8
views
0
recommends
+1 Recommend
0 collections
    0
    shares
      • Record: found
      • Abstract: found
      • Article: found
      Is Open Access

      On the understanding of current-induced spin polarization of three-dimensional topological insulators

      letter

      Read this article at

      Bookmark
          There is no author summary for this article yet. Authors can add summaries to their articles on ScienceOpen to make them more accessible to a non-specialist audience.

          Abstract

          Arising from C. H. Li et al. Nature Communications 10.1038/ncomms13518 (2016). In a recent article 1 , Li et al. reported the current-induced spin polarization (CISP) on topological insulator (TI) Bi2Se3 and on InAs (100) samples by spin potentiometry and compared the sign of the measured signals using a theoretical model to conclude the origin of the observed CISP in their TIs. Spin potentiometry has been used to electrically measure CISP in spin-orbital coupled (SOC) materials, such as TIs 2–8 , and semiconductor-based two-dimensional electron gases (2DEG) 9–11,12 , where a ferromagnetic (FM) voltage probe with magnetization (M) collinear to the induced non-equilibrium spins is used to determine the CISP. The theoretical model by Li et al. consists of two key components: (a) a spin-dependent electrochemical potential diagram and (b) an argument that FM probe with magnetization along up (down) spin direction will measure the down (up) spin electrochemical potential. However, we bring attention to the inconsistencies in their paper. First, their (spatially varying) spin-dependent electrochemical potential diagram is incorrect and inconsistent with their topological surface states (TSS) band diagram to reflect the same experimental condition. The corrected potential diagram in conjunction with their model argument (b) is inconsistent with their assigned origin of CISP based on the sign of the measured signal. Second, they incorrectly stated that the experiment on (Bi,Sb)2Te3 reported by Lee et al. 7 gave the same sign as their measurements, whereas it is opposite. Finally, we point out that the comparison of the sign of the measured signal from TI with that of InAs may not be sufficient to draw a conclusion on the origin of CISP in their measurements. Li et al. 1 adopted a model similar to our prior model 5,13,14 to explain their observed CISP in terms of the spin-momentum locking (SML) of TSS (see Fig. 1). However, there is a qualitative difference regarding the electrochemical potentials between the model described by Li et al. in ref. 1 (Fig. 1b, c taken from Fig. 5 of ref. 1 ) and our model (Fig. 1d, e). First, their electrochemical potential diagram (see Fig. 1c) is inconsistent with their band diagram (see Fig. 1b) which is supposed to reflect the same experimental situation (see Fig. 1a). In Fig. 1c, the chemical potential of up-spin (μ ↑) is lower than that of down-spin (μ ↓), whereas μ ↑ is higher than μ ↓ in Fig. 1b. The relative order and the magnitude of the two opposite spin-dependent electrochemical potentials should be self-consistent between the band diagram (energy versus momentum) and their spatial distribution plot (energy versus distance), namely μ ↑ should be higher than μ ↓ in both Fig. 1b, c. We have drawn the corrected versions of Fig. 1b, c in Fig. 1d (band diagram) and Fig. 1e (spatial variation), respectively, under the same bias condition as Fig. 1b, c. Furthermore, Li et al. 1 used the absolute values of the electrochemical potentials (|μ ↑| and |μ ↓|) in the band diagram (see Fig. 1b), which is also incorrect since it changes the actual physics depending on the choice of the ground or zero energy level. Generally speaking, which spin state of electrons is more occupied should be governed by the difference (μ ↑ − μ ↓) between μ ↑ and μ ↓, not by their absolute values. Moreover, unlike what is depicted in Fig. 1c, the choice of the reference position for zero potential (µ = 0) is not important for determining which spin state is more occupied (or which spin chemical potential is higher). Since the electrons are injected by the right contact and flow from right to left in Fig. 1c, the chemical potential (μ R) of the right contact is higher than that (μ L) of the left contact (see Fig. 1c). Meanwhile, this also means that in the channel there should be more occupation of the left-going electron states (corresponding to up-spin states in a TI channel with the SML of TSS) than the right-going states (corresponding to the down-spin states), thus μ ↑ (equivalently the chemical potential of left-going electrons) should be higher than μ ↓ (the chemical potential of right-going electrons), whereas μ ↑ is incorrectly drawn to be lower than μ ↓ in Fig. 1b. Based on the above arguments, the sign of the spin voltage expected from the corrected potential diagram (seen Fig. 1d, e) in conjunction with their model argument b is opposite to the signal sign on Bi2Se3 reported by Li et al. Thus, the origin of the CISP observed in their Bi2Se3 samples is inconsistent with the expected CISP from the TSS. Fig. 1 Comparison of the models depicting the sign of the spin signals expected from TSS. a The schematic structure of a device with a charge current (I) flowing from left (L) to the right (R); b The band structure diagram of TSS, taken from the left panel shared by Fig. 5b, c in ref. 1 ; c Diagram showing spatially varying spin-dependent electrochemical potential, taken from the top panel of Fig. 5c in ref. 1 . d, e Our understanding of the spin-dependent electrochemical potentials of TSS in 3D TIs under the same bias current as that shown in b, c. The chemical potential (μ ↑) of up spins is higher than that (μ ↓) of down spins in both d and e We further note that Li et al. 1 incorrectly stated that another experiment on (Bi,Sb)2Te3 by Lee et al. 7 (which is ref. 29 in ref. 1 ) gave a consistent sign as theirs, whereas it is opposite. We point out that the sign reported by Lee et al. 7 is consistent with several other reports (see, refs. 5,6,8 ) while opposite to Li et al. 1,2 , as summarized in Table 1. Table 1 Comparison of the reported results on CISP in 3D TIs measured by spin potentiometry Li et al. (refs. 1, 2 ) Tian et al. (ref. 5 ) Dankert et al. (ref. 6 ) Lee et al. (ref. 7 ) Yang et al. (ref. 8 ) Charge current direction (I c) +x +x +x +x +x Electron current direction (I e) −x −x −x −x −x Magnetization direction of FM (M) +y +y +y +y +y Sign of spin signal (V S) − + + + + Channel spin polarization direction (s c) −y +y +y +y +y Spin polarization from TSS (s) +y +y +y +y +y Li et al. 1 also attempted to infer the origin of CISP in their Bi2Se3 by making a comparison with SOC semiconductor InAs, where a Rashba-type 2DEG normally exists on the surface. We caution that such a comparison may not be sufficient to draw a conclusion on the origin of CISP. For example, it is known that, depending on the direction of the effective electric field (potential gradient) perpendicular to the 2DEG 9–11,12 , the spin helicity of the outer Rashba band (which dominates the signals measured in transport) can be either opposite to or the same as that of TSS. Without a careful consideration of various parameters (e.g., an independent determination of the spin texture, for example by ARPES, and consideration of capping surfaces, interfaces, etc.) of their samples, spin potentiometric measurements in their InAs sample cannot provide an unambiguous “calibration” to determine the direction of the CISP measured by spin potentiometry in their Bi2Se3. Further complications can arise from the fact that in addition to the nontrivial spin-momentum-locked TSS, Bi2Se3 often contains multiple bands and conducting channels with spin-orbit coupling that can affect CISP. For example, the trivial surface 2DEG derived from bulk states and often observed by ARPES 15 in Bi2Se3 possesses strong Rashba-type spin-orbit coupling and typically has two fermi surfaces with opposite spin helicities 12 . We note that the model (Fig. 1d, e) we developed is only for TSS in the ideal case of bulk-insulating TI materials where the Fermi level is inside the bulk band gap. If the TI samples have metallic bulk with their Fermi levels located in the conduction band where the multiple bands coexist, our model may not be sufficient to determine the sources of the measured CISP. In conclusion, the model used by Li et al. 1 is erroneous and inconsistent with their TSS band diagram, and also inconsistent with the expected CISP due to TSS. The sign of the CISP signal experimentally observed in their Bi2Se3 is opposite to that predicted by a corrected model based on TSS and that observed in other experiments on various TI materials 5–8 . Owing to the existence of multiple bands in the bulk-metallic TI samples, the source of the measured CISP can be complicated, possibly involving competition between different bands (e.g., TSS and Rashba bands). Identifying the origin of CISP in bulk-metallic TI samples such as the Bi2Se3 used by Li et al. 1 may require analysis beyond the simple model (Fig. 1c, d) we developed.

