • FAQ
    • Accessibility Statement
    • Rights and Permissions
    • Site Map
  • Contact
  • Journal Club
  • Subscribe
    • Subscription Rates
    • Subscriptions FAQ
    • Open Access
    • Recommend PNAS to Your Librarian

User menu

  • Log in
  • My Cart

Search

  • Advanced search
Home
Home
  • Log in
  • My Cart

Advanced Search

  • Home
  • Articles
    • Current
    • Special Feature Articles - Most Recent
    • Special Features
    • Colloquia
    • Collected Articles
    • PNAS Classics
    • List of Issues
  • Front Matter
    • Front Matter Portal
    • Journal Club
  • News
    • For the Press
    • This Week In PNAS
    • PNAS in the News
  • Podcasts
  • Authors
    • Information for Authors
    • Editorial and Journal Policies
    • Submission Procedures
    • Fees and Licenses
  • Submit
Research Article

Abrupt change of the superconducting gap structure at the nematic critical point in FeSe1−xSx

Yuki Sato, Shigeru Kasahara, Tomoya Taniguchi, Xiangzhuo Xing, View ORCID ProfileYuichi Kasahara, Yoshifumi Tokiwa, Youichi Yamakawa, Hiroshi Kontani, Takasada Shibauchi, and View ORCID ProfileYuji Matsuda
  1. aDepartment of Physics, Kyoto University, Kyoto 606-8502, Japan;
  2. bCenter for Electronic Correlations and Magnetism, Institute of Physics, Augsburg University, 86159 Augsburg, Germany;
  3. cDepartment of Physics, Nagoya University, Nagoya 464-8602, Japan;
  4. dDepartment of Advanced Materials Science, University of Tokyo, Chiba 277-8561, Japan

See allHide authors and affiliations

PNAS February 6, 2018 115 (6) 1227-1231; first published January 23, 2018; https://doi.org/10.1073/pnas.1717331115
Yuki Sato
aDepartment of Physics, Kyoto University, Kyoto 606-8502, Japan;
  • Find this author on Google Scholar
  • Find this author on PubMed
  • Search for this author on this site
Shigeru Kasahara
aDepartment of Physics, Kyoto University, Kyoto 606-8502, Japan;
  • Find this author on Google Scholar
  • Find this author on PubMed
  • Search for this author on this site
  • For correspondence: matsuda@scphys.kyoto-u.ac.jp kasa@scphys.kyoto-u.ac.jp
Tomoya Taniguchi
aDepartment of Physics, Kyoto University, Kyoto 606-8502, Japan;
  • Find this author on Google Scholar
  • Find this author on PubMed
  • Search for this author on this site
Xiangzhuo Xing
aDepartment of Physics, Kyoto University, Kyoto 606-8502, Japan;
  • Find this author on Google Scholar
  • Find this author on PubMed
  • Search for this author on this site
Yuichi Kasahara
aDepartment of Physics, Kyoto University, Kyoto 606-8502, Japan;
  • Find this author on Google Scholar
  • Find this author on PubMed
  • Search for this author on this site
  • ORCID record for Yuichi Kasahara
Yoshifumi Tokiwa
bCenter for Electronic Correlations and Magnetism, Institute of Physics, Augsburg University, 86159 Augsburg, Germany;
  • Find this author on Google Scholar
  • Find this author on PubMed
  • Search for this author on this site
Youichi Yamakawa
cDepartment of Physics, Nagoya University, Nagoya 464-8602, Japan;
  • Find this author on Google Scholar
  • Find this author on PubMed
  • Search for this author on this site
Hiroshi Kontani
cDepartment of Physics, Nagoya University, Nagoya 464-8602, Japan;
  • Find this author on Google Scholar
  • Find this author on PubMed
  • Search for this author on this site
Takasada Shibauchi
dDepartment of Advanced Materials Science, University of Tokyo, Chiba 277-8561, Japan
  • Find this author on Google Scholar
  • Find this author on PubMed
  • Search for this author on this site
Yuji Matsuda
aDepartment of Physics, Kyoto University, Kyoto 606-8502, Japan;
  • Find this author on Google Scholar
  • Find this author on PubMed
  • Search for this author on this site
  • ORCID record for Yuji Matsuda
  • For correspondence: matsuda@scphys.kyoto-u.ac.jp kasa@scphys.kyoto-u.ac.jp
  1. Edited by Zachary Fisk, University of California, Irvine, CA, and approved December 26, 2017 (received for review October 3, 2017)

  • Article
  • Figures & SI
  • Info & Metrics
  • PDF
Loading

Significance

Electronic nematicity that spontaneously breaks the rotational symmetry of the underlying crystal lattice has been a growing issue in high-temperature superconductivity of iron pnictides/chalcogenides and cuprates. FeSe1 − xSx, in which the nematicity can be tuned by isoelectronic sulfur substitution, offers a fascinating opportunity to clarify the direct relationship between the nematicity and superconductivity. Here, we discover a dramatic change in the superconducting gap structure at the critical concentration of sulfur where the nematicity disappears, i.e., nematic critical point. Our observation provides direct evidence that the orbital-dependent nature of the critical nematic fluctuations has a strong impact on the superconducting pairing interaction.

