Dichotomy of superconductivity between monolayer FeS and FeSe
Edited by Malcolm R. Beasley, Stanford University, Stanford, CA, and approved October 24, 2019 (received for review July 25, 2019)
Significance
Monolayer FeSe on SrTiO3 shows the highest superconducting-transition temperature (Tc) among iron-based superconductors, but the origin still remains unresolved. While monolayer FeS has the same crystal structure as FeSe, it is unknown to exhibit the superconductivity because of difficulty in crystal growth and its metastable nature. We have succeeded in synthesizing monolayer FeS by the topotactic reaction and molecular-beam epitaxy method, and compared the electronic structure between the 2. Whereas a strong interfacial electron–phonon coupling and a charge transfer through the interface proposed as a source for high-Tc superconductivity in monolayer FeSe are also observed in FeS, high-Tc superconductivity is surprisingly absent in monolayer FeS. Such striking difference provides insights into the mechanism of high-Tc superconductivity in iron-based superconductors.
Abstract
The discovery of high-temperature (Tc) superconductivity in monolayer FeSe on SrTiO3 raised a fundamental question: Whether high Tc is commonly realized in monolayer iron-based superconductors. Tetragonal FeS is a key material to resolve this issue because bulk FeS is a superconductor with Tc comparable to that of isostructural FeSe. However, difficulty in synthesizing tetragonal monolayer FeS due to its metastable nature has hindered further investigations. Here we report elucidation of band structure of monolayer FeS on SrTiO3, enabled by a unique combination of in situ topotactic reaction and molecular-beam epitaxy. Our angle-resolved photoemission spectroscopy on FeS and FeSe revealed marked similarities in the electronic structure, such as heavy electron doping and interfacial electron–phonon coupling, both of which have been regarded as possible sources of high Tc in FeSe. However, surprisingly, high-Tc superconductivity is absent in monolayer FeS. This is linked to the weak superconducting pairing in electron-doped multilayer FeS in which the interfacial effects are absent. Our results strongly suggest that the cross-interface electron–phonon coupling enhances Tc only when it cooperates with the pairing interaction inherent to the superconducting layer. This finding provides a key insight to explore heterointerface high-Tc superconductors.
Sign up for PNAS alerts.
Get alerts for new articles, or get an alert when an article is cited.
Tetragonal FeSe is an iron-based superconductor with the simplest layered structure, and has received tremendous attention because high-temperature (Tc) superconductivity was discovered in 1-monolayer (1-ML) film grown on SrTiO3 substrate (1). The Tc value of 1-ML FeSe (∼65 K) (2, 3) is almost an order of magnitude higher than that (∼8 K) of bulk FeSe (4), and is the highest among iron-based superconductors. The mechanism of such Tc enhancement is a central issue in condensed-matter physics. The widely discussed scenario is that the high Tc is closely linked to the coupling of FeSe layer with SrTiO3 substrate (5). A major role of SrTiO3 substrate is an electron doping to the FeSe layer, producing a large electronlike Fermi surface centered at the M point in the Brillouin zone (2, 3), in sharp contrast to bulk FeSe where hole and electron pockets coexist at the Γ and M point, respectively (6, 7). While this excess electron doping seems to enhance Tc, the maximum Tc achieved solely by electron doping to multilayer FeSe thin films (8, 9) never reaches 65 K. Therefore, substrate must play another important role to enhance Tc in 1-ML FeSe. As such, a strong coupling of SrTiO3 phonons with FeSe electrons across the interface has been proposed (10–12). This effect has been observed as the replica band separated by the optical phonon energy from the main band (10–12). However, it remains still elusive whether the electron–phonon coupling alone is sufficient to account for the high Tc (13–18). A crucial test to clarify the contribution of interfacial electron–phonon coupling is a comparative study on a 1-ML film of other iron-based superconductors and to elucidate whether a similar high-Tc superconductivity associated with the electron–phonon coupling can be realized. Thus, the synthesis of such 1-ML system is of great importance for a crucial understanding on the origin of high-Tc superconductivity. Also, this study would provide a pathway to further exploration of high-Tc systems.
