Bismuth subhalides

Last updated

Bismuth-containing solid-state compounds pose an interest to both the physical inorganic chemists as well as condensed matter physicists due to the element's massive spin-orbit coupling, stabilization of lower oxidation states, and the inert pair effect. [1] Additionally, the stabilization of the Bi in the +1 oxidation state gives rise to a plethora of subhalide compounds with interesting electronics and 3D structures. [2] [3] [4]

Contents

Overview of subhalide bismuth solid-state chemistry

Topological insulators and the relationship to bismuth solid-state chemistry

Bismuth subhalides, such as Bi4Br4 and β-Bi4I4, have been recently reported as topological insulators. [2] [3] Topological insulators have caught attention of physical inorganic chemists as well as condensed matter physicists due to the unique physicochemical properties emerging upon transition from bulk to surface states. [5] Exhibiting an energy band gap of classic insulator, the edge/surface states of the material acquire dissipationless electric transport. The subject has been investigated by condensed matter physicists as well as mathematicians to provide a link between the experimental emerging properties and the modeled topology. Broadly, the material's physics pertains to the Quantum Hall effect relying upon two pillars: time-reversal symmetry and spin orbit coupling, the latter dependent on the elemental material composition. [5] Bismuth's heavy pnictogen nature yields a large spin-orbit coupling. [1] Additionally, when bound to heavy halogens, bismuth subhalides give rise to a low-dimensional van der Waals bonded structure, exfoliatable into nanowires. [3]

Structure of β-Bi4I4

Low dimensional van der Waals bonded materials display a fundamental material unit, usually depicted as the simplest molecular formula obeying stoichiometry. A series of such fundamental units align in the bulk material phase due to weak van der Waals interactions. Overall, key advantages conferred by the chemical structure are the ease to scale the materials down to nanostructures under simultaneous conservation of the bulk structure and the reduction in defects amount. [6]

b-Bi4I4 solid-state material structure (Bi - purple, I - red). Beta-Bi4I4.png
β-Bi4I4 solid-state material structure (Bi - purple, I - red).
a-Bi4I4 solid-state material structure (Bi - green, I - orange). Alpha-Bi4I4 1.png
α-Bi4I4 solid-state material structure (Bi - green, I - orange).

Belonging to the larger class of quasi 1-dimensional van der Waals bonded materials, β-Bi4I4 has been recently reported as a novel topological insulator. [2] The binary bismuth-iodine family class includes the known bismuth(III) iodide along with additional representatives such as α-Bi4I4, Bi14I4, Bi16I4, and Bi18I4. [2] Having the same stoichiometric chemical formula, α-Bi4I4 and β-Bi4I4 show similar solid-state structures yet critically different physicochemical properties. [8] Specifically, α-Bi4I4 represents the trivial insulator phase, while stacking of the bismuth atoms along the b crystallographic axis in the β-Bi4I4 phase yield a different topological insulator phase. Both isoforms crystallyse in the C2/m space group, with α-Bi4I4 having a unit cell volume almost double of its topological insulator counterpart. [7] The β crytallographic angle is higher in the β-Bi4I4: 107.87 vs 92.96, making the β-Bi4I4 more tilted (see images above). [7]

Synthesis

b-Bi4I4 solid-state synthesis Bi4I4 synthesis.png
β-Bi4I4 solid-state synthesis

Crystal growth of β-Bi4I4 was achieved through a solid-state reaction between Bi and HgI2 in a ratio of 1:2. The mixture of solid-state precursors was sealed under dynamic vacuum in a quartz ampoule and subjected to a temperature gradient of 250°C - 210°C in a two-zone furnace for 20 days. [2] [8] Needle-like blue crystals were obtained with sizes varying from a couple of mm in length and tenths of mm in diameter. [2]

DFT Calculations

Key to modeling the topology of a material are the special points along k-vector of the Brillouin zone, accounting for the accurate depiction of the density of states emerging from the electronics of the material. Density functional theory (DFT) analyses predicted an indirect band gap of 0.158 eV in the β-Bi4I4 phase with the valence and conduction band maxima localized at the Γ and M k-space points, respectively. [2] Interestingly enough, the major contributors to the band structure around the Fermi level are bismuth's p orbitals of even and odd parity, thus giving the gerade and ungerade points of symmetry.