Abstract

The emergence of the nematic electronic state that breaks rotational symmetry is one of the most fascinating properties of the iron-based superconductors, and has relevance to cuprates as well. FeSe has a unique ground state in which superconductivity coexists with a nematic order without long-range magnetic ordering, providing a significant opportunity to investigate the role of nematicity in the superconducting pairing interaction. Here, to reveal how the superconducting gap evolves with nematicity, we measure the thermal conductivity and specific heat of FeSe1 − xSx, in which the nematicity is suppressed by isoelectronic sulfur substitution and a nematic critical point (NCP) appears at xc≈0.17. We find that, in the whole nematic regime (0≤x≤0.17), the field dependence of two quantities consistently shows two-gap behavior; one gap is small but highly anisotropic with deep minima or line nodes, and the other is larger and more isotropic. In stark contrast, in the tetragonal regime (x=0.20), the larger gap becomes strongly anisotropic, demonstrating an abrupt change in the superconducting gap structure at the NCP. Near the NCP, charge fluctuations of dxz and dyz orbitals are enhanced equally in the tetragonal side, whereas they develop differently in the orthorhombic side. Our observation therefore directly implies that the orbital-dependent nature of the nematic fluctuations has a strong impact on the superconducting gap structure and hence on the pairing interaction.

  • superconductivity
  • iron-based superconductors
  • nematicity
  • pairing interaction
  • superconducting gap structure

Spin fluctuations are widely discussed as a primary driving force of various unconventional superconductors, whereas, in iron-based superconductors, spin and orbital degrees of freedoms are closely intertwined because of the multiple d-orbital characters at the Fermi level (1, 2). In most iron-based superconductors, tetragonal–orthorhombic structural (nematic) and magnetic transition lines follow closely each other. These orders have been suggested to play crucial roles in superconductivity, and thus strong spin and/or orbital fluctuations have been proposed to mediate the pairing (3⇓–5). However, despite tremendous efforts in the past years, elucidating the exact pairing mechanism still remains a great challenge.

The iron chalcogenide superconductor FeSe (6), comprised only an Fe-Se layer, offers a novel platform to investigate the pairing mechanism of iron-based superconductors, because it displays several remarkable properties. The superconducting transition temperature of Tc≈9 K dramatically increases up to 38 K by the application of hydrostatic pressure (7). The superconductivity at ambient pressure coexists with a nematic order, whose properties are distinctly different from the other iron-based superconductors. The nematic transition occurs at Ts≈90 K, which is accompanied by the energy splitting of the Fe d orbits (8⇓⇓⇓⇓⇓⇓–13). Although Ts is comparable to other iron-based superconductors, no sizable low-energy spin fluctuations are observed above Ts and no long-range magnetic order occurs below Ts at ambient pressure (14⇓⇓–17). These results have raised questions regarding the spin fluctuation scenario envisaged in other iron-based superconductors. Although there is an argument that the magnetic fluctuation mechanism is still applicable (18⇓–20), an alternative scenario where fluctuations stemming from orbital degree of freedom play a primary role has aroused great interest (16, 17, 21⇓–23).

As shown in Fig. 1A, the Fermi surface in the nematic phase consists of an elliptical hole pocket at the Brillouin zone center (h1), elongated along the Γ–My line, and compensated electron pockets near the zone boundary (e1 and e2) (23). It has been reported that the e1 pocket is divided into two Dirac-like electrons in the presence of large orbital splitting (24⇓–26), although the detailed structure of the Fermi surface is still controversial (8⇓⇓⇓⇓–13, 27, 28). The size of all of the pockets is extremely small, occupying only 1 to 3% of the whole Brillouin zone (29⇓–31). Since the superconducting gap structure is intimately related to the pairing interaction, its elucidation is crucially important. The superconducting gap of FeSe has been reported to be highly anisotropic with deep minima or line nodes (31⇓–33).

Fig. 1.
  • Download figure
  • Open in new tab
  • Download powerpoint
Fig. 1.

(A and B) Schematic illustrations of the Fermi surface in the nematic and the tetragonal phases (23). Green, red, and blue areas represent the Fermi surface regimes dominated by dxz, dyz, and dxy orbital characters, respectively. (C) T–x phase diagram of FeSe1 − xSx. (D–G) Temperature dependence of resistivity (thick lines) and its temperature derivative (thin lines) for single crystals of FeSe1 − xSx for (D) x=0.08, (E) x = 0.13, (F) x = 0.16, and (G) x = 0.20. Thick arrows indicate the nematic transition temperatures.

The large anisotropy of the superconducting gap in FeSe is highly unusual because it directly implies that the pairing interaction strongly depends on the position of a tiny Fermi surface. However, the relationship between the nematicity and pairing interaction remains largely elusive. To tackle this key issue, it is of primary importance to clarify how the nematicity affects the superconducting gap structure. Isoelectronic sulfur substitution in FeSe provides the most suitable route to study this issue. In the series of FeSe1 − xSx, the density of states at the Fermi level (or bandwidth) can be tuned significantly through chemical pressure, leading to a change of the electron correlation effect (34⇓⇓⇓–38). Indeed, with increasing x, the nematic transition temperature is suppressed without inducing long-range magnetic order, and the system can be tuned to a nonmagnetic tetragonal regime (39), whose band structure is shown in Fig. 1B (23). The T - x phase diagram of FeSe1 − xSx is depicted in Fig. 1C. The elastoresistance measurements reveal that, as x is increased in the nematic regime, the nematic fluctuations are strongly enhanced with x. and, near xc≈0.17, where Ts is suppressed to zero, the nematic susceptibility diverges toward absolute zero, indicating a nematic critical point (NCP) (39). In the nematic regime, the energy splitting of d orbitals is suppressed with x and elliptical h1 pocket becomes more circular while keeping its volume nearly constant (36). FeSe1 − xSx, therefore, offers a fascinating opportunity to investigate the role of nematic fluctuation on superconductivity. Here we report the superconducting gap structure of FeSe1 − xSx in a wide x range from the nematic to tetragonal regime, which is determined by the thermal conductivity κ and specific heat C.

