Journal of Materials Science & Technology  2020 , 44 (0): 223-228 https://doi.org/10.1016/j.jmst.2019.10.033

Research Article

Structure and topological transport in Pb-doping topological crystalline insulator SnTe (001) film

C.H. Yanab, F. Weiab, Y. Baiab, F. Wangab, A.Q. Zhangac, S. Maa*, W. Liua, Z.D. Zhanga

a Shenyang National Laboratory for Materials Science, Institute of Metal Research, Chinese Academy of Sciences, Shenyang 110016, China
b University of Chinese Academy of Sciences, Beijing 100049, China
c School of Materials Science and Engineering, University of Science and Technology of China, Hefei 230026, China

Corresponding authors:   * E-mail address: songma@imr.ac.cn (S. Ma).

Received: 2019-08-2

Revised:  2019-10-17

Accepted:  2019-10-17

Online:  2020-05-01

Copyright:  2020 Editorial board of Journal of Materials Science & Technology Copyright reserved, Editorial board of Journal of Materials Science & Technology

More

Abstract

Topological crystalline insulator (TCI) as a new type of topological materials has attracted extensive research interests for its tunable topological properties. Due its symmetry topological protection essence, the structure investigation provides a solid basement for tuning its topological transport properties. On SrTiO3 (111) substrate, the SnTe film was found to be epitaxial growth only along [001] while not [111] direction. The detailed structural study was performed and a structural model was proposed to elucidate epitaxial growth of the SnTe (001) film. The transport properties of SnTe (001) film were further investigated and a typical weak anti-localization effect was observed. By Pb-doping into SnTe, the bulk carriers were inhibited and its topological surface states were strengthened to induce the enhanced surface transport contribution. With tunable multiple transport channels from the even Dirac cones, the TCI SnTe film systems will have the potential application in future spintronics devices.

Keywords: Topological crystalline insulator ; Surface state ; SnTe

0

PDF (1846KB) Metadata Metrics Related articles

Cite this article Export EndNote Ris Bibtex

C.H. Yan, F. Wei, Y. Bai, F. Wang, A.Q. Zhang, S. Ma, W. Liu, Z.D. Zhang. Structure and topological transport in Pb-doping topological crystalline insulator SnTe (001) film[J]. Journal of Materials Science & Technology, 2020, 44(0): 223-228 https://doi.org/10.1016/j.jmst.2019.10.033

1. Introduction

A new quantum matter, that is topological crystalline insulators (TCI), supplies a great application expectation to future quantum devices, because of its novel surface conductive states being immune to surface impurities or defects to avoid the possible thermal dissipation [1], [2]. The surface conductive sates of TCI arise from the crystal symmetries protection and its band structure usually consists of even number of Dirac cones in contrast to conventional topological insulator with odder Dirac cones [3], [4], [5], [6]. The first theoretically predicted TCI is the compound SnTe and its metallic surface states, originating from the even number of Dirac cones on high-symmetry crystal surfaces {001}, {110} and {111}, can be demonstrated by the band structural observation and electronic transport. The angle-resolved photoemission spectroscopy (ARPES) and scanning tunneling microscopy (STM) experiments firstly confirmed the presence of topological surface states (TSSs). The Dirac cones with linear dispersion relation have been clearly observed in single crystalline SnTe and their doping alloys SnxPb1-xTe and Pb1-xSnxSe [7], [8]. Based on the band structure confirmation of the topological surface state, the SnTe system become a significant platform to study the valley-degenerate topological systems in transport experiments for promoting TCI to be a potential quantum electronic device in future. The transport properties of SnTe system is focused on the high-symmetry crystal surfaces {001}, {110} and {111} with typical topological surface state and they are generally realized by the epitaxial growth on the corresponding single crystalline substrate SrTiO3(111), BaF2(001), BaF2(111) via the molecular beam epitaxy (MBE) [9], [10], [11], [12], [13], [14]. However, on the SrTiO3 (111) substrate, the epitaxial film SnTe (or SnSe) only growing along [001] while not [111] indicates a novel match relation between film and substrate, but this match relation is still not to be clarified. To construct the lattice match mechanism, the SnTe epitaxial film grown on SrTiO3 (111) is needed to be as a model system for presenting the corresponding structural detail and formation mechanism [15], [16], [17]. It will pave an important structural basement to understand its transport properties. Furthermore, the effect of element doping and film thickness on the transport behavior of the novel epitaxial SnTe (001) film on SrTiO3 (111) is significant for understanding the topological characteristics of TCI SnTe. The modulation of doping or film thickness will open up a continuously tunable band gap on the surface, which may lead to wide-ranging applications in thermoelectric and tunable electronics.

