Next Article in Journal
Secure the IoT Networks as Epidemic Containment Game
Next Article in Special Issue
Electronic Origin of Tc in Bulk and Monolayer FeSe
Previous Article in Journal
The Symmetric Nature of the Position Distribution of the Human Body Center of Gravity during Propelling Manual Wheelchairs with Innovative Propulsion Systems
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Unusual Temperature Evolution of Quasiparticle Band Dispersion in Electron-Doped FeSe Films

by
Kosuke Nakayama
1,2,*,
Koshin Shigekawa
1,
Katsuaki Sugawara
1,2,3,4,
Takashi Takahashi
1,3,4 and
Takafumi Sato
1,3,4
1
Department of Physics, Tohoku University, Sendai 980-8578, Japan
2
Precursory Research for Embryonic Science and Technology (PRESTO), Japan Science and Technology Agency (JST), Tokyo 102-0076, Japan
3
WPI Research Center, Advanced Institute for Materials Research, Tohoku University, Sendai 980-8577, Japan
4
Center for Spintronics Research Network, Tohoku University, Sendai 980-8577, Japan
*
Author to whom correspondence should be addressed.
Symmetry 2021, 13(2), 155; https://doi.org/10.3390/sym13020155
Submission received: 28 December 2020 / Revised: 15 January 2021 / Accepted: 18 January 2021 / Published: 20 January 2021
(This article belongs to the Special Issue Understanding Iron Superconductors and Isostructural Materials)

Abstract

:
The discovery of high-temperature (high-Tc) superconductivity in one-monolayer FeSe on SrTiO3 has attracted tremendous attention. Subsequent studies suggested the importance of cooperation between intra-FeSe-layer and interfacial interactions to enhance Tc. However, the nature of intra-FeSe-layer interactions, which would play a primary role in determining the pairing symmetry, remains unclear. Here we have performed high-resolution angle-resolved photoemission spectroscopy of one-monolayer and alkaline-metal-deposited multilayer FeSe films on SrTiO3, and determined the evolution of quasiparticle band dispersion across Tc. We found that the band dispersion in the superconducting state deviates from the Bogoliubov-quasiparticle dispersion expected from the normal-state band dispersion with a constant gap size. This suggests highly anisotropic pairing originating from small momentum transfer and/or mass renormalization due to electron–boson coupling. This band anomaly is interpreted in terms of the electronic interactions within the FeSe layers that may be related to the high-Tc superconductivity in electron-doped FeSe.

1. Introduction

Iron selenide (FeSe) is structurally the simplest iron-based superconductor with the superconducting-transition temperature (Tc) of ~9 K [1]. Intriguingly, one-monolayer (1 ML) film of FeSe grown on SrTiO3 substrate exhibits exceptionally high Tc [2]. The Tc value reported by transport measurements reaches 40 K [2,3], which is about five times higher than the bulk counterpart. In addition, Cooper pairing at a higher temperature of 65 K, which exceeds the highest Tc (56 K) ever achieved in iron-based superconductors, has been suggested from a gap-closing temperature by angle-resolved photoemission spectroscopy (ARPES) [4,5,6,7] and Meissner effect by mutual conductance measurements [8]. These observations triggered fierce debates on the origin of the Tc enhancement in 1 ML-FeSe film. One key ingredient is a novel cross-interface electron–phonon coupling. The strong coupling between electrons in the FeSe layer and optical phonons of SrTiO3 has been verified via the observation of replica bands by ARPES and theoretically proposed to enhance Tc in most of possible pairing symmetries [9,10]. Later, the close link between strong electron–phonon coupling and Tc enhancement has been supported experimentally, e.g., by isotope effects [11,12]. With these findings as a guiding principle, the search for high Tc in atomically thin films of other iron-based superconductors interfaced with SrTiO3 [13,14,15] has been accelerated. Another key ingredient for Tc enhancement is a charge transfer from SrTiO3. Heavy electron doping to the FeSe layer leads to unique electronic structure consisting only of electron-like Fermi surfaces [4,5,16], in contrast to the semimetallic nature of bulk FeSe [17,18]. The electron doping is essential for the high-Tc superconductivity, as established by the observation of high Tc above 40 K even in multilayer and bulk FeSe by doping electron carriers [19,20,21]. Therefore, there is accumulated evidence that the FeSe layer has the capability of inducing 40-K superconductivity through electron doping and the interfacial electron–phonon coupling will assist further Tc or pairing enhancement. However, little is known about why electron doping leads to the high Tc’s above 40 K. In particular, interactions within the FeSe layer, which would primarily determine the pairing symmetry, remain unclear.
In this study, we performed a comparative ARPES experiment on the surface of 1 ML- and Cs-deposited 20 ML-FeSe films on SrTiO3, where the interfacial effects were present and absent, respectively. We demonstrated anomaly in the quasiparticle-band dispersions in the superconducting state, which is not expected from the Bogoliubov-quasiparticle (BQP) dispersion induced by a simple s-wave-gap opening. We discuss implications of our observation in relation to intra-FeSe-layer interactions.