Tetragonal FeS and FeTe are particularly suited to resolve this issue because they are isostructural to FeSe. However, in FeTe, the robust antiferromagnetism (19) seems to kill the superconductivity, so that the high-Tc superconductivity has not been observed in either bulk or 1-ML film on SrTiO3 (a signature of interfacial electron–phonon coupling is also absent) (20). On the other hand, it was recently reported that FeS shows a bulk superconductivity with Tc = 4.5 K (21), comparable to that of bulk FeSe, providing a hopeful platform to explore the high-Tc superconductivity in the 1-ML system. However, a high-quality tetragonal FeS thin film (Fig. 1A) is known to be hardly obtained by the molecular-beam epitaxy (MBE) technique which is widely used for synthesizing 1-ML FeSe and FeTe, and hence the presence or absence of high-Tc superconductivity as well as the interfacial effects has not been verified (22, 23). This is mainly due to the high vapor pressure of S atoms, and more critically, to the existence of the hexagonal phase (Fig. 1B) that is thermodynamically more stable than the tetragonal phase. Despite these difficulties, it is strongly desired to find a new route to synthesize a high-quality tetragonal FeS film.
Fig. 1.
Here we demonstrate an efficient synthesis route of high-quality tetragonal 1-ML FeS. We adopted the topotactic reaction, which can convert the crystalline phase with keeping the lattice symmetry of parent crystals. Using this method with FeTe as a starting material, we have successfully grown a 1-ML tetragonal FeS film on a SrTiO3 substrate. By angle-resolved photoemission spectroscopy (ARPES), we have revealed that 1-ML FeS definitely experiences an interfacial effect. But, to our surprise, 1-ML FeS does not show high-Tc superconductivity in sharp contrast to FeSe. We show that the dichotomy of superconductivity between FeS and FeSe monolayers is the key to understand the high Tc mechanism of Fe chalcogenides.
Results
As schematically shown in Fig. 1C, the topotactic substitution of Te atoms with S atoms is promoted by heating an MBE-grown FeTe film at 270 °C while showering S molecular beams onto the film surface in a vacuum for 5 min (see Materials and Methods for details of the FeTe-growth and -sulfurization conditions). The resultant film has a clean and flat surface, as recognized from a sharp streak reflection high-energy electron diffraction (RHEED) pattern in Fig. 1D. Importantly, the tetragonal lattice is maintained, as evidenced by the 4-fold symmetry of the low-energy electron diffraction (LEED) pattern (Fig. 1E). In addition, scanning transmission electron microscopy (STEM) measurements revealed 3 atomic layers (Fig. 1F), consistent with the formation of a 1-ML film. Furthermore, we performed energy-dispersive X-ray spectroscopy (EDX) measurements to estimate the S content in the film. As visible in Fig. 1G, the composition line profile signifies the presence of S atoms with no trace of Te atoms, showing an almost 100% substitution of Te atoms with S atoms. This is corroborated by the absence of Te-derived core-level peaks in the X-ray photoemission spectroscopy (XPS) spectrum (SI Appendix, Fig. S1). All these characterizations unambiguously demonstrate a successful synthesis of tetragonal 1-ML FeS film on SrTiO3 by in situ topotactic reaction. It is noted that the spatial resolution of the EDX analysis in the present experimental setup is estimated to be about 2 nm. This is comparable to the full probed width of FeS/SrTiO3 (∼2.5 nm, Fig. 1G), meaning that even a sharp step edge in the composition is substantially broadened in the EDX line profile due to the spatial resolution. Thus, we think that the apparent diffusion of Ti in the EDX line profile is due to the finite spatial resolution in the EDX measurement.