ARPES measurements

The allowed electron energies in the topological insulator were probed with the well-employed angle-resolved photoemission spectroscopy (ARPES). [2] Γ and M space points were found to exhibit binding energies of 0.3 eV and 0.8 eV, respectively. [2] ARPES also probed the Fermi electron velocities along the x and y axes to be 0.1(1)×106 m*s-1 and 0.60(4)×106 m*s-1. [2] The emerging non-trivial states of the topological insulator are expected to show at the space point where the conductive and valence bands almost cross or, in other words, display the smallest band gap. This point indeed showed a binding energy of 0.06 eV as measured by ARPES. [2] ARPES measurements on a different β-Bi4Br4 topological insulator phase show similarity to its iodine counterpart. [9]

Subhalide complexes

Structure of RhBi7Br8 subhalide complex RhBi7Br8.jpg
Structure of RhBi7Br8 subhalide complex

A ternary rhodium-centered bipyramidal dibismuth complex is an example of subhalide complexes with interesting geometry and unusual electronic properties, particularly what has been reported as an example of Möbius aromaticity. [4] [10] The complex exhibits a 4-electron-5-centered bond in the central plane occupied by a Bi5 equatorial pentagon with the rhodium center in the middle. [10] Based on the electronic analysis carried out by Ruck (2003), the bismuth bonding consists of 2-centered-2-electron bonds, namely, Bi-Rh and Bi-Br one (see structure on the right).

The electronic analysis was carried out starting with counting the available skeletal electrons. Each of the 7 bismuth atoms contribute a total of 3x7=21 electrons (3 per each atom), while Rh gives all of its 9 electrons and the 8 bridging bromide atoms yield 3 electrons each. The total skeletal electron count is thus 54. The total skeletal electron count gets distributed as follows: 2 electrons per each of the 16 2c-2e- Bi-Br bond, 2 electrons per each of the 7 2c-2e- Rh-Bi metallic bond, 2 rhodium lone pairs remaining on the Rh centre (total of 4e-), and 4 electrons for 5c-4e- bond pertaining to the central pentagon. The sum of electrons used in bonding is therefore 54. Hence, the subhalide complex is electron-precise, i.e., with all of its skeletal electrons involved in chemical bonding. [10]

Orbital bonding in RhBi7Br8 subhalide complex compared to C5H5 cyclopentadienyl unit Orbital overlap in aromatic systems.jpg
Orbital bonding in RhBi7Br8 subhalide complex compared to C5H5 cyclopentadienyl unit

The bonding in such a system was compared to the aromatic cyclopentadienyl aromatic anion. Contrary to the π-type all in-phase orbital overlap exhibited by the organic cyclopentadienyl anion, σ-type bonding of the RhBi5 unit yields a phase change for an orbital pair (see figure). [10]

Mobius- and Huckel-type aromaticity of the Rh-Bi subhalide complex and cyclopentadienyl anion, respectively Aromaticity.png
Möbius- and Hückel-type aromaticity of the Rh-Bi subhalide complex and cyclopentadienyl anion, respectively

The relative orbital energy diagram is rationalized for each of the systems relying on the Frost-Musulin mnemonic. [11] The two lone pairs stemming from the rhodium metallic center are localized on the lowest-lying twicely degenerate set of molecular orbitals, consistent with the Möbius-type aromaticity. For reference, the electronics of the aromatic organice cyclopentadienyl unit is shown to the right of the rhodium-centered pentagonal Bi5 unit. As can be seen, Hückel rules dictate the molecular orbital splitting is inverted compared to its metallic counterpart, the highest-occupied molecular orbitals this time being twicely degenerate.

See also

Related Research Articles

<span class="mw-page-title-main">Photoemission spectroscopy</span> Examining a substance by measuring electrons emitted in the photoelectric effect

Photoemission spectroscopy (PES), also known as photoelectron spectroscopy, refers to energy measurement of electrons emitted from solids, gases or liquids by the photoelectric effect, in order to determine the binding energies of electrons in the substance. The term refers to various techniques, depending on whether the ionization energy is provided by X-ray, XUV or UV photons. Regardless of the incident photon beam, however, all photoelectron spectroscopy revolves around the general theme of surface analysis by measuring the ejected electrons.

<span class="mw-page-title-main">Cadmium arsenide</span> Chemical compound

Cadmium arsenide (Cd3As2) is an inorganic semimetal in the II-V family. It exhibits the Nernst effect.