2. Materials and Methods

The molecular beam epitaxy method was used to obtain 1 ML- and 20 ML-FeSe films; the films were grown on a TiO2-terminated Nb(0.05 wt%)-SrTiO3 substrate (SHINKOSHA) by simultaneously evaporating Fe and Se atoms while keeping a substrate temperature at 430 °C with a deposition rate of 0.01 ML/s [19]. Electron doping to 20 ML-FeSe was realized by evaporating Cs atoms onto the film surface at room temperature using a Cs dispenser (SAES Getters) [22]. After the growth, the film was transferred to the ARPES-measurement chamber without exposure to air. ARPES measurements were performed with a SES2002 spectrometer (Scienta Omicron) with the He-Iα resonance line ( = 21.218 eV) at Tohoku University. The film was kept under an ultrahigh vacuum of 5 × 10−11 Torr during the ARPES measurement, and no remarkable surface degradation was observed for a typical measurement time of 1 day. The energy and angular resolutions were set to be 7–12 meV and 0.2°, respectively. A gold film which made electrical contact with the film was referenced to calibrate the Fermi level (EF).

3. Results

First, we present the electronic structure of Cs-deposited 20 ML-FeSe film measured at T = 50 K. As shown in Figure 1a,b, there was a circular-shaped large Fermi surface at the Brillouin-zone corner (M point) which originated from EF crossing of an electron band with the bottom of the dispersion around 50 meV below EF. The top of a hole-like band around the zone center (Γ point) was about 50 meV below EF, resulting in the absence of a hole-like Fermi surface in contrast to the as-grown multilayer FeSe film [4,5] or bulk FeSe [17,18]. These observations confirmed a successful electron doping by Cs deposition onto the FeSe surface. The electron carrier concentration (ne) calculated from the Fermi-surface volume was ~0.11 electrons/Fe, which corresponds to the optimal doping level with Tc value of ~40 K [22]. To investigate how the band structure changes by the superconducting transition, we performed high-resolution measurements across Tc (50 and 13 K) along a momentum (k) cut A indicated by a blue line in Figure 1a. The results displayed in Figure 1c,d show that while the electron band above Tc crossed EF at the Fermi wave vector (kF) of ~0.17 π/a, the band dispersion below Tc had a local maximum below EF so as not to cross EF due to a superconducting-gap opening. It is noted that the k location of the electron-band top below Tc coincided with the kF point above Tc, consistent with the Cooper-pairing origin of the observed gap. The superconducting-gap opening is also clearly seen in energy distribution curves (EDCs) in Figure 1e, in which the peak position at kF (k3 defined in Figure 1c,d) was shifted from EF to a high binding energy by ~10 meV with decreasing the temperature to form a superconducting gap. Since the nodeless s-wave superconductivity is realized in electron-doped multilayer FeSe [19,20], one can see the superconducting-gap opening below Tc irrespective of the k cut, e.g., along the k cut crossing the M point (cut B), as shown in Figure 1f–h, where essentially the same behavior with cut A was recognized.
An important finding manifests itself when we compare the band dispersions of the normal and superconducting states. Figure 2a displays a direct comparison of the experimental band dispersions extracted from the peak position of EDCs in cut A. As mentioned above, the band dispersion below Tc exhibited an opening of the superconducting gap and resultant bending-back behavior with the top of the dispersion at kF. Such a characteristic band dispersion below Tc was qualitatively consistent with the dispersion relation of BQPs in the Bardeen–Cooper–Schrieffer (BCS) theory, where BQP dispersion (Ek) is expressed as E k = ε k 2 + | Δ | 2 k and ∆ are the normal-state band dispersion and the superconducting-gap size, respectively) [24]. For a quantitative comparison, we determined εk by performing a polynomial fitting to the ARPES data above Tc (magenta curve) and simulated Ek by assuming a k-independent superconducting-gap size of 10 meV (light blue curve). Intriguingly, the band dispersion below Tc shows a clear deviation from the simulated BQP dispersion Ek; specifically, although the simulation predicted a finite downward energy shift of BQP dispersion compared with εk even in the k region far away from kF (at least down to ky = 0.05 π/a) because of a large ∆ value with respect to the shallow electron-band bottom, the experimental dispersion below Tc became nearly identical to εk as soon as it moved away from kF. Almost temperature-insensitive band position away from kF was also clearly visible in the comparison of raw EDCs in Figure 1e (see EDCs at k1 and k2). The same behavior was observed at different momentum, e.g., we found that the band dispersion measured below Tc along cut B deviated from the simulated BQP dispersion over a wide k region (see Figure 2b; also see a comparison of EDCs in Figure 1h).
To clarify whether the energy difference between the experimental and simulated BQP dispersions below Tc was an essential ingredient of electron-doped high-Tc FeSe films, we investigated the band-structure evolution in 1 ML-FeSe (Figure 3). For this purpose, we performed high-resolution measurements on slightly underdoped 1 ML-FeSe (ne = 0.09) with Tc~40 K because a sharp spectral line shape compared with the heavily doped sample (Tc~65 K; ne~0.12) [4] is suited for accurately determining the quasiparticle band dispersion. As is well known, 1 ML-FeSe has a large electron-like Fermi surface centered at the M point. The electron band which forms the Fermi surface showed an opening of the superconducting gap (∆~10 meV) below Tc, as highlighted by the characteristic bending-back behavior with the minimum-gap locus at kF (see Figure 3b). A direct comparison of the band dispersions above and below Tc in Figure 3c demonstrates that an energy shift due to the superconducting-gap opening was limited to the k region around kF~0.16 π/a (compare red and blue circles), in sharp contrast to a clear downward shift of the simulated BQP dispersion for ky ≤ 0.1 π/a (light blue curve). Similarly, the case of Cs-deposited 20 ML-FeSe suggested that the deviation of the experimental band dispersion from the simulated BQP dispersion below Tc is a common feature of electron-doped FeSe films irrespective of film thickness.

