Skip to main content
Proceedings of the National Academy of Sciences of the United States of America logoLink to Proceedings of the National Academy of Sciences of the United States of America
. 2014 Nov 6;111(46):16309–16313. doi: 10.1073/pnas.1413477111

Field-induced superconducting phase of FeSe in the BCS-BEC cross-over

Shigeru Kasahara a,1, Tatsuya Watashige a,b,1, Tetsuo Hanaguri b, Yuhki Kohsaka b, Takuya Yamashita a, Yusuke Shimoyama a, Yuta Mizukami a,c, Ryota Endo a, Hiroaki Ikeda a,2, Kazushi Aoyama a,d, Taichi Terashima e, Shinya Uji e, Thomas Wolf f, Hilbert von Löhneysen f, Takasada Shibauchi a,c, Yuji Matsuda a,3
PMCID: PMC4246302  PMID: 25378706

Significance

The BCS-BEC (Bardeen–Cooper–Schrieffer––Bose–Einstein-condensate) cross-over bridges the two important theories of bound particles in a unified picture with the ratio of the attractive interaction to the Fermi energy as a tuning parameter. A key issue is to understand the intermediate regime, where new states of matter may emerge. Here, we show that the Fermi energy of FeSe is extremely small, resulting in that this system can be regarded as an extraordinary “high-temperature” superconductor located at the verge of a BCS-BEC cross-over. Most importantly, we discover the emergence of an unexpected superconducting phase in strong magnetic fields, demonstrating that the Zeeman splitting comparable to the Fermi energy leads to a strong modification of the properties of fermionic systems in such a regime.

Keywords: BCS-BEC cross-over, Fermi energy, quasiparticle interference, iron-based superconductors, exotic superconducting phase

Abstract

Fermi systems in the cross-over regime between weakly coupled Bardeen–Cooper–Schrieffer (BCS) and strongly coupled Bose–Einstein-condensate (BEC) limits are among the most fascinating objects to study the behavior of an assembly of strongly interacting particles. The physics of this cross-over has been of considerable interest both in the fields of condensed matter and ultracold atoms. One of the most challenging issues in this regime is the effect of large spin imbalance on a Fermi system under magnetic fields. Although several exotic physical properties have been predicted theoretically, the experimental realization of such an unusual superconducting state has not been achieved so far. Here we show that pure single crystals of superconducting FeSe offer the possibility to enter the previously unexplored realm where the three energies, Fermi energy εF, superconducting gap Δ, and Zeeman energy, become comparable. Through the superfluid response, transport, thermoelectric response, and spectroscopic-imaging scanning tunneling microscopy, we demonstrate that εF of FeSe is extremely small, with the ratio Δ/εF1(0.3) in the electron (hole) band. Moreover, thermal-conductivity measurements give evidence of a distinct phase line below the upper critical field, where the Zeeman energy becomes comparable to εF and Δ. The observation of this field-induced phase provides insights into previously poorly understood aspects of the highly spin-polarized Fermi liquid in the BCS-BEC cross-over regime.


Superconductivity in most metals is well explained by the weak-coupling Bardeen–Cooper–Schrieffer (BCS) theory, where the pairing instability arises from weak attractive interactions in a degenerate fermionic system. In the opposite limit of Bose–Einstein condensate (BEC), composite bosons consisting of strongly coupled fermions condense into a coherent quantum state (1, 2). In BCS superconductors, the superconducting transition temperature is usually several orders of magnitude smaller than the Fermi temperature, Tc/TF=105104, whereas in the BEC limit Tc/TF is of the order of 101. Even in the high-Tc cuprates, Tc/TF is merely of the order of 102 at optimal doping. Of particular interest is the BCS-BEC cross-over regime with intermediate coupling strength. In this regime the size of interacting pairs (ξ), which is known as the coherence length, becomes comparable to the average distance between particles (1/kF), i.e., kFξ1 (35), where kF is the Fermi momentum. This regime is expected to have the highest values of Tc/TF=0.10.2 and Δ/εF0.5 ever observed in any fermionic superfluid.