In this work, we systematically investigate the lattice structure of the SnTe (001) film grown on the SrTiO3 (111) substrate and construct a structural model for illustrating the lattice match correlation between SnTe film and SrTiO3 substrate. To suppress the bulk transport to enhance the surface conductive states contribution, the decrease of the film thickness and Pb-doping was used. The SnTe (001) film with thinner film thickness 15 nm presents larger sheet resistance indicating the enhancement of the TSSs in the general transport of the film. By introducing the Pb doping, an enhancement of phase coherent length and mobility but a decrease of the carrier density shows an increase of the TSSs and a suppression of the bulk carrier in general transport of the Pb-doping film.

2. Experimental

The thin films were prepared by molecular beam epitaxy system (MBE) with the based vacuum better than 1.5 × 10-10 mbar. The SnTe and Sn0.7Pb0.3Te films were grown on SrTiO3 (111) substrate by simultaneously evaporating pure Sn (99.999 wt.%), Te (99.999 wt.%) and Pb (99.999 wt.%) sources from the Knudsen cells. Before growth process, the SrTiO3 (111) substrate was performed degas circulation under 600 °C and then the substrate temperature was kept at 300 °C during film growth. The relatively narrow diffraction stripe from the in-situ reflection high energy electron diffraction (RHEED) indicates perfect single crystalline substrate (Fig. 1(a)) and good epitaxial growth of the thin film (Fig. 1(b)). In Fig. 1(c), the X-ray diffraction (XRD) spectrum of SnTe film further proves the grown orientation along [001] direction. The controllable thickness of SnTe film was achieved by controlling the growing rate with 1 nm/min. The Pb doping was realized by exactly controlling its evaporating ratio with Sn and Te elements whose exact content was detected by X-ray photoelectron spectroscopy (XPS) analysis at different depth of film (Fig. S1 in Supplementary materials). The cross-sectional atomic structure of the film was detected by the high resolution transmission microscopy (HRTEM). The transport behavior of the grown films was measured by a standard four-terminal Hall bar device, in which the schematic is shown in the inset of Fig. 1(d).

Fig. 1.   (a) RHEED pattern of substrate SrTiO3 (111) and (b) SnTe (001) film with thickness 15 nm; (c) XRD pattern of SnTe film in which the inset shows the lattice structure of SnTe; (d) transport measurement geometry with the Hall bar of the SnTe film.

3. Results and discussion

In Fig. 1(b), the RHEED pattern of a represented SnTe film with 15 nm thickness indicates good epitaxial growth. The faint RHEED streaks between three bright streaks indicate the multi-domain structure of the SnTe (001) film. The low-energy electron diffraction pattern of this film proves the existence of the multi-domain structure (Fig. S2). The similar domain structure was also observed in literature [13]. The diffraction peak in the XRD pattern is indexed as (002) reflection indicates the [001] growth of the SnTe film on SrTiO3 (111). The similar situation was observed in SnSe film on SrTiO3 (111), which suggests that SnTe or SnSe is easy to epitaxially grow along [001] on the (111) plane of SrTiO3 substrate [10], [18]. The atomic match correlation between the (001) plane of SnTe and the (111) plane of SrTiO3 is definitely responsible for this special epitaxial growth. In Fig. 2(a), the cross-sectional HRTEM image shows the interface structure of SnTe (001)/SrTiO3 (111). The characterized (002) and (200) crystalline planes of SnTe film were confirmed by measuring their typical d-spacing of 3.15 Å in two directions. The d-spacing of 2.76 Å and 2.25 Å correspond to (110) and (111) crystalline planes of SrTiO3 substrate. After the annealing procedure, the Ti-O plane always appears on the surface of SrTiO3 (111) and becomes the grown plane for SnTe film [19], therefore, the hexagonal Ti-O lattice on the (111) plane decides the atomic match of the SnTe film. In Fig. 2(c), the hexagonal atomic arrangement of Ti-O plane shows two different atomic distances between two Ti atoms along [110] and [$\bar{1}$10] directions that are a1 = 9.5 Å and a2 = 5.52 Å respectively. For SnTe film with a cubic rock salt structure, its (001) plane displays a square lattice with the shortest atomic distance b1 = 3.15 Å while its (111) plane shows a triangle lattice with the shortest atomic distance b2 = 4.48 Å. The lattice match between of SnTe and SrTiO3 is decided by their integer or integer multiple relation of lattice parameters, whose detailed geometry configurations are shown in Fig. 2(c) and (d).