Results and Discussion

Fig. 1 D–G depicts the T dependences of the resistivity ρ and dρ/dT for x=0.08, 0.13, 0.16, and 0.20, respectively. The nematic transition temperatures determined by the jump of dρ/dT are Ts≈75, 60, and 35 K for x=0.08, 0.13, and 0.16, respectively. These values are consistent with the previous report (39). At x=0.20, no anomaly is observed in dρ/dT, indicating that the system is in the tetragonal regime. Tc and upper critical field Hc2 for x=0.20 are rapidly suppressed from those in the nematic regime.

Fig. 2 shows the T dependences of the electronic component of specific heat divided by temperature, Ce/T, for x=0, 0.08, 0.13, and 0.20, respectively. We obtained Ce by subtracting the change of Cn(μ0H=14 T) in the normal state from the value at Tc, Ce=C(T)−ΔCn, and ΔCn=Cn(T)−γTc. For x=0, 0.08, 0.13, and 0.20, the values of μ0Hc2 are ∼16, 20, 18, and 3 T, respectively. Since μ0Hc2(T) in the nematic regime exceeds the maximum field of our experimental setup, 14 T, at low temperatures, Cn(T) below Tc(14T) is estimated by extrapolating a curve obtained by the fitting of C above Tc with Cn(T)=γT+βT3+A5T5. At Tc, Ce/T exhibits a sharp jump for all x, showing good homogeneity of S substitution. The Sommerfeld coefficient γ is 7 mJ/mol⋅K2 to 9 mJ/mol⋅K2 for all crystals in the nematic regime, suggesting that the electron correlation is little influenced by S substitution. The ratio of specific heat jump and normal state-specific heat, ΔCe/γTc = 1.5 for x=0, is larger than the Bardeen–Cooper–Schrieffer (BCS) value of 1.43, while, for 0.08, 0.13, and 0.2, ΔCe/γTc=1.3, 1.1, and 0.82, respectively, are smaller than the BCS value. This may be due to the multigap nature of the superconductivity. For x=0.20 in the tetragonal regime, Tc determined by Ce/T is slightly lower than that determined by zero resistivity. Ce/T below Tc shows a concave-downward curvature, which also supports the multigap superconductivity.

Fig. 2.
  • Download figure
  • Open in new tab
  • Download powerpoint
Fig. 2.

The electronic component of the specific heat divided by temperature, Ce/T, vs. T in FeSe1 − xSx for (A) x=0, (B) x = 0.08, (C) x = 0.13, and (D) x = 0.20. Dotted horizontal lines indicate the normal state values of Ce/T.

Fig. 3 shows the H dependences of C/T at around 450 mK for x=0, 0.08, 0.13, and 0.20, respectively. In conventional fully gapped superconductors, C(H)/T increases linearly with H due to the induced quasiparticles inside vortex cores. In stark contrast, as shown in Fig. 3, Insets, C(H)/T increases with H for all x at low fields. In superconductors with a highly anisotropic gap, the Doppler shift of the delocalized quasiparticle spectrum induces remarkable field dependence of density of states with H dependence for line node. For x=0, 0.08, and 0.13 in the nematic regime, C(H)/T deviates from the H dependence at H∗, shown by arrows. For x=0.08 and 0.13, C(H)/T exhibits a kink at H∗. Above H∗, C(H)/T increases slowly as C(H)/T∝Hα with α≳ 1. The slight upward curvature of C(H)/T above H∗ for x=0 and 0.13 is attributed to the Pauli paramagnetic effect on the superconductivity (40). The initial steep increase of C(H)/T below H∗ indicates that a substantial portion of the quasiparticles is already restored at a magnetic field far below Hc2. The slope change at H∗ provides evidence for multigap superconductivity; H∗ is interpreted as a virtual upper critical field that determines the H dependence of the smaller gap. The H behavior below H∗ indicates the presence of a Fermi pocket, whose superconducting gap is small and highly anisotropic with line node or deep minima. Moreover, Hα dependences with α≳1 above H∗ indicate the presence of another Fermi pocket, whose gap is much larger and isotropic.

Fig. 3.
  • Download figure
  • Open in new tab
  • Download powerpoint
Fig. 3.

Magnetic field dependence of C/T for (A) x=0, (B) 0.08, (C) 0.13, and (D) 0.20. Insets show the same data plotted as a function of μ0H. H* represents a magnetic field at which C/T deviates from H dependence. The black arrows in the main panels indicate H*. The gray arrow in D indicates upper critical field.