The conversion from FeTe to FeS is also confirmed by a comparison of the electronic structure by ARPES (Fig. 1 H–K). As seen from Fig. 1 H and I, an as-grown 1-ML FeTe film has a holelike Fermi surface centered at the Γ point. In addition, a significant spectral broadening is observed as in the previous study (24), possibly due to the strong correlation effect (25). In contrast, sulfurization leads to a development of sharp quasiparticle peaks (Fig. 1K), enabling a precise determination of the electronic structure: The top of the holelike band (α) at the Γ-point is located at ∼80 meV below the Fermi level (EF), and the Fermi surface consists only of large electron pockets (γ/δ) centered at the M point (Fig. 1 J and K). The observed drastic change in the electronic structure strongly supports our successful sulfurization (we found that the same method works also for the conversion from FeTe to FeSe and from FeSe to FeS; SI Appendix, Fig. S2). Another important finding is that the electron carrier concentration estimated from the Fermi-surface volume in Fig. 1J is very high (∼0.15 electrons per Fe), indicating that the heavy electron doping takes place in 1-ML FeS. This would be a result of charge transfer from the SrTiO3 substrate, similarly to the case of 1-ML FeSe (2, 3). We verified the charge transfer by performing thickness-dependent ARPES measurements (SI Appendix, Fig. S3). It is worth noting that the doping level of 0.15 electrons/Fe is well inside the high-Tc phase in the case of 1-ML FeSe (26).
A careful look at the ARPES intensity further reveals an intriguing subband feature (γ′) indicated by a white arrow in Fig. 1K. To discuss the origin of this band in more detail, we show in Fig. 2 the ARPES spectra and intensity measured around the M point. Along cut A in Fig. 2A, one can clearly see the γ and γ′ bands (Fig. 2 B–D). The observed γ′-band dispersion is well reproduced by an ∼100-meV downward shift of the γ-band dispersion (see dots in Fig. 2B, dashed curves in Fig. 2D, and schematic in Fig. 2E), and the energy separation between the γ′ and γ bands is close to the energy of optical phonons in SrTiO3. These characteristics of the γ′-band are qualitatively similar to those of the main-band replica produced by the phonon shake-off effect in 1-ML FeSe (10–12). Therefore, the γ′-band is likely attributed to the replica band caused by the coupling with SrTiO3 phonons. It is noted that a possibility that the replica band is simply the outer electron pocket (27) is ruled out because the γ′-band is not connected to the outer electron pocket (δ) as observed in cut B (Fig. 2 F and G). The interface origin of the γ′ band is also supported by the fact that it disappears in multilayer films (Fig. 2H and SI Appendix, Fig. S4).
Fig. 2.
To estimate the strength of electron–phonon coupling, we performed numerical fittings to the ARPES spectrum at the M point (Fig. 2I) by following the previous studies on FeSe that proposed the intensity ratio of γ′ to γ band (η = Iγ'/Iγ) to be proportional to the coupling constant (10–12). The obtained η of 0.17 is comparable to that reported for 1-ML FeSe (10–12), suggesting that the interfacial electron–phonon coupling in 1-ML FeS is as strong as that in 1-ML FeSe. From the numerical fittings, we also estimated the precise energy separation between the γ and γ′ bands to be 110 meV, which is comparable to that in 1-ML FeSe (∼100 meV).
Having established the presence of the electron doping and the interfacial electron–phonon coupling, it is tempting to search for high-Tc superconductivity in 1-ML FeS. Fig. 3A shows high-resolution ARPES spectra measured at several Fermi wave-vector (kF) points of the γ-band at T = 10 K. One can immediately see that the leading-edge midpoint is always located at around EF irrespective of the kF points. Corresponding symmetrized spectra in Fig. 3B show a single peak at EF, indicative of the absence of a superconducting gap. This suggests that 1-ML FeS does not show high-Tc superconductivity, in sharp contrast to 1-ML FeSe (Fig. 3 C and D) where the ARPES spectrum obviously shows a depletion of spectral weight at EF. We also found that the multilayer FeS films do not show high-Tc superconductivity even when electrons are sufficiently doped (SI Appendix, Fig. S4), again in contrast to the case of K-coated multilayer FeSe with the Tc over 40 K (8, 9).
