DOI:
10.1039/D5QI01401H
(Review Article)
Inorg. Chem. Front., 2025, Advance Article
The rise of functional organoarsenic chemistry
Received
30th June 2025
, Accepted 7th October 2025
First published on 8th October 2025
Abstract
Organoarsenic compounds have long been viewed through the lens of toxicity, yet recent advances in their synthesis, reactivity, and molecular design have begun to redefine their role in modern chemistry. This review highlights the emerging field of functional organoarsenic chemistry, which spans molecular frameworks, polymeric materials, metal coordination complexes, and reactive intermediates. Key developments include the design of π-conjugated arsole-based polymers with tunable optoelectronic properties, the construction of coordination and redox-active frameworks by harnessing the soft Lewis basicity of arsenic, and the exploitation of As(III)/As(V) redox cycles for catalysis. Furthermore, the development of arsonium-based ionic liquids and hypervalent arsenic-centered dications demonstrates the expanding versatility of arsenic across diverse chemical contexts. Collectively, these findings illustrate how arsenic's unique electronic and structural attributes—distinct from its phosphorus congener—enable new functionalities that enrich the toolbox of molecular and materials science. The rise of functional organoarsenic chemistry represents both a revival and a reimagination of this historically overlooked element, paving the way for future applications in catalysis, photonics, sensing, and soft materials.
 Hiroaki Imoto | Prof. Hiroaki Imoto received his Ph.D. from Kyoto University in 2012. From 2009 to 2012, he was a JSPS Research Fellow in the group of Prof. Yoshiki Chujo at Kyoto University. In 2012, he joined JSR Corporation as a researcher. He began his academic career at the Kyoto Institute of Technology (KIT) as an Assistant Professor in 2014 and was promoted to Associate Professor in 2019. Since 2023, he has also served as a Research Fellow of the JST FOREST Program. His research interests include organic–inorganic hybrid molecules, polymers, and supramolecular materials. |
 Chihiro Okochi | Chihiro Okochi received his Bachelor of Engineering degree from the Kyoto Institute of Technology (KIT), where he is currently pursuing a Master's degree in Materials Synthesis under the supervision of Prof. Kensuke Naka. His research focuses on the development of synthetic methodologies for novel arsenic compounds and the exploration of their unique reactivity. |
 Kazuma Kikuchi | Kazuma Kikuchi received his B.Eng. degree in 2021 and his M.Eng. degree in 2023 from the Kyoto Institute of Technology (KIT), where he is currently pursuing his Ph.D. under the supervision of Prof. Kensuke Naka. Since 2023, he has been a recipient of a JST Fellowship under the program “The Establishment of University Fellowships Toward the Creation of Science Technology Innovation”. His research interests include the design and construction of supramolecular architectures based on arsenic complexes. |
 Kensuke Naka | Prof. Kensuke Naka received his Ph.D. from Kyoto University in 1991. He began his academic career as a Research Assistant at Kagoshima University in 1990 and later moved to Kyoto University in 1996, where he became an Associate Professor in 2000. In 2007, he was promoted to Full Professor at the Kyoto Institute of Technology (KIT). He received the SPSJ Wiley Award in 2007 and the Award of The Adhesion Society of Japan in 2020. He currently serves as Director of the Materials Innovation Laboratory at KIT. |
1. Background
The integration of inorganic elements into organic frameworks has emerged as a powerful strategy to surpass the inherent limitations of traditional hydrocarbon-based materials.1–5 In the pursuit of novel functionalities, researchers have progressively incorporated heavier main-group elements, including those from the late p-block. Notably, even bismuth—regarded as the heaviest non-radioactive element—has found applications in phosphorescent materials, X-ray shielding agents, and high-refractive-index polymers.6–9 However, despite this expanding scope, arsenic remains conspicuously underexplored in the context of functional materials based on organo-element hybrids.
Arsenic, named after the yellow mineral arsenikon, is an abundant element in the Earth's crust and has been widely employed in the field of inorganic materials chemistry.10 During the 18th and 19th centuries, arsenic-containing pigments such as Paris Green (Cu(C2H3O2)2·3Cu(As2)2) and Scheele's Green (AsCuHO3) were developed and widely used. In the 1950s, gallium arsenide (GaAs) was identified as a compound with semiconducting properties, a discovery that laid the foundation for its subsequent role as a key material in modern electronics. On the other hand, the notorious toxicity of arsenic compounds has long shaped their societal perception. In France, arsenic trioxide (As2O3) was colloquially termed poudre de succession (“inheritance powder”) owing to its use as a poison in clandestine murders.
In contrast to inorganic arsenic chemistry, the history of organoarsenic compounds is relatively recent (Fig. 1).11–16 The first known organoarsenic compound, cacodyl (Me2AsAsMe2), was synthesized by Louis Claude Cadet de Gassicourt in 1760, with its molecular structure elucidated in the 19th century. A landmark development in medicinal chemistry was the synthesis of Salvarsan in 1907,17 the first effective treatment for syphilis—the exact structure of Salvarsan remained ambiguous for nearly a century, until its cyclic nature was finally confirmed in 2005.18 Unfortunately, the early promise of organoarsenic chemistry was overshadowed by its misuse in warfare: compounds such as Adamsite (diphenylaminechlorarsine), Lewisite (2-chlorovinyldichloroarsine), and Sneezing Gas (chlorodiphenylarsine) were deployed as chemical weapons during World War I.19–21 The long-lasting environmental damage caused by these arsenic-based munitions continues to pose serious ecological threats. These historical misapplications have strongly shaped the negative perception of organoarsenic chemistry, leading to a significant decline in research interest, primarily owing to toxicity concerns.
 |
| | Fig. 1 Chemical structures of historically representative organoarsenic compounds. | |
As a result, organoarsenic chemistry has increasingly shifted toward computational investigations, which provide insights into bonding, reactivity, and electronic structure without the inherent hazards of experimental work. Nevertheless, one of the central challenges in experimental organoarsenic research remains the safe synthesis of these compounds. Historically, highly volatile and toxic arsenic precursors have been employed, placing researchers at substantial risk of exposure. Therefore, the development of safer and more accessible synthetic methodologies is a crucial prerequisite for revitalizing the field.
This review begins by highlighting recent advances in practical and safer synthetic methods for organoarsenic compounds. It then explores the functional potential of these molecules across diverse domains, including π-conjugated systems, polymeric materials, metal-coordinating ligands, reactive intermediates, and ionic species.
2. Synthetic strategy
2.1. Preparation of electrophiles
To construct As–C bonds, conventional and recently developed methods have been comprehensively summarized by Tay and Pullarkat.22 Traditionally, arsenic chlorides such as trichloroarsine (AsCl3), dichlorophenylarsine (PhAsCl2), and chlorodiphenylarsine (Ph2AsCl) have served as standard electrophilic arsenic sources (Scheme 1a). However, these compounds are highly volatile—for instance, AsCl3 exhibits a vapor pressure of 10 mmHg at 23.5 °C—posing substantial safety risks, although its volatility is lower than that of the phosphorus analogue PCl3 (120 mmHg at 25 °C). Moreover, the synthesis of PhAsCl2 is particularly cumbersome, requiring the reaction of phenylarsonic acid (PhAsO3H2), hydrochloric acid (HCl), and sulfur dioxide (SO2), followed by vacuum distillation for purification (Scheme 1b).23
 |
| | Scheme 1 Preparation methods for arsenic electrophiles. (a) Traditional electrophiles, (b) synthesis of PhAsCl2, (c) As–As bond cleavage by I2, (d) in situ generation of RAsI2 (R = Me, Ph) from cyclooligoarsines, (e) generation of AsX3 (X = Br, I), (f) synthesis of dithiaarsoles, and (g) As–As bond cleavage by IX (X = Cl, I). | |
An efficient alternative involves the oxidative cleavage of As–As bonds to generate arsenic halides. For example, treatment of Me2AsAsMe2 with iodine (I2) yields iododimethylarsine (Me2AsI) (Scheme 1c).24 In this context, cyclooligoarsines such as pentamethylpentaarsine (Me5As5) and hexaphenylhexaarsine (Ph6As6) have emerged as practical precursors for generating arsenic electrophiles without resorting to volatile reagents.25,26 Upon treatment with I2, these compounds undergo selective As–As bond cleavage to afford diiodoarsines in situ: Me5As5 yields diiodomethylarsine (MeAsI2), while Ph6A6 gives diiodophenylarsine (PhAsI2) (Scheme 1d).27 Notably, these electrophiles are generated cleanly, without byproducts, and can be used directly in subsequent reactions without isolation.
As safer and more manageable alternatives to AsCl3, we proposed the use of tribromoarsine (AsBr3) and triiodoarsine (AsI3), both of which are solids at room temperature. These compounds can be synthesized directly from As2O3 using aqueous hydrobromic acid (HBr) or hydroiodic acid (HI), respectively (Scheme 1e).28 AsBr3 is conveniently extracted using n-hexane, affording an isolated yield of 92%. By contrast, owing to the poor solubility of AsI3, filtration is required to isolate the solid product, resulting in a lower yield (33%). Importantly, AsBr3 is readily soluble in common organic solvents such as n-hexane, diethyl ether, tetrahydrofuran, and toluene, making it a particularly convenient and non-volatile precursor for organoarsenic synthesis. This reagent is particularly suited for the preparation of A3-type (C3-symmetric) compounds bearing three identical substituents.
The selective synthesis of AB2-type (Cs-symmetric) arsines bearing two different substituents presents a greater challenge. In the case of phosphines, stepwise substitution of trichlorophosphine (PCl3) is often effective; however, for arsenic, the longer As–C bond reduces steric hindrance and favors over-substitution to yield A3-type products. A novel strategy has been developed to circumvent this limitation by exploiting the high affinity of arsenic halides for thiols. For instance, AsBr3 reacts with 1,2-benzenedithiol to form a brominated dithiaarsole intermediate (Scheme 1f).29 This species can undergo sequential nucleophilic substitutions: first at the As–Br bond, then via displacement of the chelating dithiol unit by a second nucleophile. This strategy enables precise control over substitution, affording AB2-type arsines with high selectivity—including chiral arsa-Buchwald-type ligands (see section 5.3 for details).
Another useful approach involves the oxidative cleavage of As–C bonds to generate arsenic halides.30–32 In 1989, Pazik et al. transformed trineopentylarsine into arsenic bromide via reaction with bromine (Br2). This method was later extended to other organoarsenic species such as 9-methyl-9-arsafluorene.33 Treatment with I2 or iodine monochloride (ICl) results in As–C bond cleavage, forming As–I or As–Cl bonds with the concurrent release of iodomethane (MeI), as illustrated in Scheme 1g. The resulting arsenic halides readily undergo nucleophilic substitution with Grignard or organolithium reagents to furnish various 9-arsafluorene derivatives. This strategy is also applicable to methyldiphenylarsine (AsMePh2), providing access to iododiphenylarsine (Ph2AsI).
2.2. Preparation of nucleophiles
Organoarsenic nucleophiles can be generated by deprotonation of As–H bonds, as well as by cleavage of As–C and As–As bonds. Traditionally, alkali metals such as lithium (Li), sodium (Na), and potassium(K) have been employed to generate arsenic nucleophiles (Scheme 2a). Tzschach et al. reported a method for obtaining diphenylarsenyllithium (Ph2AsLi) by reacting diphenylarsine (Ph2AsH) with phenyllithium (PhLi).34 Furthermore, reacting triphenylarsine (AsPh3) with an alkali metal (e.g., Li, Na, or K) has been reported to generate the corresponding arsenic nucleophiles.35,36 Martín and colleagues reported the transformation of triphenylarsine (AsPh3) into sodium diphenylarsenide (NaAsPh2), which subsequently reacts with tributyltin chloride (Bu3SnCl) to afford stannyldiphenylarsine (Bu3Sn–AsPh2) (Scheme 2b).37–42 This stannylated species serves as a valuable precursor for Pd-catalyzed cross-coupling with aryl iodides (ArI), enabling the formation of unsymmetrical triarylarsines (AsArPh2).
 |
| | Scheme 2 Preparation methods for arsenic nucleophiles. (a) Deprotonation of Ph2AsH, (b) As–C bond cleavage to generate arsenosodium species, (c) As–As bond cleavage of diarsane by phenyllithium, and (d) As–As bond cleavage of Ph6As6 by organolithium reagents. | |
An alternative approach involves the cleavage of As–As bonds to access nucleophilic arsenic species. Kauffmann and colleagues demonstrated that diarsane compounds undergo selective bond scission upon treatment with phenyllithium (PhLi), leading to the formation of arsenolithium intermediates (Scheme 2c).43,44 One phenyl group is introduced onto one arsenic center, while the other is converted into an organolithium species. This strategy can be extended to cyclic oligoarsines. For example, sequential cleavage of the As–As bonds in hexaphenylhexaarsine (Ph6As6) by PhLi generates diphenylarsenyllithium (Ph2AsLi) (Scheme 2d), which can be subsequently alkylated with benzyl bromide to furnish benzyldiphenylarsine.45
2.3. Other arsination methods
Hydroarsination of alkenes and alkynes remains a foundational method for synthesizing organoarsenic compounds.46–55 The required arsenic hydride precursors can be obtained by reducing arsenic halides or arsonic acids.56,57 One classic example is the hydroarsination of 1,3-diynes with phenylarsine (PhAsH2), which yields arsole derivatives (Scheme 3a).55 Leung and colleagues developed a broad range of hydroarsination protocols, particularly those involving asymmetric induction.48–51,53,54 Using chiral palladium complexes as templates, they enantioselectively synthesized arsenic-containing ligands (Scheme 3b),48 including arsenic–nitrogen and arsenic–phosphorus bidentate scaffolds. In addition, they established nickel-catalyzed asymmetric hydroarsination of nitrostyrenes (Scheme 3c).51 As a complementary approach, phosphine-promoted hydroarsination offers a metal-free route: triphenylphosphine (PPh3) facilitates the addition of diphenylarsine (Ph2AsH) across C
C double bonds.53
 |
| | Scheme 3 Arsination reactions. (a) Synthesis of arsole via hydroarsination and (b) chiral template for hydroarsination. (c) Asymmetric hydroarsination, (d) bisarsination using Me5As5, and (e) Pd-mediated aryl group scrambling reaction between aryl triflates and AsPh3. | |
Diarsanes bearing As–As bonds also serve as useful precursors for bisarsination reactions, enabling the simultaneous incorporation of two arsenic units. For example, reaction of tetraphenyldiarsane (Ph2AsAsPh2) with phenylacetylene under azobis(isobutyronitrile) (AIBN) initiation and photoirradiation yields 1,2-bis(arsino)ethylenes.58 Cyclooligoarsines such as Me5As5 can also undergo efficient bisarsination. When irradiated in the presence of 2,3-dimethyl-1,3-butadiene, Me5As5 furnishes 1,2,4,5-tetramethyl-tetrahydrodiarsinine in 95% yield (Scheme 3d).59 Related polymerization and cycloaddition reactions involving cyclooligoarsines and alkynes are discussed in sections 4.1 and 5.1, respectively.
Chan and colleagues reported a unique Pd-mediated aryl group scrambling reaction, wherein aryl groups from Pd–Ar species are exchanged with phenyl groups from AsPh3. This process generates Cs-symmetric AsArPh2 via Pd–Ar/As–Ph transmetalation (Scheme 3e).60,61 Notably, this strategy is compatible with various aryl triflates and has been employed to synthesize chiral arsenic–nitrogen bidentate ligands and their corresponding metal complexes. The Friedel–Crafts reaction is one of the few methods available for the direct introduction of arsenic into aromatic rings.62 While several examples of intramolecular Friedel–Crafts reactions have been reported, intermolecular cases remain rare. Lewis and colleagues obtained phenoxarsine by refluxing diphenyl ether with trihaloarsine and subsequently adding aluminum chloride.
Recently, yellow arsenic (As4) has been employed as a precursor and converted into a variety of arsenic compounds incorporating transition metal complexes.63 The As–As bond rearrangements give rise to oligomeric arsenic species, further expanding the scope of organoarsenic chemistry. These advances highlight the unique reactivity of As4 compared with its lighter homologue white phosphorus (P4). Moreover, the development of storage materials and transfer reagents for As4 has opened new opportunities for safer handling and broader synthetic applications.