For x=0.20 in the tetragonal regime, H behavior is observed in the whole H regime below Hc2, which is determined by the resistivity. As shown in Figs. 2D and 3D, large C/T at H=0 indicates that a substantial number of quasiparticles are excited even at T/Tc≈0.1. Since entropy balance imposes ∫0Tc{(Ce/T)−Cn/Tc)}dT=0, Ce/T for x=0.20 is expected to decrease rapidly with decreasing T below 0.4 K. Therefore, the remaining C/T arises from the Fermi pockets with extremely small superconducting gaps. The H dependence of C(H)/T for x=0.20 suggests the presence of Fermi pocket(s) with very small gap and other pocket(s), whose gap is larger and highly anisotropic. These results lead us to conclude that the gap structure in the tetragonal regime is essentially different from that in the nematic regime.

The thermal conductivity provides additional pivotal information on the superconducting gap structure, because the heat transport detects only the delocalized quasiparticles, insensitive to the localized quasiparticles. Fig. 4A depicts κ/T plotted as a function of T2 in zero field. At low temperature, κ/T is well fitted by κ/T=κ0/T+bT2, where b is a constant. We confirmed that the ratio of κ0/T and the electrical conductivity σ0 at T→0 above μ0Hc2 is (κ0/T)/σ0=(1.04±0.02)L0 for x=0.16 and 0.20, where L0=π2/3(kB/e) is the Lorenz number, indicating that the Wiedemann–Franz law holds. At zero field, the presence of a residual value in κ/T at T→0, κ00/T, indicates the presence of normal fluid, which can be attributed to the presence of line nodes in the gap function. Finite κ00/T is clearly resolved in x=0.08, 0.16, and 0.20, indicating the presence of line node. On the other hand, κ00/T for x=0.13 is much smaller or vanishes at T→0.

Fig. 4.
  • Download figure
  • Open in new tab
  • Download powerpoint
Fig. 4.

(A) Temperature dependence of thermal conductivity divided by T, κ/T, plotted as a function of T2 in zero field. Magnetic field dependence of κ(H)/T for (B) x= 0.08, (C) x = 0.13, (D) x = 0.16, and (E) x = 0.20. The gray arrows indicate Hc2. Insets show κ/T plotted as a function of μ0H. H* represents a magnetic field at which κ(H)/T deviates from the H dependence. The black arrows in the main panels indicate H*.

Fig. 4 B–E depicts the H dependences of κ(H)/T for x=0.08, 0.13, 0.16, and 0.20. Similar to C(H)/T, the application of small magnetic fields causes a steep increase of κ(H)/T for all x; as shown in Fig. 4 B–E, Insets, κ(H)/T increases with H at low fields. Similarly to the specific heat, the H dependence of κ(H)/T appears as a result of Doppler shift of quasiparticle spectra in the presence of line nodes. For x=0.08, 0.16, and 0.20, κ(H)/T increases immediately when the magnetic field is applied. [We note that the lower critical field Hc1 is much smaller than the field scale of interest (34).] This H dependence, along with the presence of finite κ00/T, indicates the presence of line nodes. For x=0.13, on the other hand, κ/T(H) is insensitive to H at very low fields even above Hc1, suggesting that, although the gap function has a deep minimum at certain directions, it is finite, i.e., no nodes. This is consistent with very small or absent κ00/T. As shown in Fig. 4 B–D, Insets, κ(H)/T deviates from the H dependence above H∗ for x=0.08, 0.13, and 0.16. The values of H∗ for x=0.08 and 0.13 are close to the ones observed in C(H)/T in Fig. 3 B and C. Above H∗, κ(H)/T shows much weaker H dependence than below H∗. In particular, κ(H)/T is nearly H-independent for x=0.08 and 0.16. Since thermal conductivity is insensitive to localized quasiparticles inside vortices, κ(H)/T in fully gapped superconductors is independent of H except in the vicinity of Hc2. Thus the observed initial steep increase with H dependence, followed by much weaker H dependence of κ(H)/T, provides evidence for the multigap superconductivity, in which a small gap has line nodes or deep minima and a large gap is nearly isotropic (41). This is consistent with the conclusion drawn from the specific heat. For x=0.20, κ/T increases with H nearly up to Hc2, which is again consistent with the specific heat.

Next, we compare our results with other experimental observations. It has been reported by angle-resolved photoemission spectroscopy (ARPES) and quasiparticle interference (QPI) for x<0.07 that the superconducting gap of the h1 pocket is highly anisotropic, with deep minima or nodes (42, 43). Therefore, it is natural to assume that the observed H dependences of C(H)/T and κ(H)/T at low fields for x<0.08 come from the anisotropic gap of the h1 pocket. In our experiments for a wider x range, this initial H dependence persists in the whole nematic regime. These observations suggest that the superconducting gap in the h1 pocket is always highly anisotropic in the whole nematic regime. The gap structure of the electron pockets has been less clear. In fact, no gap has been observed on the electron pockets in ARPES measurements (42). In the QPI measurements, anisotropic gap is inferred for the e1 pocket, but the gap structure of e2 pocket has not been resolved (43). However, the fact that H∗ is much smaller than Hc2 implies that the gap of the e2 pocket is larger than that of the h1 pocket. Moreover, H dependences of C/T and κ/T above H∗ suggest that the gap of the e2 pocket is much more isotropic than that of the h1 pocket. It should be stressed that the line nodes in the h1 pocket are accidental, not symmetry-protected, because, as directly revealed by the scanning tunneling microscopy measurements, the nodes are lifted near the twin boundaries (44). Moreover, the presence of line nodes has been reported by thermal conductivity measurements on some crystals for x=0 (31), while a small but finite gap has been observed in different crystals (33), which may be attributed to the difference in the amount of impurities and twin boundaries.