Fig. 2.   (a) Cross section HRTEM image of SnTe(001)/SrTiO3(111) along the lattice axis [112]; (b) atomic arrangement and atomic distance of the (001) and (111) planes of SnTe film; (c) lattice structural model for the match correlation between SnTe (001), SnTe (111) and SrTiO3 (111); (d) cross-section model image between SnTe (001) and SrTiO3 (111).

For realizing appropriate atomic match, the atomic arrangement of SnTe (001) tends to along the [110] and [$\bar{1}$10] direction on the Ti-O plane, because the distance 9.5 Å of two Ti atoms at [110] direction corresponds to three times atomic distance (3 × 3.15 Å = 9.45 Å) on (001) plane of SnTe film; the distance 11.07 Å of three Ti atoms at [$\bar{1}$10] direction corresponds to four times atomic distance (4 × 3.15 Å = 12.6 Å) on (001) plane of SnTe film (Fig. 2(c). Through present atomic arrangement, the square lattice of SnTe (001) can be epitaxially grown on the hexagonal Ti-O plane of SrTiO3 (111). The different integer time’s atomic arrangement on two directions will induce a dislocations on the interface between SrTiO3 substrate and SnTe film. According to above atomic match, the distance of three Ti atoms corresponding to four Sn(Te) atoms on [$\bar{1}$10] direction means that the four Sn(Te) atoms epitaxial match with three Ti atoms and it obviously lead to the dislocations on [$\bar{1}$10] direction. The interfacial HRTEM image of SnTe (001)/SrTiO3 (111) shown in Fig. 2(a) provides a solid evidence for the existence of the dislocations on the interface. For clearly understanding the produce of the dislocations, the interfacial model was accordingly constructed in Fig. 2(d) and the dislocations are indicated by the red arrows.

There is no related report on the SnTe (111) films directly epitaxial growth on the SrTiO3 (111). The reason is due to the improper atomic match correlation between the (111) planes of SnTe and SrTiO3. Referring above atomic arrangement model, the lattice geometry of (111) of SnTe is an equilateral triangle with side length 8.95 Å, in which the Te atom is located on the vertex and the Sn atom is located on the center position of the SnTe (111) triangle lattice (Fig. 2(b)). For realizing an epitaxial match, the triangle geometry of SnTe (111) should have integer or integer multiple relations with the SrTiO3 (111). On the Ti-O plane of SrTiO3 (111), two types of triangle lattice geometry are presented that are equilateral triangle with side length 5.52 Å and isosceles triangle with long side length 9.5 Å and short side length 5.52 Å (Fig. 2(c)). When comparing the triangle lattices of SnTe and SrTiO3, there is no proper match for the atomic arrangement of these triangle geometries, although the long side length 9.5 Å is almost two times of the distance 4.48 Å between two Te atoms in triangle lattice of the SnTe (111). Therefore, the severe mismatch between SnTe (111) and SrTiO3 (111) finally leads to SnTe film cannot be epitaxially grown on the (111) plane of SrTiO3 substrate.

The temperature dependence of sheet resistance (R) of SnTe (001) films with different thickness was shown in Fig. 3(a). All the films display metallic conductance and their conductive behaviors are similar to the SnTe (111) films grown on Bi2Se3 and BaF2 (111) [14]. Generally, the transport of the topological film is simultaneously contributed by its bulk carriers and TSSs. The metallic transport in SnTe (001) film was considered to arise from the charged Sn vacancies in bulk which dominate the general electronic transport properties [20]. Taken SnTe (001) film with thickness 25 nm as a representative, the sheet resistance was fitted as perfect linear correlation from 300K to 100K, but a nonlinear decrease from 100 K to 4 K was observed (Fig. 3(b)). It suggests that the resistance mechanism above 100 K is dominated by the electron-phonon interaction (linear), but the slowly decreased resistance at low temperature range tending to a constant means the gradually enhanced contribution from the electron-electron interactions [21]. When the temperature tends to 4 K, the enhanced electron-electron interaction may induce more obvious surface conductive state contribution at low temperature quantum limitation [11], [14]. For investigating the film thickness influence on the general transport and TSSs, the film thickness was decreased to 15 nm. The resistance magnitude of 15 nm SnTe (001) film quickly increases 5 times over entire temperature range than 25 nm SnTe film. In suggests that the bulk carrier (Sn vacancy) contribution to general transport is decreased when decreasing the film thickness. The TSSs has more possibility to be shown in the SnTe (001) film. With the thickness increase from 25 nm to 60 nm, the resistance magnitude almost keeps unchanged and it indicates the completely bulk transport domination in the SnTe film and the TSSs will be suppressed.