4. Discussion and Conclusions

Now we are going to discuss the origin of the observed anomaly in quasiparticle dispersion. To simulate BQP dispersion E k = ε k 2 + | Δ | 2 , we assumed that εk is the same as the dispersion above Tc and ∆ is k-independent. It would be natural to consider that one or both of these assumptions are incorrect, rather than thinking that the BQP picture was broken in the electron-doped FeSe. For simplicity, we consider in the following the two extreme cases that the deviation was induced by a change in either εk or Δ. First, to examine the k dependence of Δ as the origin, we put the experimental band dispersions below and above Tc into Ek and εk, respectively, and estimated ∆(k) which reproduces the experimental band dispersion below Tc. The obtained ∆(k) was strongly k-dependent as seen from Figure 4a,b for 1 ML- and Cs-deposited 20 ML-FeSe, respectively; namely, ∆(k) was finite only in the narrow k region centered at kF (within ±0.02 π/a of kF), so that band dispersion only around EF was shifted toward high binding energies by the superconducting transition, consistent with our observations. An unusual Cooper pairing in the limited k space near kF may be caused by pairing interactions which have small momentum transfer q [25,26,27]. For instance, it has been proposed by Migdal–Eliashberg theory for 1 ML-FeSe that forward scattering with small q phonons produces highly anisotropic superconducting gap peaked at kF and also leads to temperature-independent band structure away from kF [25], in qualitative agreement with ∆(k) in Figure 4a as well as band dispersion in Figure 3. Although this theory assumes a cross-interface coupling between small q phonons of SrTiO3 and electrons in 1 ML-FeSe as the key pairing interactions, our observation of anisotropic ∆(k) in 20 ML-FeSe (Figure 4b) where interfacial effects are negligible suggests that small-q interactions within the FeSe layers may be also responsible for superconductivity if the k-dependent pairing was indeed a source of the deviation from the simulated BQP dispersion.
Next, we consider another possibility that εk is temperature-dependent whereas k dependence of ∆(k) is small. To explain the observed deviation between the experimental and simulated BQP dispersions, εk below Tc must be shifted toward EF compared with the normal-state band dispersion above Tc while keeping the same kF position. Such an energy shift would be a consequence of mass renormalization linked to the superconducting transition, likely due to coupling with bosonic modes as reported for bulk crystals of high-Tc superconductors [28,29,30,31,32,33]. Here we defined the energy difference between the experimental and simulated BQP dispersions as ∆E (see black arrow in Figure 3c) and plotted it in Figure 4c,d for 1 ML- and 20 ML-FeSe films, respectively. As seen from Figure 4c,d, ∆E showed a broad peak around 20 meV in 1 ML- and 20 ML-FeSe. The obtained ∆E value can be used as a measure of the mass enhancement similarly to the real part of self-energies. By analogy with the fact that the peak position in the real part of self-energies below Tc corresponded to ∆ + Ω, where Ω is the energy of bosonic modes coupled to electrons, the observed peak structure at ~20 meV suggests a coupling to low-energy modes with Ω~10 meV (here, ∆~10 meV). The origin of the corresponding modes is an important open question; candidates include phonons [34] and magnetic resonance [35]. Nevertheless, one important outcome from our observation is that low-energy modes intrinsic to the FeSe layer must be involved because the mass renormalization was found not only in 1 ML-FeSe, but also in 20 ML-FeSe.
In summary, we reported the evolution of low-energy band dispersion across Tc in 1 ML- and Cs-deposited 20 ML-FeSe films on SrTiO3. We found deviation of the band dispersion below Tc from the simple BQP dispersion simulated with the temperature-independent εk and k-independent ∆. We proposed two possible scenarios as the origin of this observation; (i) anisotropic ∆(k) peaked around kF due to the superconducting pairing by small q transfer and (ii) enhancement in the effective mass in the superconducting state due to the coupling to low-energy bosonic modes. In either scenario, the observed similarity between 1 ML- and 20 ML-FeSe suggested intra-FeSe-layer nature of the interactions. Our result lays a foundation for understanding the mechanism of high-Tc superconductivity in electron-doped FeSe.