Fig. 3.
The observed contrasting behavior between the electron-doped FeS and FeSe thin films are summarized in Fig. 3E, where we see that even in the absence of interfacial electron–phonon coupling (η = 0, corresponding to the alkali-metal-coated multilayer films), FeSe shows high-Tc superconductivity (40–48 K) with a superconducting-gap size Δ of ∼8 meV (8, 9), while in FeS the magnitude of Tc and Δ appears to be below our detectable limits (SI Appendix, Fig. S4). It has been discussed that an inclusion of the interfacial electron–phonon coupling (η > 0, corresponding to 1-ML films) leads to an enhancement in Tc and Δ by about 20 K and 5 meV, respectively, in FeSe (2, 3), while in FeS such an enhancement in Tc or Δ is not observed (Fig. 3 A and B).
Discussion
One may think that the absence of high-Tc superconductivity in 1-ML FeS may be due to some extrinsic effects such as disorder in the crystal. We have examined the possibility of disorder in the FeS film from the ARPES measurements. It is well known that the disorder in the crystal causes a considerable broadening of quasiparticle peaks in the ARPES spectrum due to the shortened quasiparticle lifetime. If a substantial disorder is present in our 1-ML FeS and suppresses the superconductivity, the quasiparticle peak would be broader than that of superconducting 1-ML FeSe. However, as seen in Fig. 3 A and C, the quasiparticle peak in 1-ML FeS is much sharper than high-Tc 1-ML FeSe. This indicates that our 1-ML FeS has a negligible disorder in the crystal. Since one may also wonder whether the lattice strain may suppress the superconductivity, we compared the in-plane lattice constant extracted from the LEED pattern. The results give a = 3.87 ± 0.04 Å for 1-ML FeS on SrTiO3, which is about 5% larger than that of bulk FeS (3.68 Å), suggesting that there is a tensile strain in the 1-ML FeS film due to the lattice mismatch with the SrTiO3 substrate (a = 3.91 Å). Since high-Tc superconductivity occurs in 1-ML FeSe with a similar in-plane lattice constant (a ∼ 3.9 Å) or a similar magnitude of tensile strain (∼5%) (3, 28), the absence of high-Tc superconductivity in 1-ML FeS would not be directly linked to the tensile strain.
The observed striking difference between FeS and FeSe rather suggests that while the electron doping and the resultant change in the Fermi-surface topology may be important to realize superconductivity in FeSe (e.g., by suppressing a nematic order), they are not sufficient conditions for the occurrence of high-Tc superconductivity in FeS. In other words, the pairing interaction inherent in the electron-doped FeSe layer is much stronger than that in the electron-doped FeS layer. Then, the important question is what kind of interaction in FeSe layer triggers superconductivity. To clarify this point, we compared the band dispersions between 1-ML FeS and FeSe at a similar doping level and found a clear difference in the γ-bandwidth (Fig. 3 F and G). The bandwidth is twice larger in FeS than in FeSe although both bands originate from the same Fe 3d orbital, but not from the chalcogen p orbital (note that a similar difference in the bandwidth is observed in multilayer films). This result highlights the strongly correlated nature of FeSe, and implies that the superconductivity in the absence of interfacial electron–phonon coupling may be associated with spin and/or orbital fluctuations enhanced by the strong electron correlation. The strong correlation in FeSe may be also related to the emergence of a Mott-like insulating phase in nondoped 1-ML FeSe (29). In order to get a deeper insight into this issue, the growth of nondoped 1-ML FeS and its ARPES study are important in future.