Fig. 3.   (a) Temperature dependence of the sheet resistance of SnTe (001) films with different thickness and (b) R vs. T of the SnTe (001) films (25 nm thick) with fitting line above 100 K in which the inset shows the Hall curve of SnTe (001) film with 15 nm; (c) ΔG-B curves at the range of 4-40 K and (d) transfer channel of α and phase coherence length of SnTe (001) film with 15 nm.

To clearly characterize the TSSs of the SnTe (001) film, the magnetic field dependence of the conductance difference (ΔG = G(B) - G(0)) ΔG-B curves of the film with 15 nm were measured in low temperature range from 4 K to 40 K. The typical cusp-like curves with rapid conductance decrease when applying a small magnetic field are shown in Fig. 3(c). The cusp-like magneto-conductivity (MC) is attributed to the weak anti-localization (WAL) effect of the carrier transport related to the topological surface sates (TSSs) [22]. The WAL effect is recognized as one of the features of the quantum transport in the diffusive transport regime, arising from destructive interference. During transport of the TSSs, a shift in the Berry phase arises from the spin-momentum locking effect for an electron (or a hole) traveling along a self-intersecting path, which causes a destructive interference of the wave-function, leading to an enhancement of the conductivity [23]. When applying a magnetic field perpendicular to the closed path, it tends to break the destructive interference, giving rise to negative MC to indicate a key signature of WAL [24]. With temperature increasing to 10 K, the absolute values of ΔG progressively decrease and the cusp curve disappears, suggesting the temperature sensitive feature of the WAL in SnTe (001) film. The surface states are quickly suppressed by increased temperature and the enhanced phonon effect is responsible for the gradual disappearance of cusp in ΔG-B curves.

For TCI SnTe (001), its TSS has four Dirac cones along Γ-X line of the first Brillouin zone and the surface transport channel always was determined by the number of Dirac cones on the (001) surface of SnTe film. When combined with top and bottom surface states of SnTe film, the total surface channels should be eight that means the fitting value α should be four when using the classic Hikami-Larkin-Nagaoka (HLN) function (Eq. (1)) to fit the ΔG-B curve (one surface channels should be fitted as 0.5). In Eq. (1), e is the charge of an electron, ћ is Planck’s constant, lф is the phase coherent length, Ψ(x) is the digamma function, and α represents a coefficient related to the number of coherent transport channels. In temperature dependence of α curve (α~T) of Fig. 3(d), the fitting value α tends to be α=3 in the range of 4-8 K which is near to the theoretical value α = 4 representing the eight surface transport channels originated from the eight Dirac cones on the top and bottom surface of the SnTe (001) film. The little difference between fitted value and theoretical value is possibly attribute to the interaction among these channels [25]. When the temperature is increased to 10 K, the ΔG-B curves changes from cusp to parabolic indicating the disappearance of the WAL effect, then the enhanced phonon scattering gradually destroy the surface state quantum transport. For another parameter lф, it represents the phase coherence length of the Dirac electrons originating from the surface sate. The decreased tendency of the lф with the decreased temperature shows a consistence with the diffusive transport theory (Fig. 3(d)) [11], [26], [27]. For analyzing the contribution of TSSs and bulk to general transport of SnTe (001) films at low temperature, the Hall resistance was measured at 4 K and shown in the inset of Fig. 4(b). The positive slope indicates the p-type conductance. It means that the bulk holes as the main carriers is responsible for the general transport of the SnTe (001) films at low temperature. The large carriers density (4.6 × 1021 cm-3) and small mobility (17.38 (cm2/(V s))) derived from Hall curve indicate the holes mainly contribute to the general transport to the SnTe (001) film, although the typical WAL is observed in ΔG-B curve at 4 K.