One intriguing issue concerns the role of spin imbalance: whether it will lead to a strong modification of the properties of the Fermi system in the cross-over regime. This problem has been of considerable interest not only in the context of superconductivity but also in ultracold-atom physics (68). However, such Fermi systems have been extremely hard to access. In superconductors, the spin imbalance can be introduced through the Zeeman effect by applying a strong magnetic field. Again, in the high-Tc cuprates, the Zeeman energy at the upper critical field at TTc is of the order of only 102εF. In ultracold atoms, although several exotic superfluid states have been proposed (9, 10), cooling the systems down to sufficiently low temperature (TTc) is not easily attained.

FeSe provides an ideal platform for studying a highly spin-polarized Fermi system in the cross-over regime. FeSe is the simplest iron-based layered superconductor (Fig. 1A, Inset) with Tc of 9 K (11). The structural transition from tetragonal to orthorhombic crystal symmetry occurs at Ts 90 K and a large electronic in-plane anisotropy appears. In contrast with the other iron-based compounds, no magnetic order occurs below Ts. A prominent feature of the pseudobinary “11” family (FeSe1−x Tex) is the presence of very shallow pockets, as reported by angle-resolved photoemission spectroscopy (ARPES). Although a possible BCS-BEC cross-over has been suggested in the bands around the Γ-point (12, 13), it is still an open question whether all bands are in such a cross-over regime. Moreover, it should be noted that high-quality single crystals are highly requisite for the study of the cross-over regime, as exotic superconductivity often is extremely sensitive to impurities. Previous FeSex single crystals are strongly disordered, as indicated by large residual resistivity ρ0 and small residual resistivity ratio RRR, typically 0.1 mΩ cm and 5, respectively (14).

Fig. 1.

Fig. 1.

Normal and superconducting properties of high-quality single crystals of FeSe. (A) Temperature dependence of the in-plane resistivity ρ of FeSe. The structural transition occurs at Ts 90 K. (Upper Inset) Crystal structure. (Lower Inset) ρ(T) in magnetic field (Hc). From bottom to top, μ0H= 0, 2, 4, 6, 8, 10, 12, and 14 T is applied. (B) Temperature dependence of the London penetration depth. (Inset) Superfluid density normalized by the zero temperature value ρsλL2(0)/λL2(T). (C) Tunneling conductance spectrum at T=0.4 K. The peaks at ±2.5 meV (arrows) and shoulder structures at ±3.5 meV (dashed arrows) indicate the multiple superconducting gaps.

Results and Discussion

BCS-BEC Cross-Over.

By using high-quality single crystals of FeSe (SI Text, section 1 and Figs. S1 and S2) which have become available recently, we have measured the transport properties (Fig. 1A). In zero field, the temperature dependence of the resistivity ρ can be described by ρ=ρ0+ATα with α=1.051.2 below 25 K. Taking ρ(Tc+)10μΩcm as the upper limit of ρ0 leads to RRR>40, i.e., a factor of 10 higher than previous samples. In the present crystals Tc defined by the zero resistivity is 9.5 K, which is higher than Tc8 K of the low RRR samples (14). A remarkably large magnetoresistance (Fig. 1A, Inset) not observed in previously studied low-RRR crystals (14) supports that the crystal is very clean (SI Text, section 2 and Fig. S3A). The strongly T-dependent Hall coefficient RH below 60 K indicates that the electron and hole mobilities are of the same order (SI Text, section 2 and Fig. S3B). The London penetration depth λL shows a quasi–T-linear dependence, λL(T)T1.4, for T/Tc<0.2, suggesting the presence of line nodes in the superconducting gap (Fig. 1B). Fig. 1C shows the tunneling conductance, which is proportional to the density of states, measured with a scanning tunneling microscope at 0.4 K. The V-shaped spectrum at low bias voltages likewise indicates the presence of line nodes, which is consistent with previously reported observations (15). We note that the line nodes are accidental, not symmetry protected, i.e., the gap function is extended s wave, because the nodes are absent in samples with low RRR (16). Distinct peaks and shoulder structures in the spectra indicate the presence of (at least) two superconducting gaps (Δ2.5 and 3.5 meV), reflecting the multiband nature.