The absence of high-Tc superconductivity in 1-ML FeS which exhibits a clear signature of replica band also puts a strong constraint on microscopic theories to explain how the interfacial electron–phonon coupling contributes to high-Tc superconductivity. While the experimental evidence for the importance of interfacial electron–phonon coupling for the superconducting pairing is accumulated thus far, there are controversies among the theoretical proposals regarding the magnitude of Tc enhancement; some theories have proposed that the electron–phonon coupling alone explains high-Tc superconductivity (16, 18), while others suggested that a cooperation of interfacial electron–phonon coupling with other primary pairing mechanisms (likely spin/orbital fluctuations) is essential to achieve high Tc (13–15). The absence of high Tc in monolayer FeS demonstrates that the interfacial electron–phonon coupling is not strong enough to induce high-Tc superconductivity by itself, supporting the latter scenario. The electron–phonon coupling seems to enhance Tc only when it cooperates with the pairing interaction inherent to the superconducting layer. This result provides a useful strategy for exploring atomic-layer high-Tc systems; we should at first choose starting materials that inherently have a potential for high-Tc superconductivity and then increase Tc by using the cooperation with interfacial effects.
Conclusions
In summary, we have succeeded in synthesizing a high-quality 1-ML film of metastable tetragonal FeS by the combined method of the topotactic reaction and MBE. The comparative ARPES measurements between FeS and FeSe have revealed marked similarities (electron doping and electron–phonon coupling across the interface) and differences (presence/absence of high-Tc superconductivity and electron correlation strength). The present results clearly show that the excess electron doping and the electron–phonon coupling across the interface may be necessary conditions, but are not sufficient conditions for high-Tc superconductivity. The remarkable difference in the γ-band at EF between FeS and FeSe suggests an important role of strong correlation for high Tc. The present study has also established the high usefulness of topotactic reaction to synthesize metastable 2D materials. This method would be applied to a wide range of 2D materials such as transition-metal dichalcogenides and topological materials.
Materials and Methods
An FeTe thin film was grown on a Nb (0.05 wt %)-doped SrTiO3 substrate (Shinkosha) with the MBE method in a vacuum better than 2 × 10−10 Torr. The substrate was first degassed at 600 °C for 2 h, and then annealed at 900 °C for 30 min. An FeTe film was grown by coevaporating Fe and Te in a Te-rich condition while keeping the substrate temperature at 270 °C. The film thickness was controlled by varying the deposition time with keeping the constant deposition rate (∼0.01 ML/s). After the evaporation, the film was postannealed at 300 °C for 2 h in a vacuum. Next, the film was heated to 270 °C and exposed to a sulfur (S) molecular beam (the partial pressure of ∼10−8 Torr) for 5 min to obtain an FeS film. Then, the FeS film was postannealed at 530 °C for 1 h in a vacuum better than 2 × 10−10 Torr. After the growth, the film was immediately transferred to the ARPES-measurement chamber which is directly connected to the MBE chamber. RHEED and LEED measurements were employed to characterize the surface cleanness and structure of the sample. EDX and XPS measurements were performed to confirm the substitution of Te with S atoms. The spatial resolution in the present EDX setup (∼50-nm sample thickness and 200-keV incident beam energy) is estimated to be about 2 nm by referring to the textbook (30). Deposition of cesium (Cs) atoms was carried out at room temperature with a Cs dispenser (SAES Getters).
ARPES measurements were performed in an ultrahigh vacuum better than 5 × 10−11 Torr using a Scienta-Omicron SES2002 spectrometer with a high-flux He discharge lamp (hν = 21.218 eV) at Tohoku University. The energy and angular resolutions were set to be 2–30 meV and 0.2°, respectively. The Fermi level of sample was referenced to that of a gold film evaporated onto the sample holder.
Data Availability
Data deposition: The data that support the findings of this study are available from the corresponding author on request.
Acknowledgments
We thank K. Fujiwara for discussion. This work was supported by Japan Society for the Promotion of Science KAKENHI Grants JP17H04847, JP17H01139, JP18H01160, JP18K18986, and JP18H01821; The Ministry of Education, Culture, Sports, Science and Technology of Japan (Innovative Area “Topological Materials Science” Grant JP15H05853); the Japan Science and Technology Agency Precursory Research for Embryonic Science and Technology Grant JPMJPR18L7; Core Research for Evolutional Science and Technology Grant JPMJCR18T1; and High Energy Accelerator Research Organization (KEK) Photon Factory Proposal 2018S2-001.