<span class="mw-page-title-main">Bismuth(III) oxide</span> Chemical compound

Bismuth(III) oxide is a compound of bismuth, and a common starting point for bismuth chemistry. It is found naturally as the mineral bismite (monoclinic) and sphaerobismoite, but it is usually obtained as a by-product of the smelting of copper and lead ores. Dibismuth trioxide is commonly used to produce the "Dragon's eggs" effect in fireworks, as a replacement of red lead.

<span class="mw-page-title-main">SIESTA (computer program)</span>

SIESTA is an original method and its computer program implementation, to efficiently perform electronic structure calculations and ab initio molecular dynamics simulations of molecules and solids. SIESTA uses strictly localized basis sets and the implementation of linear-scaling algorithms. Accuracy and speed can be set in a wide range, from quick exploratory calculations to highly accurate simulations matching the quality of other approaches, such as the plane-wave and all-electron methods.

The Hückel method or Hückel molecular orbital theory, proposed by Erich Hückel in 1930, is a simple method for calculating molecular orbitals as linear combinations of atomic orbitals. The theory predicts the molecular orbitals for π-electrons in π-delocalized molecules, such as ethylene, benzene, butadiene, and pyridine. It provides the theoretical basis for Hückel's rule that cyclic, planar molecules or ions with π-electrons are aromatic. It was later extended to conjugated molecules such as pyridine, pyrrole and furan that contain atoms other than carbon and hydrogen (heteroatoms). A more dramatic extension of the method to include σ-electrons, known as the extended Hückel method (EHM), was developed by Roald Hoffmann. The extended Hückel method gives some degree of quantitative accuracy for organic molecules in general and was used to provide computational justification for the Woodward–Hoffmann rules. To distinguish the original approach from Hoffmann's extension, the Hückel method is also known as the simple Hückel method (SHM).

<span class="mw-page-title-main">Bismuth telluride</span> Chemical compound

Bismuth telluride is a gray powder that is a compound of bismuth and tellurium also known as bismuth(III) telluride. It is a semiconductor, which, when alloyed with antimony or selenium, is an efficient thermoelectric material for refrigeration or portable power generation. Bi2Te3 is a topological insulator, and thus exhibits thickness-dependent physical properties.

<span class="mw-page-title-main">Möbius aromaticity</span>

In organic chemistry, Möbius aromaticity is a special type of aromaticity believed to exist in a number of organic molecules. In terms of molecular orbital theory these compounds have in common a monocyclic array of molecular orbitals in which there is an odd number of out-of-phase overlaps, the opposite pattern compared to the aromatic character to Hückel systems. The nodal plane of the orbitals, viewed as a ribbon, is a Möbius strip, rather than a cylinder, hence the name. The pattern of orbital energies is given by a rotated Frost circle (with the edge of the polygon on the bottom instead of a vertex), so systems with 4n electrons are aromatic, while those with 4n + 2 electrons are anti-aromatic/non-aromatic. Due to incrementally twisted nature of the orbitals of a Möbius aromatic system, stable Möbius aromatic molecules need to contain at least 8 electrons, although 4 electron Möbius aromatic transition states are well known in the context of the Dewar-Zimmerman framework for pericyclic reactions. Möbius molecular systems were considered in 1964 by Edgar Heilbronner by application of the Hückel method, but the first such isolable compound was not synthesized until 2003 by the group of Rainer Herges. However, the fleeting trans-C9H9+ cation, one conformation of which is shown on the right, was proposed to be a Möbius aromatic reactive intermediate in 1998 based on computational and experimental evidence.

<span class="mw-page-title-main">Topological insulator</span> State of matter with insulating bulk but conductive boundary

A topological insulator is a material whose interior behaves as an electrical insulator while its surface behaves as an electrical conductor, meaning that electrons can only move along the surface of the material.

<span class="mw-page-title-main">Rhodocene</span> Organometallic chemical compound

Rhodocene is a chemical compound with the formula [Rh(C5H5)2]. Each molecule contains an atom of rhodium bound between two planar aromatic systems of five carbon atoms known as cyclopentadienyl rings in a sandwich arrangement. It is an organometallic compound as it has (haptic) covalent rhodium–carbon bonds. The [Rh(C5H5)2] radical is found above 150 °C (302 °F) or when trapped by cooling to liquid nitrogen temperatures (−196 °C [−321 °F]). At room temperature, pairs of these radicals join via their cyclopentadienyl rings to form a dimer, a yellow solid.