Author Contributions

Conceptualization, K.N.; formal analysis, K.S. (Koshin Shigekawa); investigation, K.N., K.S. (Koshin Shigekawa) and K.S. (Katsuaki Sugawara); data curation, K.N. and K.S. (Koshin Shigekawa); writing—original draft preparation, K.N.; writing—review and editing, T.T. and T.S.; visualization, K.N. and K.S. (Koshin Shigekawa); project administration, K.N. and T.S.; funding acquisition, K.N., K.S. (Katsuaki Sugawara), T.T. and T.S. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by JST-CREST (No. JPMJCR18T1) and Grant-in-Aid for Scientific Research (JSPS KAKENHI Grant Numbers JP17H04847, JP17H01139, JP18H01160, and JP20H01847).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

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

Acknowledgments

We thank G. Phan and M. Kuno for their assistance in the ARPES measurements.

Conflicts of Interest

The authors declare no conflict of interests.

References

  1. Hsu, F.C.; Luo, J.-Y.; Yeh, K.-W.; Chen, T.-K.; Huang, T.-W.; Wu, P.M.; Lee, Y.-C.; Huang, Y.-L.; Chu, Y.-Y.; Yan, D.-C.; et al. Superconductivity in the PbO-type structure α-FeSe. Proc. Natl. Acad. Sci. USA 2009, 105, 14262–14264. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Wang, Q.Y.; Wang, Q.-Y.; Li, Z.; Zhang, W.-H.; Zhang, Z.-C.; Zhang, J.-S.; Li, W.; Ding, H.; Ou, Y.-B.; Deng, P.; et al. Interface-induced high-temperature superconductivity in single unit-cell FeSe films on SrTiO3. Chin. Phys. Lett. 2012, 29, 037402. [Google Scholar] [CrossRef] [Green Version]
  3. Sun, Y.; Zhang, W.; Xing, Y.; Li, F.; Zhao, Y.; Xia, Z.; Wang, L.; Ma, X.; Xue, Q.-K.; Wang, J. High temperature superconducting FeSe films on SrTiO3 substrates. Sci. Rep. 2014, 4, 6040. [Google Scholar] [CrossRef] [PubMed]
  4. He, S.L.; He, S.; Zhang, W.; Zhao, L.; Liu, D.; Liu, X.; Mou, D.; Ou, Y.-B.; Wang, Q.-Y.; Li, Z.; et al. Phase diagram and electronic indication of high-temperature superconductivity at 65 K in single-layer FeSe films. Nat. Mater. 2013, 12, 605–610. [Google Scholar] [CrossRef]
  5. Tan, S.Y.; Zhang, Y.; Xia, M.; Ye, Z.; Chen, F.; Xie, X.; Peng, R.; Xu, D.; Fan, Q.; Xu, H.; et al. Interface-induced superconductivity and strain-dependent spin density wave in FeSe/SrTiO3 thin films. Nat. Mater. 2013, 12, 634–640. [Google Scholar] [CrossRef] [Green Version]
  6. Peng, R.; Xu, H.C.; Tan, S.Y.; Cao, H.Y.; Xia, M.; Shen, X.P.; Huang, Z.C.; Wen, C.H.P.; Song, Q.; Zhang, T.; et al. Tuning the band structure and superconductivity in single-layer FeSe by interface engineering. Nat. Commun. 2014, 5, 5044. [Google Scholar] [CrossRef] [Green Version]
  7. Peng, R.; Shen, X.P.; Xie, X.; Xu, H.C.; Tan, S.Y.; Xia, M.; Zhang, T.; Cao, H.Y.; Gong, X.G.; Hu, J.P.; et al. Measurement of an Enhanced Superconducting Phase and a Pronounced Anisotropy of the Energy Gap of a Strained FeSe Single Layer in FeSe/Nb:SrTiO3/KTaO3 Heterostructures Using Photoemission Spectroscopy. Phys. Rev. Lett. 2014, 112, 107001. [Google Scholar] [CrossRef] [Green Version]
  8. Zhang, Z.; Wang, Y.-H.; Song, Q.; Liu, C.; Peng, R.; Moler, K.A.; Feng, D.; Wang, Y. Onset of the Meissner effect at 65 K in FeSe thin film grown on Nb-doped SrTiO3 substrate. Sci. Bull. 2015, 60, 1301–1304. [Google Scholar] [CrossRef] [Green Version]