Since the elliptical h1 pocket becomes more circular with increasing x, the highly anisotropic gap in the h1 pocket in the whole nematic regime implies that the anisotropic pairing interaction is little influenced by the elliptical distortion of the h1 pocket. This immediately excludes the possibility of the intraband pairing, in which the superconductivity is mediated by fluctuations with very small momentum. This is because, in such a case, the gap anisotropy should be sensitive to the shape of the Fermi surface. As displayed by the green area in the h1 pocket in Fig. 1A, the gap node/minimum locates at the area dominated by dxz orbital character for x=0 and 0.07 (42, 43) To explain such a highly anisotropic gap in a tiny Fermi pocket, a pairing interaction which is strongly orbital dependent has been proposed (26, 45). In this scenario, the gap minimum/node appears as a result of the strong nesting properties of dyz orbit area, shown by red in Fig. 1A, between the h1 and e1 pockets. Since the pairing interaction is dominated by dyz orbital, the gap minimum/node can appear in the area with dxz orbital character in the h1 pocket. In fact, strong nesting properties between dyz orbitals has been discussed in BaFe2As2 with stripe-type magnetic order.

Although the detailed superconducting gap structure in the tetragonal regime requires further investigation, the present results reveal a dramatic change in the gap function across the NCP (Fig. 1C, SC1 and SC2). Near the NCP, charge fluctuations of dxz and dyz orbitals are enhanced equally in the tetragonal side. On the other hand, they develop differently in the orthorhombic side. Thus, the nature of the nematic charge fluctuations drastically changes at the NCP. In the tetragonal phase (x>xc), strong charge fluctuations of dxz and dyz orbitals develop near x=xc, due to the Aslamazov–Larkin vertex correction (22, 23). In the orthorhombic phase (x<xc), the orbital splitting ΔE=Eyz−Exz increases rapidly in proportion to (xc−x)1/2, by which the ferro-nematic fluctuations are suppressed. At the same time, the emergence of ΔE causes a large imbalance of charge fluctuations between dxz and dyz orbitals, reflecting the improvement of the dyz orbital nesting condition with ΔE>0 (26). Such a change in the orbital fluctuations gives rise to the abrupt change of the superconducting gap structure at x=xc, which may also be relevant to the change of Tc at xc. The present results therefore strongly indicate that the orbital selectivity of the nematic fluctuations plays an essential role for the superconductivity of FeSe. Intriguingly, a nodal superconducting state has also been reported in tetragonal FeS (46, 47). We also note that a possible dx2−y2 pairing is proposed based on spin fluctuation theory (48). Pinning down the position of nodes and effect of orbital fluctuations in this end material would provide further clues to elucidate the pairing mechanism in iron-based superconductors.

In summary, the thermal conductivity and specific heat measurements on FeSe1 − xSx in a wide x range provide bulk evidence for the presence of deep minima or line nodes in the superconducting gap function in both the whole nematic and tetragonal regimes. Moreover, the multigap nature of the superconductivity is commonly observed in both regimes. These results imply that the pairing interaction is significantly anisotropic in both the nematic and tetragonal regimes. We find that the gap structure dramatically changes when crossing the NCP. This demonstrates that the orbital-dependent nature of the nematic fluctuations has a strong impact on the superconducting pairing interaction, which should provide a clue to understanding a pairing mechanism of highly unusual superconductivity in FeSe.

Materials and Methods

Single crystals of FeSe1 − xSx (x=0, 0.08, 0.13, 0.16, and 0.20) were grown by chemical vapor transport technique (39, 49). Observation of quantum oscillations, even in the heavily substituted sample with x=0.2 (50), the nearly 100% Meissner signal, and the sharp jump in specific heat all demonstrate the high quality of the samples. Specific heat was measured for x=0, 0.08, 0.13, and 0.20 by the quasi-adiabatic method in 3He cryostat. The thermal conductivity was measured on crystals with the same x values by the standard steady-state method by applying the thermal current in the 2D plane in a dilution refrigerator. In addition to these crystals, we measured κ for x=0.16 in the vicinity of NCP. Since the physical properties of the crystals near NCP are sensitive to the inhomogeneous distribution of sulfur, we carefully selected a tiny crystal with a sharp superconducting transition. For both C and κ measurements, we applied a magnetic field perpendicular to the 2D plane (H ∥c).

Acknowledgments

We thank T. Watashige for experimental support, and T. Hanaguri for helpful discussion. This work was supported by Grants-in-Aid for Scientific Research (KAKENHI) 25220710, 15H02106, 15H03688, and 15KK0160, and Grant-in-Aid for Scientific Research on Innovative Areas “Topological Materials Science” 15H05852 from Japan Society for the Promotion of Science.

Footnotes

  • ↵1To whom correspondence may be addressed. Email: matsuda{at}scphys.kyoto-u.ac.jp or kasa{at}scphys.kyoto-u.ac.jp.
  • Author contributions: S.K., T.S., and Y.M. designed research; Y.S., S.K., T.T., X.X., Y.K., Y.T., Y.Y., H.K., and Y.M. performed research; Y.S., S.K., T.T., and X.X. analyzed data; and S.K., T.S., and Y.M. wrote the paper.