Supporting Information
Appendix (PDF)
- Download
- 2.59 MB
References
1
Q. Y. Wang et al., Interface-induced high-temperature superconductivity in single unit-cell FeSe films on SrTiO3. Chin. Phys. Lett. 29, 037402 (2012).
2
S. He et al., Phase diagram and electronic indication of high-temperature superconductivity at 65 K in single-layer FeSe films. Nat. Mater. 12, 605–610 (2013).
3
S. Tan et al., Interface-induced superconductivity and strain-dependent spin density waves in FeSe/SrTiO3 thin films. Nat. Mater. 12, 634–640 (2013).
4
F. C. Hsu et al., Superconductivity in the PbO-type structure α-FeSe. Proc. Natl. Acad. Sci. U.S.A. 105, 14262–14264 (2008).
5
D.-H. Lee, Routes to high-temperature superconductivity: A lesson from FeSe/SrTiO3. Annu. Rev. Condens. Matter Phys. 9, 261–282 (2018).
6
J. Maletz et al., Unusual band renormalization in the simplest iron-based superconductor FeSe1-x. Phys. Rev. B 89, 220506 (2014).
7
K. Nakayama et al., Reconstruction of band structure induced by electronic nematicity in an FeSe superconductor. Phys. Rev. Lett. 113, 237001 (2014).
8
Y. Miyata, K. Nakayama, K. Sugawara, T. Sato, T. Takahashi, High-temperature superconductivity in potassium-coated multilayer FeSe thin films. Nat. Mater. 14, 775–779 (2015).
9
C. H. P. Wen et al., Anomalous correlation effects and unique phase diagram of electron-doped FeSe revealed by photoemission spectroscopy. Nat. Commun. 7, 10840 (2016).
10
J. J. Lee et al., Interfacial mode coupling as the origin of the enhancement of Tc in FeSe films on SrTiO3. Nature 515, 245–248 (2014).
11
C. Zhang et al., Ubiquitous strong electron-phonon coupling at the interface of FeSe/SrTiO3. Nat. Commun. 8, 14468 (2017).
12
Q. Song et al., Evidence of cooperative effect on the enhanced superconducting transition temperature at the FeSe/SrTiO3 interface. Nat. Commun. 10, 758 (2019).
13
Y. Y. Xiang, F. Wang, D. Wang, Q. H. Wang, D. H. Lee, High-temperature superconductivity at the FeSe/SrTiO3 interface. Phys. Rev. B 86, 134508 (2012).
14
Z. X. Li, F. Wang, H. Yao, D. H. Lee, What makes the Tc of monolayer FeSe on SrTiO3 so high: A sign-problem-free quantum Monte Carlo study. Sci. Bull. (Beijing) 61, 925–930 (2016).
15
Y. Wang, A. Linscheid, T. Berlijn, S. Johnston, Ab initio study of cross- interface electron-phonon couplings in FeSe thin films on SrTiO3 and BaTiO3. Phys. Rev. B 93, 134513 (2016).
16
L. Rademaker, Y. Wang, T. Berlijn, S. Johnston, Enhanced superconductivity due to forward scattering in FeSe thin films on SrTiO3 substrates. New J. Phys. 18, 022001 (2016).
17
Y. J. Zhou, A. J. Millis, Dipolar phonons and electronic screening in monolayer FeSe on SrTiO3. Phys. Rev. B 96, 054516 (2017).
18
M. L. Kulić, O. V. Dolgov, The electron-phonon interaction with forward scattering peak is dominant in high Tc superconductors of FeSe films on SrTiO3 (TiO2). New J. Phys. 19, 013020 (2017).
19
S. Li et al., First-order magnetic and structural phase transitions in Fe1+ySexTe1-x. Phys. Rev. B Condens. Matter Mater. Phys. 79, 054503 (2009).
20
F. Li et al., Interface-enhanced high-temperature superconductivity in single-unit-cell FeTe1-xSex films on SrTiO3. Phys. Rev. B 91, 220503 (2015).
21
X. Lai et al., Observation of superconductivity in tetragonal FeS. J. Am. Chem. Soc. 137, 10148–10151 (2015).
22
K. Zhao et al., Molecular beam epitaxy growth of tetragonal FeS films on SrTiO3(001) substrates. Chin. Phys. Lett. 34, 087401 (2017).
23
H. Lin et al., Growth of atomically thick transition metal sulfide films on graphene/6H-SiC(0001) by molecular beam epitaxy. Nano Res. 11, 4722–4727 (2018).
24
Z.-Q. Han, X. Shi, X.-L. Peng, Y.-J. Sun, S.-C. Wang, High-quality FeTe1-xSex monolayer films on SrTiO3(001) substrates grown by molecular beam epitaxy. Chin. Phys. Lett. 34, 107401 (2017).
25
E. Ieki et al., Evolution from incoherent to coherent electronic states and its implications for superconductivity in FeTe1-xSex. Phys. Rev. B 89, 140506 (2014).
26
X. Shi et al., FeTe1-xSex monolayer films: Towards the realization of high-temperature connate topological superconductivity. Sci. Bull. (Beijing) 62, 503–507 (2017).
27
I. A. Nekrasov, N. S. Pavlov, M. V. Sadovskii, On the origin of the shallow and “replica” bands in FeSe monolayer superconductors. JETP Lett. 105, 370–374 (2017).
28
Z. H. Chuang et al., Electronic structure and superconductivity of single-layer FeSe on Nb:SrTiO3/LaAlO3 with varied tensile strain. 2D Mater. 3, 014005 (2016).
29
Y. Hu et al., Insulating parent phase and distinct doping evolution to superconductivity in single-layer FeSe/SrTiO3 films. arXiv:1904.04662 (9 April 2019).
30
J. Hutchison, A. Kirkland, Eds., Nanocharacterisation (Royal Society of Chemistry, 2014).
Information & Authors
Information
Published in
Classifications
Copyright
Copyright © 2019 the Author(s). Published by PNAS. This open access article is distributed under Creative Commons Attribution-NonCommercial-NoDerivatives License 4.0 (CC BY-NC-ND).
Data Availability
Data deposition: The data that support the findings of this study are available from the corresponding author on request.
Submission history
Published online: November 19, 2019
Published in issue: December 3, 2019
Keywords
Acknowledgments
We thank K. Fujiwara for discussion. This work was supported by Japan Society for the Promotion of Science KAKENHI Grants JP17H04847, JP17H01139, JP18H01160, JP18K18986, and JP18H01821; The Ministry of Education, Culture, Sports, Science and Technology of Japan (Innovative Area “Topological Materials Science” Grant JP15H05853); the Japan Science and Technology Agency Precursory Research for Embryonic Science and Technology Grant JPMJPR18L7; Core Research for Evolutional Science and Technology Grant JPMJCR18T1; and High Energy Accelerator Research Organization (KEK) Photon Factory Proposal 2018S2-001.
Notes
This article is a PNAS Direct Submission.
Authors
Competing Interests
The authors declare no competing interest.
Metrics & Citations
Metrics
Citation statements
Altmetrics
Citations
Cite this article
116 (49) 24470-24474,
Export the article citation data by selecting a format from the list below and clicking Export.
Cited by
Loading...
View Options
View options
PDF format
Download this article as a PDF file
DOWNLOAD PDFLogin options
Check if you have access through your login credentials or your institution to get full access on this article.
Personal login Institutional LoginRecommend to a librarian
Recommend PNAS to a LibrarianPurchase options
Purchase this article to access the full text.