  9. Lee, J.J.; Schmitt, F.T.; Moore, R.G.; Johnston, S.; Cui, Y.-T.; Li, W.; Yi, M.; Liu, Z.K.; Hashimoto, M.; Zhang, Y.; et al. Interfacial mode coupling as the origin of the enhancement of Tc in FeSe films on SrTiO3. Nature 2014, 515, 245. [Google Scholar] [CrossRef]
  10. Xiang, Y.Y.; Wang, F.; Wang, D.; Wang, Q.H.; Lee, D.H. High-temperature superconductivity at the FeSe/SrTiO3 interface. Phys. Rev. B 2012, 86, 134508. [Google Scholar] [CrossRef]
  11. Song, Q.; Yu, T.L.; Lou, X.; Xie, B.P.; Xu, H.C.; Wen, C.H.P.; Yao, Q.; Zhang, S.Y.; Zhu, X.T.; Guo, J.D.; et al. Evidence of cooperative effect on the enhanced superconducting transition temperature at the FeSe/SrTiO3 interface. Nat. Commun. 2019, 10, 758. [Google Scholar] [CrossRef] [PubMed]
  12. Rebec, S.N.; Jia, T.; Zhang, C.; Hashimoto, M.; Lu, D.-H.; Moore, R.G.; Shen, Z.-X. Coexistence of Replica Bands and Superconductivity in FeSe Monolayer Films. Phys. Rev. Lett. 2017, 118, 067002. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Li, F.; Ding, H.; Tang, C.; Peng, J.; Zhang, Q.; Zhang, W.; Zhou, G.; Zhang, D.; Song, C.-L.; He, K.; et al. Interface-enhanced high-temperature superconductivity in single-unit-cell FeTe1xSex films on SrTiO3. Phys. Rev. B 2015, 91, 220503. [Google Scholar] [CrossRef]
  14. Shi, X.; Han, Z.-Q.; Richard, P.; Wu, X.-X.; Peng, X.-L.; Qian, T.; Wang, S.-C.; Hu, J.-P.; Sun, Y.-J.; Ding, H. FeTe1−xSex monolayer films: Towards the realization of high-temperature connate topological superconductivity. Sci. Bull. 2017, 62, 503–507. [Google Scholar] [CrossRef] [Green Version]
  15. Shigekawa, K.; Nakayama, K.; Kuno, M.; Phan, G.N.; Owada, K.; Sugawara, K.; Takahashi, T.; Sato, T. Dichotomy of superconductivity between monolayer FeS and FeSe. Proc. Natl. Acad. Sci. USA 2019, 116, 24470–24474. [Google Scholar] [CrossRef] [Green Version]
  16. Liu, D.F.; Zhang, W.; Mou, D.; He, J.; Ou, Y.-B.; Wang, Q.-Y.; Li, Z.; Wang, L.; Zhao, L.; He, S.; et al. Electronic origin of high-temperature superconductivity in single-layer FeSe superconductor. Nat. Commun. 2012, 3, 931. [Google Scholar] [CrossRef] [Green Version]
  17. Maletz, J.; Zabolotnyy, V.B.; Evtushinsky, D.V.; Thirupathaiah, S.; Wolter, A.U.B.; Harnagea, L.; Yaresko, A.N.; Vasiliev, A.N.; Chareev, D.A.; Böhmer, A.E.; et al. Unusual band renormalization in the simplest iron-based superconductor FeSe1−x. Phys. Rev. B 2014, 89, 220506. [Google Scholar] [CrossRef] [Green Version]
  18. Nakayama, K.; Miyata, Y.; Phan, G.N.; Sato, T.; Tanabe, Y.; Urata, T.; Tanigaki, K.; Takahashi, T. Reconstruction of Band Structure Induced by Electronic Nematicity in an FeSe Superconductor. Phys. Rev. Lett. 2014, 113, 237001. [Google Scholar] [CrossRef] [Green Version]
  19. Miyata, Y.; Nakayama, K.; Sugawara, K.; Sato, T.; Takahashi, T. High-temperature superconductivity in potassium-coated multilayer FeSe thin films. Nat. Mater. 2015, 14, 775–779. [Google Scholar] [CrossRef]
  20. Wen, C.H.P.; Xu, H.C.; Chen, C.; Huang, Z.C.; Lou, X.; Pu, Y.J.; Song, Q.; Xie, B.P.; Abdel-Hafiez, M.; Chareev, D.A.; et al. Anomalous correlation effects and unique phase diagram of electron-doped FeSe revealed by photoemission spectroscopy. Nat. Commun. 2016, 7, 10840. [Google Scholar] [CrossRef]
  21. Shiogai, J.; Ito, Y.; Mitsuhashi, T.; Nojima, T.; Tsukazaki, A. Electric-field-induced superconductivity in electrochemically etched ultrathin FeSe films on SrTiO3 and MgO. Nat. Phys. 2016, 12, 42–46. [Google Scholar] [CrossRef]