  • The authors declare no conflict of interest.

  • This article is a PNAS Direct Submission.

Published under the PNAS license.

References

  1. ↵
    1. Chubukov AV,
    2. Khodas M,
    3. Fernandes RM
    (2016) Magnetism, superconductivity, and spontaneous orbital order in iron-based superconductors: Which comes first and why? Phys Rev X 6:041045.
    OpenUrl
  2. ↵
    1. Hirschfeld PJ,
    2. Korshunov MM,
    3. Mazin II
    (2011) Gap symmetry and structure of Fe-based superconductors. Rep Prog Phys 74:124508.
    OpenUrlCrossRef
  3. ↵
    1. Mazin II,
    2. Singh DJ,
    3. Johannes MD,
    4. Du MH
    (2008) Unconventional superconductivity with a sign reversal in the order parameter of LaFeAsO1−xFx. Phys Rev Lett 101:057003.
    OpenUrlCrossRefPubMed
  4. ↵
    1. Kuroki K,
    2. Onari S,
    3. Arita S,
    4. Usui H,
    5. Tanaka Y,
    6. Kontani H,
    7. Aoki H
    (2008) Unconventional pairing originating from the disconnected Fermi surfaces of superconducting LaFeAsO1 − xFx. Phys Rev Lett 101:087004.
    OpenUrlCrossRefPubMed
  5. ↵
    1. Kontani H,
    2. Onari S
    (2010) Orbital-fluctuation-mediated superconductivity in iron pnictides: Analysis of the five-orbital Hubbard-Holstein model. Phys Rev Lett 104:157001.
    OpenUrlCrossRefPubMed
  6. ↵
    1. Hsu F-C, et al.
    (2008) Superconductivity in the PbO-type structure α-FeSe. Proc Natl Acad Sci USA 105:14262–14264.
    OpenUrlAbstract/FREE Full Text
  7. ↵
    1. Sun JP, et al.
    (2016) Dome-shaped magnetic order competing with high-temperature superconductivity at high pressures in FeSe. Nat Commun 7:12146.
    OpenUrl
  8. ↵
    1. Nakayama K, et al.
    (2014) Reconstruction of band structure induced by electronic nematicity in an FeSe superconductor. Phys Rev Lett 113:237001.
    OpenUrl
  9. ↵
    1. Shimojima T, et al.
    (2014) Lifting of xz/yz orbital degeneracy at the structural transition in detwinned FeSe. Phys Rev B 90:121111.
    OpenUrl
  10. ↵
    1. Watson MD, et al.
    (2015) Emergence of the nematic electronic state in FeSe. Phys Rev B 91:155106.
    OpenUrl
  11. ↵
    1. Suzuki Y, et al.
    (2015) Momentum-dependent sign inversion of orbital order in superconducting FeSe. Phys Rev B 92:205117.
    OpenUrl
  12. ↵
    1. Fedorov A, et al.
    (2016) Effect of nematic ordering on electronic structure of FeSe. Sci Rep 6:36834.
    OpenUrlCrossRefPubMed
  13. ↵
    1. Coldea AI,
    2. Watson MD
    (2017) The key ingredients of the electronic structure of FeSe. arXiv:1706.00338.
  14. ↵
    1. Imai T, et al.
    (2009) Why does undoped FeSe become a high-Tc superconductor under pressure? Phys Rev Lett 102:177005.
    OpenUrlCrossRefPubMed
  15. ↵
    1. McQueen TM, et al.
    (2009) Tetragonal-to-orthorhombic structural phase transition at 90 K in the superconductor Fe1.01Se. Phys Rev Lett 103:057002.
    OpenUrlCrossRefPubMed
  16. ↵
    1. Böhmer AE, et al.
    (2015) Origin of the tetragonal-to-orthorhombic phase transition in FeSe: A combined thermodynamic and NMR study of nematicity. Phys Rev Lett 114:027001.
    OpenUrlCrossRefPubMed
  17. ↵
    1. Baek S-H, et al.
    (2015) Orbital-driven nematicity in FeSe. Nat Mater 14:210–214.
    OpenUrlPubMed
  18. ↵
    1. Chubukov AV,
    2. Fernandes RM,
    3. Schmalian J
    (2015) Origin of nematic order in FeSe. Phys Rev B 91:201105.
    OpenUrl
  19. ↵
    1. Wang F,
    2. Kivelson SA,
    3. Lee D-H
    (2015) Nematicity and quantum paramagnetism in FeSe. Nat Phys 11:959–963.
    OpenUrlCrossRef
  20. ↵
    1. Yu R,
    2. Si Q
    (2015) Antiferroquadrupolar and Ising-nematic orders of a frustrated bilinear-biquadratic Heisenberg model and implications for the magnetism of FeSe. Phys Rev Lett 115:116401.
    OpenUrlCrossRefPubMed
  21. ↵
    1. Massat P, et al.
    (2016) Charge-induced nematicity in FeSe. Proc Natl Acad Sci USA 113:9177–9181.
    OpenUrlAbstract/FREE Full Text
  22. ↵
    1. Yamakawa Y,
    2. Onari S,
    3. Kontani H
    (2016) Nematicity and magnetism in FeSe and other families of Fe-based superconductors. Phys Rev X 6:021032.
    OpenUrl
  23. ↵
    1. Onari S,
    2. Yamakawa Y,
    3. Kontani H
    (2016) Sign-reversing orbital polarization in the nematic phase of FeSe due to the C2 symmetry breaking in the self-energy. Phys Rev Lett 116:227001.
    OpenUrl
  24. ↵
    1. Tan SY, et al.
    (2016) Observation of Dirac cone band dispersions in FeSe thin films by photoemission spectroscopy. Phys Rev B 93:104513.
    OpenUrl
  25. ↵
    1. Zhang Y, et al.