The high quality of the single crystals enables us to estimate the Fermi energies εFe and εFh from the band edges of electron and hole sheets, respectively, by using several techniques; all of them consistently point to extremely small Fermi energies. In 2D systems εF is related to λL(0) as εF=(π2d/μ0e2)λL2(0), where d is the interlayer distance and μ0 is the vacuum permeability. From the T dependence of λL(T), we obtain λL(0)400 nm (Fig. 1B and SI Text, section 3 and Fig. S4) (17). Very recent quantum oscillation measurements on the present FeSe crystals revealed that the Fermi surface consists of one hole sheet and one (compensating) electron sheet (Fig. 2A) (18). Then λL can be written as 1/λL2=1/(λLe)2+1/(λLh)2, where λLe and λLh represent the contribution from the electron and hole sheets, respectively. Assuming that two sheets have similar effective masses as indicated by the Hall effect (see below and SI Text, section 2), we estimate εFeεFh78 meV. The magnitude of the Fermi energy can also be inferred from the thermoelectric response in the normal state (SI Text, section 2). From the Seebeck coefficient S, the upper limit of εFe is deduced to be ∼10 meV (SI Text, section 2 and Fig. S3C). Moreover, the sign change of RH(T) at 60 K (SI Text, section 2 and Fig. S3B) indicates that the Fermi energies εFe and εFh are of similar size, a feature also observed in underdoped cuprate superconductors with small electron and hole pockets (19). In contrast with the cuprate case, however, RH in FeSe almost vanishes at high temperatures, which sheds light on the unique feature of FeSe with extremely small Fermi energy.

Fig. 2.

Fig. 2.

Band dispersions obtained from the QPI pattern. (A) Schematic figure of the Fermi surface of FeSe inferred from QPI. (B and C) QPI dispersions at 12 T obtained by taking linecuts from the energy-dependent Fourier-transformed normalized conductance images (SI Text, section 4 and Fig. S5) along qa and qb, respectively. A q-independent feature at ∼+10 meV is associated with defect states (SI Text, section 4 and Fig. S6). Peak positions of the representative branches are fitted with parabolic function to obtain Fermi energies and effective masses (solid lines). The top (bottom) of the hole (electron) band EHT (EEB) is indicated by white bars. Expected intraband scattering vectors associated with the α-band detected by ARPES (22) are plotted in B by yellow circles. (D and E) QPI dispersions at H=0. A pair of sharp intensity peaks (±Δ) appears at E±2 meV due to the opening of the superconducting gap. Superconducting gap defined by the positions of the coherence peaks (white bars) are comparable to the Fermi energies. Note that the feature attributed to defect states is independent of field.

We can determine the electron and hole Fermi energies directly by measuring the electronic dispersion curves in momentum space yielding εFh=EHTEF (εFe=EFEEB) for the hole (electron) band. Here EHT (EEB) is the energy of the top (bottom) of the hole (electron) band and EF is the electrochemical potential. For this assignment, we exploit spectroscopic-imaging scanning tunneling microscopy to observe the quasiparticle interference (QPI) patterns (20, 21) associated with electron waves scattered off by defects. By taking the Fourier transform of energy-dependent normalized conductance images, characteristic wave vectors of electrons at different energies reflecting the band dispersion can be determined (SI Text, section 4). The observed QPI patterns of FeSe at 1.5 K (SI Text, section 4 and Fig. S5) consist of hole- and electron-like branches that disperse along the crystallographic b and a(<b) directions, respectively. These branches can naturally be ascribed to the hole and electron sheets illustrated in Fig. 2A. The QPI signals exhibit a strong in-plane anisotropy. Such an anisotropy is consistent with the largely elongated vortex core structure (15). The origin of the strong in-plane anisotropy of the QPI signals is unclear, but a possible cause may be the orbital ordering in the orthorhombic phase.