$\Delta G=G(B)-G(0)=-\alpha\frac{e^{2}}{2 \pi^{2}\hbar}[in(\frac{\hbar}{4Bel^{2}_{\Phi}})-\Psi(\frac{1}{2}+\frac{\hbar}{4Bel^{2}_{\Phi}})]$ (1)

For further decreasing the influence of bulk carriers on general transport and inducing stronger surface state in SnTe (001) film, the n-type Pb-doping was introduced to inhibit the formation of Sn vacancy and decrease the number of holes. The transport study of Sn0.7Pb0.3Te (001) film was carried out to check whether more TSSs contribution to general electronic transport is obvious. In Fig. 4(a), the R-T curve of Sn0.7Pb0.3Te (001) film with 30 nm thickness indicates a semi-conductive behavior in which the sheet resistance increases with decreasing temperature and a platform appears when the temperature decrease to 30 K. It indicates that Pb-doping in SnTe may open a gap at the Dirac point on (001) to induce a transition from metallic conductive SnTe to semi-conductive Sn0.7Pb0.3Te. The appearance of the resistance platform below 30 K indicates the gradually increased contribution from the electron-electron interaction coming from the TSSs of Pb-doping Sn0.7Pb0.3Te(001) film. Furthermore, the carrier density and mobility was obtained from the Hall curve of Sn0.7Pb0.3Te (001) film, as shown in the inset of Fig. 4(a). The positive slope shows that the p-type carriers still maintain the major contribution to the general transport, but the carrier density is really decreased to 9.3 × 1019 cm-3 which is decreased two orders comparing to the SnTe (001) film. The mobility of the film is also quickly increased to 361.64 cm2/(V s) with almost two orders increase when comparing the SnTe (001) film. It indicates that Pb-doping really decreases the holes number in bulk and leads the enhancement of the surface state in present Sn0.7Pb0.3Te (001) film. The typical WAL effect in Sn0.7 Pb0.3Te (001) film from 4 K to 15 K are observed in Fig. 4(b) and the critical temperature for WAL effect disappearing is raised to 15 K when comparing with un-doped SnTe film. These facts suggest that the Pb-doping really suppresses the formation of p-type vacancy and strengthens the topological non-trivial state in the present film. With HLN functions, the parameters, including the transfer channel α and phase coherent length lф, were also fitted from the ΔG-B curves of Sn0.7Pb0.3Te film. From 4 K to 15 K, the α value is kept as reasonable value and the existing temperature for the reasonable α value is maintained to be little higher than the SnTe film, which suggests that the strengthened TSSs can weaken the phonon effect when temperature increasing. At the same time, the phase coherent length lф of present Sn0.7Pb0.3Te film was found to be quickly enhanced than SnTe film (inset of Fig. 4(c)). It indicates that the Pb-doping not only decreases the influence of bulk carrier density on the surface transport of the film, but also enhances the lф to induce the increase of the surface mobility of the TCI film.

Fig. 4.   (a) Temperature dependence of sheet resistance of Sn0.7Pb0.3Te(001) film; (b) ΔG-B curves at range of 4-30 K of Sn0.7Pb0.3Te(001) film; (c) transfer channel of α and coherent length of Sn0.7Pb0.3Te(001).

4. Conclusion

The SnTe (001) film was prepared on the single crystalline substrate SrTiO3 (111) by MBE system. The epitaxial growth of SnTe (001) film on SrTiO3 (111) was confirmed by the XRD pattern HRTEM analysis. The corresponding structural model was proposed according to their atomic match correlation. The severely atomic mismatch between the SnTe (111) and SrTiO3 (111) explains the non-epitaxial growth of SnTe (111). The WAL correlated the TSSs was observed in the SnTe (001) films, but the metallic transport indicates the bulk holes mainly contribute to the general transport of the film. Through Pb-doping to reduce the bulk holes in SnTe film, the general transport is induced to semi-conductive behavior in Sn0.7Pb0.3Te (001) film and the TSSs is enhanced. With the strengthen TSSs in Sn0.7Pb0.3Te (001) film, the WAL existing temperature and phase coherent length are simultaneously enhanced, which indicates that Pb-doping is the effective strategy for suppressing the bulk transport and promote the TSSs in TCI SnTe system.

Acknowledgment

This work was supported financially by the National Natural Science Foundation of China (Nos. 51571195 and 51590883) and the National Key R&D Program of China (No. 2017YFA0206301).

Appendix A. Supplementary data

Supplementary material related to this article can be found, in the online version, at doi: https://doi.org/10.1016/j.jmst.2019.10.033.


Reference

[1] H.T. Hsieh, L. Hsin, J.W. Liu, W.H. Duan, A. Bansil, L. Fu, Nat. Commun. 3(2012) 982.

DOI      URL      PMID      [Cited within: 1]      Abstract

Topological crystalline insulators are new states of matter in which the topological nature of electronic structures arises from crystal symmetries. Here we predict the first material realization of topological crystalline insulator in the semiconductor SnTe by identifying its non-zero topological index. We predict that as a manifestation of this non-trivial topology, SnTe has metallic surface states with an even number of Dirac cones on high-symmetry crystal surfaces such as {001}, {110} and {111}. These surface states form a new type of high-mobility chiral electron gas, which is robust against disorder and topologically protected by reflection symmetry of the crystal with respect to {110} mirror plane. Breaking this mirror symmetry via elastic strain engineering or applying an in-plane magnetic field can open up a continuously tunable band gap on the surface, which may lead to wide-ranging applications in thermoelectrics, infra-red detection and tunable electronics. Closely related semiconductors PbTe and PbSe also become topological crystalline insulators after band inversion by pressure, strain and alloying.
[2] L. Fu,Phys. Rev. Lett. 106(2011), 106802.

DOI      URL      PMID      [Cited within: 1]      Abstract

The recent discovery of topological insulators has revived interest in the band topology of insulators. In this Letter, we extend the topological classification of band structures to include certain crystal point group symmetry. We find a class of three-dimensional "topological crystalline insulators" which have metallic surface states with quadratic band degeneracy on high symmetry crystal surfaces. These topological crystalline insulators are the counterpart of topological insulators in materials without spin-orbit coupling. Their band structures are characterized by new topological invariants. We hope this work will enlarge the family of topological phases in band insulators and stimulate the search for them in real materials.
[3] M.Z. Hasan, C.L. Kane, Rev. Mod. Phys. 82(2010) 3045-3067.

[Cited within: 1]     

[4] X.L. Qi, S.C. Zhang, Rev. Mod. Phys. 83(2011) 1057.

[Cited within: 1]     

[5] P. Dziawa, B.J. Kowalski, K. Dybko, R. Buczko, A. Szczerbakow, M. Szot, E. Lusakowska, T. Balasubramanian, B.M. Wojek, M.H. Berntsen, O. Tjernberg, Nat. Mater. 11(2012) 1023-1027.

DOI      URL      PMID      [Cited within: 1]      Abstract

Topological insulators are a class of quantum materials in which time-reversal symmetry, relativistic effects and an inverted band structure result in the occurrence of electronic metallic states on the surfaces of insulating bulk crystals. These helical states exhibit a Dirac-like energy dispersion across the bulk bandgap, and they are topologically protected. Recent theoretical results have suggested the existence of topological crystalline insulators (TCIs), a class of topological insulators in which crystalline symmetry replaces the role of time-reversal symmetry in ensuring topological protection. In this study we show that the narrow-gap semiconductor Pb(1-x)Sn(x)Se is a TCI for x  =  0.23. Temperature-dependent angle-resolved photoelectron spectroscopy demonstrates that the material undergoes a temperature-driven topological phase transition from a trivial insulator to a TCI. These experimental findings add a new class to the family of topological insulators, and we anticipate that they will lead to a considerable body of further research as well as detailed studies of topological phase transitions.
[6] Y. Tanaka, Z. Ren, T. Sato, K. Nakayama, S. Souma, T. Takahashi, K. Segawa, Y. Ando, Nat. Phys. 8(2012) 800-803.

DOI      URL      [Cited within: 1]     

[7] Y. Okada, M. Serbyn, H. Lin, D. Walkup, W. Zhou, C. Dhital, M. Neupane, S. Xu, Y.J. Wang, R. Sankar, F. Chou, A. Bansil, M.Z. Hasan, S.D. Wilson, L. Fu, V. Madhavan, Science 341 (2013) 1496-1499.

DOI      URL      PMID      [Cited within: 1]      Abstract

In topological crystalline insulators (TCIs), topology and crystal symmetry intertwine to create surface states with distinct characteristics. The breaking of crystal symmetry in TCIs is predicted to impart mass to the massless Dirac fermions. Here, we report high-resolution scanning tunneling microscopy studies of a TCI, Pb(1-x)Sn(x)Se that reveal the coexistence of zero-mass Dirac fermions protected by crystal symmetry with massive Dirac fermions consistent with crystal symmetry breaking. In addition, we show two distinct regimes of the Fermi surface topology separated by a Van-Hove singularity at the Lifshitz transition point. Our work paves the way for engineering the Dirac band gap and realizing interaction-driven topological quantum phenomena in TCIs.
[8] C.H. Yan, J.W. Liu, J.F. Wang, Z.Y. Wang, P. Wang, Z.D. Zhang, L.L. Wang, X.C. Ma, S.H. Ji, K. He, L. Fu, W.H. Duan, Q.K. Xue, X. Chen, Phys. Rev. Lett. 112(2014) 186801.

DOI      URL      PMID      [Cited within: 1]      Abstract

The surface of a topological crystalline insulator (TCI) carries an even number of Dirac cones protected by crystalline symmetry. We epitaxially grew high-quality Pb(1-x)Sn(x)Te(111) films and investigated the TCI phase by in situ angle-resolved photoemission spectroscopy. Pb(1-x)Sn(x)Te(111) films undergo a topological phase transition from a trivial insulator to TCI via increasing the Sn/Pb ratio, accompanied by a crossover from n-type to p-type doping. In addition, a hybridization gap is opened in the surface states when the thickness of the film is reduced to the two-dimensional limit. The work demonstrates an approach to manipulating the topological properties of TCI, which is of importance for future fundamental research and applications based on TCI.
[9] B.A. Assaf, F. Katmis, P. Wei, B. Satpati, Z. Zhang, S.P. Bennett, V.G. Harris, J.S. Moodera, D. Heiman,Appl. Phys. Lett. 105(2014), 102108.

DOI      URL      [Cited within: 1]     

[10] C. Zhang, Y.W. Liu, X. Yuan, W.Y. Wang, S.H. Liang, F.X. Xiu, Nano Lett. 15(2014) 2161-2167.

DOI      URL      PMID      [Cited within: 2]      Abstract

Recently, rock-salt IV-VI semiconductors, such as Pb(1-x)Sn(x)Se(Te) and SnTe, have been observed to host topological crystalline insulator (TCI) states. The nontrivial states have long been believed to exhibit ambipolar field effects and possess massive Dirac Fermions in two-dimension (2D) limit due to the surface hybridization. However, these exciting attributes of TCI remain previously inaccessible owing to the complicated control over composition and thickness. Here, we systematically investigate doping and thickness-induced topological phase transitions by electrical transport. We demonstrate the first evidence of the ambipolar properties in Pb(1-x)Sn(x)Se thin films. Surface gap opening is observed in 10 nm TCI originated from the strong finite-size effect. Importantly, magnetoconductance hosts a competition between weak antilocalization and weak localization, suggesting a strikingly tunable Berry phase evolution and strong electron-electron interaction. Our findings serve as a new probe to study electron behavior and pave the way for further exploring and manipulating this novel 2D TCI phase.
[11] R. Akiyama, K. Fujisawa, T. Yamaguchi, R. Ishikawa, S. Kuroda, Nano Res. 9(2016) 490-498.

[Cited within: 3]     

[12] C.H. Yan, H. Guo, J. Wen, Z.D. Zhang, L.L. Wang, K. He, X.C. Ma, S.H. Ji, X. Chen, Q.K. Xue, Surf. Sci. 621(2014) 104-108.

[Cited within: 1]     

[13] H. Guo, C.H. Yan, J.W. Liu, Z.Y. Wang, R. Wu, Z.D. Zhang, L.L. Wang, K. He, X.C. Ma, S.H. Ji, W.H. Duan, X. Chen, Q.K. Xue,APL Mater. 2(2014), 056106.

[Cited within: 2]     

[14] A.A. Taskin, F. Yang, S. Sasaki, K. Segawa, Y. Ando,Phys. Rev. B 89 (2014) ,121302.

[Cited within: 3]     

[15] H. Ozawa, A. Yamakage, M. Sato, Y. Tanaka,Phys. Rev. B 90 (2014), 045309.

[Cited within: 1]     

[16] Y. Zhang, K. He, C.Z. Chang, C.L. Song, L.L. Wang, X. Chen, J.F. Jia, Z. Fang, X. Dai, W.Y. Shan, Nat. Phys. 6(2010) 584.

[Cited within: 1]     

[17] C.X. Liu, H. Zhang, B. Yan, X.L. Qi, T. Frauenheim, X. Dai, Z. Fang, S.C. Zhang,Phys. Rev. B 81 (2010), 041307.

[Cited within: 1]     

[18] Z.Y. Wang, J.F. Wang, Y.Y. Zang, Q.H. Zhang, J.A. Shi, T. Jiang, G. Yan, C.L. Song, S.H. Ji, L.L. Wang, L. Gu, K. He, W.H. Duan, X.C. Ma, X. Chen, Q.K. Xue, Adv. Mater. 27(2015) 4150.

DOI      URL      PMID      [Cited within: 1]     

[19] X. Chen, S.N. Choing, D.J. Aschaffenburg, C.D. Pemmaraju, D. Prendergast, T. Cuk, J. Am. Chem. Soc. 139(2017) 1830-1841.

DOI      URL      PMID      [Cited within: 1]      Abstract

3, these one-electron intermediates are radicals located in Ti-O• (oxyl) and Ti-O•-Ti (bridge) groups arranged perpendicular and parallel to the surface respectively, and form electronic states in the band gap of SrTiO3. Using an ultrafast sub band gap probe of 400 nm and white light, we excited transitions between these radical states and the conduction band. By measuring the time evolution of surface reflectivity following the pump pulse of 266 nm light, we determined an initial radical formation time of 1.3 ± 0.2 ps, which is identical to the time to populate the surface with titanium oxyl (Ti-O•) radicals. The oxyl was separately observed by a subsurface vibration near 800 cm-1 from Ti-O located in the plane right below Ti-O•. Second, a polarized transition optical dipole allows us to assign the 1.3 ps time constant to the production of both O-site radicals. After a 4.5 ps delay, another distinct surface species forms with a time constant of 36 ± 10 ps with a yet undetermined structure. As would be expected, the radicals' decay, specifically probed by the oxyl's subsurface vibration, parallels that of the photocurrent. Our results led us to propose a nonadiabatic kinetic mechanism for generating radicals of the type Ti-O• and Ti-O•-Ti from valence band holes based on their solvation at aqueous interfaces.]]>
[20] N. Wang, D.M. West, J.W. Liu, J. Li, Q.M. Yan, B.L. Gu, S.B. Zhang,Phys. Rev. B 89 (2014), 045142.

[Cited within: 1]     

[21] J. Wang, A.M. DaSilva, C.Z. Chang, K. He, J.K. Jain, N. Samarth, X.C. Ma, Q.K. Xue, M.H.W. Chan,Phys. Rev. B 83 (2011), 245438.

[Cited within: 1]     

[22] S. Hikami, A.I. Larkin, Y. Nagaoka, Prog. Theor. Phys. Suppl. 63(1980) 707-710.

[Cited within: 1]     

[23] E. McCann, K. Kechedzhi, V.I. Falko, H. Suzuura, T. Ando, B.L. Altshuler,Phys. Rev. Lett. 97(2006), 146805.

DOI      URL      PMID      [Cited within: 1]      Abstract

Because of the chiral nature of electrons in a monolayer of graphite (graphene) one can expect weak antilocalization and a positive weak-field magnetoresistance in it. However, trigonal warping (which breaks p-->-p symmetry of the Fermi line in each valley) suppresses antilocalization, while intervalley scattering due to atomically sharp scatterers in a realistic graphene sheet or by edges in a narrow wire tends to restore conventional negative magnetoresistance. We show this by evaluating the dependence of the magnetoresistance of graphene on relaxation rates associated with various possible ways of breaking a "hidden" valley symmetry of the system.
[24] H.Z. Lu, J.R. Shi, S.Q. Shen,Phys. Rev. Lett. 107(2011), 076801.

DOI      URL      PMID      [Cited within: 1]      Abstract

A magnetoconductivity formula is presented for the surface states of a magnetically doped topological insulator. It reveals a competing effect of weak localization and weak antilocalization in quantum transport when an energy gap is opened at the Dirac point by magnetic doping. It is found that, while random magnetic scattering always drives the system from the symplectic to the unitary class, the gap could induce a crossover from weak antilocalization to weak localization, tunable by the Fermi energy or the gap. This crossover presents a unique feature characterizing the surface states of a topological insulator with the gap opened at the Dirac point in the quantum diffusion regime.
[25] Z.H. Wang, X.P.A. Gao, Z.D. Zhang,Chin. Phys.B 27 (2018), 107901.

[Cited within: 1]     

[26] P.W. Anderson, E. Abrahams, T.V. Ramakrishnan, Phys. Rev. Lett. 43(1979) 718-720.

[Cited within: 1]     

[27] E. Abrahams, P.W. Anderson, P.A. Lee, T.V. Ramakrishnan, Phys. Rev. B 24 (1981) 6783-6789.

[Cited within: 1]     

/