  22. Phan, G.N.; Nakayama, K.; Kanayama, S.; Kuno, M.; Sugawara, K.; Sato, T.; Takahashi, T. ARPES study of cesium-coated FeSe thin films on SrTiO3. J. Phys. Conf. Ser. 2017, 871, 012017. [Google Scholar] [CrossRef] [Green Version]
  23. Norman, M.R.; Randeria, M.; Ding, H.; Campuzano, J.C. Phenomenology of the low-energy spectral function in high-Tc superconductors. Phys. Rev. B 1998, 57, R11093–R11096. [Google Scholar] [CrossRef] [Green Version]
  24. Matsui, H.; Sato, T.; Takahashi, T.; Wang, S.-C.; Yang, H.-B.; Ding, H.; Fujii, T.; Watanabe, T.; Matsuda, A. BCS-Like Bogoliubov Quasiparticles in High-Tc Superconductors Observed by Angle-Resolved Photoemission Spectroscopy. Phys. Rev. Lett. 2003, 90, 217002. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Wang, Y.; Nakatsukasa, K.; Rademaker, L.; Berlijn, T.; Johnston, S. Aspects of electron–phonon interactions with strong forward scattering in FeSe Thin Films on SrTiO3 substrates. Supercond. Sci. Technol. 2016, 29, 054009. [Google Scholar] [CrossRef] [Green Version]
  26. Yamase, H.; Zeyher, R. Superconductivity from orbital nematic fluctuations. Phys. Rev. B 2013, 88, 180502. [Google Scholar] [CrossRef] [Green Version]
  27. Agatsuma, T.; Yamase, H. Structure of the pairing gap from orbital nematic fluctuations. Phys. Rev. B 2016, 94, 214505. [Google Scholar] [CrossRef] [Green Version]
  28. Norman, M.R.; Ding, H.; Campuzano, J.C.; Takeuchi, T.; Randeria, M.; Yokoya, T.; Takahashi, T.; Mochiku, T.; Kadowaki, K. Unusual Dispersion and Line Shape of the Superconducting State Spectra of Bi2Sr2CaCu2O8+δ. Phys. Rev. Lett. 1997, 79, 3506–3509. [Google Scholar] [CrossRef] [Green Version]
  29. Valla, T.; Fedorov, A.V.; Johnson, P.D.; Wells, B.O.; Hulbert, S.L.; Li, Q.; Gu, G.D.; Koshizuka, N. Evidence for Quantum Critical Behavior in the Optimally Doped Cuprate Bi2Sr2CaCu2O8+δ. Science 1999, 285, 2110. [Google Scholar] [CrossRef]
  30. Kaminski, A.; Randeria, M.; Campuzano, J.C.; Norman, M.R.; Fretwell, H.; Mesot, J.; Sato, T.; Takahashi, T.; Kadowaki, K. Renormalization of Spectral Line Shape and Dispersion below Tc in Bi2Sr2CaCu2O8+δ. Phys. Rev. Lett. 2001, 86, 1070. [Google Scholar] [CrossRef] [Green Version]
  31. Lanzara, A.; Bogdanov, P.V.; Zhou, X.J.; Kellar, S.A.; Feng, D.L.; Lu, E.D.; Yoshida, T.; Eisaki, H.; Fujimori, A.; Kishio, K.; et al. Evidence for ubiquitous strong electron–phonon coupling in high-temperature superconductors. Nature 2001, 412, 510. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Richard, P.; Sato, T.; Nakayama, K.; Souma, S.; Takahashi, T.; Xu, Y.-M.; Chen, G.F.; Luo, J.L.; Wang, N.L.; Ding, H. Angle-Resolved Photoemission Spectroscopy of the Fe-Based Ba0.6K0.4Fe2As2 High Temperature Superconductor: Evidence for an Orbital Selective Electron-Mode Coupling. Phys. Rev. Lett. 2009, 102, 047003. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Koitzsch, A.; Inosov, D.S.; Evtushinsky, D.V.; Zabolotnyy, V.B.; Kordyuk, A.A.; Kondrat, A.; Hess, C.; Knupfer, M.; Büchner, B.; Sun, G.L.; et al. Temperature and Doping-Dependent Renormalization Effects of the Low Energy Electronic Structure of Ba1−xKxFe2As2 Single Crystals. Phys. Rev. Lett. 2009, 102, 167001. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Tang, C.; Liu, C.; Zhou, G.; Li, F.; Ding, H.; Li, Z.; Zhang, D.; Li, Z.; Song, C.; Ji, S.; et al. Interface-enhanced electron-phonon coupling and high-temperature superconductivity in potassium-coated ultrathin FeSe films on SrTiO3. Phys. Rev. B 2016, 93, 020507. [Google Scholar] [CrossRef] [Green Version]