3. Conjugated molecules
3.1. Arsoles
Arsoles were first reported in the 1920s.64,65 Traditionally, their synthesis has relied on volatile and toxic arsenic chlorides or hydrides (Scheme 4a).55,66–70 Owing to serious safety concerns, many studies in this area remained computational, focusing on estimating structural and electronic properties. Consequently, experimental investigations into arsole derivatives were scarce until our group developed a synthetic strategy in 2015 that avoided the use of hazardous precursors.27 This method utilizes diiodoarsines generated in situ from cyclooligoarsines, which react with organometallic reagents to afford a variety of arsole derivatives (Scheme 4b).
 |
| | Scheme 4 Synthesis of arsoles: (a) conventional and (b) practical methods. | |
One notable class of arsoles is arsafluorenes, which are arsenic analogues of carbazole. These compounds are synthesized from RAsI2 (R = Me, Ph) and 2,2′-dilithiobiphenyl.27 Importantly, access to 9-phenylarsafluorene enabled the first direct experimental comparison across group 15 analogs of carbazole (Fig. 2a)—namely carbazole (N), phosphafluorene (P), arsafluorene (As), stibafluorene (Sb), and bismafluorene (Bi).71 While the nitrogen atom in carbazole exhibits a delocalized lone pair that enforces a trigonal planar geometry, the heavier pnictogen analogs adopt trigonal pyramidal geometries owing to limited lone pair delocalization. Reactivity studies using oxidants such as O2, I2, and AuCl revealed contrasting behaviors among the group 15 elements. Carbazole was unreactive owing to the inertness of the C–N bond and lone-pair delocalization. Phosphafluorene easily oxidized in air while maintaining its molecular framework. By contrast, arsafluorene demonstrated excellent oxidative stability and strong coordination ability, readily forming complexes with AuCl. The heavier congeners displayed increased reactivity: 9-phenylstibafluorene lost the phenyl group, while 9-phenylbismafluorene underwent complete cleavage of the Bi–C bond. These studies revealed that arsafluorene uniquely balances structural stability with coordination reactivity. Photophysical measurements further elucidated group-specific behavior. All derivatives—carbazole, arsafluorene, stibafluorene, and bismafluorene—showed dual emission (fluorescence and phosphorescence), arising from lone-pair or heavy-atom-mediated intersystem crossing (ISC). We also demonstrated that bridging arsafluorene with silicon or germanium enhances phosphorescence via increased spin–orbit coupling.72
 |
| | Fig. 2 Arsole derivatives. (a) Group 15 analogs of carbazole. (b–e) Emissions of (b) 1,2,5-triphenylarsole and 1,2,5-triphenylphosphole, (c) 1,2,3,4,5-pentaphenylarsole, (d) 4-phenyldithieno[3,2-b:2′,3′-d]arsoles, and (e) 4-phenyldithieno[3,4-b:3′,4′-d]arsole. (f) Selective transformation of dipyridinoarsole. (g–i) Chemical structures of (g) As-heteropentacenes, (h) benzoarsole, and (i) dithiazoloarsole. | |
Additional arsole derivatives were synthesized by reacting PhAsI2 with organometallic precursors such as titanacyclopentadienes, which enabled regioselective access to 1,2,5-triarylarsoles.73,74 Unlike arsafluorene, these 1,2,5-triarylarsoles exhibit strong room-temperature emission. For instance, 1,2,5-triphenylarsole emits at 458 nm with a quantum yield (Φ) of 0.59, comparable to that of 1,2,5-triphenylphosphole (emission maximum: λem = 465 nm, Φ = 0.56) (Fig. 2b).
Solid-state luminescence properties of arsoles have also been explored. 1,2,3,4,5-Pentaarylarsoles show aggregation-induced emission.75 While the Φ of 1,2,3,4,5-pentaphenylarsole in solution is low (0.01) owing to phenyl group rotation, restricted motion in the solid state boosts Φ to 0.61 (Fig. 2c)—comparable to that of pentaphenylsilole. The stability of cycloalka[c]arsoles, synthesized via reactions of PhAsI2 with zirconacyclopentadienes, is highly dependent on ring size.76 For example, cyclopentane-fused derivatives are stable under ambient conditions, while cyclohexane-fused derivatives decompose in solution via photoinduced reaction with singlet oxygen (1O2), regenerating phenylarsonic acid.77 This unique degradation pathway enables chemical recycling of arsenic units back into arsoles.
π-Extended arsoles such as 4-phenyldithieno[3,2-b:2′,3′-d]arsoles were developed by Heeney et al. in 2016.78 Dibromination using N-bromosuccinimide, followed by Pd-catalyzed coupling, provided access to π-extended dithienoarsoles and their polymers (see section 4.1).79 These compounds exhibit strong emission, tunable via oxidation or coordination at the As center. Computational analyses revealed arsenic contributions to the lowest unoccupied molecular orbital (LUMO) through σ*–π* interactions, similar to phospholes.80 Oxidation or metal coordination lowers the LUMO energy, resulting in bathochromic shifts in absorption and emission (Fig. 2d). Substituent effects were also significant: sterically bulky groups distorted the π-system, producing red-shifted emission owing to structural relaxation in the excited state via As–C bond elongation.81
Dithieno[3,4-b:3′,4′-d]arsoles, isomers of dithieno[3,2-b:2′,3′-d]arsoles, displayed different photophysical behavior:82 while dithieno[3,2-b:2′,3′-d]arsoles showed phosphorescence at 77 K, dithieno[3,2-b:2′,3′-d]arsoles exhibited fluorescence only (Fig. 2e). Dipyridinoarsoles demonstrated chemoselective redox and coordination chemistry—arsenic centers selectively oxidize or coordinate, while pyridyl nitrogens are quaternized to form viologen-type structures (Fig. 2f).83 These species exhibit reversible electrochromism. This behavior is likely associated with the oxidative resistance of As(III), which contributes to electrochemical stability. Notably, Baumgartner and co-workers have reported reversible electrochromism in phosphaviologens, specifically in the oxidized forms of phosphoryl-bridged viologens.84,85 In addition, the hard basicity of nitrogen further enables selective coordination to Zn2+, allowing metal–organic framework (MOF) construction without involving the arsenic center (see section 4.3).86
Additional arsole derivatives include benzothiophene-fused As-heteropentacenes (Fig. 2g), which show higher charge mobility than their benzofuran analogs owing to intermolecular As⋯S interactions in thin films.87 Substituent-dependent fluorescence behavior was also observed in 2-arylbenzo[b]arsoles (Fig. 2h): electron-donating groups promote ISC, populating T1 states that deactivate non-radiatively.88 Heeney and colleagues synthesized dithiazoloarsole from PhAsCl2 and studied its aromaticity, air stability, and photophysics (Fig. 2i).89 Comparative studies of arene-fused benzoarsoles with silole, germole, and phosphole analogs revealed that the trivalent arsenic atom enhances ISC through its lone pair and heavy atom effect, contributing to relatively efficient phosphorescence.90 Notably, the corresponding arsole oxide, in which the arsenic lone pair is absent, displayed efficient fluorescence rather than phosphorescence, underscoring the critical role of the lone pair in facilitating ISC.
In response to toxicity concerns, the cytotoxicity of π-conjugated organoarsenic compounds was evaluated using HCT-116 human colon cancer cells.91 Nanoparticle formulations of these compounds exhibited low-to-negligible cytotoxicity (IC50 > 10 μM), in stark contrast to the well-known toxin phenylarsine oxide (IC50 = 2.8 μM). Notably, phenylarsonic acid—a key synthetic precursor—also showed minimal cytotoxicity, further supporting the safety of these advanced materials.
3.2. Excited state dynamics
Luminescent dyes that undergo substantial structural relaxation upon photoexcitation have attracted considerable attention owing to their large Stokes shifts and environmentally responsive emission behavior. A representative example is cyclooctatetraene, in which excited-state planarization is driven by Baird aromaticity—an electronic phenomenon in which 4n π-electron systems become aromatic in the excited state, in contrast to Hückel's rule governing ground-state aromaticity.92,93 In this context, the seven-membered ring system of arsepin is particularly intriguing, as it possesses an 8π-electron configuration suitable for Baird aromaticity in the excited state. By contrast, arsoles, with their 6π-electron systems, follow conventional Hückel-type aromaticity.94 Arsepins adopt bent geometries in the ground state in accordance with Hückel's rule but planarize upon photoexcitation owing to the onset of Baird aromaticity. Dibenzoarsepin and its gold(I) chloride complex have been synthesized and studied in detail. Dibenzoarsepin displays dual fluorescence at 371 nm and 557 nm, corresponding to emission from the Franck–Condon geometry and the planarized excited state, respectively (Fig. 3a). By contrast, the gold(I) chloride complex exhibits only the shorter-wavelength emission (375 nm), attributed to coordination of the arsenic lone pair, which suppresses planarization and thus precludes long-wavelength emission. Temperature-dependent photoluminescence measurements revealed that the energy barrier between the two excited-state conformers of dibenzoarsepin is 3.2 kJ mol−1. Following this discovery, various heteropins, including oxepins and azepins, have been developed to explore excited-state aromaticity phenomena.95,96
 |
| | Fig. 3 Excited state dynamics and photoluminescence spectra of (a) dibenzoarsepin and (b) arsenic-bridged diphenyl sulfone. | |
Arsenic-bridged diphenyl sulfones present another example of unusual excited-state dynamics.97 Like triarylphosphines, triarylarsines undergo planarization upon excitation owing to a reduction in electron density around the pnictogen atom.98 This effect enhances the Lewis acidity of the pnictogen center, enabling photo-induced Lewis pair formation when an internal Lewis base is present. In arsenic-bridged diphenyl sulfone, the As⋯O distance shortens from 3.18 Å in the ground state to 2.27 Å in the singlet excited state, leading to significant bending of the central six-membered ring (Fig. 3b). An energy barrier between the Franck–Condon and relaxed geometries results in dual fluorescence, with emission properties sensitive to environmental factors such as temperature and solvent viscosity. The excited-state relaxation behavior of other pnictogen-bridged systems—such as diphenyl ethers, diphenyl sulfides, and diphenyl sulfoximines—has also been examined.99,100 Both the identity and oxidation state of the bridging pnictogen atom strongly influence structural changes upon excitation. In the case of pnictogen-bridged sulfoximines, substituents on the sulfoximine nitrogen also play a critical role in tuning excited-state dynamics.
3.3. Other conjugated molecules
Orthaber and colleagues reported the synthesis of an arsaalkene featuring an exocyclic arsenic atom on a cyclopentadithiophene core (Fig. 4a).101 Compared to its phosphaalkene analog, the arsaalkene displays a more stabilized LUMO, highlighting the electronic influence of arsenic relative to phosphorus. Our group recently synthesized a series of arene-substituted arsines, in which the phenyl groups of AsPh3 were replaced by expanded π-systems such as naphthyl, phenanthryl, and pyrenyl groups (Fig. 4b).102 The naphthyl and phenanthryl derivatives phosphoresced at 77 K, whereas the pyrenyl analogs primarily fluoresced. (Diphenylarsino)quinolines also phosphoresced at low temperature, and their solid-state emission properties were found to depend critically on the substitution pattern of the diphenylarsino group on the quinoline ring (Fig. 4c).103
 |
| | Fig. 4 (a) Synthesis of alsaalkene. (b and c) Chemical structures of (b) triarylarsines and (c) AsPh2-substituted quinoline derivatives. | |
Designing planar π-conjugated molecules incorporating arsenic is particularly challenging owing to the inherently pyramidal geometry of tertiary arsines. One strategy to achieve planarity involves the development of heavier analogs of pyridine, such as phosphinines and arsinines. However, these heavier heterocycles often suffer from instability and dimerization, which necessitate the use of bulky substituents that can, in turn, compromise planarity.104,105 To circumvent this, Yamada and Matsuo introduced benzothiophene-fused phosphinines, which demonstrated high stability without the need for steric protection.106 Inspired by this, we designed and synthesized thiophene- and benzothiophene-fused arsinines (Fig. 5a), whose planarity was confirmed by structural analysis.107 Among them, benzothiophene-fused arsinines exhibited remarkable ambient stability. Their emission properties vary depending on the fused ring and fusion mode. Notably, α-type benzothiophene-fused arsinines exhibit intermolecular As⋯As interactions in the solid state—interactions not observed in their phosphorus analogs—owing to the larger atomic radius of arsenic. These interactions offer novel design opportunities for arsenic-based optoelectronic materials.
 |
| | Fig. 5 Chemical structures and emission behaviors of (a) arsinines and (b) arsaborines. | |
The incorporation of other heteroatoms into the π-conjugated framework is another promising strategy to diversify arsenic-containing materials. Dibenzoheteraborins are particularly attractive owing to their rigid, planar architectures, which promote electronic communication between the heteroatoms and the π-system. Shuford reported theoretical studies on the aromaticity of arsaborins.108 We synthesized the first dibenzoarsaborins and evaluated their photophysical properties (Fig. 5b).109 At 298 K, these compounds fluoresced with large Stokes shifts, attributed to significant structural relaxation. At 77 K, strong phosphorescence was observed, driven by the heavy-atom effect of arsenic. A thiophene-fused arsaborin derivative displayed temperature-dependent emission color changes in the solid state, resulting from shifts in the contributions of monomer fluorescence, excimer fluorescence, and phosphorescence.
4. Polymer materials
4.1. Conjugated polymers
The first structurally characterized arsenic-containing conjugated polymer is poly(vinylene arsine).110–113 This polymer is obtained by radical-initiated copolymerization of cyclooligoarsines (As5Me5 or As6Ph6) with terminal alkynes in the presence of AIBN. The resulting poly(vinylene arsine)s feature well-defined alternating sequences owing to a mechanism involving radical chain ring-opening alternating copolymerization (RCRAC) (Scheme 5). The AIBN-derived radicals cleave As–As bonds in the cyclic oligomers, forming reactive arsenic radicals. Ring opening leads to further homolytic bond cleavage, and these arsenic radicals add to terminal alkynes, generating vinyl radicals that propagate the copolymerization. Notably, the resulting polymers are soluble in common organic solvents, enabling full structural characterization by nuclear magnetic resonance spectroscopy, size exclusion chromatography, and other methods.
 |
| | Scheme 5 Mechanism of RCRAC. | |
Following these early reports, arsenic-containing conjugated polymers remained largely unexplored until 2016, when three independent studies, including our own, reignited interest in the field.78,79,114 Heeney and colleagues synthesized a dithienoarsole dibromide monomer and polymerized it via Stille coupling with trans-1,2-bis(tributylstannyl)ethene to produce a dithienoarsole-based polymer (Scheme 6a).78 This polymer exhibited promising hole mobility (0.08 cm2 V−1 s−1) owing to efficient solid-state aggregation and crystallinity, along with notable ambient stability. Simultaneously, our group developed a dithienoarsole polymer via Suzuki–Miyaura polycondensation of dithienoarsole dibromide with fluorene diboronic acid (Scheme 6b).79 The resulting polymer displayed strong yellow fluorescence in solution with high quantum yield (Φ = 0.44).
 |
| | Scheme 6 Synthesis of arsole polymers reported in 2016. (a) Microwave-assisted Stille coupling, (b) Suzuki–Miyaura coupling, and (c) post-polymerization functionalization with titanocyclopentadiene polymer. | |
We also collaborated with Tomita to synthesize another type of arsole polymer via post-polymerization functionalization.114 A polymer bearing titanocyclopentadiene units was reacted with PhAsI2 to afford an arsole polymer (Scheme 6c). The same method successfully produced a phosphole analog for comparison.115 Cyclic voltammetry revealed that the arsole polymer exhibited quasi-reversible redox behavior, while the corresponding phosphole polymer irreversibly degraded, highlighting the greater oxidative robustness of the As(III) center.
In 2017, Kuehne and Orthaber independently expanded the family of arsenic-containing π-conjugated polymers.116,117 Kuehne and colleagues synthesized heterofluorene–fluorene copolymers incorporating various heteroatoms (Si, Ge, N, As, Se, and Te) into the central fluorene unit (Scheme 7a).116 The arsafluorene-based polymer, synthesized from arsafluorene dibromide, emitted blue light at 458 nm with a moderate quantum yield (Φ = 0.11). Remarkably, among the series, only the arsafluorene and germafluorene copolymers demonstrated amplified spontaneous emission (ASE) upon nanosecond laser excitation, positioning arsenic-based systems as strong candidates for photonic and laser applications. Orthaber and colleagues further synthesized a thiophene-substituted arsaalkene monomer and electropolymerized it (Scheme 7b).117 Chronoamperometry of the resulting 2 μm thick film revealed a reversible electrochromic response at 600 nm with a high contrast ratio of 58%, maintained over multiple cycles. The doped polymer remained stable for several days under inert conditions.
 |
| | Scheme 7 Synthesis of (a) arsafluorene polymer, (b) arsaalkene polymer via electropolymerization, (c) dithienoarsole polymers (TBAB = tetrabutylammonium bromide), (d) dithienoarsole polymer via electropolymerization, (e) dithienoarsole polymer for OPV, and (f) NIR-emissive polymers. | |
A key advantage of arsenic-based π-conjugated polymers lies in their low propensity for metal contamination. In typical transition-metal-catalyzed polycondensations (e.g., Suzuki–Miyaura, Sonogashira–Hagihara), residual metal catalysts can be problematic, particularly owing to the coordinating ability of As(III). However, we found that palladium and copper residues can be effectively removed using scavenger resins post-polymerization (Scheme 7c).118 X-ray fluorescence spectroscopy confirmed negligible metal content, indicating that As(III) exhibits weaker coordination than phosphorus, facilitating purification.
The electrochemical stability of arsenic-containing conjugated units has also been validated. Hayashi and our group demonstrated that while dithienoarsole is electrochemically inert in the presence of tetraethylammonium tetrafluoroborate (Et4NBF4), boron trifluoride diethyl etherate (BF3·OEt3) lowers its oxidation potential, enabling electropolymerization to produce a uniform homopolymer film on an ITO substrate (Scheme 7d).119
Dithienoarsole–benzodithiophene copolymers have been evaluated as donor materials in organic photovoltaic (OPV) devices (Scheme 7e).120 Synthesized via Stille coupling between dithienoarsole dibromide and stannylated benzodithiophene, these polymers exhibited good ambient stability—an essential property for OPV materials. The power conversion efficiency (PCE) reached 3.90%, with open-circuit voltages as high as 0.91 V. Variations in alkyl chain length influenced PCE, solubility, and morphology. Further functionalization of dithienoarsole monomers with long alkyl chains enhanced solubility and backbone planarity.121 Copolymerization with electron-deficient units such as benzothiadiazole or benzoxadiazole afforded donor–acceptor polymers, exhibiting near-infrared emission (Scheme 7e). The benzoxadiazole copolymer emitted at 886 nm.
Singlet oxygen (1O2) photosensitizers are a particularly noteworthy application of dithienoarsole–fluorene copolymers (Fig. 6).122 The polymer showed an exceptional 1O2 quantum yield of 0.54, the highest reported for a single-component π-conjugated polymer, and outperformed arsenic-free analogs (e.g., 0.14 for bithiophene–fluorene copolymer). The 1O2-generation efficiency is attributed to the heavy atom effect of arsenic. Importantly, the polymer backbone remains photostable under 1O2-generating conditions. Under continuous pulsed laser excitation, two dithienoarsole polymers exhibited sustained ASE for over 15 hours at pump energies up to 28.9 μJ.123 Under similar conditions, rhodamine 6G rapidly degraded, highlighting the superior photostability of the arsenic-based systems.
 |
| | Fig. 6 Singlet oxygen generation using dithienoarsole polymer as a photosensitizer. | |
Additionally, we synthesized a novel arsole monomer, 2,3-diarylbenzoarsole dibromide, via a zirconacycle intermediate.124 This monomer was copolymerized with fluorene diboronic acid using Suzuki–Miyaura coupling to yield a main-chain polymer (Scheme 8a). A side-chain polymer analog was also synthesized via ring-opening metathesis polymerization of a norbornyl-substituted monomer using Grubbs’ third-generation catalyst (Scheme 8b). In the solid-state, the main-chain polymer exhibited aggregation-caused quenching owing to π–π stacking, while the side-chain polymer showed aggregation-induced emission enhancement owing to restricted intramolecular motion. Furthermore, 2,3-diarylbenzoarsoles possess As-stereogenic centers, and optical resolution of 2,3-diphenylbenzoarsole was achieved, indicating that their chiroptical properties warrant further exploration in polymeric systems.
 |
| | Scheme 8 Synthesis and quantum yields of (a) main-chain- and (b) side-chain-type 2,3-diphenylbenzoarsole polymers. | |
4.2. Reactive polymers
Polymers bearing arsonic acid functionalities have been studied since the 1970s, primarily as metal-binding agents and polyelectrolytes owing to their ionic character.125–127 In this minireview, we highlight recent developments in side-chain-type polymers that exploit the chemical reactivity of arsenic atoms. Work in this area has been pioneered by Wilson and colleagues.128–131 They synthesized diblock copolymers comprising poly(ethylene glycol) methyl acrylate (PEGA), N-isopropylacrylamide (NIPAM), and phenylarsonic acid-functionalized acrylamide (AsAm) segments via living radical polymerization. Upon reduction of the arsonic acid groups with phosphinic acid (H3PO2), cross-linking occurs between NIPAM and AsAm segments through As–As bond formation, resulting in the formation of polymeric nanoparticles (Scheme 9a).129 A similar approach was applied to prepare hydrogels: random copolymers of NIPAM and AsAm were synthesized by free-radical polymerization, and their reduction with H3PO2 generated As–As-linked networks, producing hydrogels.130 Furthermore, p-arsanilic acid could be incorporated into the hydrogel matrices and released under oxidative conditions, demonstrating potential as a redox-responsive drug delivery system.131
 |
| | Scheme 9 Reactive polymers having arsenic side chains for (a) cross-linking and (b) catalytic arsa-Wittig reaction. VA-044: 2,2′-azobis[2-(2-imidazolin-2-yl)propane]dihydrochloride; TCP = tetra(p-chlorophenyl)porphyrinate; PMHS = polymethylhydrosiloxane. | |
Tang and colleagues developed a different class of reactive polymers by introducing diphenylarsine (AsPh2) moieties into polymer side chain for use in catalytic arsa-Wittig reactions (see section 6.1 and Scheme 9b).132 While free AsPh3 is weakly nucleophilic and requires activation with Fe(TCP)Cl (TCP = tetra(p-chlorophenyl)porphyrinate) for ylide generation from diazo compounds, immobilizing the AsPh2 units on a polymer backbone enhances recyclability. Ethylene–diphenyl(undec-10-enyl)arsine copolymers were synthesized by coordination polymerization and exhibited high catalytic activity and E-selectivity in the arsa-Wittig reaction of ketones with diazo compounds in the presence of Fe(TCP)Cl and polymethylhydrosiloxane (PMHS). These polymer-supported catalysts retained their activities over at least five cycles.
4.3. Coordination polymers
MOFs are a class of crystalline porous materials with wide-ranging applications in gas storage, separation, catalysis, and sensing. Post-synthetic functionalization (PSF) is a powerful strategy to enhance MOF functionality and complexity.133,134 MOFs that feature coordination-free donor sites within their pores are particularly attractive for incorporating metal catalysts. However, protecting groups are often required during MOF synthesis to prevent premature coordination. Since deprotection post-assembly can be challenging, protection-free strategies are highly desirable. Phosphines have been explored for PSF owing to their soft Lewis basicity and limited interaction with hard metal nodes. However, their air sensitivity and tendency to oxidize during MOF synthesis necessitate stringent anaerobic conditions. By contrast, arsenic offers a promising alternative: it has soft Lewis basicity like phosphorus but exhibits significantly lower oxophilicity, making it more compatible with aerobic MOF construction.
To demonstrate this concept, we designed a novel ligand, cis-1,4-dihydro-1,4-diarsinine (cis-DHDA) tetracarboxylic acid, which bears four carboxylate groups for MOF construction and two arsenic centers reserved for post-synthetic metal coordination (Scheme 10a).135 Although a copper(II)-based MOF was successfully formed from this ligand, its crystalline structure collapsed upon solvent removal, likely owing to insufficient framework rigidity. More successful implementation of arsenic-centered PSF was reported by Humphrey and colleagues.136 They synthesized ACM-1, a coordination polymer formed from a pyridyl-functionalized triarylarsine ligand and Ni(II) cations (Scheme 10b). In this framework, two Ni(II) cations are bridged by carboxylate anions and a μ2-OH, while the N-donor sites of the arsenic-containing ligand coordinate to the Ni(II) centers. Notably, the uncoordinated arsenic atoms within the framework remained accessible and were successfully functionalized with gold(I) chloride in a post-synthetic modification step.
 |
| | Scheme 10 Synthesis and crystal structures of MOFs bearing (a) DHDA, (b) pyridyl-functionalized triarylarsine, and (c) dipyridinoarsole ligands (guest: solvent (DMF) or gas (N2 and CO2) molecules). The hydrogen atoms are omitted for clarity in the crystal structures. | |
In another example, Zn(II)-based MOFs incorporating dipyridinoarsole ligands were constructed using the two pyridyl groups for framework formation,86 leaving the arsenic atoms uncoordinated and available for PSF (Scheme 10c). This study revealed notable structural flexibility and gas adsorption behavior. Remarkably, a phenyl-substituted dipyridinoarsole formed a “breathing” MOF, exhibiting reversible pore opening and closing in response to guest molecules adsorption/desorption. By contrast, the MOFs using methyl-substituted dipyridinoarsole or phenyl-substituted dipyridinophosphole molecules partially amorphized upon solvent removal and did not exhibit breathing behavior. These results indicate that the steric bulk of the phenyl group at the arsenic atom plays a critical role in stabilizing the closed-pore conformation.
5. Ligands for metal complexes
5.1. Ligand library
Unlike the broad availability and commercial accessibility of phosphine ligands, structurally diverse arsine ligands remain scarce. Beyond AsPh3, most arsine ligands are synthetically challenging to access, limiting the systematic investigation of structure–property relationships in arsine-ligated metal complexes. As a result, the design of functional coordination compounds leveraging the unique electronic and steric properties of arsenic has long been underexplored. However, the recent development of practical and modular synthetic methods for organoarsenic compounds has significantly advanced the field.
Cyclooligoarsines Me5As5 and Ph6As6 serve as key precursors for the synthesis of structurally diverse arsine ligands. During the RCRAC process, vinyl radicals can be stabilized by electron-withdrawing substituents (e.g., esters), promoting Z-to-E isomerization.137,138 Subsequent dimerization of the E-form radicals yields cyclic diarsenic compounds known as cis-1,4-dihydro-1,4-diarsinines (cis-DHDAs). These cyclic ligands have been coordinated with a variety of transition metals, including Pt(II), Pd(II), Au(I), and Cu(I) (Scheme 11).139–146 The cis-DHDA framework holds two arsenic donor atoms at a fixed short distance (∼3 Å), offering a rigid template for the construction of di- or multinuclear metal complexes.
 |
| | Scheme 11 Synthesis of cis-DHDA-ligated transition metal complexes. | |
The arsenic electrophile PhAsI2, which is readily generated from Ph6As6, has been used to prepare diarylphenylarsines (Fig. 7a).147 Moreover, dithiaarsole can provide further versatile Cs-symmetric ligands such as arsa-Buchwald ligands (Fig. 7b).29 In turn, nucleophilic substitution using organolithium reagents (RLi) with Ph6As6 furnishes RPhAsLi species, which can be elaborated into C1-symmetric arsines bearing a stereogenic arsenic center (Fig. 7c).45 These nucleophiles are also suitable for preparing bidentate arsine ligands. Additionally, Ph2AsI, accessible via oxidative cleavage of Ph2AsMe, provides a route to chelating diphenylarsine-type ligands. These synthetic platforms have enabled access to arsenic analogs of well-established phosphine ligands such as dppe, xantphos, and dppf, thereby extending the structural diversity of arsine ligands available for coordination chemistry (Fig. 7d).148
 |
| | Fig. 7 Arsenic ligand library. (a) Diarylphenylarsines, (b) arsa-Buchwald ligands, (c) C1-symmetric ligand, (d) bidentate ligands, (e) NAN-ligand, (f) arsafluorene dimer, and (g) arsanylborane. | |
A new class of chelating ligands has been established through the synthesis of phenyldiquinolinylarsine, a tridentate N–As–N (NAN) pincer-type ligand (Fig. 7e).149 Copper(I) halide complexes formed with this ligand display highly distorted geometries. The steric strain typically associated with tridentate ligand binding is alleviated around the arsenic center, which exhibits less directional coordination behavior than phosphorus, reflecting the softer and more diffuse lone pair of As(III).
Orthaber and colleagues also reported arsafluorene dimers linked via phenylene bridges on the arsenic centers (Fig. 7f).150 The flexibility of these ligand backbones permits the formation of both mono- and dinuclear metal complexes. Remarkably, a unique rearrangement of an allylic intermediate was observed during complexation, leading to the formation of an arsapalladacycle—a structural motif rarely seen in organoarsenic chemistry.
Scheer and co-workers demonstrated the versatility of arsanylborane complexes (Fig. 7g) through selective functionalization and coordination. For instance, Lewis acid/base-stabilized phosphanyl- and arsanylboranes undergo selective dihalogenation at the pnictogen atom with CX4 (X = Cl, Br), affording novel complexes comprehensively characterized by spectroscopy and X-ray crystallography.151 Moreover, gold(I) complexes bearing phosphanyl- and arsanylborane ligands exhibit unsupported Au⋯Au interactions and one-dimensional chain formation in the solid state, leading to intriguing luminescence properties such as strong temperature-dependent red shifts correlated with decreasing Au⋯Au distances.152 In addition, a tert-butyl-substituted arsanylborane has been developed as a versatile building block, showing diverse reactivity toward Lewis acids and transition metals, and enabling the first synthesis of oligomeric arsanylboranes under thermal conditions.153
Functionally tailored ligands have also been developed using cyclooligoarsine-based methodologies. In particular, luminescent metal complexes and transition-metal catalysts derived from these ligands are discussed in subsequent sections. We designed dibenzoarsacrowns, in which an oxygen atom in conventional crown ethers is replaced with an arsenic atom (Fig. 8a).154 The soft Lewis basicity of arsenic allows for selective coordination to late transition metals, while the oxygen atoms retain their strong binding affinity for alkali metal cations (Fig. 8b). Intriguingly, pre-coordination of the arsenic atom to AuCl enhances the alkali metal binding constant, demonstrating a positive allosteric effect. Moreover, weak secondary interactions between the arsenic atom and the encapsulated alkali metal modulate the redox behavior of the crown structure, enabling electrochemical sensing of alkali metal ions.155 Additionally, arsafluorene dimers linked via oligo(ethylene glycol) chains can form enchained-type metallacrown architectures through coordination to Pt(II) dihalides (Fig. 8c).156 The geometric structure around the Pt(II) center and its phosphorescent properties were found to be strongly influenced by the coordinated halide anions.
 |
| | Fig. 8 (a) Chemical structures of dibenzoarsacrowns. (b) Coordination behaviors of 21-dibenzoarsacrown-7 and crystalline structures of the resultant complexes. (c) Construction of enchained-type metallacrown ethers with arsafluorene-dimer. | |
5.2. Luminescent materials
Luminescent transition metal complexes bearing arsine or arsine oxide ligands have been extensively investigated. Recent advances in arsine ligand synthesis have enabled systematic studies on the relationship between ligand structure and photophysical properties.157 Complexes of platinum(II), gold(I), silver(I), and copper(I) with structurally diverse arsine ligands have been synthesized, and in some cases, they exhibit highly efficient room-temperature phosphorescence or thermally activated delayed fluorescence.144,145,158–168 This is likely attributed to the heavy-atom effect of arsenic, which enhances spin–orbit coupling and promotes ISC from the singlet to the triplet excited state. For example, the platinum(II) complex trans-[PtI2(9-phenyl-9-arsafluorene)2] has a phosphorescence quantum yield of 0.52 in the solid state at room temperature.163 Substituent variation on the arsenic center via a straightforward substitution route (Scheme 1g) revealed that both the emissive properties and the cis/trans geometry of the complexes are strongly affected.158 Notably, only trans-isomers with specific ligand/halogen combinations exhibited significant emission, whereas all cis-isomers were non-emissive.
Oligomeric and polymeric architectures have also been proven to exhibit strong phosphorescence. A tetranuclear complex incorporating Au(I)⋯Cu(I) double salts, stabilized by homo- and heterometallophilic interactions, was constructed using the bidentate ligand dpam (for dpam, please see Fig. 7d).167 This complex formed three distinct crystalline polymorphs depending on the recrystallization solvent. All polymorphs exhibited remarkably strong room-temperature phosphorescence, with quantum yields reaching up to 0.97, and demonstrated a broad range of emission colors, including near-infrared and white-light emission. Interestingly, these polymorphs could be reversibly interconverted by external stimuli. Additionally, coordination polymers provide a valuable strategy for achieving intense luminescence. Dinuclear rhomboid Cu2I2 clusters coordinated by AsPh3 were linked through various bidentate N-heteroaromatic ligands, yielding strongly emissive coordination polymers with quantum yields up to 0.60.159 The emission color could be finely tuned across the green-to-red spectrum by altering the bridging ligands.
Beyond emission efficiency, arsine-ligated metal complexes have also demonstrated stimui-responsive luminescence, particularly in crystalline states. The relatively weak and less directional coordination ability of arsine ligands, compared to phosphines, is thought to impart greater structural flexibility, which facilitates crystal-to-crystal transformations. This behavior is particularly pronounced in solvent vapor-responsive crystals, as exemplified below. Crystals of trans-[PtBr2(9-phenyl-9-arsafluorene)2] form nonporous molecular crystals (NMCs) without apparent 1D channels.169 Crystals grown from chlorobenzene (PhCl) are non-emissive owing to encapsulation of PhCl, which acts as a quencher. Exposure to volatile organic compound (VOC) vapors triggers crystal-to-crystal transformation, releasing the PhCl molecules and restoring phosphorescence (Fig. 9a). Remarkably, the NMC exhibits shape-selective molecular recognition: VOCs with smaller minimum cross-sectional diameters induce emission recovery, while bulkier VOCs do not. A broad range of VOCs—including alcohols, ethers, haloalkanes, and alkanes—are applicable.
 |
| | Fig. 9 Stimuli-responsive crystals. (a) Molecular shape recognition by NMC, (b) switching between porous and non-porous molecular crystals, and (c) MeOH vapor sensing. | |
A structurally distinct example involves the compound trans-[PtI2(9-pentafluorophenyl-9-arsafluorene)2].170 When crystallized from o-dichlorobenzene, it forms a molecular porous crystal (MPC), while an NMC is obtained from a CH2Cl2/hexane mixture. The MPC is non-emissive, whereas the NMC is emissive. These two forms are reversibly interconvertible via solvent vapor exposure, enabling visible luminescence switching (Fig. 9b). Notably, the porous framework of the MPC remains intact even after complete solvent removal, meaning that reversible pore opening and closing is possible without structural collapse.
Another notable case is trans-[PtCl2(21-dibenzoarsacrown-7)2], which functions as a turn-on luminescent methanol (MeOH) sensor.171 Crystals obtained from CH2Cl2/n-hexane mixtures are non-emissive, while those crystallized from CH2Cl2/MeOH are emissive. X-ray crystallography indicates that MeOH molecules are incorporated via hydrogen bonding with the crown ether moieties, restricting molecular motions and thereby suppressing non-radiative decay. The crystal shows fast and reversible luminescence switching upon exposure to MeOH vapor (Fig. 9c): emission turns on within ∼30 s and off within ∼90 s. This MeOH selectivity is exceptionally high, and the sensing performance remains stable over at least five cycles.
In addition to As(III) ligands, arsine oxides (As(V)) have also been explored as metal-coordinating ligands. Phosphine oxides are widely recognized as effective antenna ligands for sensitizing lanthanide (Ln3+) emission owing to their low vibrational frequencies (∼1150–1160 cm−1), which reduce non-radiative decay. Furthermore, they induce asymmetric coordination environments, promoting radiative f–f transitions. In this context, arsine oxides (As
O) are attractive alternatives. Their vibrational frequencies (900–920 cm−1) are comparable to that of P
O, and their higher polarizabilities, stemming from the lower electronegativity of arsenic, enhances ligand-to-metal charge transfer according to the dynamic coupling theory.
We synthesized europium(III) complexes bearing arsine oxide ligands.172–175 For instance, a Eu(NO3)3 complex with triphenylarsine oxide (O
AsPh3) exhibited a 7.9-fold enhancement in photosensitized energy-transfer efficiency compared with its O
PPh3 analog.172 Photophysical and computational studies revealed that the T1 state of O
AsPh3 dominates the energy-transfer process, supported by the heavy-atom effect of arsenic. These complexes also show exceptional thermal durability, maintaining luminescence at elevated temperatures. In parallel, Layfield reported the synthesis and structural characterization of yttrium complexes bearing arsine, arsenide, and the first rare-earth arsinidene ligand, obtained through stepwise deprotonation of a primary arsine precursor.176
Bidentate arsine oxide ligands have also been applied to lanthanide coordination polymers (Fig. 10).175 Eu(hfa)3 (hfa = hexafluoroacetylacetonate) polymers bridged by bisarsine oxide linkers displayed high quantum yields, enhanced energy transfer efficiency, and excellent thermal stability. Emission intensity increased upon heating from 300 K to 400 K, indicating heat-induced luminescence enhancement—a rare phenomenon attributable to the robust polymeric structure and favorable coordination environment supported by the arsine oxide ligands. Notably, intense luminescence was sustained, even at 550 K.
 |
| | Fig. 10 Thermally durable emission of bisarsine oxide-bridged Eu(hfa)3 complex. | |
5.3. Transition metal catalysts
In transition metal catalysis, arsine ligands have been anticipated to offer significant advantages over traditional phosphine-based systems, potentially enabling breakthroughs in reactivity and selectivity. A landmark study by Farina demonstrated that a palladium catalyst ligated with AsPh3 accelerated the Stille coupling of iodobenzene and tributylvinyltin by three orders of magnitude compared with its PPh3 counterpart (Scheme 12a).177 This dramatic rate enhancement is likely due to the inherently weaker σ-donating ability of arsine ligands, which facilitates more efficient reductive elimination. In the field of hydroformylation, van Leeuwen and co-workers reported that newly designed wide-bite-angle arsine ligands based on the xantphos scaffold exhibited unprecedented activity and selectivity in platinum/tin-catalyzed hydroformylation of terminal alkenes (Scheme 12b).178 This study provided the first clear demonstration that arsine-based ligands can outperform their phosphine counterparts in industrially relevant hydroformylation processes. Shibasaki et al. developed an axially chiral arsine ligand, BINAAs, the arsenic analog of BINAP (2,2′-bis(diphenylphosphino)-1,1′-binaphthyl).179 In the intramolecular asymmetric Heck reaction, BINAAs provided both higher yields and enantioselectivity compared with phosphine analogs, highlighting the potential of chiral arsine ligands in asymmetric catalysis (Scheme 12c). Nishibayashi reported an arsenic–nitrogen–arsenic (ANA) pincer ligand as a structural analog of phosphorus-based PNP pincers.180 Ruthenium complexes bearing the ANA framework showed high efficiency and selectivity in dehydrogenation reactions (Scheme 12d), expanding the scope of pincer-type catalyst design.181 Dong and colleagues, known for developing Pd/norbornene cooperative catalysis platforms, discovered that AsPh3 was essential to promoting vicinal difunctionalization of thiophenes (Scheme 12e). Again, this underscores the unique electronic contributions of arsine ligands in catalytic cycles that require delicate control over elementary steps such as oxidative addition and reductive elimination.
 |
| | Scheme 12 Reactions with arsine-ligated transition metal catalysts. (a) Stille coupling reaction, (b) hydroformylation, (c) asymmetric Heck reaction, (d) dehydrogenative coupling reaction of alcohol and amine, and (e) C–H difunctionalization of thiophene. | |
The expanding arsine ligand library, encompassing monodentate, bidentate, and even tridentate structures, provides a valuable platform for investigating the structure–reactivity relationships in transition metal catalysis. Numerous C–C bond-forming reactions—such as Mizoroki–Heck, Stille, Suzuki–Miyaura, and C–H activation—have been explored using structurally diverse arsine ligands.28,29,147,148,182–184 A particularly illustrative example is the development of arsa-Buchwald ligands, arsenic analogs of the widely used Buchwald-type phosphines. In the Suzuki–Miyaura coupling of sterically hindered substituents, these ligands exhibited excellent performance (Scheme 13).182 This behavior can be rationalized in terms of steric parameters, particularly the percent buried volume (%Vbur), which quantifies the steric congestion around the metal center. The %Vbur of SArs—a structural analog of SPhos—is 27.9%, significantly smaller than that of SPhos (35.1%). This indicates that arsine ligands afford a more sterically open catalytic site, which can facilitate the approach and transformation of bulky substrates.
 |
| | Scheme 13 Suzuki–Miyaura coupling reaction using arsa-Buchwald ligand. | |
In summary, the incorporation of arsine ligands into transition metal catalytic systems introduces a distinct steric and electronic landscape, enabling reactivity and selectivity patterns that are challenging to achieve using conventional phosphine ligands. As the arsine ligand toolbox continues to expand, new applications in homogeneous catalysis are expected.
6. Reactive species
6.1. Catalytic Wittig reaction
The Wittig olefination is a cornerstone transformation in organic synthesis, enabling the construction of C
C bonds to access highly functionalized or structurally complex molecules. Since the development of the catalytic variant by O'Brien in 2009, which relies on the reductive regeneration of phosphine oxides, extensive efforts have been made to design efficient phosphine catalysts.185 However, notably, the arsa-Wittig reaction—based on As(III)/As(V) redox cycling—was explored prior to O'Brien's report. Early systems employed simple aryl and alkyl arsines such as AsPh3 (Scheme 14a)186 and tributylarsine (AsBu3) (Scheme 14b),187 but these were limited by low reactivity and poor stability. Owing to their weak nucleophilicity, arsines like AsPh3 often require iron co-catalysts, and AsBu3 suffers from volatility and air sensitivity, hindering practical applications.
 |
| | Scheme 14 Catalytic arsa-Wittig reactions using (a) AsPh3, (b) AsBu3, and (c) tris(p-(dimethylamino)phenyl)arsine as arsine catalysts. | |
Recent advances have revitalized interest in arsine-mediated olefination. We designed synthetically accessible arsines that enable the development of electron-rich cyclic and acyclic arsine catalysts.188,189 These new catalysts promote arsa-Wittig reactions under mild, metal-free conditions, even at room temperature. Among them, tris(p-(dimethylamino)phenyl)arsine demonstrated high catalytic efficiency and excellent E-selectivity, offering a promising platform for arsine-based catalytic systems (Scheme 14c).
6.2. Redox behavior
Phosphine-mediated redox reactions, such as the Wittig, Appel, Staudinger, and Mitsunobu reactions, are fundamental to synthetic organic chemistry.190–192 These transformations rely on the high oxophilicity of phosphorus, which facilitates the formation of phosphine oxides during the oxidative step. However, the reduction of phosphine oxides is often energetically demanding and typically requires strong reductants (e.g., silanes). Cyclic phosphine designs help overcome this barrier by enhancing nucleophilicity and introducing ring strain that destabilizes the phosphine oxide.
By contrast, arsenic exhibits significantly lower oxophilicity, making it a compelling alternative for redox-mediated transformations. In 1986, Abalonin reported that O
AsPh3 could oxidize benzyl bromide to benzaldehyde, concomitantly regenerating AsPh3193—behavior not observed for O
PPh3. Inspired by this precedent, we developed an arsine-based oxygen-transfer system involving complementary reactions (Scheme 15): the arsa-Appel reaction, where As(III) is oxidized to As(V), and benzyl bromide oxidation, which reduces As(V) back to As(III).194 Although the process currently exhibits modest efficiency, it provides a conceptual framework for future arsine-mediated redox catalysis.
 |
| | Scheme 15 Sequential oxygen atom transfer using arsines. | |
6.3. Frustrated Lewis pairs
Lewis acid–base chemistry is classically categorized into classical Lewis adducts (CLAs) and frustrated Lewis pairs (FLPs).195,196 In 1942, Brown reported that steric hindrance between lutidine and trimethylborane prevented adduct formation—a phenomenon later generalized by Stephan in 2006, who showed that sterically encumbered FLPs could activate dihydrogen (H2). This seminal discovery catalyzed rapid growth in FLP research, particularly in catalysis, small-molecule activation, and CO2 capture. While most FLP systems utilize phosphines and boranes, FLPs involving arsines as the Lewis base are much less explored. Ketkov et al. reported the crystal structure and theoretical analysis of the CLA formed between AsPh3 and tris(pentafluorophenyl)borane (B(C6F5)3)—one of the few well-characterized arsine-containing adducts.197
We synthesized a series of tertiary arsines with varying steric and electronic properties to systematically investigate their interactions with B(C6F5)3.198 Depending on the ligand structure, the resulting Lewis pairs exhibited behavior consistent with either CLAs or FLPs (Scheme 16a). For example, trimesitylarsine and B(C6F5)3 form an FLP-like encounter complex in solution. These systems were reactive toward small molecules such as phenylacetylene and diphenyl disulfide, yielding arsonium borate and thioarsonium thioborate salts, respectively (Scheme 16b and c). It should be emphasized that the relatively weak Lewis basicity of arsines generally precludes H2 activation, which is a hallmark reaction of FLP chemistry. Nevertheless, this limitation also opens opportunities: the modest reactivity of arsenic-based FLPs may enable controlled, selective transformations that would be unachievable with more reactive phosphorus analogs. Further exploration of reactivity modulation through ligand design and cooperative effects is currently underway.
 |
| | Scheme 16 (a) Lewis pair formation behavior between arsines and B(C6F5)3. Reactions of tricyclohexylarsine (AsCy3) with (b) phenylacetylene and (c) diphenyl disulfide in the presence of B(C6F5)3. | |
7. Ionic species
7.1. Arsonium cations
Quaternary arsonium cations occur naturally in compounds such as arsenobetaine, a well-known non-toxic organoarsenic species.199 However, synthetic arsonium cations have received relatively little attention in the field of materials science, such as ionic liquids (ILs)200–202 and conjugated molecules.203 Phosphonium cations tend to reduce the viscosity of ILs relative to their ammonium counterparts owing to the larger atomic radius of phosphorus.204–206 By extension, arsonium cations, which are even larger and less directional, are expected to further lower the viscosity and improve fluidity in IL systems. May et al. reported the electrochemical properties of tetramethylarsonium bis(trifluoromethylsulfonyl)amide (TFSA), although the salt exhibited a relatively high melting point (∼140 °C), limiting its applicability.207
To expand on this concept, we synthesized a novel arsonium-based IL, trihexylmethylarsonium TFSA, and compared its physicochemical properties with the corresponding trihexylmethylphosphonium IL.208 The arsonium IL showed a lower glass transition temperature, reduced viscosity, and higher ionic conductivity than the phosphonium analog (Fig. 11). These enhancements are attributed to the greater conformational flexibility of the arsonium cation, which reduces lattice energy and promotes ion mobility. Although the arsonium IL exhibited a slightly narrower electrochemical window, its thermal stability and alkali resistance were comparable to those of the phosphonium counterpart. These findings highlight the potential of arsonium cations as promising building blocks for high-performance, low-viscosity ILs.
 |
| | Fig. 11 Properties of ionic liquids containing arsonium and phosphonium cations. (a) Variable-temperature viscosity. (b) Arrhenius plots of ionic conductivities. | |
7.2. Dications
Main-group-element polycations have gained increasing interest owing to their applications in synthesis, Lewis acid catalysis, and organocatalysis. A common stabilization strategy involves coordination by neutral donor ligands, which facilitate hypercoordination at the polycationic center. While polycations of heavier pnictogens, such as Sb and Bi, have been extensively studied—exemplified by donor-stabilized species such as [Ph3Sb]2+ and [Ph3Bi]2+, reported by Burford209,210—the corresponding arsenic-centered dications have remained underexplored owing to the smaller atomic radius of arsenic, which limits its ability to accept additional ligands.
Weigand and colleagues addressed this challenge by synthesizing [Ph3As]2+ species stabilized through tetracoordination with DMAP and IDipp (1,3-bis(2,6-diisopropylphenyl)imidazol-2-ylidene) (Fig. 12a).211 These arsenic dications pentacoordinated through two triflate (TfO−) anions. Pentacoordinated arsenic-centered dications were isolated using neutral donor ligands such as DMAP and pyridine (py), marking the first examples of such species stabilized without anionic frameworks (Fig. 12b).212 Furthermore, Stephan synthesized a toluene-coordinated pentamethylcyclopentadienyl arsenic dication exhibiting η5- and π-coordination (Fig. 12c).213
 |
| | Fig. 12 Arsenic dications with (a) tetracoordination, (b) pentacoordination, (c) η5- and π-coordination, (d) divinyldiarsene motif, and (e) hexacoordination. | |
Ghadwal and co-workers developed the first crystalline diarsene radical cations and dications (Fig. 12d) through the stepwise one-electron oxidation of NHC-stabilized divinyldiarsenes with gallium trichloride (GaCl3).214 The radical cations [{(NHC)C(Ph)}As]2(GaCl4) and the corresponding dications [{(NHC)C(Ph)}As]2(GaCl4)2 were isolated as stable crystalline solids. Structural analyses revealed sequential elongation of the As
As bond and contraction of the C–As bonds upon oxidation, accompanied by distinct red-shifted absorption bands and characteristic EPR signals. Computational studies indicated that the unpaired electron in the radical cations is delocalized over the conjugated CAs2C-framework, with spin density mainly located on the arsenic and vinylic carbon atoms. These findings establish diarsene radical cations and dications as rare and well-defined main-group radical species, expanding the frontier of arsenic-centered redox chemistry.
For hexacoordinated arsenic dictations, an intramolecular chelation strategy was employed using three 2-(2-pyridyl)phenyl (ppy) ligands (Fig. 12e). Sakabe, Sato, and our group successfully synthesized hexacoordinated arsenic-centered dication salts [(ppy)3As]Cl2 and [(ppy)3As](SbCl6)2.215 These compounds exhibit exceptional thermal and hydrolytic stabilities, remaining intact at elevated temperatures and in aqueous media. Both experimental characterization and computational analyses indicate that the remarkable stability originates from strong As–N coordination, which effectively rigidifies the complex. Surprisingly, these As-centered dications were more thermally stable than their heavier analogs, such as [(ppy)3Sb](SbCl6)2,216 despite antimony's larger size and common association with hypercoordination. These findings reveal that with appropriate ligand design, arsenic-centered dications can match or even outperform their heavier congeners in terms of stability and structural robustness, broadening the scope of polycationic main-group chemistry.
8. Final remarks
Over the past decades, organoarsenic chemistry has undergone a profound transformation—from a field once stigmatized by its association with toxicity and warfare to a rapidly evolving area of functional molecular science. The recent development of practical, safe, and modular synthetic methods has opened the door to a broad range of structurally and electronically diverse organoarsenic compounds. As highlighted in this review, these advances have given rise to functional organoarsenic materials with emerging applications in catalysis, photophysics, redox chemistry, molecular recognition, and ionic materials. Arsenic's distinctive chemical features—including its moderate Lewis basicity, low oxophilicity, heavy-atom character, and flexible coordination geometry—are not merely alternatives to those of phosphorus but offer orthogonal design elements that expand the chemical space of molecular and materials chemistry. Notably, arsine-based ligands are now recognized for enabling catalytic transformations that are inaccessible or inefficient with traditional phosphines. In photofunctional materials, arsenic contributes to tunable and stimuli-responsive emission behaviors, as well as to thermal robustness. Furthermore, the development of arsonium-based ionic species and hypervalent arsenic-centered dications illustrates that the versatility of arsenic spans all oxidation states and structural types.
The rise of functional organoarsenic chemistry thus reflects more than a technical progression—it represents a paradigm shift in how we view the element arsenic: not solely as a hazard to be avoided, but as a resourceful and adaptable component of next-generation molecular design. With growing interest in element diversity, sustainable catalysis, and functional hybrid materials, arsenic is poised to occupy a strategic position in the design of innovative molecular systems. It is anticipated that future progress will be driven by interdisciplinary collaborations that integrate synthetic, computational, and application-oriented approaches. As new arsenic-based platforms continue to emerge, we expect the functional potential of this historically underutilized element to be further unlocked, advancing not only fundamental science but also real-world technologies.
Author contributions
H. Imoto: conceptualization, visualization, writing – original draft, writing – review and editing, project administration, supervision; C. Okochi: visualization, writing – original draft; K. Kikuchi: visualization, writing – original draft; K. Naka: writing – original draft, writing – review and editing, supervision.
Conflicts of interest
There are no conflicts to declare.
Data availability
No primary research results, software or code have been included and no new data were generated or analyzed as part of this review.
Acknowledgements
The project on functional organoarsenic chemistry has been supported by JST FOREST Program, Grant Number JPMJFR221K to HI. In addition, a part of this project has been supported by the establishment of university fellowships towards the creation of science technology innovation, Grant Number JPMJFS2124 to KK.
References
- M. Gon, K. Tanaka and Y. Chujo, Recent progress in the development of advanced element-block materials, Polym. J., 2018, 50, 109–126 CrossRef CAS.
- M. Gon, K. Tanaka and Y. Chujo, Creative synthesis of organicinorganic molecular hybrid materials, Bull. Chem. Soc. Jpn., 2017, 90, 463–474 CrossRef CAS.
- Y. Chujo and K. Tanaka, New polymeric materials based on element-blocks, Bull. Chem. Soc. Jpn., 2015, 88, 633–643 CrossRef CAS.
- F. Vidal and F. Jäkle, Functional Polymeric Materials Based on Main-Group Elements, Angew. Chem., Int. Ed., 2019, 58, 5846–5870 CrossRef CAS PubMed.
- D. P. Gates, Inorganic and organometallic polymers, Annu. Rep. Prog. Chem., Sect. A: Inorg. Chem., 2006, 102, 449–468 RSC.
- B. Ochiai, K. Kikuta, Y. Matsumura, H. Horikoshi, K. Furukawa, M. Miyamoto and Y. Nishimura, Transparent and Photochromic Material Prepared by Copolymerization of Bismuth(III) Methacrylate, ACS Appl. Polym. Mater., 2021, 3, 4419–4423 CrossRef CAS.
- S. M. Parke, M. A. B. Narreto, E. Hupf, R. McDonald, M. J. Ferguson, F. A. Hegmann and E. Rivard, Understanding the origin of phosphorescence in bismoles: A synthetic and computational study, Inorg. Chem., 2018, 57, 7536–7549 CrossRef CAS PubMed.
- S. M. Parke, E. Hupf, G. K. Matharu, I. de Aguiar, L. Xu, H. Yu, M. P. Boone, G. L. C. de Souza, R. McDonald, M. J. Ferguson, G. He, A. Brown and E. Rivard, Aerobic Solid State Red Phosphorescence from Benzobismole Monomers and Patternable Self-Assembled Block Copolymers, Angew. Chem., Int. Ed., 2018, 57, 14841–14846 CrossRef CAS PubMed.
- Y. Matsumura, H. Horikoshi, K. Furukawa, M. Miyamoto, Y. Nishimura and B. Ochiai, Synthesis of Bismuth-Containing Polymer Films with High Refractive Index and X-ray Shielding Property by Radical Polymerization of Styrylbismuthine Derivatives, ACS Macro Lett., 2022, 11, 723–726 CrossRef CAS PubMed.
- J. Emsley, Nature's building blocks: an AZ guide to the elements, Oxford University Press, 2011 Search PubMed.
- H. Imoto and K. Naka, The Dawn of Functional Organoarsenic Chemistry, Chem. – Eur. J., 2019, 25, 1883–1894 CrossRef CAS PubMed.
- H. Imoto, Development of macromolecules and supramolecules based on silicon and arsenic chemistries, Polym. J., 2018, 50, 837–846 CrossRef CAS.
- H. Imoto and K. Naka, Recent progress on arsenic-containing functional polymers, Polymer, 2022, 241, 124464 CrossRef CAS.
- W. Levason and G. Reid, in Comprehensive Coordination Chemistry II, Elsevier, 2003, pp. 377–389 Search PubMed.
- J. L. Wardell, Arsenic, Antimony and Bismuth, Compr. Organomet. Chem., 1982, 2, 681–707 Search PubMed.
- W. R. Cullen, Organoarsenic Chemistry, Adv. Organomet. Chem., 1966, 4, 145–242 CrossRef CAS.
- H. Schweitzer, Ehrlich's Chemotherapy—A New Science, Science, 1910, 32, 809–823 CrossRef CAS PubMed.
- N. C. Lloyd, H. W. Morgan, B. K. Nicholson and R. S. Ronimus, The composition of Ehrlich's Salvarsan: Resolution of a century-old debate, Angew. Chem., Int. Ed., 2005, 44, 941–944 CrossRef CAS PubMed.
- H. Q. Le and S. J. Knudsen, Exposure to a first world war blistering agent, Emerg. Med. J., 2006, 23, 296–299 CrossRef CAS PubMed.
- G. J. Fitzgerald, Chemical Warfare and Medical Response During World War I, Am. J. Public Health, 2008, 98, 611–625 CrossRef PubMed.
- J. Henriksson, A. Johannisson, P. A. Bergqvist and L. Norrgren, The toxicity of organoarsenic-based warfare agents: In vitro and in vivo studies, Arch. Environ. Contam. Toxicol., 1996, 30, 213–219 CrossRef CAS PubMed.
- W. S. Tay and S. A. Pullarkat, C–As Bond Formation Reactions for the Preparation of Organoarsenic(III) Compounds, Chem.
– Asian J., 2020, 15, 2428–2436 CrossRef CAS PubMed.
- R. Betz, M. M. Reichvilser, E. Schumi, C. Miller and P. Klüfers, Tetrachlorophenyl-λ5-arsane and tetramethoxyphenyl- λ5-arsane: Crystal structures, NMR spectra and bonding situation, Z. Anorg. Allg. Chem., 2009, 635, 1204–1208 CrossRef CAS.
- C. T. Mortimer and H. A. Skinner, 828. The thermochemistry of organo-arsenic compounds. Part I. Cacodyl, As2Me4, J. Chem. Soc., 1952, 4331–4334 RSC.
- J. W. B. Reesor and G. F. Wright, Arsenobenzene-Dimetal Adducts, J. Org. Chem., 1957, 22, 382–385 CrossRef CAS.
- P. Elmes, S. Middleton and B. West, Cyclic phosphines and arsines. II. Cyclic arsines, Aust. J. Chem., 1970, 23, 1559–1570 CrossRef CAS.
- T. Kato, S. Tanaka and K. Naka, In situ iodination of organoarsenic homocycles: Facile synthesis of 9-arsafluorene, Chem. Lett., 2015, 44, 1476–1478 CrossRef CAS.
- S. Tanaka, M. Konishi, H. Imoto, Y. Nakamura, M. Ishida, H. Furuta and K. Naka, Fundamental Study on Arsenic(III) Halides (AsX3; X = Br, I) toward the Construction of C3-Symmetrical Monodentate Arsenic Ligands, Inorg. Chem., 2020, 59, 9587–9593 CrossRef CAS PubMed.
- A. Sumida, H. Imoto and K. Naka, Synthetic Strategy for AB2-Type Arsines via Bidentate Dithiolate Leaving Groups, Inorg. Chem., 2022, 61, 17419–17426 CrossRef CAS PubMed.
- J. Ellermann, L. Brehm, E. Lindner, W. Hiller, R. Fawzi, F. L. Dickert and M. Waidhas, Chemistry of polyfunctional molecules. Part 86. Organoarsenic compounds derived from 1,3-bis(di-iodoarsino)propane with monocyclic, tricyclic, and straight-chain structures; X-ray crystal structure of 2,8,13,14-tetraoxa-1,3,7,9-tetra-arsatricyclo[7.3.1.1], J. Chem. Soc., Dalton Trans., 1986, 997–1001 RSC.
- J. C. Pazik and C. George, Neopentylarsenic chemistry: synthesis and characterization of As(CH2CMe3)nBr3-n (n = 1–3) and As(CH2CMe3)nBr5-n (n = 2 or 3). Crystal and molecular structure of As(CH2CMe3)3 and As(CH2CMe3)3Br2, Organometallics, 1989, 487, 482–487 CrossRef.
- D. Lu, A. D. Rae, G. Salem, M. L. Weir, A. C. Willis and S. B. Wild, Arsenicin A, A natural polyarsenical: Synthesis and crystal structure, Organometallics, 2010, 29, 32–33 CrossRef CAS.
- S. Tanaka, H. Imoto, T. Yumura and K. Naka, Arsenic Halogenation of 9-Arsafluorene and Utilization for As-C Bond Formation Reaction, Organometallics, 2017, 36, 1684–1687 CrossRef CAS.
- A. Tzschach and W. Deylig, Arsen–organo–Verbindungen, VI. Zur Reaktion der Alkaliarside MeAsR 2 mit cyclischen Äthern, Chem. Ber., 1965, 98, 977–982 CrossRef CAS.
- T. D. DuBois and F. T. Smith, Ligand-bridged five-coordinate nickel(II) complexes, Inorg. Chem., 1973, 12, 735–740 CrossRef CAS.
- A. M. Aguiar and T. G. Archibald, Retention of configuration in nucleophilic vinylic halide substitution. II. Stereospecific preparation of vinylarsines, J. Org. Chem., 1967, 32, 2627–2628 CrossRef CAS.
- M. Bonaterra, S. E. Martín and R. A. Rossi, One-pot palladium-catalyzed cross-coupling reaction of aryl iodides with stannylarsanes and stannylstibanes, Org. Lett., 2003, 5, 2731–2734 CrossRef CAS PubMed.
- M. Bonaterra, R. A. Rossi and S. E. Martín, Organoheteroatom stannanes in palladium-catalyzed cross-coupling reactions with 1-naphthyl triflate, Organometallics, 2009, 28, 933–936 CrossRef CAS.
- P. M. Uberman, M. N. Lanteri and S. E. Martin, Highly efficient palladium-catalyzed arsination. Synthesis of a biphenyl arsine ligand and its application to obtain perfluoroalkylarsines, Organometallics, 2009, 28, 6927–6934 CrossRef CAS.
- P. M. Uberman, M. N. Lanteri, S. C. P. Puenzo and S. E. Martín, Synthesis of biphenyl-based arsine ligands by Suzuki-Miyaura coupling and their application to Pd-catalyzed arsination, Dalton Trans., 2011, 40, 9229–9237 RSC.
- P. M. Uberman, M. R. Caira and S. E. Martín, A chiral bis(arsine) ligand: Synthesis and applications in palladium-catalyzed asymmetric allylic alkylations, Organometallics, 2013, 32, 3220–3226 CrossRef CAS.
- G. J. Quinteros, P. M. Uberman and S. E. Martín, Bulky Monodentate Biphenylarsine Ligands: Synthesis and Evaluation of Their Structure Effects in the Palladium-Catalyzed Heck Reaction, Eur. J. Org. Chem., 2015, 2015, 2698–2705 CrossRef CAS.
- T. Kauffmann and J. Ennen, Vielelektronenliganden, IX. Synthese von drei Kronenarsanen, Chem. Ber., 1985, 118, 2692–2702 CrossRef CAS.
- T. Kauffmann, J. Ennen and K. Berghus, Titankomplexe offenkettiger und cyclischer arsanliganden: synthese und katalytische wirkung (1), Tetrahedron Lett., 1984, 25, 1971–1974 CrossRef CAS.
- S. Tanaka, H. Imoto, T. Kato and K. Naka, A practical method for the generation of organoarsenic nucleophiles towards the construction of a versatile arsenic library, Dalton Trans., 2016, 45, 7937–7940 RSC.
- K. Maitra, V. J. Catalano, J. Clark and J. H. Nelson, Synthesis and Characterization of Unsymmetrical (cis-(Phosphino–arsino)ethene)tetracarbonylmolybdenum Complexes: Hydroarsination Reactions of (H3CC⋮CPPh2)Mo(CO)5 and [H2CCCH(DBP)]Mo(CO)5 and Hydrophosphination Reactions of (Ph2AsCH2C⋮CH)Mo(CO)5, Inorg. Chem., 1998, 37, 1105–1111 CrossRef CAS.
- M. Scheer, The coordination chemistry of group 15 element ligand complexes - A developing area, Dalton Trans., 2008, 4372–4386 RSC.
- M. L. Bungabong, W. T. Kien, Y. Li, S. V. Selvaratnam, K. G. Dongol and P. H. Leung, A novel asymmetric hydroarsination reaction promoted by a chiral organopalladium complex, Inorg. Chem., 2007, 46, 4733–4736 CrossRef CAS PubMed.
- F. Liu, S. A. Pullarkat, Y. Li, S. Chen and P. H. Leung, Novel enantioselective synthesis of functionalized pyridylarsanes by a chiral palladium template promoted asymmetric hydroarsanation reaction, Eur. J. Inorg. Chem., 2009, 4134–4140 CrossRef CAS.
- Y. L. Cheow, S. A. Pullarkat, Y. Li and P. H. Leung, Asymmetric hydroarsination reactions toward synthesis of alcohol functionalised C-chiral As-P ligands promoted by chiral cyclometallated complexes, J. Organomet. Chem., 2012, 696, 4215–4220 CrossRef CAS.
- W. S. Tay, X. Y. Yang, Y. Li, S. A. Pullarkat and P. H. Leung, Nickel catalyzed enantioselective hydroarsination of nitrostyrene, Chem. Commun., 2017, 53, 6307–6310 RSC.
- C. A. Bange and R. Waterman, Zirconium-catalyzed hydroarsination with primary arsines, Polyhedron, 2018, 156, 31–34 CrossRef CAS.
- W. S. Tay, Y. Li, X. Y. Yang, S. A. Pullarkat and P. H. Leung, Air-stable phosphine organocatalysts for the hydroarsination reaction, J. Organomet. Chem., 2020, 914, 121216 CrossRef CAS.
- W. S. Tay, Y. Li, Y. Lu, S. A. Pullarkat and P. H. Leung, Chemoselective Synthesis and Evaluation of β-Oxovinylarsines as an Arsenic Synthetic Precursor, Organometallics, 2020, 39, 271–278 CrossRef CAS.
- G. Märkl, H. Hauptmann and A. Merz, Synthese von 1-phenyl-2,5-diaryl(dialkyl)-arsolen; umsetzung der arsole mit alkalimetallen und lithiumorganylen, J. Organomet. Chem., 1983, 249, 335–363 CrossRef.
- C. S. Palmer and R. Adams, The reactions of the arsines. II. Condensation of aromatic primary arsines with aldehydes 1, J. Am. Chem. Soc., 1922, 44, 1356–1382 CrossRef CAS.
- J. D. Smith, Organometallic Compounds, Chem. Arsenic, Antimony Bismuth, 1973, 24, 622–639 Search PubMed.
- A. Tzschach and S. Baensch, Organoarsen–Verbindungen. XVII. Zur Umsetzung von Biphosphinen und Biarsinen mit Phenylacetylen, J. Prakt. Chem., 1971, 313, 254–258 CrossRef CAS.
- U. Schmidt, I. Boie, C. Osterroht, R. Schröer and H.-F. Grützmacher, Über Phosphinidene, 4. Versuche zum Nachweis von Phosphiniden R—P, Chem. Ber., 1968, 101, 1381–1397 CrossRef CAS.
- F. Y. Kwong, C. W. Lai, M. Yu, D. M. Tan, F. L. Lam, A. S. C. Chan and K. S. Chan, Convenient palladium-catalyzed arsination: Direct synthesis of functionalized aryl arsines, optically active As,N ligands, and their metal complexes, Organometallics, 2005, 24, 4170–4178 CrossRef CAS.
- F. Y. Kwong, C. W. Lai and K. S. Chan, Catalytic Solvent-Free Arsination: First Catalytic Application of Pd–Ar/As–Ph Exchange in the Syntheses of Functionalized Aryl Arsines, J. Am. Chem. Soc., 2001, 123, 8864–8865 CrossRef CAS PubMed.
- W. L. Lewis, C. D. Lowry and F. H. Bergeim, Some Derivatives of Phenoxarsin. 1, J. Am. Chem. Soc., 1921, 43, 891–896 CrossRef CAS.
- M. Seidl, G. Balázs and M. Scheer, The Chemistry of Yellow Arsenic, Chem. Rev., 2019, 119, 8406–8434 CrossRef CAS PubMed.
- J. A. Aeschlimann, N. D. Lees, N. P. McCleland and G. N. Nicklin, X.—Organic compounds of arsenic. Part II. Derivatives of the arsenic analogue of carbazole, J. Chem. Soc., Trans., 1925, 127, 66–69 RSC.
- H. Gottlieb-Billroth, Derivatives of Dibenzo-Arsenole, J. Am. Chem. Soc., 1927, 49, 482–486 CrossRef CAS.
- A. J. Ashe and F. J. Drone, Synthesis of 1-phenylarsole and 1-phenylstibole, Organometallics, 1985, 4, 1478–1480 CrossRef CAS.
- F. F. Blicke, O. J. Weinkauff and G. W. Hargreaves, Diarsyls. II. Tetra-Aryldiarsyls 1, J. Am. Chem. Soc., 1930, 52, 780–786 CrossRef CAS.
- E. H. Braye, W. Hübel and I. Caplier, New Unsaturated Heterocyclic Systems. I, J. Am. Chem. Soc., 1961, 83, 4406–4413 CrossRef CAS.
- D. Sartain and M. R. Truter, 845. The crystal structure of 9-phenyl-9-arsafluorene (monoclinic form), J. Chem. Soc., 1963, 4414–4423 RSC.
- P. J. Faran and W. A. Nugent, Transfer from Zirconium, J. Am. Chem. Soc., 1988, 110, 2310–2312 CrossRef.
- H. Imoto, K. Naka, R. Inaba, K. Oka, T. Iwami, Y. Miyake and K. Tajima, Systematic Study of Pnictogen-Fused Heterofluorenes, Inorg. Chem., 2022, 61, 7318–7326 CrossRef PubMed.
- H. Sasaki, I. Akioka, H. Imoto and K. Naka, Arsenic-Bridged Silafluorene and Germafluorene as a Novel Class of Mixed-Heteroatom-Bridged Heterofluorenes, Eur. J. Org. Chem., 2021, 2021, 1390–1395 CrossRef CAS.
- M. Ishidoshiro, Y. Matsumura, H. Imoto, Y. Irie, T. Kato, S. Watase, K. Matsukawa, S. Inagi, I. Tomita and K. Naka, Practical Synthesis and Properties of 2,5-Diarylarsoles, Org. Lett., 2015, 17, 4854–4857 CrossRef CAS PubMed.
- M. Ishidoshiro, H. Imoto, S. Tanaka and K. Naka, An experimental study on arsoles: Structural variation, optical and electronic properties, and emission behavior, Dalton Trans., 2016, 45, 8717–8723 RSC.
- H. Imoto, A. Urushizaki, I. Kawashima and K. Naka, Peraryl Arsoles: Practical Synthesis, Electronic Structures, and Solid-State Emission Behaviors, Chem. – Eur. J., 2018, 24, 8797–8803 CrossRef CAS PubMed.
- S. Iwasaki, Y. Miyake, H. Imoto and K. Naka, 1,2,5-Triarylcycloalka[c]arsoles: Structural Effects of Fused-cycloalkanes on Stability and Photophysical Properties, Chem. – Eur. J., 2022, 28, e202202084 CrossRef CAS PubMed.
- C. Takahara, S. Iwasaki, Y. Miyake, T. Yumura, H. Imoto and K. Naka, Element recycling in arsoles, Bull. Chem. Soc. Jpn., 2024, 97, uoad028 CrossRef CAS.
- J. P. Green, Y. Han, R. Kilmurray, M. A. McLachlan, T. D. Anthopoulos and M. Heeney, An Air-Stable Semiconducting Polymer Containing Dithieno[3,2-b:2′,3′-d]arsole, Angew. Chem., Int. Ed., 2016, 55, 7148–7151 CrossRef CAS PubMed.
- T. Kato, H. Imoto, S. Tanaka, M. Ishidoshiro and K. Naka, Facile synthesis and properties of dithieno[3,2-: B:2′,3′- d] arsoles, Dalton Trans., 2016, 45, 11338–11345 RSC.
- H. Imoto, I. Kawashima, C. Yamazawa, S. Tanaka and K. Naka, Multi-mode emission color tuning of dithieno[3,2-: B:2′,3′- d] arsoles, J. Mater. Chem. C, 2017, 5, 6697–6703 RSC.
- C. Takahara, S. Iwasaki, H. Kihara, Y. Miyake, H. Imoto and K. Naka, 4-Aryldithieno[3,2-b:2′,3′-d]arsoles: effects of the As-substituent on the structure, photophysical properties, and stability, Dalton Trans., 2022, 51, 13734–13741 RSC.
- A. Urushizaki, T. Yumura, Y. Kitagawa, Y. Hasegawa, H. Imoto and K. Naka, Dithieno[3,4-b:3′,4′-d]arsole: A Novel Class of Hetero[5]radialenes, Eur. J. Org. Chem., 2020, 2020, 3965–3970 CrossRef CAS.
- T. Fujii, S. Tanaka, S. Hayashi, H. Imoto and K. Naka, Dipyridinoarsole: a new class of stable and modifiable heteroatom-bridged bipyridines, Chem. Commun., 2020, 56, 6035–6038 RSC.
- M. Stolar, J. Borau-Garcia, M. Toonen and T. Baumgartner, Synthesis and tunability of highly electron-accepting, N-benzylated ‘phosphaviologens’, J. Am. Chem. Soc., 2015, 137, 3366–3371 CrossRef CAS PubMed.
- C. Reus, M. Stolar, J. Vanderkley, J. Nebauer and T. Baumgartner, A Convenient N-Arylation Route for Electron-Deficient Pyridines: The Case of π-Extended Electrochromic Phosphaviologens, J. Am. Chem. Soc., 2015, 137, 11710–11717 CrossRef CAS PubMed.
- K. Kikuchi, H. Sei, K. Okubo, N. Tohnai, K. Oka, S. Dekura, T. Kikuchi, H. Imoto and K. Naka, Breathing Metal-Organic Frameworks Supported by an Arsenic-Bridged 4,4′-Bipyridine Ligand, Inorg. Chem., 2024, 63, 4337–4343 CrossRef CAS PubMed.
- H. Imoto, T. Fujii, S. Tanaka, S. Yamamoto, M. Mitsuishi, T. Yumura and K. Naka, As-Heteropentacenes: An Experimental and Computational Study on a Novel Class of Heteroacenes, Org. Lett., 2018, 20, 5952–5955 CrossRef CAS PubMed.
- K. Ishijima, S. Tanaka, H. Imoto and K. Naka, 2-Arylbenzo[: B] arsoles: An experimental and computational study on the relationship between structural and photophysical properties, Dalton Trans., 2020, 49, 15612–15621 RSC.
- J. P. Green, S. J. Cryer, J. Marafie, A. J. P. White and M. Heeney, Synthesis of a Luminescent Arsolo[2,3-d:5,4-d′]bis(thiazole) Building Block and Comparison to Its Phosphole Analogue, Organometallics, 2017, 36, 2632–2636 CrossRef CAS.
- A. Sumida, K. Naka and H. Imoto, Heteroarene–Fused Benzo[ b ]arsoles: Structure, Photophysical Properties, and Effects of the Bridging Element, Chem. – Asian J., 2025, 20, e202401767 CrossRef CAS PubMed.
- T. Yasui, A. Shibata, T. Ishikawa, Y. Fujita, C. Okochi, Y. Koseki, H. Kasai, K. Oka, K. Naka and H. Imoto, Cytotoxicity Study of π-Conjugated Arsenic Compounds, ChemistrySelect, 2025, 10, e02502 CrossRef CAS.
- C. Yuan, S. Saito, C. Camacho, S. Irle, I. Hisaki and S. Yamaguchi, A π-conjugated system with flexibility and rigidity that shows environment-dependent RGB luminescence, J. Am. Chem. Soc., 2013, 135, 8842–8845 CrossRef CAS PubMed.
- M. Rosenberg, C. Dahlstrand, K. Kilså and H. Ottosson, Excited state Aromaticity and Antiaromaticity: Opportunities for Photophysical and photochemical rationalizations, Chem. Rev., 2014, 114, 5379–5425 CrossRef CAS PubMed.
- I. Kawashima, H. Imoto, M. Ishida, H. Furuta, S. Yamamoto, M. Mitsuishi, S. Tanaka, T. Fujii and K. Naka, Dibenzoarsepins: Planarization of 8π-Electron System in the Lowest Singlet Excited State, Angew. Chem., Int. Ed., 2019, 58, 11686–11690 CrossRef CAS PubMed.
- J. Toldo, O. El Bakouri, M. Solà, P. O. Norrby and H. Ottosson, Is Excited-State Aromaticity a Driving Force for Planarization of Dibenzannelated 8π-Electron Heterocycles?, ChemPlusChem, 2019, 84, 712–721 CrossRef CAS PubMed.
- R. Kotani, L. Liu, P. Kumar, H. Kuramochi, T. Tahara, P. Liu, A. Osuka, P. B. Karadakov and S. Saito, Controlling the S1energy profile by tuning excited-state aromaticity, J. Am. Chem. Soc., 2020, 142, 14985–14992 CrossRef CAS PubMed.
- T. Fujii, T. Kusukawa, H. Imoto and K. Naka, Pnictogen-Bridged Diphenyl Sulfones as Photoinduced Pnictogen Bond Forming Emission Motifs, Chem. – Eur. J., 2023, 29, e202202572 CrossRef CAS PubMed.
- T. Machida, T. Iwasa, T. Taketsugu, K. Sada and K. Kokado, Photoinduced Pyramidal Inversion Behavior of Phosphanes Involved with Aggregation–Induced Emission Behavior, Chem. – Eur. J., 2020, 26, 8028–8034 CrossRef CAS PubMed.
- T. Fujii, A. Urushizaki, H. Imoto and K. Naka, Syntheses and Photophysical Properties of Pnictogen- and Chalcogen-Containing Diheteroanthracenes, Asian J. Org. Chem., 2023, 12, e202300067 CrossRef CAS.
- K. Mizuta, H. Shimoji, T. Fujii, S. Yamamoto, H. Ohkita, K. Naka and H. Imoto, Pnictogen–Bridged Sulfoximines: Effects of Element and N –Substituent on Excited–State Dynamics, ChemPhotoChem, 2025, 9, e202400336 CrossRef CAS.
- A. El Nahhas, M. A. Shameem, P. Chabera, J. Uhlig and A. Orthaber, Synthesis and Characterization of Cyclopentadithiophene Heterofulvenes: Design Tools for Light-Activated Processes, Chem. – Eur. J., 2017, 23, 5673–5677 CrossRef CAS PubMed.
- H. Kihara, H. Imoto and K. Naka, Practical Syntheses and Luminescent Properties of Arene-substituted Arsines, Asian J. Org. Chem., 2021, 10, 2682–2689 CrossRef CAS.
- S. Iwasaki, H. Kihara, H. Imoto and K. Naka, Arsinoquinolines as a Novel Class of Luminophores, Asian J. Org. Chem., 2021, 10, 2618–2624 CrossRef CAS.
- V. Kremláček, M. Erben, R. Jambor, A. Růžička, J. Turek, E. Rychagova, S. Ketkov and L. Dostál, From a 2,1-Benzazaarsole to Elusive 1-Arsanaphthalenes in One Step, Chem. – Eur. J., 2019, 25, 5668–5671 CrossRef PubMed.
- A. J. Ashe, X. Fang and J. W. Kampf, 1-Arsanaphthalene. The Structure of Tricarbonyl(2-trimethylsilyl-1-arsanaphthalene)molybdenum, Organometallics, 2001, 20, 2109–2113 CrossRef CAS.
- K. Matsuo, R. Okumura, H. Hayashi, N. Aratani, S. Jinnai, Y. Ie, A. Saeki and H. Yamada, Phosphaacene as a structural analogue of thienoacenes for organic semiconductors, Chem. Commun., 2022, 58, 13576–13579 RSC.
- A. Sumida, A. Saeki, K. Matsuo, K. Naka and H. Imoto, Dithienoarsinines: stable and planar π-extended arsabenzenes, Chem. Sci., 2025, 16, 1126–1135 RSC.
- P. A. Brown, C. D. Martin and K. L. Shuford, Aromaticity of unsaturated BEC4 heterocycles (E = N, P, As, Sb, O, S, Se, Te), Phys. Chem. Chem. Phys., 2019, 21, 18458–18466 RSC.
- A. Sumida, T. Onishi, H. Imoto and K. Naka, Synthesis, structures, and photophysical properties of π-extended arsaborins, Dalton Trans., 2023, 53, 1706–1713 RSC.
- K. Naka, T. Umeyama and Y. Chujo, Synthesis of poly(vinylene-arsine)s: Alternating radical copolymerization of arsenic atomic biradical equivalent and phenylacetylene, J. Am. Chem. Soc., 2002, 124, 6600–6603 CrossRef CAS PubMed.
- T. Umeyama, K. Naka, A. Nakahashi and Y. Chujo, Radical Copolymerization of Acetylenic Compounds with Phenyl-Substituted Cyclooligoarsine: Substituent Effect and Optical Properties, Macromolecules, 2004, 37, 1271–1275 CrossRef CAS.
- T. Umeyama, K. Naka and Y. Chujo, Synthesis of poly(vinylene arsine)s through the ring-collapsed radical alternating copolymerization of an organoarsenic homocycle with aliphatic acetylenes and their properties, J. Polym. Sci., Part A: Polym. Chem., 2004, 42, 3604–3611 CrossRef CAS.
- K. Naka, Synthesis of polymers containing group 15 elements via bismetallation of acetylenic compounds, Polym. J., 2008, 40, 1031–1041 CrossRef CAS.
- Y. Matsumura, M. Ishidoshiro, Y. Irie, H. Imoto, K. Naka, K. Tanaka, S. Inagi and I. Tomita, Arsole-Containing π-Conjugated Polymer by the Post-Element-Transformation Technique, Angew. Chem., Int. Ed., 2016, 55, 15040–15043 CrossRef CAS PubMed.
- Y. Matsumura, M. Ueda, K. Fukuda, K. Fukui, I. Takase, H. Nishiyama, S. Inagi and I. Tomita, Synthesis of π-conjugated polymers containing phosphole units in the main chain by reaction of an organometallic polymer having a titanacyclopentadiene unit, ACS Macro Lett., 2015, 4, 124–127 CrossRef CAS PubMed.
- V. H. K. Fell, A. Mikosch, A. K. Steppert, W. Ogieglo, E. Senol, D. Canneson, M. Bayer, F. Schoenebeck, A. Greilich and A. J. C. Kuehne, Synthesis and Optical Characterization of Hybrid Organic-Inorganic Heterofluorene Polymers, Macromolecules, 2017, 50, 2338–2343 CrossRef CAS.
- D. M. Salazar, E. Mijangos, S. Pullen, M. Gao and A. Orthaber, Functional small-molecules & polymers containing P=C and As=C bonds as hybrid π-conjugated materials, Chem. Commun., 2017, 53, 1120–1123 RSC.
- C. Yamazawa, H. Imoto and K. Naka, Syntheses of dithienoarsole-containing polymers via suzuki-miyaura and sonogashira-hagihara coupling reactions, Chem. Lett., 2018, 47, 887–890 CrossRef CAS.
- H. Imoto, C. Yamazawa, S. Hayashi, M. Aono and K. Naka, Electropolymerization of Dithieno[3,2-b:2′,3′-d]arsole, ChemElectroChem, 2018, 5, 3357–3360 CrossRef CAS.
- J. P. Green, H. Cha, M. Shahid, A. Creamer, J. R. Durrant and M. Heeney, Dithieno[3,2-b:2′,3′- d] arsole-containing conjugated polymers in organic photovoltaic devices, Dalton Trans., 2019, 48, 6676–6679 RSC.
- C. Takahara, M. Nakamura, Y. Aoyama, T. Yanagihara, S. Ito, K. Tanaka, H. Imoto and K. Naka, Versatile and Practical Design of Dithieno[3,2-b:2′,3′-d]arsole Polymers, Macromolecules, 2023, 56, 6758–6763 CrossRef CAS.
- S. Tanaka, T. Enoki, H. Imoto, Y. Ooyama, J. Ohshita, T. Kato and K. Naka, Highly Efficient Singlet Oxygen Generation and High Oxidation Resistance Enhanced by Arsole-Polymer-Based Photosensitizer: Application as a Recyclable Photooxidation Catalyst, Macromolecules, 2020, 53, 2006–2013 CrossRef CAS.
- C. Yamazawa, Y. Hirano, H. Imoto, N. Tsutsumi and K. Naka, Superior light-resistant dithieno[3,2-b:2′,3′-d]arsole-based polymers exhibiting ultrastable amplified spontaneous emission, Chem. Commun., 2021, 57, 1595–1598 RSC.
- K. Ishijima, S. Tanaka, H. Imoto and K. Naka, 2,3-Diarylbenzo[b]arsole: Structural Modification and Polymerization for Tuning of Photophysical Properties, Chem. – Eur. J., 2021, 27, 4676–4682 CrossRef CAS PubMed.
- J. S. Fritz and E. M. Moyers, Concentration and separation of trace metals with an arsonic acid resin, Talanta, 1976, 23, 590–593 CrossRef CAS PubMed.
- T. Zayas, M. J. Percino, J. Cardoso and V. M. Chapela, Novel water-soluble polyelectrolytes with arsonic acid group for flocculation application, Polymer, 2000, 41, 5505–5512 CrossRef CAS.
- J. García-Serrano, A. M. Herrera, F. Pérez-Moreno, M. A. Valdez and U. Pal, Synthesis of Novel Ionic Polymers Containing Arsonic Acid Group, J. Polym. Sci., Part B:Polym. Phys., 2006, 44, 1627–1634 CrossRef.
- J. Tanaka, S. Tani, R. Peltier, E. H. Pilkington, A. Kerr, T. P. Davis and P. Wilson, Synthesis, aggregation and responsivity of block copolymers containing organic arsenicals, Polym. Chem., 2018, 9, 1551–1556 RSC.
- J. Tanaka, G. Moriceau, A. Cook, A. Kerr, J. Zhang, R. Peltier, S. Perrier, T. P. Davis and P. Wilson, Tuning the Structure, Stability, and Responsivity of Polymeric Arsenical Nanoparticles Using Polythiol Cross-Linkers, Macromolecules, 2019, 52, 992–1003 CrossRef CAS.
- J. Tanaka, J. I. Song, A. M. Lunn, R. A. Hand, S. Häkkinen, T. L. Schiller, S. Perrier, T. P. Davis and P. Wilson, Polymeric arsenicals as scaffolds for functional and responsive hydrogels, J. Mater. Chem. B, 2019, 7, 4263–4271 RSC.
- J. Tanaka, A. Evans, P. Gurnani, A. Kerr and P. Wilson, Functionalisation and stabilisation of polymeric arsenical nanoparticles prepared by sequential reductive and radical cross-linking, Polym. Chem., 2020, 11, 2519–2531 RSC.
- P. Wang, C. R. Liu, X. L. Sun, S. S. Chen, J. F. Li, Z. Xie and Y. Tang, A newly-designed PE-supported arsine for efficient and practical catalytic Wittig olefination, Chem. Commun., 2012, 48, 290–292 RSC.
- W. He, D. Lv, Y. Guan and S. Yu, Post-synthesis modification of metal-organic frameworks: synthesis, characteristics, and applications, J. Mater. Chem. A, 2023, 11, 24519–24550 RSC.
- M. Kalaj and S. M. Cohen, Postsynthetic Modification: An Enabling Technology for the Advancement of Metal-Organic Frameworks, ACS Cent. Sci., 2020, 6, 1046–1057 CrossRef CAS PubMed.
- M. Ishidoshiro, H. Imoto and K. Naka, A metal-organic framework containing arsenic atoms with a free lone pair, Bull. Chem. Soc. Jpn., 2016, 89, 1057–1062 CrossRef CAS.
- R. E. Sikma, P. Kunal, S. G. Dunning, J. E. Reynolds, J. S. Lee, J. S. Chang and S. M. Humphrey, Organoarsine Metal-Organic Framework with cis-Diarsine Pockets for the Installation of Uniquely Confined Metal Complexes, J. Am. Chem. Soc., 2018, 140, 9806–9809 CrossRef CAS PubMed.
- A. Nakahashi, K. Naka and Y. Chujo, 1,4-dihydro-1,4-diarsinine: Facile synthesis via nonvolatile arsenic intermediates by radical reactions, Organometallics, 2007, 26, 1827–1830 CrossRef CAS.
- M. Arita, K. Naka, Y. Morisaki, A. Nakahashi and Y. Chujo, Synthesis and characterization of stereoisomers of 1,4-dihydro-1,4- diarsinines, Organometallics, 2009, 28, 6109–6113 CrossRef CAS.
- H. Unesaki, T. Kato, S. Watase, K. Matsukawa and K. Naka, Polymorph control of luminescence properties in molecular crystals of a platinum and organoarsenic complex and formation of stable one-dimensional nanochannel, Inorg. Chem., 2014, 53, 8270–8277 CrossRef CAS PubMed.
- K. Naka, A. Nakahashi, M. Arita and Y. Chujo, Stoichiometric complexation of palladium(II) with 1,4-dihydro-1,4- diarsinine as a rigid symmetrical bidentate ligand, Organometallics, 2008, 27, 1034–1036 CrossRef CAS.
- M. Arita, K. Naka, T. Shimamoto, T. Yumura, A. Nakahashi, Y. Morisaki and Y. Chujo, 1,4-dihydro-1,4-diarsinine-bridged dinuclear trans-dihaloplatinum(II) complexes: Synthesis and controlled Pt-Pt interaction by halogen substitution induced conformational change, Organometallics, 2010, 29, 4992–5003 CrossRef CAS.
- M. Arita, K. Naka, Y. Morisaki and Y. Chujo, Structural diversity in the coordination of 1,4-dihydro–1,4-diarsinine as a cyclic ditopic organoarsenic ligand to metal ions, Heteroat. Chem., 2012, 23, 16–26 CrossRef CAS.
- H. Adachi, H. Imoto, S. Watase, K. Matsukawa and K. Naka, As-stereogenic C2-symmetric organoarsines: Synthesis and enantioselective self-assembly into a dinuclear triple-stranded helicate with copper iodide, Dalton Trans., 2015, 44, 15372–15376 RSC.
- H. Imoto, S. Nishiyama, T. Yumura, S. Watase, K. Matsukawa and K. Naka, Control of aurophilic interaction: Conformations and electronic structures of one-dimensional supramolecular architectures, Dalton Trans., 2017, 46, 8077–8082 RSC.
- H. Imoto, S. Nishiyama and K. Naka, Substituent-dependent stimuli recognition of luminescent gold(I) chloride complexes based on diarsenic ligands, Bull. Chem. Soc. Jpn., 2018, 91, 349–354 CrossRef CAS.
- K. Naka, T. Kato, S. Watase and K. Matsukawa, Organic vapor triggered repeatable on-off crystalline-state luminescence switching, Inorg. Chem., 2012, 51, 4420–4422 CrossRef CAS PubMed.
- H. Imoto, C. Yamazawa, S. Tanaka, T. Kato and K. Naka, A practical screening strategy of arsenic ligands for a transition-metal-catalyzed reaction, Chem. Lett., 2017, 46, 821–823 CrossRef CAS.
- H. Imoto, M. Konishi, S. Nishiyama, H. Sasaki, S. Tanaka, T. Yumura and K. Naka, Construction of a bidentate arsenic ligand library starting from a cyclooligoarsine, Chem. Lett., 2019, 48, 1312–1315 CrossRef CAS.
- H. Kihara, S. Tanaka, H. Imoto and K. Naka, Phenyldiquinolinylarsine as a Nitrogen-Arsenic-Nitrogen Pincer Ligand, Eur. J. Inorg. Chem., 2020, 2020, 3662–3665 CrossRef CAS.
- A. K. Gupta, S. Akkarasamiyo and A. Orthaber, Rich Coordination Chemistry of π-Acceptor Dibenzoarsole Ligands, Inorg. Chem., 2017, 56, 4504–4511 CrossRef PubMed.
- K. C. Schwan, A. Adolf, C. Thoms, M. Zabel, A. Y. Timoshkin and M. Scheer, Selective halogenation at the pnictogen atom in Lewis-acid/base-stabilised phosphanylboranes and arsanylboranes, Dalton Trans., 2008, 5054–5058 RSC.
- J. Braese, A. Schinabeck, M. Bodensteiner, H. Yersin, A. Y. Timoshkin and M. Scheer, Gold(I) Complexes Containing Phosphanyl- and Arsanylborane Ligands, Chem. – Eur. J., 2018, 24, 10073–10077 CrossRef CAS PubMed.
- F. Lehnfeld, M. Seidl, A. Y. Timoshkin and M. Scheer, Synthesis and Reactivity of a Lewis-Base-Stabilized tert-Butyl Arsanylborane: A Versatile Building Block for Arsenic-Boron Oligomers, Eur. J. Inorg. Chem., 2022, 2022, e202100930 CrossRef CAS.
- A. Sumida, R. Kobayashi, T. Yumura, H. Imoto and K. Naka, Dibenzoarsacrowns: an experimental and computational study on the coordination behaviors, Chem. Commun., 2021, 57, 2013–2016 RSC.
- K. Ogawa, A. Sumida, T. Ami, K. Oka, N. Tohnai, Y. Miyake, H. Imoto and K. Naka, Cation Recognition by Dibenzoarsacrowns, Asian J. Org. Chem., 2023, 12, e202300525 CrossRef CAS.
- K. Kikuchi, A. Sumida, H. Imoto and K. Naka, Phosphorescent Metallacrown Ethers Enchained Through Coordination of Arsafluorene to Platinum(II) Dihalide, Eur. J. Inorg. Chem., 2021, 2021, 4463–4469 CrossRef CAS.
- M. K. Pandey, D. Mondal, B. S. Kote and M. S. Balakrishna, Synthesis and Photophysical Properties of Heavier Pnictogen Complexes, ChemPlusChem, 2023, 88, e202200460 CrossRef CAS PubMed.
- H. Imoto, H. Sasaki, S. Tanaka, T. Yumura and K. Naka, Platinum(II) Dihalide Complexes with 9-Arsafluorenes: Effects of Ligand Modification on the Phosphorescent Properties, Organometallics, 2017, 36, 2605–2611 CrossRef CAS.
- K. Kikuchi, H. Imoto and K. Naka, Robust and highly emissive copper(I) halide 1D-coordination polymers with triphenylarsine and a series of bridging N-heteroaromatic co-ligands, Dalton Trans., 2023, 52, 11168–11175 RSC.
- Y. V. Demyanov, E. H. Sadykov, M. I. Rakhmanova, A. S. Novikov, I. Y. Bagryanskaya and A. V. Artem'ev, Tris(2-Pyridyl)Arsine as a New Platform for Design of Luminescent Cu(I) and Ag(I) Complexes, Molecules, 2022, 27, 1–15 CrossRef PubMed.
- R. Kobayashi, T. Fujii, H. Imoto and K. Naka, Dinuclear Gold(I) Chloride Complexes with Diarsine Ligands, Eur. J. Inorg. Chem., 2021, 2021, 217–222 CrossRef CAS.
- R. Kobayashi, R. Inaba, H. Imoto and K. Naka, Multi-mode switchable luminescence of tetranuclear cubic copper(I) iodide complexes with tertiary arsine ligands, Bull. Chem. Soc. Jpn., 2021, 94, 1340–1346 CrossRef CAS.
- H. Imoto, S. Tanaka, T. Kato, S. Watase, K. Matsukawa, T. Yumura and K. Naka, Highly Efficient Solid-State Phosphorescence of Platinum Dihalide Complexes with 9-Phenyl-9-arsafluorene Ligands, Organometallics, 2016, 35, 364–369 CrossRef CAS.
- R. Kobayashi, H. Kihara, T. Kusukawa, H. Imoto and K. Naka, Dinuclear rhombic copper(I) iodide complexes with rigid bidentate arsenic ligands, Chem. Lett., 2021, 50, 382–385 CrossRef CAS.
- H. Shimoji, H. Kihara, A. Sumida, H. Imoto and K. Naka, Synthesis of Phosphorus/Arsenic Tridentate Ligands and Their Application for Luminescent Copper(I) Halide Complexes, Eur. J. Inorg. Chem., 2024, 27, e202400090 CrossRef CAS.
- J. Yukiyasu, A. Sumida, H. Shimoji, A. Urushizaki, T. Yanagihara, H. Ogoshi, M. Nakamura, S. Ito, K. Tanaka, H. Imoto and K. Naka, Structures and Luminescent Properties of Platinum(II) Dihalide Complexes with Bridged Triphenylarsine Derivatives, Eur. J. Inorg. Chem., 2024, 27, e202400036 CrossRef CAS.
- R. Kobayashi, T. Yumura, H. Imoto and K. Naka, Homo- And hetero-metallophilicity-driven synthesis of highly emissive and stimuli-responsive Au(i)-Cu(i) double salts, Chem. Commun., 2021, 57, 5382–5385 RSC.
- R. Kobayashi, H. Imoto and K. Naka, Stimuli-Responsive Emission of Dinuclear Rhombic Copper(I) Iodide Complexes Having Triphenylarsine and N-Heteroaromatic Co-Ligands, Eur. J. Inorg. Chem., 2020, 2020, 3548–3553 CrossRef CAS.
- H. Imoto, S. Tanaka, T. Kato, T. Yumura, S. Watase, K. Matsukawa and K. Naka, Molecular Shape Recognition by Using a Switchable Luminescent Nonporous Molecular Crystal, Organometallics, 2016, 35, 3647–3650 CrossRef CAS.
- H. Sasaki, H. Imoto, T. Kitao, T. Uemura, T. Yumura and K. Naka, Fluorinated porous molecular crystals: Vapor-triggered on-off switching of luminescence and porosity, Chem. Commun., 2019, 55, 6487–6490 RSC.
- A. Sumida, H. Imoto and K. Naka, Turn-on type sensing of methanol vapor by a luminescent platinum(II) dichloride complex with 21-dibenzoarsacrown-7, Dalton Trans., 2021, 50, 6682–6687 RSC.
- T. Fujii, Y. Kitagawa, Y. Hasegawa, H. Imoto and K. Naka, Drastic Enhancement of Photosensitized Energy Transfer Efficiency of a Eu(III) Complex Driven by Arsenic, Inorg. Chem., 2021, 60, 8605–8612 CrossRef CAS PubMed.
- T. Fujii, Y. Kitagawa, Y. Hasegawa, H. Imoto and K. Naka, Emission Properties of Eu(III) Complexes Containing Arsine and Phosphine Ligands with Annulated Structures, Inorg. Chem., 2022, 61, 17662–17672 CrossRef CAS PubMed.
- H. Shimoji, T. Fujii, A. Sumida, Y. Kitagawa, Y. Hasegawa, H. Imoto and K. Naka, Intensive emission of Eu(iii) β-diketonate complexes with arsine oxide ligands, J. Mater. Chem. C, 2023, 11, 15608–15615 RSC.
- H. Shimoji, Y. Aoyama, K. Inage, M. Nakamura, T. Yanagihara, K. Yuhara, Y. Kitagawa, Y. Hasegawa, S. Ito, K. Tanaka, H. Imoto and K. Naka, Highly Efficient and Thermally Durable Luminescence of 1D Eu3+ Coordination Polymers with Arsenic Bridging Ligands, Chem. – Eur. J., 2024, 30, e202400615 CrossRef CAS PubMed.
- T. Pugh, A. Kerridge and R. A. Layfield, Yttrium complexes of arsine, arsenide, and arsinidene ligands, Angew. Chem., Int. Ed., 2015, 54, 4255–4258 CrossRef CAS PubMed.
- V. Farina and B. Krishnan, Large Rate Accelerations in the Stille Reaction with Tri-2-furylphosphine and Triphenylarsine as Palladium Ligands: Mechanistic and Synthetic Implications, J. Am. Chem. Soc., 1991, 113, 9585–9595 CrossRef CAS.
- L. A. D. Van Der Veen, P. K. Keeven, P. C. J. Kamer and P. W. N. M. Van Leeuwen, Novel arsine ligands for selective hydroformylation of alk-1-enes employing platinum/tin catalysts, Chem. Commun., 2000, 333–334 RSC.
- A. Kojima, C. D. J. Boden and M. Shibasaki, Synthesis and evaluation of a new chiral arsine ligand; 2,2′-bis(diphenylarsino)-1,1′-binaphthyl (BINAs), Tetrahedron Lett., 1997, 38, 3459–3460 CrossRef CAS.
- Y. Tanabe, S. Kuriyama, K. Arashiba, Y. Miyake, K. Nakajima and Y. Nishibayashi, Preparation and reactivity of molybdenum-dinitrogen complexes bearing an arsenic-containing ANA-type pincer ligand, Chem. Commun., 2013, 49, 9290–9292 RSC.
- Y. Tanabe, S. Kuriyama, K. Arashiba, K. Nakajima and Y. Nishibayashi, Synthesis and Reactivity of Ruthenium Complexes Bearing Arsenic-Containing Arsenic-Nitrogen-Arsenic-Type Pincer Ligand, Organometallics, 2014, 33, 5295–5300 CrossRef CAS.
- A. Sumida, K. Ogawa, H. Imoto and K. Naka, Steric and electronic effects of arsa-Buchwald ligands on Suzuki-Miyaura coupling reaction, Dalton Trans., 2023, 52, 2838–2844 RSC.
- K. Ogawa, A. Sumida, T. Yumura, H. Imoto and K. Naka, Tertiary Arsine Ligands for Pd-Catalyzed Direct Arylation, Organometallics, 2024, 43, 322–329 CrossRef CAS.
- A. Chishiro, M. Konishi, R. Inaba, T. Yumura, H. Imoto and K. Naka, Tertiary arsine ligands for the Stille coupling reaction, Dalton Trans., 2022, 51, 95–103 RSC.
- C. J. O'Brien, J. L. Tellez, Z. S. Nixon, L. J. Kang, A. L. Carter, S. R. Kunkel, K. C. Przeworski and G. A. Chass, Recycling the waste: The development of a catalytic wittig reaction, Angew. Chem., Int. Ed., 2009, 48, 6836–6839 CrossRef PubMed.
- P. Cao, C. Y. Li, Y. B. Kang, Z. Xie, X. L. Sun and Y. Tang, Ph3As-catalyzed Wittig-type olefination of aldehydes with diazoacetate in the presence of Na2S2O4, J. Org. Chem., 2007, 72, 6628–6630 CrossRef CAS PubMed.
- L. Shi, W. Wang, Y. Wang and Y. Z. Huang, The First Example of a Catalytic Wittig-Type Reaction. Tri-n-butylarsine-Catalyzed Olefination in the Presence of Triphenyl Phosphite, J. Org. Chem., 1989, 54, 2027–2028 CrossRef CAS.
- R. Inaba, I. Kawashima, T. Fujii, T. Yumura, H. Imoto and K. Naka, Systematic Study on the Catalytic Arsa-Wittig Reaction, Chem. – Eur. J., 2020, 26, 13400–13407 CrossRef CAS PubMed.
- J. Yukiyasu, R. Inaba, T. Yumura, H. Imoto and K. Naka, Rational design of arsine catalysts for arsa-Wittig reaction, Org. Chem. Front., 2022, 9, 6786–6794 RSC.
- H. A. Van Kalkeren, F. L. Van Delft and F. P. J. T. Rutjes, Organophosphorus catalysis to bypass phosphine oxide waste, ChemSusChem, 2013, 6, 1615–1624 CrossRef CAS PubMed.
- H. Guo, Y. C. Fan, Z. Sun, Y. Wu and O. Kwon, Phosphine Organocatalysis, Chem. Rev., 2018, 118, 10049–10293 CrossRef CAS PubMed.
- J. M. Lipshultz, G. Li and A. T. Radosevich, Main Group Redox Catalysis of Organopnictogens: Vertical Periodic Trends and Emerging Opportunities in Group 15, J. Am. Chem. Soc., 2021, 143, 1699–1721 CrossRef CAS PubMed.
- Z. M. Abalonin, B. E. Lokhotskaya and L. A. Izmailova, Elimination in reactions of tertiary arsine oxides and sulfides with halogen derivatives, Zh. Obshch. Khim., 1986, 56, 228–229 Search PubMed.
- C. Okochi, R. Inaba, T. Yumura, K. Naka and H. Imoto, Sequential Oxygen Transfer via Oxidation of Arsine and Reduction of Arsine Oxide, Asian J. Org. Chem., 2025, e202500213 CrossRef CAS.
- D. W. Stephan, Frustrated lewis Pairs, J. Am. Chem. Soc., 2015, 137, 10018–10032 CrossRef CAS PubMed.
- D. W. Stephan and G. Erker, Frustrated Lewis pair chemistry: Development and perspectives, Angew. Chem., Int. Ed., 2015, 54, 6400–6441 CrossRef CAS PubMed.
- S. Ketkov, E. Rychagova, R. Kather and J. Beckmann, Pnictogen effects on the electronic interactions in the Lewis pair complexes Ph3EB(C6F5)3 (E = P, As, Sb), J. Organomet. Chem., 2021, 949, 121944 CrossRef CAS.
- T. Onishi, A. Sumida, C. Okochi, Y. Miyake, K. Kanaori, T. Iwamoto, K. Naka and H. Imoto, Combination of Arsines and Tris(pentafluorophenyl)borane Toward Frustrated Lewis Pair, Chem. – Eur. J., 2025, 31, e202501086 CrossRef CAS PubMed.
- W. R. Cullen and K. J. Reimer, Arsenic Speciation in the Environment, Chem. Rev., 1989, 89, 713–764 CrossRef CAS.
- Z. Lei, B. Chen, Y. M. Koo and D. R. Macfarlane, Introduction: Ionic Liquids, Chem. Rev., 2017, 117, 6633–6635 CrossRef PubMed.
- F. Philippi and T. Welton, Targeted modifications in ionic liquids - From understanding to design, Phys. Chem. Chem. Phys., 2021, 23, 6993–7021 RSC.
- F. Philippi, D. Rauber, B. Kuttich, T. Kraus, C. W. M. Kay, R. Hempelmann, P. A. Hunt and T. Welton, Ether functionalisation, ion conformation and the optimisation of macroscopic properties in ionic liquids, Phys. Chem. Chem. Phys., 2020, 22, 23038–23056 RSC.
- K. Górski, Ł. W. Ciszewski, A. Wrzosek, A. Szewczyk, A. L. Sobolewski and D. T. Gryko, An efficient method for the synthesis of π-expanded phosphonium salts, Org. Chem. Front., 2025, 12, 5484–5495, 10.1039/d5qo00708a.
- K. Tsunashima and M. Sugiya, Physical and electrochemical properties of low-viscosity phosphonium ionic liquids as potential electrolytes, Electrochem. Commun., 2007, 9, 2353–2358 CrossRef CAS.
- S. Seki, K. Hayamizu, S. Tsuzuki, K. Fujii, Y. Umebayashi, T. Mitsugi, T. Kobayashi, Y. Ohno, Y. Kobayashi, Y. Mita, H. Miyashiro and S. I. Ishiguro, Relationships between center atom species (N, P) and ionic conductivity, viscosity, density, self-diffusion coefficient of quaternary cation room-temperature ionic liquids, Phys. Chem. Chem. Phys., 2009, 11, 3509–3514 RSC.
- L. K. Scarbath-Evers, P. A. Hunt, B. Kirchner, D. R. MacFarlane and S. Zahn, Molecular features contributing to the lower viscosity of phosphonium ionic liquids compared to their ammonium analogues, Phys. Chem. Chem. Phys., 2015, 17, 20205–20216 RSC.
- A. I. Bhatt, I. May, V. A. Volkovich, M. E. Hetherington, B. Lewin, R. C. Thied and N. Ertok, Group 15 quaternary alkyl bistriflimides: Ionic liquids with potential application in electropositive metal deposition and as supporting electrolytes, J. Chem. Soc., Dalton Trans., 2002, 4532–4534 RSC.
- R. Inaba, T. Imai, S. Kitajima, H. Kasai, K. Oka, R. Hifumi, I. Tomita, M. Yoshizawa-Fujita, K. Naka and H. Imoto, An ionic liquid containing arsonium cation, Chem. Commun., 2024, 60, 14022–14025 RSC.
- A. P. M. Robertson, S. S. Chitnis, H. A. Jenkins, R. McDonald, M. J. Ferguson and N. Burford, Establishing the coordination chemistry of antimony(V) Cations: Systematic assessment of Ph4Sb(OTf) and Ph3Sb(OTf)2 as Lewis acceptors, Chem. – Eur. J., 2015, 21, 7902–7913 CrossRef CAS PubMed.
- A. P. M. Robertson, N. Burford, R. McDonald and M. J. Ferguson, Coordination complexes of Ph3Sb2+ and Ph 3Bi2+: Beyond pnictonium cations, Angew. Chem., Int. Ed., 2014, 53, 3480–3483 CrossRef CAS PubMed.
- M. Donath, M. Bodensteiner and J. J. Weigand, Versatile reagent Ph3As(OTf)2: One-pot synthesis of [P7(AsPh3)3][OTf]3 from PCl3, Chem. – Eur. J., 2014, 20, 17306–17310 CrossRef CAS PubMed.
- M. Donath, K. Schwedtmann, T. Schneider, F. Hennersdorf, A. Bauzá, A. Frontera and J. J. Weigand, Direct conversion of white phosphorus to versatile phosphorus transfer reagents via oxidative onioation, Nat. Chem., 2022, 14, 384–391 CrossRef CAS PubMed.
- J. Zhou, L. L. Liu, L. L. Cao and D. W. Stephan, The Arene-Stabilized η 5 -Pentamethylcyclopentadienyl Arsenic Dication [(η 5 -Cp*)As(toluene)] 2+, Angew. Chem., Int. Ed., 2019, 58, 5407–5412 CrossRef CAS PubMed.
- M. K. Sharma, S. Blomeyer, B. Neumann, H. G. Stammler, M. van Gastel, A. Hinz and R. S. Ghadwal, Crystalline Divinyldiarsene Radical Cations and Dications, Angew. Chem., Int. Ed., 2019, 58, 17599–17603 CrossRef CAS PubMed.
- T. Yasui, A. Chishiro, M. Sakabe, S. Sato, T. Yumura, K. Kanaori, K. Naka and H. Imoto, A Stable Hexacoordinated Arsenic(V) Dication, Eur. J. Inorg. Chem., 2025, 28, e202400819 CrossRef CAS.
- M. Sakabe and S. Sato, Isolation and Structural Determination of a Hexacoordinated Antimony(V) Dication, Chemistry, 2021, 27, 5658–5665 CrossRef CAS PubMed.
|
| This journal is © the Partner Organisations 2025 |
Click here to see how this site uses Cookies. View our privacy policy here.