          Related collections

          Most cited references15

          • Record: found
          • Abstract: found
          • Article: not found

          Electrical detection of charge-current-induced spin polarization due to spin-momentum locking in Bi2Se3.

          Topological insulators exhibit metallic surface states populated by massless Dirac fermions with spin-momentum locking, where the carrier spin lies in-plane, locked at right angles to the carrier momentum. Here, we show that a charge current produces a net spin polarization via spin-momentum locking in Bi2Se3 films, and this polarization is directly manifested as a voltage on a ferromagnetic contact. This voltage is proportional to the projection of the spin polarization onto the contact magnetization, is determined by the direction and magnitude of the charge current, scales inversely with Bi2Se3 film thickness, and its sign is that expected from spin-momentum locking rather than Rashba effects. Similar data are obtained for two different ferromagnetic contacts, demonstrating that these behaviours are independent of the details of the ferromagnetic contact. These results demonstrate direct electrical access to the topological insulators' surface-state spin system and enable utilization of its remarkable properties for future technological applications.
            Bookmark
            • Record: found
            • Abstract: found
            • Article: not found

            Electrical detection of spin-polarized surface states conduction in (Bi(0.53)Sb(0.47))2Te3 topological insulator.

            Strong spin-orbit interaction and time-reversal symmetry in topological insulators enable the spin-momentum locking for the helical surface states. To date, however, there has been little report of direct electrical spin injection/detection in topological insulator. In this Letter, we report the electrical detection of spin-polarized surface states conduction using a Co/Al2O3 ferromagnetic tunneling contact in which the compound topological insulator (Bi0.53Sb0.47)2Te3 was used to achieve low bulk carrier density. Resistance (voltage) hysteresis with the amplitude up to about 10 Ω was observed when sweeping the magnetic field to change the relative orientation between the Co electrode magnetization and the spin polarization of surface states. The two resistance states were reversible by changing the electric current direction, affirming the spin-momentum locking in the topological surface states. Angle-dependent measurement was also performed to further confirm that the abrupt change in the voltage (resistance) was associated with the magnetization switching of the Co electrode. The spin voltage amplitude was quantitatively analyzed to yield an effective spin polarization of 1.02% for the surface states conduction in (Bi0.53Sb0.47)2Te3. Our results show a direct evidence of spin polarization in the topological surface states conduction. It might open up great opportunities to explore energy-efficient spintronic devices based on topological insulators.
              Bookmark
              • Record: found
              • Abstract: not found
              • Article: not found

              Emergent quantum confinement at topological insulator surfaces

                Bookmark

                Author and article information

                Contributors
                jtian@uwyo.edu
                Journal
                Nat Commun
                Nat Commun
                Nature Communications
                Nature Publishing Group UK (London )
                2041-1723
                1 April 2019
                1 April 2019
                2019
                : 10
                : 1461
                Affiliations
                [1 ]ISNI 0000 0004 1937 2197, GRID grid.169077.e, Department of Physics and Astronomy, , Purdue University, ; West Lafayette, IN 47907 USA
                [2 ]ISNI 0000 0004 1937 2197, GRID grid.169077.e, Birck Nanotechnology Center, , Purdue University, ; West Lafayette, IN 47907 USA
                [3 ]ISNI 0000 0001 2109 0381, GRID grid.135963.b, Department of Physics and Astronomy, , University of Wyoming, ; Laramie, WY 82071 USA
                [4 ]ISNI 0000 0004 1937 2197, GRID grid.169077.e, School of Electrical and Computer Engineering, , Purdue University, ; West Lafayette, IN 47907 USA
                [5 ]ISNI 0000000121053345, GRID grid.35541.36, Center for spintronics, Post-silicon Semiconductor Institute, , Korea Institute of Science and Technology (KIST), ; Seoul, 02792 South Korea
                [6 ]ISNI 0000 0001 2181 7878, GRID grid.47840.3f, Electrical Engineering and Computer Science, , University of California Berkeley, ; CA, 94720 USA
                [7 ]ISNI 0000 0001 2097 4281, GRID grid.29857.31, Department of Physics, , The Pennsylvania State University, ; University Park, Pennsylvania, 16802 USA
                [8 ]ISNI 0000 0004 1937 2197, GRID grid.169077.e, Purdue Quantum Science and Engineering Institute, , Purdue University, ; West Lafayette, IN 47907 USA
                [9 ]ISNI 0000 0001 2248 6943, GRID grid.69566.3a, WPI-AIMR International Research Center for Materials Sciences, , Tohoku University, ; Sendai, 980-8577 Japan
                Author information
                http://orcid.org/0000-0003-2921-470X
                http://orcid.org/0000-0002-1001-1462
                http://orcid.org/0000-0003-2599-346X
                Article
                9271
                10.1038/s41467-019-09271-1
                6443799
                30931922
                0016b683-b610-406d-a59f-a15034c01fc1
                © The Author(s) 2019

                Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons license, and indicate if changes were made. The images or other third party material in this article are included in the article’s Creative Commons license, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons license and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this license, visit http://creativecommons.org/licenses/by/4.0/.

                History
                : 23 April 2018
                : 28 February 2019
                Categories
                Matters Arising
                Custom metadata
                © The Author(s) 2019

                Uncategorized
                Uncategorized

                Comments

                Comment on this article