As shown in Fig. 2 B and C, full dispersion curves of hole- and electron-like branches are clearly identified by taking linecuts from the series of Fourier-transformed conductance images (SI Text, section 4 and Fig. S5). Here, a magnetic field μ0H=12 T is applied parallel to the c axis (Hc) to mostly suppress superconductivity. Multiple hole-like branches are identified in Fig. 2B. Because QPI signals include both intra- and interband scattering processes, it is difficult to disentangle all of the QPI branches to resolve the bare band structure. Nevertheless, the top of the hole band can be faithfully determined to yield εFh 10 meV from the highest energy of the topmost branch. This branch is quantitatively consistent with the intraband scattering associated with the α-band detected by ARPES (22). In the case of the electron-like branch, an even smaller band bottom of εFe 2–3 meV is estimated (Fig. 2C). These small values are consistent with the ones estimated from the superfluid and thermoelectric responses. The effective mass of electron (hole) determined by QPI assuming parabolic dispersion is 2.5 m0 (3.5 m0), where m0 is the free-electron mass. The observation of comparable effective masses of electrons and holes is consistent with the Hall-effect data (SI Text, section 2). We stress that the electronic structure obtained from QPI, including masses of electron and hole, the size and the number of each pocket, and the magnitude of the Fermi energies, is consistent with the values recently reported by the quantum oscillations in the quantitative level (18). Remarkably, the superconducting gaps are of the same order as the Fermi energy of each band, Δ/εFe1 (Fig. 2D) and Δ/εFh0.3 (Fig. 2E), implying the BCS-BEC cross-over regime. Additional strong support of the cross-over is provided by extremely small kFξ1–4. Here, kF of the electron (hole) sheet obtained from Fig. 2C (Fig. 2B) is roughly 0.3 (0.75) nm−1, and ξ determined from the upper critical field (17 T) is roughly 5 nm.

Field-Induced Superconducting Phase.

So far we discussed the relation between εF and Δ. How does the Zeeman energy scale μBH, where μB is the Bohr magneton, enter the game? The thermal conductivity κ is well suited to address the issue of how the magnetic field affects the extraordinary pairing state by probing quasiparticle excitations out of the superconducting condensate, as the Cooper pair condensate does not contribute to heat transport. Figure 3A shows the T dependence of κ/T in zero field. Below Tc, κ is enhanced due to the suppression of quasiparticle scattering rates owing to the gap formation. As shown in Fig. 3A (Inset), κ/T at low temperatures is well fitted as κ/T=κ0/T+βT, similar to Tl2Ba2CuO6+δ (23). The presence of the residual κ0/T at T0 is consistent with line nodes in the gap.

Fig. 3.

Fig. 3.

Field-induced transition revealed by the thermal conductivity. (A) Temperature dependence of the in-plane thermal conductivity divided by temperature, κ/T. Arrow marked Tc indicates the onset temperature of the superconductivity determined by zero resistivity (Fig. 1A) and zero thermoelectric power (SI Text, section 2 and Fig. S3C). (Inset) κ/T at low temperatures. (B) Magnetic-field dependence of κ/T at low temperatures (Hc). No hysteresis with respect to the field-sweep direction is observed. κ/T shows a plateau-like behavior in a wide field range. At H, κ/T shows a cusp-like peak, suggestive of a nearly temperature-independent transition (dashed line). At 1.5 K, the cusp disappears and a weak structure (within the accuracy of the measurement) is observed at lower field. (C) Magnetic-field dependence of κ/T in the zero-temperature limit, κ0/T, obtained by linear extrapolation of κ/T versus T at low temperatures.

Fig. 3B shows the H dependence of κ/T for Hc well below Tc obtained after averaging over many field sweeps at constant temperatures. Beyond the initial steep drop at low fields, likely caused by the suppression of the quasiparticle mean-free path e through the introduction of vortices, κ(H)/T becomes nearly H-independent. Similar behavior has been reported for Bi2Sr2CaCu2O8 (24) and CeCoIn5 (25). It has been suggested that the nearly H-independent κ reflects a compensation between the enhancement of the density of states by magnetic field in nodal superconductors (Doppler shift) and the concomitant reduction in e due to increased scattering from vortices (26). At high fields, above the smoothly varying background, κ(H)/T exhibits a cusp-like feature at a field H that is practically independent of T. The height of the cusp-like peak decreases with increasing T. To further analyze our data, the κ/T values at different temperatures are extrapolated to T=0 for each field value measured to yield κ0(H)/T as shown in Fig. 3C, corroborating the robustness of the cusp. In particular, the cusp of κ0/T is unrelated to phonon heat transport because phonons do not contribute to κ/T for T0. Because the thermal conductivity has no fluctuation corrections (27), the cusp of κ usually corresponds to the mean-field phase transition. We note that at H the field dependence of magnetic torque shows no discernible anomaly (SI Text, section 5 and Fig. S7). However, such a difference of the sensitivity to the transition in different measurements has been reported for the field-induced transition between two superconducting phases in CeCoIn5, which is hardly resolved in magnetization (28), despite clear anomaly in some other bulk probes (29). Moreover, the hysteresis in the magnetization due to vortex pinning may smear out a possible torque anomaly at H.