  35. Ma, M.; Bourges, P.; Sidis, Y.; Xu, Y.; Li, S.; Hu, B.; Li, J.; Wang, F.; Li, Y. Prominent Role of Spin-Orbit Coupling in FeSe Revealed by Inelastic Neutron Scattering. Phys. Rev. X 2017, 7, 021025. [Google Scholar] [CrossRef] [Green Version]
Figure 1. (a) ARPES intensity map at EF as a function of two-dimensional wave vector for Cs-deposited 20 monolayer (ML)-FeSe film obtained at T = 50 K with = 21.218 eV. Intensity at EF was obtained by integrating the spectral intensity within ±10 meV of EF. Green circle is a guide for the eyes to trace the Fermi surface. (b) Plot of ARPES intensity along the ΓM cut at 50 K as a function of binding energy and wave vector. (c,d) Near-EF ARPES intensity along cut A in (a) at T = 50 and 13 K, respectively, divided by the Fermi–Dirac distribution (FD) function at each temperature convoluted with the resolution function. Intensity above EF is displayed up to 3kBT. Red and blue circles in (c,d), respectively, are the band dispersion determined by fitting the energy distribution curves (EDCs) with Bardeen–Cooper–Schrieffer (BCS) spectral function [23]. (e) Comparison of EDCs between T = 50 K (red) and 13 K (blue) taken at representative ky points [k1, k2, and k3 indicated by magenta lines in (c,d)]. Red and blue dots indicate the local maxima corresponding to the peak position. (f–h) Same as (cf) but obtained along cut B in (a).
Figure 1. (a) ARPES intensity map at EF as a function of two-dimensional wave vector for Cs-deposited 20 monolayer (ML)-FeSe film obtained at T = 50 K with = 21.218 eV. Intensity at EF was obtained by integrating the spectral intensity within ±10 meV of EF. Green circle is a guide for the eyes to trace the Fermi surface. (b) Plot of ARPES intensity along the ΓM cut at 50 K as a function of binding energy and wave vector. (c,d) Near-EF ARPES intensity along cut A in (a) at T = 50 and 13 K, respectively, divided by the Fermi–Dirac distribution (FD) function at each temperature convoluted with the resolution function. Intensity above EF is displayed up to 3kBT. Red and blue circles in (c,d), respectively, are the band dispersion determined by fitting the energy distribution curves (EDCs) with Bardeen–Cooper–Schrieffer (BCS) spectral function [23]. (e) Comparison of EDCs between T = 50 K (red) and 13 K (blue) taken at representative ky points [k1, k2, and k3 indicated by magenta lines in (c,d)]. Red and blue dots indicate the local maxima corresponding to the peak position. (f–h) Same as (cf) but obtained along cut B in (a).
Symmetry 13 00155 g001
Figure 2. (a) Comparison of the near-EF band dispersions in Cs-deposited 20 ML-FeSe at T = 50 K (red circles) and 13 K (blue circles) along cut A in Figure 1a. Magenta curve is the normal-state band dispersion εk extracted from a polynomial fitting to the red circles. Light blue curve is the calculated Bogoliubov-quasiparticle (BQP) dispersion based on the BCS formula E k = ε k 2 + | Δ | 2 with a constant ∆ of 10 meV. (b) Same as (a) but for cut B in Figure 1a.
Figure 2. (a) Comparison of the near-EF band dispersions in Cs-deposited 20 ML-FeSe at T = 50 K (red circles) and 13 K (blue circles) along cut A in Figure 1a. Magenta curve is the normal-state band dispersion εk extracted from a polynomial fitting to the red circles. Light blue curve is the calculated Bogoliubov-quasiparticle (BQP) dispersion based on the BCS formula E k = ε k 2 + | Δ | 2 with a constant ∆ of 10 meV. (b) Same as (a) but for cut B in Figure 1a.