    (2016) Distinctive orbital anisotropy observed in the nematic state of a FeSe thin film. Phys Rev B 94:115153.
    OpenUrl
  26. ↵
    1. Yamakawa Y,
    2. Kontani H
    (2016) Nematicity, magnetism and superconductivity in FeSe under pressure: Unified explanation based on the self-consistent vertex correction theory. arXiv:1609.09618.
  27. ↵
    1. Watson MD, et al.
    (2106) Evidence for unidirectional nematic bond ordering in FeSe. Phys Rev B 94:201107(R).
    OpenUrl
  28. ↵
    1. Watson MD,
    2. Haghighirad AA,
    3. Rhodes LC,
    4. Hoesch M,
    5. Kim TK
    (2017) Electronic anisotropies in the nematic phase of FeSe. arXiv:1705.02286.
  29. ↵
    1. Terashima T, et al.
    (2014) Anomalous Fermi surface in FeSe seen by Shubnikov-de Haas oscillation measurements. Phys Rev B 90:144517.
    OpenUrl
  30. ↵
    1. Watson MD, et al.
    (2015) Dichotomy between the hole and electron behaviour in multiband superconductor FeSe probed by ultrahigh magnetic fields. Phys Rev Lett 115:027006.
    OpenUrlCrossRefPubMed
  31. ↵
    1. Kasahara S, et al.
    (2014) Field-induced superconducting phase of FeSe in the BCS-BEC cross-over. Proc Natl Acad Sci USA 111:16309–16313.
    OpenUrlAbstract/FREE Full Text
  32. ↵
    1. Song C-L, et al.
    (2011) Direct observation of nodes and twofold symmetry in FeSe superconductor. Science 332:1410–1413.
    OpenUrlAbstract/FREE Full Text
  33. ↵
    1. Bourgeois-Hope P, et al.
    (2016) Thermal conductivity of the iron-based superconductor FeSe: Nodeless gap with a strong two-band character. Phys Rev Lett 117:097003.
    OpenUrlCrossRefPubMed
  34. ↵
    1. Abdel-Hafiez M, et al.
    (2015) Superconducting properties of sulfur-doped iron selenide. Phys Rev B 91:165109.
    OpenUrl
  35. ↵
    1. Moore SA, et al.
    (2015) Evolution of the superconducting properties in FeSe1 − xSx. Phys Rev B 92:235113.
    OpenUrl
  36. ↵
    1. Watson MD, et al.
    (2015) Suppression of orbital ordering by chemical pressure in FeSe1 − xSx. Phys Rev B 92:121108(R).
    OpenUrl
  37. ↵
    1. Miao J, et al.
    (2017) Electronic structure of FeS. Phys Rev B 95:205127.
    OpenUrl
  38. ↵
    1. Reiss P, et al.
    (2017) Suppression of electronic correlations by chemical pressure from FeSe to FeS. arXiv:1705.11139.
  39. ↵
    1. Hosoi S, et al.
    (2016) Nematic quantum critical point without magnetism in FeSe1 − xSx superconductors. Proc Natl Acad Sci USA 113:8139–8143.
    OpenUrlAbstract/FREE Full Text
  40. ↵
    1. Tsutsumi Y,
    2. Machida K,
    3. Ichioka M
    (2015) Hidden crossover phenomena in strongly Pauli-limited multiband superconductors: Application to CeCu2Si2. Phys Rev B 92:020502.
    OpenUrl
  41. ↵
    1. Watashige T, et al.
    (2017) Quasiparticle excitations in the superconducting state of FeSe probed by thermal Hall conductivity in the vicinity of the BCS-BEC crossover. J Phys Soc Jpn 86:014707.
    OpenUrl
  42. ↵
    1. Xu HC, et al.
    (2016) Highly anisotropic and twofold symmetric superconducting gap in nematically ordered FeSe0.93S0.07. Phys Rev Lett 117:157003.
    OpenUrlCrossRefPubMed
  43. ↵
    1. Sprau PO, et al.
    (2017) Discovery of orbital-selective Cooper pairing in FeSe. Science 357:75–80.
    OpenUrlAbstract/FREE Full Text
  44. ↵
    1. Watashige T, et al.
    (2015) Evidence for time-reversal symmetry breaking of the superconducting state near twin-boundary interfaces in FeSe revealed by scanning tunnelling spectroscopy. Phys Rev X 5:031022.
    OpenUrl
  45. ↵
    1. Kreisel A, et al.
    (2017) Orbital selective pairing and gap structures of iron-based superconductors. Phys Rev B 95:174504.
    OpenUrl
  46. ↵
    1. Xing J, et al.
    (2016) Nodal superconducting gap in tetragonal FeS. Phys Rev B 93:104520.
    OpenUrl
  47. ↵
    1. Ying TP, et al.
    (2016) Nodal superconductivity in FeS: Evidence from quasiparticle heat transport. Phys Rev B 94:100504(R).
    OpenUrl
  48. ↵
    1. Yang Y,
    2. Wang W-S,
    3. Lu H-Y,
    4. Xiang Y-Y,
    5. Wang Q-H
    (2016) Electronic structure anddx2−y2-wave superconductivity in FeS. Phys Rev B 93:104514.
    OpenUrl
  49. ↵
    1. Böhmer AE, et al.
    (2013) Lack of coupling between superconductivity and orthorhombic distortion in stoichiometric single-crystalline FeSe. Phys Rev B 87:180505.
    OpenUrl
  50. ↵
    1. Coldea AI, et al.
    (2016) Evolution of the Fermi surface of the nematic superconductors FeSe1 − xSx. arXiv:1611.07424.
PreviousNext
Back to top
Article Alerts
Email Article

