This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

\UseRawInputEncoding
thanks: These authors contributed equally.thanks: These authors contributed equally.thanks: These authors contributed equally.

Room-Temperature Topological Phase Transition in Quasi-One-Dimensional Material Bi4I4

Jianwei Huang Department of Physics and Astronomy, Rice University, Houston, TX 77005, USA    Sheng Li Department of Physics, The University of Texas at Dallas, 800 West Campbell Road, Richardson, Texas 75080-3021, USA    Chiho Yoon Department of Physics, The University of Texas at Dallas, 800 West Campbell Road, Richardson, Texas 75080-3021, USA Department of Physics and Astronomy, Seoul National University, Seoul 08826, Korea    Ji Seop Oh Department of Physics, University of California, Berkeley, Berkeley, California 94720, USA Department of Physics and Astronomy, Rice University, Houston, TX 77005, USA    Han Wu Department of Physics and Astronomy, Rice University, Houston, TX 77005, USA    Xiaoyuan Liu Department of Physics, The University of Texas at Dallas, 800 West Campbell Road, Richardson, Texas 75080-3021, USA    Nikhil Dhale Department of Physics, The University of Texas at Dallas, 800 West Campbell Road, Richardson, Texas 75080-3021, USA    Yan-Feng Zhou Department of Physics, The University of Texas at Dallas, 800 West Campbell Road, Richardson, Texas 75080-3021, USA    Yucheng Guo Department of Physics and Astronomy, Rice University, Houston, TX 77005, USA    Yichen Zhang Department of Physics and Astronomy, Rice University, Houston, TX 77005, USA    Makoto Hashimoto Stanford Synchrotron Radiation Lightsource, SLAC National Accelerator Laboratory, Menlo Park, California 94025, USA    Donghui Lu Stanford Synchrotron Radiation Lightsource, SLAC National Accelerator Laboratory, Menlo Park, California 94025, USA    Jonathan Denlinger Advanced Light Source, Lawrence Berkeley National Laboratory, Berkeley, California 94720, USA    Xiqu Wang Department of Chemistry, University of Houston, Houston, Texas 77204 USA    Chun Ning Lau Department of Physics, The Ohio State University, Columbus, OH 43210, USA    Robert J. Birgeneau [email protected] Department of Physics, University of California, Berkeley, Berkeley, California 94720, USA Materials Science Division, Lawrence Berkeley National Laboratory, Berkeley, California 94720, USA    Fan Zhang [email protected] Department of Physics, The University of Texas at Dallas, 800 West Campbell Road, Richardson, Texas 75080-3021, USA    Bing Lv [email protected] Department of Physics, The University of Texas at Dallas, 800 West Campbell Road, Richardson, Texas 75080-3021, USA    Ming Yi [email protected] Department of Physics and Astronomy, Rice University, Houston, TX 77005, USA
Abstract

Quasi-one-dimensional (1D) materials provide a superior platform for characterizing and tuning topological phases for two reasons: i) existence for multiple cleavable surfaces that enables better experimental identification of topological classification, and ii) stronger response to perturbations such as strain for tuning topological phases compared to higher dimensional crystal structures. In this paper, we present experimental evidence for a room-temperature topological phase transition in the quasi-1D material Bi4I4, mediated via a first order structural transition between two distinct stacking orders of the weakly-coupled chains. Using high resolution angle-resolved photoemission spectroscopy on the two natural cleavable surfaces, we identify the high temperature β\beta phase to be the first weak topological insulator with gapless Dirac cones on the (100) surface and no Dirac crossing on the (001) surface, while in the low temperature α\alpha phase, the topological surface state on the (100) surface opens a gap, consistent with a recent theoretical prediction of a higher-order topological insulator beyond the scope of the established topological materials databases that hosts gapless hinge states. Our results not only identify a rare topological phase transition between first-order and second-order topological insulators but also establish a novel quasi-1D material platform for exploring unprecedented physics.

I INTRODUCTION

Refer to caption
Figure 1: Crystal structure and sample characterization. (a) Temperature dependent resistivity data for both cooling and warming curves. The inset is the enlarged view showing the hysteresis behavior of the first order transition. (b) Images of a Bi4I4 crystal showing the cleavable (001) and (100) surfaces where the grid is 1 mm. (c) and (f) The real space crystal structure and the reciprocal space lattice from the h0l layer precession images integrated from X-ray diffraction data for the α\alpha-Bi4I4 phase. The shaded region is the structural unit cell. (d) and (g) The unit cell volume and refined lattice parameter aa from X-ray diffraction reflecting the α\alpha to β\beta structural transition. (e,h) Same as (c,f) but for the β\beta-Bi4I4 phase. (i) and (j) The bulk and projected surface Brillouin zones (BZ) of α\alpha-Bi4I4 and β\beta-Bi4I4.