Fig. 4 displays the H-T phase diagram for Hc. The irreversibility fields Hirr at low temperatures extend to fields well above H, indicating that H is located inside the superconducting state. The anomaly at H is not caused by some changes of the flux-line lattice, such as melting transition, because the peak field determined by the torque is strongly T-dependent and well below H (SI Text, section 5 and Fig. S7), indicating that the flux-line lattice is already highly disordered at H.

Fig. 4.

Fig. 4.

H-T phase diagram of FeSe for Hc. Solid blue and open red circles represent Hirr determined by zero resistivity and the onset of hysteresis loops of the magnetic torque, respectively. The mean-field upper critical field is above Hirr. Solid red circles represent H determined by the cusp of κ(H)/T shown in Fig. 3 B and C, separating a high-field superconducting phase (B phase) from the low-field phase (A phase). (Inset) Overall H-T phase diagram. Superconducting fluctuation regime is determined by the temperature at which the resistivity deviates from the behavior expected in conventional semimetallic systems.

As shown in Fig. 1A (Inset), the resistivity at μ0H=14 T increases with decreasing temperature and decreases after showing a broad maximum at around 15 K. The T dependence at high temperature is a typical behavior of the very pure compensated semimetals. However, the decrease of the resistivity at low-temperature regime is not expected in conventional semimetals. This unusual decrease may be attributed to the strong superconducting fluctuations above Hirr (Fig. 4, Inset). In higher fields the fluctuation region expands to higher T>Tc. In fact, the Ginzburg number, which is given by Gi(Tc/TF)4 within the BCS framework (30), is orders of magnitude larger than in any other superconductors. This large range of fluctuations may be related to the presence of preformed pairs predicted in the BCS-BEC cross-over regime (1, 35).

The appearance of the high-field phase (B phase in Fig. 4) where three characteristic energy scales are comparable, μBHεFΔ(0), suggests a phase transition of the Fermi liquid with strong spin imbalance in the BCS-BEC cross-over regime. Whether the observed distinct phase arises from strong spin imbalance and/or a BCS-BEC cross-over, however, needs to be resolved in the future with particular attention to multiband effects. We discuss two possible scenarios. (i) The phase line might signal an electronic transition akin to a Lifshitz transition, i.e., a topology change of the Fermi surface. Indeed, the phase line would be independent of T and smeared by thermal fluctuations. However, the fact that this phase line vanishes at Hirr would be accidental. Furthermore, the absence of any discernible anomaly in torque magnetometry at H (SI Text, section 5 and Fig. S7) implies that the κ(H)/T anomaly at H is not caused by a Lifshitz transition nor, for that matter, by a spin-density–wave type of magnetic order. (ii) Comparable Fermi and Zeeman energies may lead to an unprecedented superconducting state of highly spin-polarized electrons, such as spin-triplet pairing and an admixture of even- and odd-frequency pairing (31). Comparable gap and Zeeman energies may alternatively induce a Fulde–Ferrell–Larkin–Ovchinnikov (FFLO)-like state with Cooper pairs having finite total momentum (k, k+q) owing to the pairing channel between the Zeeman–split Fermi surfaces (29). The FFLO state requires a large Maki parameter, i.e., a ratio of orbital and Pauli-paramagnetic limiting fields, αM2Hc2orb/Hc2P>1.5 in the BCS limit. In this regime, αM2m/m0Δ/εF, yielding, for FeSe, αM as large as ∼5 (2.5) in the electron (hole) pockets. This estimate may be questionable in the regime of Δ/εF for FeSe. In any case, we stress that the high-field phase is not a simple FFLO phase because in the multiband superconductor FeSe the interaction between electron and hole bands is crucial. Even in the single-band systems, it has been suggested that the FFLO state becomes unstable in the cross-over regime (32). Our work should motivate further studies in the field of strongly interacting Fermi liquids near the BCS-BEC cross-over regime and in the presence of large spin imbalance, which remains largely unexplored and might bridge the areas of condensed-matter and ultracold-atom systems.