Thank you for your interest in spreading the word on PNAS.

NOTE: We only request your email address so that the person you are recommending the page to knows that you wanted them to see it, and that it is not junk mail. We do not capture any email address.

Enter multiple addresses on separate lines or separate them with commas.
Abrupt change of the superconducting gap structure at the nematic critical point in FeSe1−xSx
(Your Name) has sent you a message from PNAS
(Your Name) thought you would like to see the PNAS web site.
CAPTCHA
This question is for testing whether or not you are a human visitor and to prevent automated spam submissions.
Citation Tools
Abrupt change of superconducting gap in FeSe1−xSx
Yuki Sato, Shigeru Kasahara, Tomoya Taniguchi, Xiangzhuo Xing, Yuichi Kasahara, Yoshifumi Tokiwa, Youichi Yamakawa, Hiroshi Kontani, Takasada Shibauchi, Yuji Matsuda
Proceedings of the National Academy of Sciences Feb 2018, 115 (6) 1227-1231; DOI: 10.1073/pnas.1717331115

Citation Manager Formats

  • BibTeX
  • Bookends
  • EasyBib
  • EndNote (tagged)
  • EndNote 8 (xml)
  • Medlars
  • Mendeley
  • Papers
  • RefWorks Tagged
  • Ref Manager
  • RIS
  • Zotero
Request Permissions
Share
Abrupt change of superconducting gap in FeSe1−xSx
Yuki Sato, Shigeru Kasahara, Tomoya Taniguchi, Xiangzhuo Xing, Yuichi Kasahara, Yoshifumi Tokiwa, Youichi Yamakawa, Hiroshi Kontani, Takasada Shibauchi, Yuji Matsuda
Proceedings of the National Academy of Sciences Feb 2018, 115 (6) 1227-1231; DOI: 10.1073/pnas.1717331115
del.icio.us logo Digg logo Reddit logo Twitter logo CiteULike logo Facebook logo Google logo Mendeley logo
  • Tweet Widget
  • Facebook Like
  • Mendeley logo Mendeley

Article Classifications

  • Physical Sciences
  • Physics
Proceedings of the National Academy of Sciences: 115 (6)
Table of Contents

Submit

Sign up for Article Alerts

Jump to section

  • Article
    • Abstract
    • Results and Discussion
    • Materials and Methods
    • Acknowledgments
    • Footnotes
    • References
  • Figures & SI
  • Info & Metrics
  • PDF

You May Also be Interested in

Smoke emanates from Japan’s Fukushima nuclear power plant a few days after tsunami damage
Core Concept: Muography offers a new way to see inside a multitude of objects
Muons penetrate much further than X-rays, they do essentially zero damage, and they are provided for free by the cosmos.
Image credit: Science Source/Digital Globe.
Water from a faucet fills a glass.
News Feature: How “forever chemicals” might impair the immune system
Researchers are exploring whether these ubiquitous fluorinated molecules might worsen infections or hamper vaccine effectiveness.
Image credit: Shutterstock/Dmitry Naumov.
Venus flytrap captures a fly.
Journal Club: Venus flytrap mechanism could shed light on how plants sense touch
One protein seems to play a key role in touch sensitivity for flytraps and other meat-eating plants.
Image credit: Shutterstock/Kuttelvaserova Stuchelova.
Illustration of groups of people chatting
Exploring the length of human conversations
Adam Mastroianni and Daniel Gilbert explore why conversations almost never end when people want them to.
Listen
Past PodcastsSubscribe
Panda bear hanging in a tree
How horse manure helps giant pandas tolerate cold
A study finds that giant pandas roll in horse manure to increase their cold tolerance.
Image credit: Fuwen Wei.

Similar Articles

Site Logo
Powered by HighWire
  • Submit Manuscript
  • Twitter
  • Facebook
  • RSS Feeds
  • Email Alerts

Articles

  • Current Issue
  • Special Feature Articles – Most Recent
  • List of Issues

PNAS Portals

  • Anthropology
  • Chemistry
  • Classics
  • Front Matter
  • Physics
  • Sustainability Science
  • Teaching Resources

Information

  • Authors
  • Editorial Board
  • Reviewers
  • Subscribers
  • Librarians
  • Press
  • Cozzarelli Prize
  • Site Map
  • PNAS Updates
  • FAQs
  • Accessibility Statement
  • Rights & Permissions
  • About
  • Contact

Feedback    Privacy/Legal

Copyright © 2021 National Academy of Sciences. Online ISSN 1091-6490