Topological phases and associated phase transitions in quantum materials have garnered tremendous interest in recent years since the discovery of the two-dimensional (2D) quantum spin Hall (QSH) effect [1, 2]. Extending this concept to three dimensions (3D) led to the idea of topological insulator (TI), which hosts gapless boundary states. Within the class of TIs, there are two schemes by which the topological class can be further distinguished. The first scheme is the distinction between strong and weak TI [3, 4]. While every surface of a strong TI possesses a gapless Dirac cone, such nontrivial surface states only appear on selected surfaces of a weak TI [4, 5, 6, 7]. The experimental identification between a strong and a weak TI therefore depends critically on the measurement of topological non-trivial surface states on multiple surfaces of a material, commonly probed via the technique of angle-resolved photoemission spectroscopy (ARPES). The challenge of such experimental efforts is that in natural materials, usually only a single preferred cleavage surface is experimentally accessible for ARPES. Most TIs that have been experimentally confirmed have been strong TIs, identified by a combination of a single-surface measurement and theoretical calculations, such as in the bismuth and antimony chalcogenides [8, 9, 10, 11]. Direct experimental verification of a weak TI, however, remains scarce due to the lack of materials with multiple cleavage surfaces where the existence of topologically nontrivial surface states could be probed [7, 12, 13, 14, 15, 16]. The second scheme to classify further a TI is by its topological order, which is distinguished by the dimension of its gapless boundary states compared to the dimension of its bulk state. This realization led to the identification of a novel phase of matter dubbed the higher-order TI [17, 18, 19, 20], which are materials that host protected gapless states on boundaries of two or three spatial dimensions lower than its bulk.

Interestingly, materials predicted to be candidate weak TIs are often in the vicinity of a topological critical point, where the exact topological classification depends sensitively upon the lattice parameters [7, 21]. On the one hand, this produces an uncertainty in theoretical predictions of their topological properties, which makes experimental determination a necessity [7, 21]. On the other hand, this renders an advantage that a topological phase transition between different topological phases can be easily induced by external perturbations such as strain and thermal effects [21, 7, 22, 23].

As a quasi-1D material, Bi4I4 provides a superior platform not only to realize but also to be experimentally examined for the aforementioned distinct TI classes [7]. Consisting of stacked quasi-1D chains, Bi4I4 naturally stabilizes in two crystal structures due to distinct stacking sequences, identified as α\alpha-Bi4I4 and β\beta-Bi4I4 [24]. The α\alpha-phase is naturally stabilized at low temperatures and transitions into the β\beta-phase via a first-order phase transition around 300 K. β\beta-Bi4I4 was theoretically predicted to be topologically nontrivial, in the vicinity of a transition point between weak and strong TIs [7, 25]. Two subsequent ARPES studies reached conflicting conclusions [25, 26]. While one study reports strong TI characteristics for a low temperature phase identified as β\beta-Bi4I4 [25], the other utilized a quenching method to stabilize the β\beta-Bi4I4 phase at low temperatures and identified it as a weak TI [26] while categorized the α\alpha-Bi4I4 phase as a trivial insulator with a distinct electronic structure [26]. Besides the controversy amongst the experimental reports, these results also differ from a recent theoretical work predicting the α\alpha-phase to be a topologically nontrivial higher-order TI hosting helical hinge modes [20]. While beyond the scope of the recently established topological materials database [27, 28, 29], this predicted rare phase may be viewed as a QSH effect [1, 2] around the top or bottom surface of a 3D insulator, i.e., the time-reversal-invariant counterpart of the long-desired 3D quantum Hall effect [30, 31]. To resolve the aforementioned controversies, we systematically carried out electrical transport, X-ray diffraction, and ARPES measurements in combination with theoretical calculations to clarify the topological classification of α\alpha-Bi4I4 and β\beta-Bi4I4 as tuned by the natural structural transition via temperature.

Refer to caption
Figure 2: Electronic structure from different surfaces. (a) Fermi surface mapping of α\alpha-Bi4I4 on the (100)’ surface, with the surface BZ marked by gray rectangles (the solid one corresponds to α\alpha-Bi4I4 and the dashed one β\beta-Bi4I4). (b) Constant energy contour of α\alpha-Bi4I4 on the (001) surface. (e-h) Measured band dispersions of α\alpha-Bi4I4 along the cut #1 to #4 indicated in (a) and (b). Related surface states (SS) and bulk states (BS) are marked by arrows. (c-d) and (i-l) Same as the top row but for β\beta-Bi4I4. All band images are divided by the Fermi-Dirac function. (m) Band image along the cut #5 except taken at ky=0k_{y}=0. Cut #5 was drawn intentionally off-center to not block the Fermi surface image. The red dashed lines are sinusoidal fittings of the band image showing a periodicity of 0.3 Å-1 due to band doubling, which corresponds to the (100)’ surface. (n) Band image along the cut #6 indicated in (b). The black arrow shows a periodicity of 0.88 Å-1 which corresponds to the (001) surface. (o) Measured dispersions in α\alpha-Bi4I4 along high symmetry cuts. (p) Calculated bulk band structure of α\alpha-Bi4I4 projected onto the (001) surface (see Supplementary Materials). (q,r) Same as (o,p) but for β\beta-Bi4I4. A clear doubling of the bulk valence bands is evident from β\beta-Bi4I4 to α\alpha-Bi4I4 in both the measured and calculated dispersions. All measurement temperatures are as indicated.
Refer to caption
Figure 3: Bulk band doubling across the phase transition. (a)-(c) Warm-up and (d)-(f) cool-down band images measured along cut #1 indicated in Fig. 2a at selected temperatures. (g)-(h) Corresponding MDCs at -1.2 eV during warm-up and cool-down, respectively. (i) Warm-up and (j) cool-down false-color images showing a more detailed temperature evolution of the MDCs indicated in (a). (k) Example of an MDC fitted by two Lorentzian peaks. (l) Full-width-at-half-max of the MDC fittings indicated in (k) as a function of temperature. The warm-up results are shifted 0.04 Å upwards. (m) The separation of the peak positions of the MDC fittings indicated in (k) as a function of temperature.