Supplementary Material

Supplementary File
pnas.201413477SI.pdf (1.1MB, pdf)

Acknowledgments

We thank A. E. Böhmer, A.V. Chubukov, I. Eremin, P. J. Hirschfeld, H. Kontani, S. S. Lee, C. Meingast, A. Nevidomskyy, L. Radzihovsky, M. Randeria, I. Vekhter, and Y. Yanase for valuable discussion. This work has been supported by Japan–Germany Research Cooperative Program, Grant-in-Aid for Scientific Research (KAKENHI) from Japan Society for the Promotion of Science and Project 56393598 from German Academic Exchange Service, and the “Topological Quantum Phenomena” (25103713) KAKENHI on Innovative Areas from Ministry of Education, Culture, Sports, Science and Technology of Japan.

Footnotes

The authors declare no conflict of interest.

This article is a PNAS Direct Submission.

This article contains supporting information online at www.pnas.org/lookup/suppl/doi:10.1073/pnas.1413477111/-/DCSupplemental.

References

  • 1.Noziéres P, Schmitt-Rink S. Bose condensation in an attractive fermion gas: From weak to strong coupling superconductivity. J Low Temp Phys. 1985;59(3-4):195–211. [Google Scholar]
  • 2.Leggett AJ. In: Modern Trends in the Theory of Condensed Matter. Pekalski A, Przystawa R, editors. Springer; Berlin: 1980. [Google Scholar]
  • 3.Chen Q, Stajic J, Tan S, Levin K. BCS-BEC crossover: From high temperature superconductors to ultracold superfluids. Phys Rep. 2005;412(1):1–88. [Google Scholar]
  • 4.Giorgini S, Pitaevskii LP, Stringari S. Theory of ultracold atomic Fermi gases. Rev Mod Phys. 2008;80(4):1215. [Google Scholar]
  • 5.Randeria M, Taylor E. Crossover from Bardeen-Cooper-Schrieffer to Bose-Einstein condensation and the unitary Fermi gas. Annu Rev Condens Matter Phys. 2014;5:209–232. [Google Scholar]
  • 6.Chevy F, Mora C. Ultra-cold polarized Fermi gases. Rep Prog Phys. 2010;73(11):112401. [Google Scholar]
  • 7.Shin YI, Schunck CH, Schirotzek A, Ketterle W. Phase diagram of a two-component Fermi gas with resonant interactions. Nature. 2008;451(7179):689–693. doi: 10.1038/nature06473. [DOI] [PubMed] [Google Scholar]
  • 8.Gubbels KB, Stoof HTC. Imbalanced Fermi gases at unitarity. Phys Rep. 2012;525(4):255–313. [Google Scholar]
  • 9.Gubankova E, Liu WV, Wilczek F. Breached pairing superfluidity: Possible realization in QCD. Phys Rev Lett. 2003;91(3):032001. doi: 10.1103/PhysRevLett.91.032001. [DOI] [PubMed] [Google Scholar]
  • 10.Yoshida N, Yip S-K. Larkin-Ovchinnikov state in resonant Fermi gas. Phys Rev A. 2007;75(6):063601. [Google Scholar]
  • 11.Hsu F-C, et al. Superconductivity in the PbO-type structure α-FeSe. Proc Natl Acad Sci USA. 2008;105(38):14262–14264. doi: 10.1073/pnas.0807325105. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 12.Lubashevsky Y, Lahoud E, Chashka K, Podolsky D, Kanigel A. Shallow pockets and very strong coupling superconductivity in FeSexTe1−x. Nat Phys. 2012;8(4):309–312. [Google Scholar]
  • 13.Okazaki K, et al. Superconductivity in an electron band just above the Fermi level: Possible route to BCS-BEC superconductivity. Sci Rep. 2014;4:4109. doi: 10.1038/srep04109. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.Braithwaite D, et al. Superconducting and normal phases of FeSe single crystals at high pressure. J Phys Condens Matter. 2009;21(23):232202. doi: 10.1088/0953-8984/21/23/232202. [DOI] [PubMed] [Google Scholar]
  • 15.Song C-L, et al. Direct observation of nodes and twofold symmetry in FeSe superconductor. Science. 2011;332(6036):1410–1413. doi: 10.1126/science.1202226. [DOI] [PubMed] [Google Scholar]
  • 16.Dong JK, et al. Multigap nodeless superconductivity in FeSex: Evidence from quasiparticle heat transport. Phys Rev B. 2009;80(2):024518. [Google Scholar]
  • 17.Hashimoto K, et al. A sharp peak of the zero-temperature penetration depth at optimal composition in BaFe2(As1-xPx)2. Science. 2012;336(6088):1554–1557. doi: 10.1126/science.1219821. [DOI] [PubMed] [Google Scholar]
  • 18.Terashima T, et al. 2014. Anomalous Fermi surface in FeSe seen by Shubnikov-de Haas oscillation measurements. Phys Rev B 90(14):144517.
  • 19.LeBoeuf D, et al. Lifshitz critical point in the cuprate superconductor YBa2Cu3Oy from high-field Hall effect measurements. Phys Rev B. 2011;83(5):054506. [Google Scholar]
  • 20.Hasegawa Y, Avouris P. Direct observation of standing wave formation at surface steps using scanning tunneling spectroscopy. Phys Rev Lett. 1993;71(7):1071–1074. doi: 10.1103/PhysRevLett.71.1071. [DOI] [PubMed] [Google Scholar]
  • 21.Crommie MF, Lutz CP, Eigler DM. Imaging standing waves in a two-dimensional electron gas. Nature. 1993;363(6429):524–527. [Google Scholar]
  • 22.Maletz J, et al. Unusual band renormalization in the simplest iron-based superconductor. Phys Rev B. 2014;89(22):220506(R). [Google Scholar]
  • 23.Hawthorn DG, et al. Doping dependence of the superconducting gap in Tl2Ba2CuO6+δ from heat transport. Phys Rev B. 2007;75(10):104518. [Google Scholar]
  • 24.Krishana K, Ong NP, Li Q, Gu GD, Koshizuka N. Plateaus observed in the field profile of thermal conductivity in the superconductor Bi2Sr2CaCu2O8. Science. 1997;277(5322):83–85. [Google Scholar]
  • 25.Kasahara Y, et al. Anomalous quasiparticle transport in the superconducting state of CeCoIn5. Phys Rev B. 2005;72(21):214515. [Google Scholar]
  • 26.Franz M. Mixed-state quasiparticle transport in high-Tc cuprates. Phys Rev Lett. 1999;82:1760. doi: 10.1103/PhysRevLett.86.316. [DOI] [PubMed] [Google Scholar]
  • 27.Vishveshwara S, Fisher MPA. Critical dynamics of thermal conductivity at the normal-superconducting transition. Phys Rev B. 2001;64(13):134507. [Google Scholar]
  • 28.Tayama T, et al. Unconventional heavy-fermion superconductor CeCoIn5: dc magnetization study at temperatures down to 50 mK. Phys Rev B. 2002;65(18):180504(R). [Google Scholar]
  • 29.Matsuda Y, Shimahara H. Fulde-Ferrell-Larkin-Ovchinnikov state in heavy fermion superconductors. J Phys Soc Jpn. 2007;76(5):051005. [Google Scholar]
  • 30.Larkin AI, Varlamov AA. 2013. Fluctuation phenomena in superconductors. arXiv:cond-mat/0109177.
  • 31.Yanase Y. FFLO superconductivity near the antiferromagnetic quantum critical point. J Phys Soc Jpn. 2008;77(6):063705. [Google Scholar]
  • 32.Radzihovsky L, Sheehy DE. Imbalanced Feshbach-resonant Fermi gases. Rep Prog Phys. 2010;73(7):076501. [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

Supplementary File
pnas.201413477SI.pdf (1.1MB, pdf)

Articles from Proceedings of the National Academy of Sciences of the United States of America are provided here courtesy of National Academy of Sciences

RESOURCES