Bismuth selenide is a gray compound of bismuth and selenium also known as bismuth(III) selenide.

In quantum mechanics, fractionalization is the phenomenon whereby the quasiparticles of a system cannot be constructed as combinations of its elementary constituents. One of the earliest and most prominent examples is the fractional quantum Hall effect, where the constituent particles are electrons but the quasiparticles carry fractions of the electron charge. Fractionalization can be understood as deconfinement of quasiparticles that together are viewed as comprising the elementary constituents. In the case of spin–charge separation, for example, the electron can be viewed as a bound state of a 'spinon' and a 'chargon', which under certain conditions can become free to move separately.

<span class="mw-page-title-main">Solid nitrogen</span> Solid form of the 7th element

Solid nitrogen is a number of solid forms of the element nitrogen, first observed in 1884. Solid nitrogen is mainly the subject of academic research, but low-temperature, low-pressure solid nitrogen is a substantial component of bodies in the outer Solar System and high-temperature, high-pressure solid nitrogen is a powerful explosive, with higher energy density than any other non-nuclear material.

Bismuth antimonides, Bismuth-antimonys, or Bismuth-antimony alloys, (Bi1−xSbx) are binary alloys of bismuth and antimony in various ratios.

<span class="mw-page-title-main">Dirac cone</span> Quantum effect in some non-metals

In physics, Dirac cones are features that occur in some electronic band structures that describe unusual electron transport properties of materials like graphene and topological insulators. In these materials, at energies near the Fermi level, the valence band and conduction band take the shape of the upper and lower halves of a conical surface, meeting at what are called Dirac points.

<span class="mw-page-title-main">Bismuth polycations</span> Polyatomic ion

Bismuth polycations are polyatomic ions of the formula Bin+
x
. They were originally observed in solutions of bismuth metal in molten bismuth chloride. It has since been found that these clusters are present in the solid state, particularly in salts where germanium tetrachloride or tetrachloroaluminate serve as the counteranions, but also in amorphous phases such as glasses and gels. Bismuth endows materials with a variety of interesting optical properties that can be tuned by changing the supporting material. Commonly-reported structures include the trigonal bipyramidal Bi3+
5
cluster, the octahedral Bi2+
6
cluster, the square antiprismatic Bi2+
8
cluster, and the tricapped trigonal prismatic Bi5+
9
cluster.

<span class="mw-page-title-main">Rashba–Edelstein effect</span>

The Rashba–Edelstein effect (REE) is a spintronics-related effect, consisting in the conversion of a bidimensional charge current into a spin accumulation. This effect is an intrinsic charge-to-spin conversion mechanism and it was predicted in 1990 by the scientist V.M. Edelstein. It has been demonstrated in 2013 and confirmed by several experimental evidences in the following years.

<span class="mw-page-title-main">Bismuth compounds</span>

Bismuth forms mainly trivalent and a few pentavalent compounds. Many of its chemical properties are similar to those of arsenic and antimony, although much less toxic.

Nano Angle-Resolved Photoemission Spectroscopy (Nano-ARPES), is a variant of the experimental technique ARPES. It has the ability to precisely determine the electronic band structure of materials in momentum space with submicron lateral resolution. Due to its demanding experimental setup, this technique is much less extended than ARPES, widely used in condensed matter physics to experimentally determine the electronic properties of a broad range of crystalline materials. Nano-ARPES can access the electronic structure of well-ordered monocrystalline solids with high energy, momentum, and lateral resolution, even if they are nanometric or heterogeneous mesoscopic samples. Nano-ARPES technique is also based on Einstein's photoelectric effect, being photon-in electron-out spectroscopy, which has converted into an essential tool in studying the electronic structure of nanomaterials, like quantum and low dimensional materials.

The stabilization of bismuth's +3 oxidation state due to the inert pair effect yields a plethora of organometallic bismuth-transition metal compounds and clusters with interesting electronics and 3D structures.