II Structural Transition between the α\alpha- and β\beta-phases

Large Bi4I4 single crystals were synthesized with the quasi-1D chains oriented along the bb axis, with large flat (001) and (100) surfaces (Fig. 1) (see Supplementary Materials). The structural phase transition can be clearly identified from the temperature-dependent resistivity (Fig. 1a) and heat capacity (Supplemental Fig. S1) measurements, where a hysteresis indicating a first order phase transition appears around 300 K. Below 270 K, the resistivity data show a typical semiconductor behavior with a broad hump at 150 K and are consistent with previous reports [26].

Structurally, the α\alpha- and β\beta-phases differ in the packing of the Bi4I4 chains. The α\alpha-Bi4I4 has double-layered Bi4I4 chains per unit cell (Fig. 1c) whereas the β\beta-Bi4I4 has only a single-layer of Bi4I4 chains per unit cell (Fig. 1e). This subtle difference in stacking causes a large change of the cc lattice and the associated doubling of the unit cell volume (Fig. 1d), as demonstrated by X-ray single crystal diffraction (Fig. 1f, h). Contrasting the single crystal diffraction patterns for the reciprocal lattice aa^{\ast}-cc^{\ast} planes taken in the α\alpha-phase at 293 K (Fig. 1f) and the β\beta-phase at 313 K (Fig. 1h), a near-doubling of the cc^{\ast} in reciprocal space from α\alpha-Bi4I4 to β\beta-Bi4I4 is visible, as seen by the additional peaks in Fig. 1f marked by a circle. This structural change is well-captured in the temperature-dependent refined lattice parameters (Fig. 1g and Fig. S2). The lattice parameters of the α\alpha-phase are refined at 300 K to be a=14.251(2)Åa=14.251(2)\AA, b=4.4304(5)Åb=4.4304(5)\AA, c=19.976(6)Åc=19.976(6)\AA, and β=92.93(2)\beta=92.93(2)^{\circ}, while those of the β\beta-phase are refined at 310 K to be a=14.394(5)Åa=14.394(5)\AA, b=4.4288(13)Åb=4.4288(13)\AA, c=10.494(5)Åc=10.494(5)\AA, and β=107.96(5)\beta=107.96(5)^{\circ}.

The large facets on both the (001) (>> 0.5 mm) and (100) (>> 0.2 mm) surfaces are naturally cleavable, exposing flat surfaces that enable conventional ARPES studies on either surface using an incident beam spot size of 50 × 20 μ\mum2 (Fig. 1b). We note the β\beta-angle difference between α\alpha-Bi4I4 and β\beta-Bi4I4 crystal structures, where the natural cleavable side (100) surface of β\beta-Bi4I4 becomes the (2¯\bar{2}01) surface in the α\alpha-phase per definition in the true primitive unit cell (Fig. 1c, e), we label it as (100)’ in α\alpha-Bi4I4 (Fig. 1i, j) to avoid confusion in the comparison with the (100) surface of the β\beta-phase.

Taking advantage of the rare multi-surface experimental access to Bi4I4, we carried out ARPES measurements on both the (001) and (100) surfaces of the α\alpha- and β\beta-phases in their respective temperature regimes. We note that such two-surface ARPES measurement has also been desirable to identify unambiguously the topological characters of weak TIs and higher-order TIs. A comparison of the complete electronic structure of the two phases are summarized in Fig. 2. Two main observations can be made: i) there is a strong distinction between the electronic structures measured on the (100) and (001) surfaces for both phases, and ii) the change between the two phases across the structural transition is subtle.

III Distinguishing the (001) and (100) Surfaces

We first examine the distinction between the (100) and (001) surfaces, which is most clearly seen in the measured Fermi surfaces (FS). On the (100) surface, both phases exhibit a quasi-1D FS along kzk_{z} (Fig. 2a, c), revealing a rather weak interlayer coupling in the cc-direction. The FS on the (001) surface, by contrast, consists of island-like bright spots for both phases (Fig. 2b, d and Fig. S9), suggesting a stronger interlayer coupling along the aa axis compared to that of the cc axis. Furthermore, the measured sample surface orientation can be unambiguously distinguished by the periodicity of the observed band dispersions, which corresponds to the clearly distinct lattice parameters associated with the (100) and (001) surfaces. The α\alpha-phase, for example, has a period of 0.3 Å-1 along the kzk_{z} direction on the (100)’ surface (Fig. 2m) and 0.88 Å-1 along the kxk_{x} direction on the (001) surface (Fig. 2n). We note that we do not observe a drastic change of the fermiology across the phase transition, in contrast to the previous report that the quasi-1D FS only exists in the β\beta-phase [26]. By clearly identifying single domains of each cleaved surface, we now clarify that the quasi-1D intensity and island-like intensity are due to the two different crystal surfaces measured rather than the two distinct structural phases.

We can furthermore examine the corresponding low-energy band dispersions. At the high symmetry points of both surfaces in both phases, an intense hole-like bulk band is observed below EFE_{F}, revealing a gap in the bulk state in all cases. The bulk nature of this band is confirmed by photon energy dependence as it disperses along kzk_{z} (Fig. S6). However, two observations can be noted. Firstly, as the electronic structure is quasi-1D along kzk_{z} on the (100) surfaces, the bands near the Γ¯\bar{{\Gamma}} and Z¯\bar{Z} points are quite similar in both phases (Fig. 2e,f,i,j) while those near the Γ¯\bar{{\Gamma}} and M¯\bar{M} points on the (001) surface are notably distinct. Secondly, while we observe a hole-like bulk valence band on all the cuts shown, within the bulk band gap on the (100) surface, a sharp Dirac surface state is observed for both the α\alpha- and β\beta-phases (Fig. 2e,f,i,j). On the contrary, no Dirac surface state is observed inside the bulk band gap on the (001) surface of either phase (Fig. 2g,h,k,l). At the Γ¯\bar{\Gamma}-point, the bulk valence band tops at -0.3 eV. No surface state is observed. At the M¯\bar{M}-point, the bulk valence band approaches EFE_{F}, with a visible gap near EFE_{F} between the bulk valence and conduction bands.

IV Electronic Reconstruction between the α\alpha and β\beta Phases

Having distinguished the two surfaces, we now focus on the transition between the two structural phases. Overall, no dramatic change of the bulk band dispersions is observed across the phase transition. There is, however, a doubling in the number of bands from the β\beta-phase to the α\alpha-phase [20] due to the doubling of the stacking layers in the unit cell. This band doubling is well-demonstrated by our theoretical calculations of the bulk bands projected onto the (001) surface for both the α\alpha and β\beta phases (Fig. 2p, r and Supplementary Materials). While the band splitting is small due to the weak interlayer coupling, it is clearly resolved in our corresponding dispersions measured at low temperatures in the α\alpha-phase (see pairs of arrows in Fig. 2o,q as well as in a larger energy window shown in Fig. S3), consistent with the doubled stacking in the low temperature phase.

Next, we demonstrate that this band doubling is indeed induced via the structural transition by systematically measuring across the structural transition near room temperature. We note that while it is challenging to resolved the small band splitting near the valence band top at elevated temperatures, this band doubling manifests as a broadening in the momentum distribution curves (MDCs) when a single band splits into two from the higher temperature β\beta-phase to the lower temperature α\alpha-phase. Clearly, this effect is in contrast to thermal broadening effects, which would have the opposite trend with respective to temperature. In Fig. 3a-c, we plot the dispersions across the Γ¯\bar{{\Gamma}} point measured on (100) at selected temperatures. A clear abrupt broadening is indeed observed from the high temperature β\beta-phase to the lower temperature α\alpha-phase as exemplified by the MDCs at -1.2 eV (Fig. 3i,j). Importantly, this abrupt change is accompanied by a hysteresis behavior revealed in the warm up (Fig. 3i) and cool down (Fig. 3j) processes. To quantify this behavior, we fit the MDCs and plot both their full-width-at-half-max (Fig. 3l) and peak positions (Fig. 3m), both of which show the hysteresis behavior reminiscent of that in the resistivity measurement, confirming the electronic structure’s response to the first order structural transition.

Refer to caption
Figure 4: Surface states evolution across the phase transition. (a) Band image along cut #4 on the (001) surface indicated in Fig. 2b of α\alpha-Bi4I4 as well as its corresponding EDC stacks. The red EDC corresponds to k = 0. The green dots track the peak positions of the bulk band which reveal a gap opening of \sim85 meV. (b) Same as (a) but of β\beta-Bi4I4. A gap opens at k = 0 of \sim100 meV. (c) Band image along cut #1 on the (100)’ surface indicated in Fig. 2a of α\alpha-Bi4I4 as well as its corresponding EDC stacks. The blue dots track the surface state which reveal a \sim35 meV gap opening. (d) Same as (c) but of β\beta-Bi4I4. A gapless Dirac surface state is observed. (e) 3D view of the electronic structure on the (100)’ surface of α\alpha-Bi4I4. A quasi-1D Dirac-like surface state is observed. (f) Same as (e) but of β\beta-Bi4I4. The quasi-1D Dirac surface state reveals the weak TI property. (g) Schematics (see Supplementary Materials) showing the higher-order TI state of α\alpha-Bi4I4 with a particular surface termination. Upper panel: all the bulk states and surface states are gapped, whereas a helical hinge state around the top surface exists in the surface state gap. Lower panel: the zoom-in (100)’ surface state features a band doubling and a small gap. (h) Schematics (see Supplementary Materials) showing the weak TI state of β\beta-Bi4I4. Upper panel: the bulk states and (001) surface state are gapped, whereas there is a quasi-1D gapless Dirac surface state on the (100) surface. Lower panel: the zoom-in (100) surface state features two gapless Dirac cones.

V Topological Characterization

Having confirmed the phase transition as observed from the electronic structure, we examine the topological properties of the two phases. In Fig. 4 we present high resolution measurements near EFE_{F} for both phases. We first examine the β\beta-phase. On the (001) surface (Fig. 2k, l and Fig. 4b), we observe a clear gap in all the bulk gaps and no additional surface Dirac crossings. This can be confirmed from the energy distribution curve (EDC) stacks of the bands across M¯\bar{M} (Fig. 4b), where the green dots track the bulk band dispersion and reveal a gap of \sim100 meV. Similar suppressed intensity can also be clearly identified in the MDC stacks (Fig. S7). On the (100) surface, a quasi-1D Dirac surface state exists inside the bulk band gap (Fig. 2i, j and Fig. 4d, f). Particularly at the Γ¯\bar{\Gamma} and Z¯\bar{Z} points, the surface states are gapless within the resolution of measurements at 350 K. These combined measurements on both the (001) and (100) surfaces demonstrate that the high temperature β\beta-Bi4I4 is a weak TI where only selected surfaces possess an even number (two here) of gapless Dirac cones [3, 7] (Fig. 4h). In the α\alpha-phase, all the observed bands on the (001) surface are also gapped (Fig. 2g, h and Fig. 4a). A gap of \sim85 meV can be further seen from the bands tracked by the EDC stacks in Fig. 4a. On the (100)’ surface, the quasi-1D Dirac-like surface state still persists (Fig. 2e, f). From a high-resolution measurement taken with 30 eV photons at 10 K (Fig. 4c), we observe a gap of \sim35 meV clearly in the EDC stacks, indicating a gapped surface state within a bulk band gap. This is consistent with our theoretical prediction and calculation of a higher-order TI scenario of α\alpha-Bi4I4 [20], illustrated in Fig. 4g, where both the bulk bands and surfaces bands are required to be gapped, hosting gapless hinge states inside the surface gap. A comparison of our calculated bulk and surface states with our measured bands show strong agreement (Fig. S11). We summarize our experimental findings and classification of the two phases in Table 1.

Table 1: Summary of the key features on α\alpha-Bi4I4 and β\beta-Bi4I4 revealed by our ARPES study
Structure Temperature Bulk bands (001) surface (100) surface Topology
α\alpha-Bi4I4 <<295 K 8N \sim85 meV (M¯\bar{M}) \sim35 meV (Γ¯\bar{\Gamma}) Higher-order TI
β\beta-Bi4I4 >>305 K 4N \sim100 meV (M¯\bar{M}) Gapless Weak TI

Finally, we comment on the inconsistencies in the previous two ARPES studies on Bi4I4. In the first report [25], since no quenching was performed, the phase measured at low temperatures was most likely the α\alpha-phase instead of the claimed β\beta-phase. The conclusion that this phase is a strong TI was based on the observation of a gapless Dirac crossing on the (001) surface. Here we show that this Dirac band is rather a bulk band, where high resolution measurements reveal a gap (Fig. 4a). In the second report with nanoARPES [26], the β\beta-phase was claimed to be stabilized at low temperature by quenching directly from high temperature. The authors ascribed point-like FS to α\alpha-Bi4I4 and quasi-1D FS to β\beta-Bi4I4 based on a superposition of signals from the (001) and (100) surfaces. Based on this assumption, they concluded that the β\beta-phase is a weak TI and the α\alpha-phase is a trivial insulator. Benefiting from the relatively large cleavable (100) and (001) surfaces free of domains, here we unambiguously show that both phases have similar quasi-1D FS on the (100) side surface, and that the distinction between the two phases is the subtle band doubling in the α\alpha-phase, which was not resolved in previous studies. Reminiscent of the bulk Peierls distortion in the Su-Schrieffer-Heeger model [32], this yields a gap opening in the Dirac surface state on the (100)’ surface, which is the prerequisite for the appearance of mid-gap hinge states in α\alpha-Bi4I4 as a higher-order TI [20].

VI Discussions

Topological properties of a variety of quantum materials have been verified by a combination of ARPES and first-principles calculations during the last decades [10, 11, 33, 34, 35, 36, 37, 38, 39, 40]. For the quasi-1D material Bi4I4, not only does the uncertainty of the theoretical prediction for the topological classification of both α\alpha-Bi4I4 and β\beta-Bi4I4 call for experimental exploration, but also the change of the electronic structures between β\beta-Bi4I4 and α\alpha-Bi4I4 is subtle such that careful ARPES measurements with considerable energy resolution is necessary. Unambiguously, the β\beta-phase Bi4I4 is shown to be the first weak TI by our ARPES measurements on large natural cleavable planes of the top and side surfaces, which resolves the controversy of the previous reports [25, 26]. For α\alpha-Bi4I4, two distinctions are clearly observed as compared with β\beta-Bi4I4: band doubling and gap opening in the side surface state. Together with the identification of the β\beta-phase as a weak TI, the existence of the Dirac-like surface state with a small gap indicates that the α\alpha-phase is consistent with the identification as a rare higher-order TI [20] instead of the trivial insulator based on symmetry indicators alone [27, 28, 29]. Since the β\beta-phase has been shown as a weak TI here, each (001) monolayer is a 2D Z2 TI [20]. As the α\alpha-phase has been shown to have two monolayers per unit cell here, i.e., a 2D trivial insulator, the 3D bulk must be Z2 trivial given the weak interlayer couplings [3, 7]; in other words, the surface states are gapped. Moreover, for an α\alpha-phase thin film with a large odd number of monolayers, it can be viewed as a 2D Z2 TI with a helical edge state. Apparently, this edge state can only appear at hinges and form a loop, demonstrating that the α-phase is a higher-order TI [20]. However, we do note that further direct evidence of the hinge states is desirable.

Fundamentally, our result implies that even without strong correlation and magnetism, there are still many topological materials that are beyond the scope of the already efficacious topological quantum chemistry or symmetry indicators [27, 28, 29]. Significantly, our work reveals the strong tunability of the quasi-1D TI Bi4I4 as a novel material platform where the abundance of topological phases could be explored. In the future, it would be exciting to investigate in this platform, for example, the strain-induced topological phases and their transitions, the spatial locations and Luttinger parameters of helical hinge states, the possible intrinsic and extrinsic topological superconductivity, as well as applications of our identified room-temperature topological phase transition as a topological switch that controls the dimensions of the boundary conduction channels.

Acknowledgement This work is mainly supported by National Science Foundation (NSF) through the DMREF program. This research used resources of the Stanford Synchrotron Radiation Lightsource and the Advanced Light Source, both U.S. DOE Office of Science User Facilities under contract Nos. AC02-76SF00515 and DE-AC02-05CH112319, respectively. We acknowledge the Texas Advanced Computing Center (TACC) for providing resources that have contributed to the research results reported in this work. The work at Rice is supported by NSF under Grant No. DMR-1921847, the Gordon and Betty Moore Foundation’s EPiQS Initiative through grant no. GBMF9470, the Robert A. Welch Foundation under Grant No. C-2024, and the Sloan Foundation FG-2019-12224. The work at UCB is supported by NSF Grant No. DMR-1921798. The work at UTD is supported by NSF under Grant Nos. DMR-1921581 and DMR-1945351, AFOSR under Grant No. FA9550-19-1-0037, and Army Research Office (ARO) under Grant No. W911NF-18-1-0416. The work at OSU is supported by NSF under Grant No. DMR-1922076.

References

  • Kane and Mele [2005] C. L. Kane and E. J. Mele, Z2 Topological Order and the Quantum Spin Hall Effect, Physical Review Letters 95, 146802 (2005).
  • Konig et al. [2007] M. Konig, S. Wiedmann, C. Brune, A. Roth, H. Buhmann, L. W. Molenkamp, X.-L. Qi, and S.-C. Zhang, Quantum Spin Hall Insulator State in HgTe Quantum Wells, Science 318, 766 (2007).
  • Fu et al. [2007] L. Fu, C. L. Kane, and E. J. Mele, Topological Insulators in Three Dimensions, Physical Review Letters 98, 106803 (2007).
  • Moore and Balents [2007] J. E. Moore and L. Balents, Topological invariants of time-reversal-invariant band structures, Physical Review B 75, 121306 (2007).
  • Hasan and Kane [2010] M. Z. Hasan and C. L. Kane, Colloquium : Topological insulators, Reviews of Modern Physics 82, 3045 (2010).
  • Qi and Zhang [2011] X.-L. Qi and S.-C. Zhang, Topological insulators and superconductors, Reviews of Modern Physics 83, 1057 (2011).
  • Liu et al. [2016] C.-C. Liu, J.-J. Zhou, Y. Yao, and F. Zhang, Weak Topological Insulators and Composite Weyl Semimetals: β\beta-Bi4X4 (X=Br, I), Physical Review Letters 116, 066801 (2016).
  • Xia et al. [2009] Y. Xia, D. Qian, D. Hsieh, L. Wray, A. Pal, H. Lin, A. Bansil, D. Grauer, Y. S. Hor, R. J. Cava, and M. Z. Hasan, Observation of a large-gap topological-insulator class with a single Dirac cone on the surface, Nature Physics 5, 398 (2009).
  • Zhang et al. [2009] H. Zhang, C.-X. Liu, X.-L. Qi, X. Dai, Z. Fang, and S.-C. Zhang, Topological insulators in Bi2Se3, Bi2Te3 and Sb2Te3 with a single Dirac cone on the surface, Nature Physics 5, 438 (2009).
  • Chen et al. [2009] Y. L. Chen, J. G. Analytis, J.-H. Chu, Z. K. Liu, S.-K. Mo, X. L. Qi, H. J. Zhang, D. H. Lu, X. Dai, Z. Fang, S. C. Zhang, I. R. Fisher, Z. Hussain, and Z.-X. Shen, Experimental Realization of a Three-Dimensional Topological Insulator, Bi2Te3Science 325, 178 (2009).
  • Hsieh et al. [2009] D. Hsieh, Y. Xia, D. Qian, L. Wray, J. H. Dil, F. Meier, J. Osterwalder, L. Patthey, J. G. Checkelsky, N. P. Ong, A. V. Fedorov, H. Lin, A. Bansil, D. Grauer, Y. S. Hor, R. J. Cava, and M. Z. Hasan, A tunable topological insulator in the spin helical Dirac transport regime, Nature 460, 1101 (2009).
  • Yan et al. [2012] B. Yan, L. Müchler, and C. Felser, Prediction of Weak Topological Insulators in Layered Semiconductors, Physical Review Letters 109, 116406 (2012).
  • Rasche et al. [2013] B. Rasche, A. Isaeva, M. Ruck, S. Borisenko, V. Zabolotnyy, B. Büchner, K. Koepernik, C. Ortix, M. Richter, and J. van den Brink, Stacked topological insulator built from bismuth-based graphene sheet analogues, Nature Materials 12, 422 (2013).
  • Tang et al. [2014] P. Tang, B. Yan, W. Cao, S.-C. Wu, C. Felser, and W. Duan, Weak topological insulators induced by the interlayer coupling: A first-principles study of stacked Bi2TeI, Physical Review B 89, 041409 (2014).
  • Yang et al. [2014] G. Yang, J. Liu, L. Fu, W. Duan, and C. Liu, Weak topological insulators in PbTe/SnTe superlattices, Physical Review B 89, 085312 (2014).
  • Li et al. [2015] X. Li, F. Zhang, Q. Niu, and J. Feng, Superlattice valley engineering for designer topological insulators, Scientific Reports 4, 6397 (2015).
  • Zhang et al. [2013] F. Zhang, C. L. Kane, and E. J. Mele, Surface State Magnetization and Chiral Edge States on Topological Insulators, Physical Review Letters 110, 046404 (2013).
  • Po et al. [2017] H. C. Po, A. Vishwanath, and H. Watanabe, Symmetry-based indicators of band topology in the 230 space groups, Nature Communications 8, 50 (2017).
  • Schindler et al. [2018] F. Schindler, A. M. Cook, M. G. Vergniory, Z. Wang, S. S. P. Parkin, B. A. Bernevig, and T. Neupert, Higher-order topological insulators, Science Advances 4, eaat0346 (2018).
  • Yoon et al. [2020] C. Yoon, C.-C. Liu, H. Min, and F. Zhang, Quasi-One-Dimensional Higher-Order Topological Insulators, arXiv 2005, 14710 (2020).
  • Weng et al. [2014] H. Weng, X. Dai, and Z. Fang, Transition-Metal Pentatelluride ZrTe5 and HfTe5: A paradigm for large-gap quantum spin hall insulators, Physical Review X 4, 011002 (2014).
  • Mutch et al. [2019] J. Mutch, W.-C. Chen, P. Went, T. Qian, I. Z. Wilson, A. Andreev, C.-C. Chen, and J.-H. Chu, Evidence for a strain-tuned topological phase transition in ZrTe5Science Advances 5, eaav9771 (2019).