Symmetry 13 00155 g002
Figure 3. (a,b) ARPES intensity divided by the FD function measured along the k cut crossing the M point in 1 ML-FeSe at T = 50 K and 13 K, respectively. Red and blue circles show the band dispersion extracted from the peak position of the EDCs. (c) Comparison of the near-EF band dispersions at T = 50 K (red circles) and 13 K (blue circles), together with εk determined by polynomial fitting to the red circles (magenta curve) and the BQP dispersion E k = ε k 2 + | Δ | 2 simulated with a constant ∆ of 10 meV (light blue curve).
Figure 3. (a,b) ARPES intensity divided by the FD function measured along the k cut crossing the M point in 1 ML-FeSe at T = 50 K and 13 K, respectively. Red and blue circles show the band dispersion extracted from the peak position of the EDCs. (c) Comparison of the near-EF band dispersions at T = 50 K (red circles) and 13 K (blue circles), together with εk determined by polynomial fitting to the red circles (magenta curve) and the BQP dispersion E k = ε k 2 + | Δ | 2 simulated with a constant ∆ of 10 meV (light blue curve).
Symmetry 13 00155 g003
Figure 4. (a) k dependence of the superconducting-gap size ∆(k) which was calculated to reproduce the experimental band dispersion at 13 K for 1 ML-FeSe (blue circles in Figure 3c) with the formula E k = ε k 2 + | Δ | 2 , where εk is the normal-state band dispersion extracted at T = 50 K (red circles in Figure 3c). (b) Same as (a) but for Cs-deposited 20 ML-FeSe. (c) Energy difference between the experimental band dispersion at 13 K (blue circles in Figure 3c) and the simulated BQP dispersion with a constant gap size of 10 meV (light blue curve in Figure 3c) in 1 ML-FeSe. (d) Same as (c) but for Cs-deposited 20 ML-FeSe.
Figure 4. (a) k dependence of the superconducting-gap size ∆(k) which was calculated to reproduce the experimental band dispersion at 13 K for 1 ML-FeSe (blue circles in Figure 3c) with the formula E k = ε k 2 + | Δ | 2 , where εk is the normal-state band dispersion extracted at T = 50 K (red circles in Figure 3c). (b) Same as (a) but for Cs-deposited 20 ML-FeSe. (c) Energy difference between the experimental band dispersion at 13 K (blue circles in Figure 3c) and the simulated BQP dispersion with a constant gap size of 10 meV (light blue curve in Figure 3c) in 1 ML-FeSe. (d) Same as (c) but for Cs-deposited 20 ML-FeSe.
Symmetry 13 00155 g004
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Nakayama, K.; Shigekawa, K.; Sugawara, K.; Takahashi, T.; Sato, T. Unusual Temperature Evolution of Quasiparticle Band Dispersion in Electron-Doped FeSe Films. Symmetry 2021, 13, 155. https://doi.org/10.3390/sym13020155

AMA Style

Nakayama K, Shigekawa K, Sugawara K, Takahashi T, Sato T. Unusual Temperature Evolution of Quasiparticle Band Dispersion in Electron-Doped FeSe Films. Symmetry. 2021; 13(2):155. https://doi.org/10.3390/sym13020155

Chicago/Turabian Style

Nakayama, Kosuke, Koshin Shigekawa, Katsuaki Sugawara, Takashi Takahashi, and Takafumi Sato. 2021. "Unusual Temperature Evolution of Quasiparticle Band Dispersion in Electron-Doped FeSe Films" Symmetry 13, no. 2: 155. https://doi.org/10.3390/sym13020155

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop