Fact-checked by Grok 2 weeks ago

Halonium ion

A halonium ion is a positively charged onium species in which a atom (typically , , or iodine) bears the formal positive charge and is bridged between two carbon atoms via a , often manifesting as a three-membered cyclic structure. These ions serve as crucial reactive intermediates in , particularly in the of to alkenes, where they enforce stereospecific anti addition by shielding one face of the . Halonium ions can be classified into cyclic variants, such as the bromonium ion formed from and , and acyclic bridged forms like the dimethylbromonium ion, both of which have been characterized through NMR spectroscopy and media. Pioneering work by George A. Olah in the enabled the isolation and direct observation of stable halonium salts using conditions, confirming their structures and distinguishing them from open alternatives. In addition to carbon-bridged forms, halonium ions can also form symmetric linear complexes [D···X···D]⁺ with two Lewis base donors (D), exhibiting strong halogen bonding interactions with energies up to 45 kcal/mol, which enhances their utility in supramolecular assemblies and as synthons in crystal engineering. In synthetic applications, halonium ions drive regioselective halocyclizations, oxidations, and arylation reactions, with anion effects modulating their reactivity and product distribution. Recent advances, including gas-phase ultrafast , have captured their transient structures in , revealing bond lengths (e.g., C-Br at 1.96 in bromonium ions) and dynamic transformations on timescales, bridging the gap between theoretical models and experimental validation. As of 2025, new syntheses of stable acyclic dialkyl halonium salts, including chloronium ions via fluoroalkylation, have expanded their accessibility for reactivity studies. Their electron-deficient nature positions them as potent electrophiles, influencing fields from asymmetric synthesis to materials design, though their high reactivity often limits isolation to specialized conditions.

Fundamentals

Definition and Properties

A halonium ion is a positively charged species in which a atom (X) serves as the central atom in a three-center, two-electron bonding arrangement, typically represented by the general formula R-X^+-R', where X is a (F, Cl, Br, or I) and R and R' are groups, atoms, or other substituents. These ions exhibit a bridged structure where the halogen shares its positive charge with the adjacent groups, distinguishing them from simple haloalkyl cations. Halonium ions are classified into cyclic and acyclic forms. Cyclic halonium ions, such as haliranium ions, feature a three-membered involving the and two carbon atoms, often from addition, and can exhibit symmetric or asymmetric bridging depending on the substituents. Acyclic forms involve linear or open-chain arrangements without closure, commonly stabilized by donor atoms like . Stability of halonium ions decreases across the halogen series from iodonium > bromonium > chloronium > fluoronium, primarily due to increasing and decreasing atomic size, which weaken the bridging interactions; fluoronium ions, in particular, are highly unstable and rarely isolated. Most halonium ions are transient intermediates with short lifetimes in solution, except for certain iodonium salts that can persist for days or longer under dry, aprotic conditions. These ions display high electrophilicity, making them potent reagents in , and exhibit solubility in aprotic polar solvents such as and , while being insoluble in nonpolar media or decomposing in protic solvents. A prototypical example is the chloronium (C_2H_4Cl^+), a cyclic formed during chlorination of , illustrating the bridged nature central to halonium reactivity.

Historical Development

The concept of the halonium ion was first proposed in 1937 by Irving Roberts and George E. Kimball to rationalize the stereospecific anti addition of halogens to alkenes, invoking a bridged that preserved Markovnikov while accounting for the observed diastereoselectivity. This theoretical construct addressed longstanding discrepancies in mechanisms, suggesting an ionic rather than a simple open . During the and , accumulating theoretical and experimental evidence bolstered the bridged halonium model. Quantum mechanical calculations, including Hückel approaches, demonstrated the stability of three-center two-electron bonds in such systems, providing a conceptual framework for hypervalent species. Spectroscopic studies, particularly NMR investigations in the mid-, offered direct support by revealing symmetric bridged structures in halonium-like intermediates generated under acidic conditions. A major milestone occurred in 1970 when George A. Olah achieved the first isolation and characterization of stable halonium ions using media, such as SbF₅ in SO₂ClF, enabling low-temperature NMR spectroscopy of species like the ethylchloronium ion. This breakthrough, part of Olah's pioneering "stable ion" technique for reactive intermediates, confirmed the bridged geometry and earned him the 1994 for contributions to chemistry. In the 1980s, structural confirmation advanced with , exemplified by the 1985 determination of the biadamantylidene-bromonium ion, which revealed a symmetric three-membered with Br–C bond lengths of approximately 1.98 , validating the in solid state. By the early , research shifted toward practical applications, emphasizing iodonium salts as thermally stable, non-metal alternatives for arylation reactions due to their tunable reactivity and ease of handling.

Structure and Bonding

Cyclic Structures

Cyclic halonium ions, particularly the prototypical three-membered ring species, are characterized by a three-center two-electron (3c-2e) bonding model in which the atom bridges two adjacent carbon atoms, forming partial bonds with each. In this arrangement, the utilizes its p-orbitals to interact with the carbon-carbon π-system, resulting in a delocalized positive charge distributed across the C-X-C framework, where X represents or . The idealized symmetric structure for the halonium ion is represented as [\ce{CH2-CH2-X}]^{+}, with X = Cl or Br, featuring equivalent C-X distances typically in the range of 2.0–2.2 Å for bromonium ions based on calculations. The C-X-C bond angle at the halogen is approximately 60°, reflecting the severe inherent to the three-membered geometry. In symmetric cases like the ethylene bromonium ion, the bridging is equitable, with both C-X bonds identical and the halogen positioned equidistant from the carbons, minimizing electronic . However, substitution on the carbons introduces steric or electronic perturbations, leading to unsymmetric rings where the shifts toward the less substituted or more electron-rich carbon, resulting in disparate C-X bond lengths (e.g., one shortened to ~1.99 and the other elongated to ~2.39 in certain derivatives). This arises from differential charge stabilization, as confirmed by NMR studies of stable cyclic halonium ions. Representative examples include the bromonium ion, a in alkene bromination reactions, where the 3c-2e bond enforces anti addition stereochemistry. In carbohydrate chemistry, ions form bridged structures with glycals, facilitating stereoselective glycosylations and highlighting the role of these species in synthesis. The pronounced in these cyclic halonium ions, exacerbated by the small C-X-C angle and compressed geometry, enhances their electrophilicity, with chloronium ions exhibiting greater reactivity than bromonium counterparts due to the smaller atom imposing higher angular distortion (~60° vs. slightly larger for Br). This strain contributes to their transient nature and propensity for nucleophilic ring opening, though isolated examples persist under conditions.

Acyclic and Hypervalent Structures

Acyclic feature a linear, open-chain structure of the form R–X⁺–R', where X is a and R, R' are alkyl or aryl groups bound directly to the central via two-coordinate bonds, without bridging or formation. These species are less common than their cyclic counterparts due to their inherent instability, as the positive charge on the leads to rapid or rearrangement, particularly for lighter like . For instance, dialkylchloronium ions such as [Cl(CH₃)₂]⁺ are highly reactive and thermally unstable, decomposing rapidly even at low temperatures, while fluorinated analogs like [Cl(CH₂CF₃)₂]⁺ exhibit slightly improved stability but still require low-temperature conditions for isolation. In these open-chain forms, the C–X–C bond angles are typically around 100–102°, as observed in structures of [Br(CH₂CF₃)₂]⁺ with C–Br bond lengths of approximately 1.96 Å. Hypervalent halonium ions, particularly iodonium species, represent a distinct class where the central exceeds the , accommodating two ligands with 10 valence electrons in a bent . Diaryliodonium salts, such as Ph₂I⁺ (diphenyliodonium), exemplify this hypervalency and are notably stable as air- and moisture-resistant crystalline solids, unlike simpler acyclic forms. The bonding in these hypervalent ions is best described by the three-center four-electron (3c-4e) model, involving hypervalent interactions that utilize s and p orbitals on the iodine without significant d-orbital participation, though earlier interpretations invoked d-orbital involvement for the expanded octet. A representative structure for diaryliodonium ions is Ar–I⁺–Ar', where the I–C bond lengths are approximately 2.1 , as measured in various salts like diphenyliodonium fluoborate (I–C ≈ 2.02 , C–I–C ≈ 94°). Fluoronium analogs of hypervalent halonium ions are rare due to fluorine's poor ability to support hypervalency, stemming from its high and limited capacity for expanded coordination. These hypervalent structures contrast with simpler acyclic ions by leveraging the larger size and lower electronegativity of iodine to stabilize the 10-electron configuration.

Synthesis and Characterization

Generation Methods

Halonium ions are commonly generated through the of or interhalogens to , forming cyclic three-membered ring intermediates. In this process, a molecule such as Br₂ approaches the π-bond of the , leading to heterolytic cleavage and formation of a bromonium with the . For example, the reaction of with Br₂ produces the ethylenebromonium , which can be represented as: \ce{CH2=CH2 + Br2 -> [CH2-CH2-Br]+ Br-} This method is widely used in organic synthesis for anti addition reactions, with conditions typically involving non-nucleophilic solvents like dichloromethane at room temperature. Stable halonium ions, including both cyclic and acyclic variants, can be generated and isolated in superacid media, pioneered by George A. Olah. Using Magic Acid (a mixture of fluorosulfuric acid, FSO₃H, and antimony pentafluoride, SbF₅), alkyl halides or dihalides are ionized at low temperatures (e.g., -60°C) to form long-lived halonium ions with weakly coordinating anions like Sb₂F₁₁⁻. For instance, 1,2-dibromoethane in Magic Acid yields the ethylenepromonium ion, allowing spectroscopic study without decomposition. This approach has been instrumental in characterizing hypervalent and bridged structures. Hypervalent halonium ions, particularly diaryliodonium salts, are synthesized via oxidation of aryl iodides with arenes in the presence of peroxides. A common one-pot protocol involves treating an iodoarene (e.g., iodobenzene) with an arene (e.g., ) and m-chloroperbenzoic acid (mCPBA) in , followed by addition of (TfOH) to form the iodonium with yields often exceeding 70%. The reaction proceeds through hypervalent iodine(III) intermediates, such as (diacetoxyiodo)benzene, oxidized to the dicationic species. oxidants like Oxone® offer sustainable alternatives, enabling scalable synthesis under mild conditions. These salts serve as transferable aryl cation equivalents, isolable with counterions like BF₄⁻ or OTf⁻. Halonium ions also form via neighboring group participation, particularly in halocyclization reactions leading to five-membered rings. For example, allylic halides or alcohols undergo intramolecular cyclization with electrophilic sources like N-bromosuccinimide (NBS) in aqueous media, where the halogen migrates to form a bromonium ion , facilitating stereospecific ring closure. Olah demonstrated this in the ionization of vicinal dihalides, such as 1,3-dibromopropane in , yielding propylenebromonium ions through 1,3-halogen shift. This method is prevalent in synthesizing tetrahydrofurans or pyrrolidines from unsaturated precursors, with conditions tuned to and solvent for selectivity.

Spectroscopic and Structural Analysis

Nuclear magnetic resonance (NMR) spectroscopy provides key evidence for the bridged structure of halonium ions in solution, particularly through the equivalence of signals for symmetric cyclic systems. In the bromonium ion generated in media, the ¹H NMR exhibits a symmetric CH₂ signal at approximately δ 5 , reflecting the rapid bridging of the atom and the resulting equivalence of the two protons. Similarly, the ¹³C NMR shows a single downfield-shifted signal for the bridged carbons at around δ 100-110 , consistent with the three-center two-electron (3c-2e) bonding and lack of distinction between the two carbons. These observations, first reported by Olah and co-workers using NMR on various cyclic halonium ions, confirm the symmetric nature of the bridge and distinguish it from open-chain haloalkylcarbenium ion alternatives. Infrared (IR) and further support the structural characterization of halonium ions by identifying characteristic vibrational modes associated with the C–X–C bridge. For bromonium ions, a prominent band at approximately 700 cm⁻¹ in the IR and Raman spectra corresponds to the asymmetric stretching of the C–Br–C unit, providing a diagnostic signature for the bridged geometry. These vibrational frequencies, observed in solutions of cyclic halonium ions, align with the expected weakening of the C–X bonds due to the delocalized 3c-2e interaction and have been used to corroborate solution-phase structures. X-ray crystallography has provided definitive solid-state evidence for halonium ion structures, with the first reported crystal structure of a bromonium ion in 1985 by Olah and colleagues confirming the predicted 3c-2e bonding motif. In this structure, the bromine-carbon distances were measured at approximately 1.98 , longer than typical C–Br single bonds (1.94 ) but indicative of partial bonding in the three-membered ring, while the C–C distance remained consistent with a single bond at about 1.50 . This seminal work on a stabilized bromonium ion derivative resolved long-standing debates about the intermediacy of such species in electrophilic additions and established benchmark geometric parameters for the class. Mass spectrometry techniques, particularly electrospray ionization mass spectrometry (ESI-MS), have enabled the observation of intact halonium ions in the gas phase, especially for more stable iodonium salts. For example, symmetric diaryliodonium ions appear as prominent molecular ions in ESI-MS spectra without fragmentation, allowing confirmation of their connectivity and stability under soft ionization conditions. These measurements, often coupled with collision-induced dissociation to probe bond strengths, complement solution studies by demonstrating the persistence of the bridged structure in isolated ions. Computational methods, such as (DFT), have been employed to validate experimental spectroscopic and structural data for halonium ions. At the B3LYP/6-31G* level, calculated geometries for cyclic bromonium ions reproduce the observed Br–C distances of ~1.98 Å and C–C bond lengths, while predicted NMR chemical shifts match experimental ¹H (δ ~5 ) and ¹³C values within 5-10 , supporting the symmetric bridged model. These simulations, benchmarked against Olah's experimental data, provide insights into the electronic structure and aid in interpreting subtle asymmetries in larger systems.

Reactivity and Mechanisms

Electrophilic Behavior

Halonium ions serve as potent electrophiles in the of , initiating the by interacting with the π-electron density of the carbon-carbon . The process begins when a dihalogen (X₂, where X = Cl, Br, or I) approaches the , forming a three-center cyclic halonium ion and releasing a halide anion (X⁻). This bridged structure, first proposed to explain the stereospecificity of halogen additions, positions the halogen atom above the plane of the , shielding one face from further attack. The electrophilic nature of the halonium ion enforces strict anti addition stereochemistry in the overall reaction. The subsequent nucleophilic attack occurs from the opposite side of the bridged , leading to trans delivery of the two substituents across the original . This backside displacement mechanism was experimentally confirmed through the stereospecific conversion of cis- and trans-2-butene to the corresponding meso and racemic dibromides, respectively, without inversion or syn addition products. A representative example is the bromination of cyclohexene with Br₂, which proceeds via a bromonium ion intermediate to yield trans-1,2-dibromocyclohexane. In the presence of water, the bromide anion is displaced by H₂O as the nucleophile, forming a bromohydrin such as trans-2-bromocyclohexanol. The reaction can be summarized as: \text{Alkene} + \ce{X2} \rightarrow [\text{halonium ion}]^{+} + \ce{X-} \quad \text{(followed by nucleophile attack)} This pathway ensures stereospecific anti addition while avoiding carbocation rearrangements common in other electrophilic additions. In unsymmetrical alkenes, the halonium ion bridge is asymmetric, with the positive charge more localized on the carbon better able to stabilize it (typically the more substituted one). This partial charge distribution directs , where the preferentially attacks the more substituted carbon, mimicking Markovnikov orientation without forming a free . For instance, in the aqueous bromination of propene, the bromonium leads predominantly to 1-bromopropan-2-ol, with the group at the secondary carbon.

Nucleophilic Reactions and Rearrangements

Halonium ions, particularly cyclic variants formed from alkenes, serve as highly electrophilic intermediates that undergo ring-opening reactions upon attack by s. The nucleophilic attack typically proceeds in an relative to the halogen bridge, resulting in stereospecific trans addition products. This mechanism combines features of SN1 and SN2 pathways, with the three-membered ring imparting partial character to the carbons while enforcing backside attack. In unsymmetrical halonium ions, the preferentially targets the more substituted carbon atom, which bears greater positive charge density, thereby dictating . Common nucleophiles include , which leads to formation, as exemplified by the reaction of a bromonium with H₂O to yield a trans-2-bromohydrin. Similarly, halide s such as Br⁻ or Cl⁻ attack to form vicinal dihalides, while alcohols like produce β-halo ethers through crossed addition. These reactions are widely observed in under aqueous or alcoholic conditions, maintaining the anti stereochemistry due to the bridged structure. For instance, the addition of Br₂ to in generates trans-2-bromocyclohexanol as the major product. Rearrangements occur when the halonium bridge collapses to an open , a process facilitated in polar solvents where stabilizes the dissociated form, or for less stable halonium ions such as chloronium or fluoronium species that exhibit greater open-ion character. This collapse enables carbocation migrations, including the Wagner-Meerwein rearrangement, observed in the of where skeletal reorganization leads to more stable polycyclic products. Temperature-dependent NMR studies reveal equilibria between cyclic halonium ions and open carbonium ions, with shifts indicating dynamic interconversion that promotes such rearrangements. In semipinacol-type processes, allylic alcohols undergo halonium-mediated 1,2-shifts to form β-halo carbonyl compounds. The simplified representation of nucleophilic ring-opening is given by: [\ce{R2C-X-CR2}]^+ + \ce{Nu^-} \rightarrow \ce{R2C(Nu)-X-CR2} where X denotes the and Nu the , highlighting the regioselective addition at the bridged carbons.

Applications and Recent Advances

Role in Organic Synthesis

Halonium ions play a pivotal role in by serving as electrophilic intermediates that enable stereoselective cyclizations and functionalizations under mild conditions. In halocyclization reactions, intramolecular formation of halonium ions from unsaturated alcohols facilitates the construction of oxygen-containing heterocycles such as tetrahydrofurans and tetrahydropyrans. For instance, treatment of homoallylic alcohols with N-bromosuccinimide (NBS) or iodine sources generates bromonium or iodonium intermediates, which are subsequently attacked by the pendant hydroxyl group, yielding exo- or endo-cyclized products with high . These transformations are particularly valuable for building fused ring systems, as the bridged halonium structure directs nucleophilic attack from the opposite face, ensuring anti . Asymmetric variants of halocyclization have advanced the field by incorporating chiral ligands or catalysts to achieve enantioselectivity. Chiral phosphoramidites or thioureas, often in catalytic amounts, coordinate with the source to bias the formation of one enantiotopic face of the halonium ion, enabling stereocontrol in additions to alkenes. For example, Lewis base catalysis with chiral Brønsted acids promotes enantioselective bromocycloetherification of 4-penten-1-ols to tetrahydrofurans with up to 95% . These methods draw inspiration from earlier asymmetric protocols but adapt them for halogen-mediated processes, providing access to enantioenriched heterocycles essential for pharmaceutical intermediates. Hypervalent iodine reagents, such as diaryliodonium salts, function as aryl cation equivalents akin to halonium ions, offering mild electrophiles for C–H activation and arylation. In palladium- or copper-catalyzed reactions, these salts transfer aryl groups to indoles or arenes via selective C–H functionalization, often in tandem with N–H arylation to form diarylated products in good yields (e.g., 65% for 1,3-diphenylindole). The mild conditions (60–110°C) and broad functional group tolerance make them superior to traditional cross-coupling methods. Similarly, Selectfluor acts as a fluoronium-like reagent for electrophilic fluorination, enabling regioselective introduction of fluorine into alkenes or aromatics under ambient conditions, as seen in the synthesis of fluorinated piperazines. In synthesis, halonium-mediated cyclizations have been instrumental in assembling complex toxins. Bromocyclization of trans-(+)-laurediol derivatives using tetrabromocyclohexa-2,5-dien-1-one constructs the ring in trans-(+)-deacetylkumausyne, a from Laurencia majuscula, with improved stereoselectivity (5:1 :syn) via intermediates. These approaches highlight the utility in forging polycyclic ethers found in ladder toxins. Compared to routes, halonium ion pathways offer superior stereocontrol through bridged intermediates that prevent rearrangements and enable diastereoselective additions (>15:1 in some cases), while operating under mild, neutral conditions compatible with sensitive substrates like polyenes in scaffolds.

Modern Developments and Examples

In 2018, researchers reported the first spectroscopic evidence for a [C–F–C]⁺ in solution, characterized through extensive ¹⁹F, ¹H, and ¹³C NMR studies that revealed a symmetric cage-like structure with equivalent carbon atoms bridged by the central , confirming its elusive divalent fluoronium nature. This breakthrough addressed long-standing challenges in observing such hypervalent species in non-solid states, building on earlier gas-phase detections. Subsequent structural confirmation came in 2021 via of a modified double-norbornyl-type fluoronium cation, which displayed near-covalent F–C bonds and a symmetric , resolving debates over its bonding and stability. During the 2020s, computational studies have provided deeper insights into halonium ion stability, particularly through quantum chemical modeling of bonding interactions. For instance, analyses in 2020 demonstrated the stabilization of iodonium ions in three-center, four-electron O–I–O bonds with oxygen ligands, revealing bond strengths comparable to bonds and highlighting their potential in supramolecular assemblies. These models emphasize the role of electrostatic and charge-transfer contributions in enhancing stability within complex molecular environments. Recent examples illustrate the integration of halonium ions into and synthetic protocols. Halogen-bonded frameworks (XOFs) based on iodonium-bridged N⋅⋅⋅I⁺⋅⋅⋅N interactions represent a type of periodic network, as reported in 2021. In 2025, stable chlorine(I)-bridged two-dimensional halogen-bonded frameworks (XOFs) were reported, demonstrating potential for . Additionally, [N···X···N]⁺-based XOFs enabled ultrafast proton conduction. Iodonium species continue to find applications in , including visible-light-mediated reactions with hypervalent iodine reagents for selective carbon-halogen couplings. These developments address gaps in by promoting solvent-free halogenations, such as mechanochemical protocols using N-halosuccinimides for efficient, catalyst-free of and anilines. Fluorine-substituted electrolytes have been explored for lithium-metal batteries, improving interfacial stability and reducing during cycling, though with a slight decrease in ionic , as shown in studies (e.g., 2.1 × 10⁻⁴ S cm⁻¹ at for Li₂ZrCl₅.₅F₀.₅).

References

  1. [1]
    Onium ions. X. Structural study of acyclic and cyclic halonium ions ...
    Structural study of acyclic and cyclic halonium ions by carbon-13 nuclear magnetic resonance spectroscopy. Question of intra- and intermolecular equilibration ...
  2. [2]
    Capturing the generation and structural transformations of molecular ...
    Jan 10, 2024 · A case in point is the halonium ions, whose highly reactive nature and ring strain make them short-lived intermediates that are readily attacked ...
  3. [3]
    [PDF] George A. Olah - Nobel Lecture
    The developed superacidic, “stable ion” methods also gained wide application in the preparation of other ionic intermediates (nitronium, halonium, oxoni- urn ...<|control11|><|separator|>
  4. [4]
    Structure and Bonding of Halonium Compounds | Inorganic Chemistry
    May 31, 2023 · A review. Due to their electron deficiency, halonium ions act as particularly strong halogen bond donors. By accepting electrons in both lobes ...
  5. [5]
    Halogen bonds of halonium ions - RSC Publishing
    Mar 25, 2020 · Following the discussion of the nature and properties of halonium ions' halogen bonds, this tutorial review provides an overview of their ...
  6. [6]
    [PDF] Carbenium and Halonium Ions with F, Cl, Br, and I - SMU
    Feb 20, 2014 · The stability of halonium ions increases with the atomic number of X, which is reflected by a strengthening of the fractional. (electron ...
  7. [7]
    Halogen Bonds of Halogen(I) Ions—Where Are We and Where to Go?
    Halonium ions are open-chain or cyclic onium ions of the form D2X+, where X is a halogen. ... Halonium ions bound to two carbons, [C–X–C]+, do not possess halogen ...<|control11|><|separator|>
  8. [8]
    Isolation and Reactivity of Trifluoromethyl Iodonium Salts
    May 13, 2016 · (10) The solubility properties of 1·HCl were consistent with an ionic compound, as the salt exhibited poor solubility in low polarity solvents ( ...
  9. [9]
    Journal of the American Chemical Society
    The crystal and molecular structure of the bromonium ion of adamantylideneadamantane ... X-ray Structures and Anionotropic Rearrangements of Di-tert-butyl ...Missing: biadamantylidene original
  10. [10]
    Halogen Bonds of Halogen(I) Ions Where Are We and Where to Go?
    Due to their electron deficiency, halonium ions act as particularly strong halogen bond donors. By accepting electrons in both lobes of their empty p-orbital, ...
  11. [11]
    Addition of Heteroatom Radicals to endo-Glycals † - MDPI
    The chloronium ion 6b is subsequently trapped by the solvent ... sugars, important building blocks for carbohydrate chemistry. Thus, various ...
  12. [12]
  13. [13]
  14. [14]
    Crystal structure of diphenyliodonium fluoborate | Russian Chemical ...
    The bond lengths have the following values with an accuracy of ± 0.03 A: I-C=2.02, C-C=1.40, and B-F=1.43 A. 3. The cation (C6H5)2I+ has an angular ...
  15. [15]
    Stable carbonium ions. LXII. Halonium ion formation via neighboring ...
    LXII. Halonium ion formation via neighboring halogen participation: ethylenehalonium, propylenehalonium, and 1,2-dimethylethylenehalonium ions.
  16. [16]
    Advancements in the Synthesis of Diaryliodonium Salts: Updated ...
    May 4, 2023 · Our group has reported several one-pot protocols for the synthesis of diaryliodonium salts, which have been recognized as attractive multi-purpose reagents.
  17. [17]
    Sustainable and scalable one-pot synthesis of diaryliodonium salts
    Mar 3, 2025 · protocol utilizing mCPBA and TfOH/TsOH,2a–c thus making it a more sustainable option in the synthesis of diaryliodonium salts. Finally, 2e ...
  18. [18]
    Stereoselective Halogenation in Natural Product Synthesis - PMC
    The intermediacy of a bromonium species generated via neighboring group participation of the bromine was proposed. ... Reversibility in halonium ion ...
  19. [19]
    [N⋅⋅⋅I+⋅⋅⋅N] Halogen‐Bonded Dimeric Capsules from ...
    Oct 6, 2016 · Electrospray ionization mass spectrometry (ESI-MS) measurements provide clear evidence of the selective formation of the halogen-bonded capsules ...
  20. [20]
    Halogen Bond of Halonium Ions: Benchmarking DFT Methods for ...
    Nov 2, 2020 · A review. Due to their electron deficiency, halonium ions act as particularly strong halogen bond donors. By accepting electrons in both lobes ...
  21. [21]
    The Halogenation of Ethylenes | Journal of the American Chemical ...
    Research Article May 1, 1937. The Halogenation of Ethylenes. Click to ... Intramolecular Halogen Transfer via Halonium Ion Intermediates in the Gas Phase.
  22. [22]
  23. [23]
    Halonium ion rearrangements. Temperature-dependent carbon-13 ...
    Halonium ion rearrangements. Temperature-dependent carbon-13 chemical shifts as indicators of equilibriums between cyclic halonium ions and open carbonium ions.
  24. [24]
    Halonium Ion - an overview | ScienceDirect Topics
    Halonium ion is defined as a highly electrophilic intermediate generated from halogenium ions that facilitates the semipinacol rearrangement of allylic alcohols ...
  25. [25]
    Halogen-Induced Controllable Cyclizations as Diverse Heterocycle ...
    In this review, innovative reactions controlled by halogen groups are discussed as a new concept in the field of organic synthesis.
  26. [26]
    Catalytic, Asymmetric Halofunctionalization of Alkenes—A Critical ...
    Sep 25, 2012 · Such processes typically proceed by the formation of an olefin/dihalogen π complex with subsequent ionization to form a cyclic halonium ( ...
  27. [27]
    Asymmetric Synthesis of Halocyclized Products by Using Various ...
    Jul 25, 2022 · This review will address all the catalytic asymmetric Halocyclization reactions such as fluorocyclization reaction, chlorocyclization reaction, ...
  28. [28]
    Tandem C–H and N–H Arylation of Indoles - ACS Publications
    Jan 14, 2015 · This paper describes a new palladium-catalyzed method for C-H activation/carbon-carbon bond formation with hypervalent iodine arylating agents.Supporting Information · Author Information · References
  29. [29]
    Recent Advances in the Application of SelectfluorTMF-TEDA-BF4 as ...
    SelectfluorTMF-TEDA-BF4 is used for electrophilic fluorination, as a mediator/catalyst in oxidations, and in functionalizations like iodination, bromination, ...
  30. [30]
    Strategies for the Synthesis of Cyclic Ethers of Marine Natural Products
    toxins,2 annonaceous acetogenins,3 and lauroxanes4 (Fig- ure 1). Many ... be biosynthesized through cyclization of the natural meth- ylene-interrupted ...
  31. [31]
    Spectroscopic Characterization of a [C−F−C]+ Fluoronium Ion in ...
    Jan 9, 2018 · We report the first spectroscopic evidence for a [C−F−C] + fluoronium ion in solution. Extensive NMR studies ( 19 F, 1 H, 13 C) characterize a symmetric cage- ...
  32. [32]
    Structural proof of a [C–F–C]+ fluoronium cation - Nature
    Sep 6, 2021 · Finally, in 2018 they supported the formation of the aforementioned fluoronium ion by NMR spectroscopy16,17; yet, the structural proof of this ...
  33. [33]
    O–I–O halogen bond of halonium ions - RSC Publishing
    Jul 13, 2020 · Our study provides new insights into both halogen bonding and 3c4e bonds. Additionally, [O–I–O]+ complexes are of interest for the ...
  34. [34]
    Halogen hydrogen-bonded organic framework (XHOF) constructed ...
    Mar 16, 2022 · A halogen hydrogen-bonded organic framework (XHOF) is fabricated by using Cl − ions as central connection nodes to connect organic ligands.
  35. [35]
    Recent advances in catalytic asymmetric haloamination and ...
    Aug 15, 2023 · ... halonium ion intermediate with a nucleophile intra- or intermolecularly [1]. As widely available nucleophilic elements, O- and N-nucleophiles ...
  36. [36]
    Fluorine-Substituted Halide Solid Electrolytes with Enhanced ...
    Aug 1, 2023 · The fluorine-substituted compound exhibits substantially lower polarization after 800 h of lithium stripping and plating owing to enhanced interfacial ...Missing: halonium | Show results with:halonium
  37. [37]
    Automated grindstone chemistry: a simple and facile way for PEG ...
    Aug 9, 2022 · A simple electrical mortar–pestle was used for the development of a green and facile mechanochemical route for the catalyst-free halogenation of phenols and ...