  • Zhang et al. [2021] P. Zhang, R. Noguchi, K. Kuroda, C. Lin, K. Kawaguchi, K. Yaji, A. Harasawa, M. Lippmaa, S. Nie, H. Weng, V. Kandyba, A. Giampietri, A. Barinov, Q. Li, G. D. Gu, S. Shin, and T. Kondo, Observation and control of the weak topological insulator state in ZrTe5Nature Communications 12, 406 (2021).
  • von Schnering et al. [1978] H. G. von Schnering, H. von Benda, and C. Kalveram, Wismutmonojodid BiJ, eine Verbindung mit Bi(O) und Bi(II), Zeitschrift für anorganische und allgemeine Chemie 438, 37 (1978).
  • Autès et al. [2016] G. Autès, A. Isaeva, L. Moreschini, J. C. Johannsen, A. Pisoni, R. Mori, W. Zhang, T. G. Filatova, A. N. Kuznetsov, L. Forró, W. Van den Broek, Y. Kim, K. S. Kim, A. Lanzara, J. D. Denlinger, E. Rotenberg, A. Bostwick, M. Grioni, and O. V. Yazyev, A novel quasi-one-dimensional topological insulator in bismuth iodide β\beta-Bi4I4Nature Materials 15, 154 (2016).
  • Noguchi et al. [2019] R. Noguchi, T. Takahashi, K. Kuroda, M. Ochi, T. Shirasawa, M. Sakano, C. Bareille, M. Nakayama, M. D. Watson, K. Yaji, A. Harasawa, H. Iwasawa, P. Dudin, T. K. Kim, M. Hoesch, V. Kandyba, A. Giampietri, A. Barinov, S. Shin, R. Arita, T. Sasagawa, and T. Kondo, A weak topological insulator state in quasi-one-dimensional bismuth iodide, Nature 566, 518 (2019).
  • Zhang et al. [2019] T. Zhang, Y. Jiang, Z. Song, H. Huang, Y. He, Z. Fang, H. Weng, and C. Fang, Catalogue of topological electronic materials, Nature 566, 475 (2019).
  • Vergniory et al. [2019] M. G. Vergniory, L. Elcoro, C. Felser, N. Regnault, B. A. Bernevig, and Z. Wang, A complete catalogue of high-quality topological materials, Nature 566, 480 (2019).
  • Tang et al. [2019a] F. Tang, H. C. Po, A. Vishwanath, and X. Wan, Comprehensive search for topological materials using symmetry indicators, Nature 566, 486 (2019a).
  • Halperin [1987] B. I. Halperin, Possible states for a three-dimensional electron gas in a strong magnetic field, Japanese Journal of Applied Physics 26, 1913 (1987).
  • Tang et al. [2019b] F. Tang, Y. Ren, P. Wang, R. Zhong, J. Schneeloch, S. A. Yang, K. Yang, P. A. Lee, G. Gu, Z. Qiao, and L. Zhang, Three-dimensional quantum Hall effect and metal–insulator transition in ZrTe5Nature 569, 537 (2019b).
  • Su et al. [1979] W. P. Su, J. R. Schrieffer, and A. J. Heeger, Solitons in Polyacetylene, Physical Review Letters 42, 1698 (1979).
  • Tanaka et al. [2012] Y. Tanaka, Z. Ren, T. Sato, K. Nakayama, S. Souma, T. Takahashi, K. Segawa, and Y. Ando, Experimental realization of a topological crystalline insulator in SnTe, Nature Physics 8, 800 (2012).
  • Xu et al. [2012] S.-Y. Xu, C. Liu, N. Alidoust, M. Neupane, D. Qian, I. Belopolski, J. Denlinger, Y. Wang, H. Lin, L. Wray, G. Landolt, B. Slomski, J. Dil, A. Marcinkova, E. Morosan, Q. Gibson, R. Sankar, F. Chou, R. Cava, A. Bansil, and M. Hasan, Observation of a topological crystalline insulator phase and topological phase transition in Pb1-xSnxTe, Nature Communications 3, 1192 (2012).
  • Liu et al. [2014a] Z. K. Liu, J. Jiang, B. Zhou, Z. J. Wang, Y. Zhang, H. M. Weng, D. Prabhakaran, S.-K. Mo, H. Peng, P. Dudin, T. Kim, M. Hoesch, Z. Fang, X. Dai, Z. X. Shen, D. L. Feng, Z. Hussain, and Y. L. Chen, A stable three-dimensional topological Dirac semimetal Cd3As2Nature Materials 13, 677 (2014a).
  • Liu et al. [2014b] Z. K. Liu, B. Zhou, Y. Zhang, Z. J. Wang, H. M. Weng, D. Prabhakaran, S.-K. Mo, Z. X. Shen, Z. Fang, X. Dai, Z. Hussain, and Y. L. Chen, Discovery of a Three-Dimensional Topological Dirac Semimetal, Na3Bi, Science 343, 864 (2014b).
  • Xu et al. [2015] S.-Y. Xu, I. Belopolski, N. Alidoust, M. Neupane, G. Bian, C. Zhang, R. Sankar, G. Chang, Z. Yuan, C.-C. Lee, S.-M. Huang, H. Zheng, J. Ma, D. S. Sanchez, B. Wang, A. Bansil, F. Chou, P. P. Shibayev, H. Lin, S. Jia, and M. Z. Hasan, Discovery of a Weyl fermion semimetal and topological Fermi arcs, Science 349, 613 (2015).
  • Lv et al. [2015] B. Q. Lv, N. Xu, H. M. Weng, J. Z. Ma, P. Richard, X. C. Huang, L. X. Zhao, G. F. Chen, C. E. Matt, F. Bisti, V. N. Strocov, J. Mesot, Z. Fang, X. Dai, T. Qian, M. Shi, and H. Ding, Observation of Weyl nodes in TaAs, Nature Physics 11, 724 (2015).
  • Lv et al. [2019] B. Lv, T. Qian, and H. Ding, Angle-resolved photoemission spectroscopy and its application to topological materials, Nature Reviews Physics 1, 609 (2019).
  • Sobota et al. [2020] J. A. Sobota, Y. He, and Z.-X. Shen, Electronic structure of quantum materials studied by angle-resolved photoemission spectroscopy, arXiv 2008, 02378 (2020).