References

  1. 1 2 Hirai, Yoshua; Yoshikawa, Naotaka; Hirose, Hana; Kawaguchi, Masashi; Hayashi, Masamitsu; Shimano, Ryo (2020-12-04). "Terahertz Emission from Bismuth Thin Films Induced by Excitation with Circularly Polarized Light". Physical Review Applied. 14 (6): 064015. Bibcode:2020PhRvP..14f4015H. doi:10.1103/PhysRevApplied.14.064015. S2CID   230642783.
  2. 1 2 3 4 5 6 7 8 9 10 11 12 Autès, Gabriel; Isaeva, Anna; Moreschini, Luca; Johannsen, Jens C.; Pisoni, Andrea; Mori, Ryo; Zhang, Wentao; Filatova, Taisia G.; Kuznetsov, Alexey N.; Forró, László; Van den Broek, Wouter; Kim, Yeongkwan; Kim, Keun Su; Lanzara, Alessandra; Denlinger, Jonathan D. (2015-12-14). "A novel quasi-one-dimensional topological insulator in bismuth iodide β-Bi4I4". Nature Materials. 15 (2): 154–158. arXiv: 1606.06192 . doi:10.1038/nmat4488. ISSN   1476-4660. PMID   26657327. S2CID   44808186.
  3. 1 2 3 Zhou, Jin-Jian; Feng, Wanxiang; Liu, Cheng-Cheng; Guan, Shan; Yao, Yugui (2014-08-13). "Large-Gap Quantum Spin Hall Insulator in Single Layer Bismuth Monobromide Bi4Br4". Nano Letters. 14 (8): 4767–4771. arXiv: 1405.3823 . Bibcode:2014NanoL..14.4767Z. doi:10.1021/nl501907g. ISSN   1530-6984. PMID   25058154. S2CID   7403617.
  4. 1 2 3 Ruck, Michael (1997-10-02). "Bi7RhBr8 : A Subbromide with Molecular [{RhBi 7 }Br 8 ] Clusters". Angewandte Chemie International Edition in English. 36 (18): 1971–1973. doi:10.1002/anie.199719711. ISSN   0570-0833.
  5. 1 2 Qi, Xiao-Liang; Zhang, Shou-Cheng (2011-10-14). "Topological insulators and superconductors". Reviews of Modern Physics. 83 (4): 1057–1110. arXiv: 1008.2026 . Bibcode:2011RvMP...83.1057Q. doi:10.1103/RevModPhys.83.1057. S2CID   118373714.
  6. Ma, Yuhang; Yi, Huaxin; Liang, Huanrong; Wang, Wan; Zheng, Zhaoqiang; Yao, Jiandong; Yang, Guowei (2023-09-14). "Low-dimensional van der Waals materials for linear-polarization-sensitive photodetection: materials, polarizing strategies and applications". Materials Futures. 3: 012301. doi: 10.1088/2752-5724/acf9ba . ISSN   2752-5724.
  7. 1 2 3 4 "Wismutmonojodid Bi J, eine Verbindung mit Bi(0) und Bi(II)". 1978.
  8. 1 2 Wang, Peipei; Tang, Fangdong; Wang, Peng; Zhu, Haipeng; Cho, Chang-Woo; Wang, Junfeng; Du, Xu; Shao, Yonghong; Zhang, Liyuan (2021-04-08). "Quantum transport properties of β-Bi4I4 near and well beyond the extreme quantum limit". Physical Review B. 103 (15): 155201. arXiv: 2103.13079 . doi:10.1103/PhysRevB.103.155201. S2CID   232335759.
  9. Noguchi, Ryo; Kobayashi, Masaru; Jiang, Zhanzhi; Kuroda, Kenta; Takahashi, Takanari; Xu, Zifan; Lee, Daehun; Hirayama, Motoaki; Ochi, Masayuki; Shirasawa, Tetsuroh; Zhang, Peng; Lin, Chun; Bareille, Cédric; Sakuragi, Shunsuke; Tanaka, Hiroaki (2021-02-11). "Evidence for a higher-order topological insulator in a three-dimensional material built from van der Waals stacking of bismuth-halide chains". Nature Materials. 20 (4): 473–479. arXiv: 2002.01134 . Bibcode:2021NatMa..20..473N. doi:10.1038/s41563-020-00871-7. ISSN   1476-4660. PMID   33398124. S2CID   230660660.
  10. 1 2 3 4 5 6 King, R. Bruce (2003-01-29). "Möbius aromaticity in bipyramidal rhodium-centered bismuth clusters". Dalton Transactions (3): 395–397. doi:10.1039/B211440M. ISSN   1477-9234.
  11. 1 2 Frost, Arthur A.; Musulin, Boris (1953-03-01). "A Mnemonic Device for Molecular Orbital Energies". The Journal of Chemical Physics. 21 (3): 572–573. Bibcode:1953JChPh..21..572F. doi:10.1063/1.1698970. ISSN   0021-9606.