Oxonium ion

Last updated

In chemistry, an oxonium ion is any cation containing an oxygen atom that has three bonds and 1+ formal charge. [1] The simplest oxonium ion is the hydronium ion (H3O+). [2]

Contents

Alkyloxonium

Hydronium is one of a series of oxonium ions with the formula RnH3−nO+. Oxygen is usually pyramidal with an sp3 hybridization. Those with n = 1 are called primary oxonium ions, an example being protonated alcohol (e.g. methanol). In acidic media, the oxonium functional group produced by protonating an alcohol can be a leaving group in the E2 elimination reaction. The product is an alkene. Extreme acidity, heat, and dehydrating conditions are usually required. Other hydrocarbon oxonium ions are formed by protonation or alkylation of alcohols or ethers (R−C−+O−R1R2).

Secondary oxonium ions have the formula R2OH+, an example being protonated ethers.

Tertiary oxonium ions have the formula R3O+, an example being trimethyloxonium. [3] Tertiary alkyloxonium salts are useful alkylating agents. For example, triethyloxonium tetrafluoroborate (Et
3
O+
)(BF
4
), a white crystalline solid, can be used, for example, to produce ethyl esters when the conditions of traditional Fischer esterification are unsuitable. [4] It is also used for preparation of enol ethers and related functional groups. [5] [6]

Oxonium-ion-2D.png Trimethyloxonium-2D-skeletal.png Trimethyloxonium-3D-balls.png Trimethyloxonium-3D-vdW.png
general pyramidal
oxonium ion
skeletal formula of the
trimethyloxonium cation
ball-and-stick model
of trimethyloxonium
space-filling model
of trimethyloxonium

Oxatriquinane and oxatriquinacene are unusually stable oxonium ions, first described in 2008. Oxatriquinane does not react with boiling water or with alcohols, thiols, halide ions, or amines, although it does react with stronger nucleophiles such as hydroxide, cyanide, and azide.

Oxocarbenium ions

Another class of oxonium ions encountered in organic chemistry is the oxocarbenium ions, obtained by protonation or alkylation of a carbonyl group e.g. R−C=+O−R′ which forms a resonance structure with the fully-fledged carbocation R−+C−O−R′ and is therefore especially stable:

Carbonyl-oxonium-resonance-2D-skeletal.png

Gold-stabilized species

Triphenylphosphinegoldoxonium.png

An unusually stable oxonium species is the gold complex tris[triphenylphosphinegold(I)]oxonium tetrafluoroborate, [(Ph3PAu)3O][BF4], where the intramolecular aurophilic interactions between the gold atoms are believed responsible for the stabilisation of the cation. [7] [8] This complex is prepared by treatment of Ph3PAuCl with Ag2O in the presence of NaBF4: [9]

3 Ph3PAuCl + Ag2O + NaBF4 → [(Ph3PAu)3O]+[BF4] + 2 AgCl + NaCl

It has been used as a catalyst for the propargyl Claisen rearrangement. [10]

Relevance to natural product chemistry

Complex bicyclic and tricyclic oxonium ions have been proposed as key intermediates in the biosynthesis of a series of natural products by the red algae of the genus Laurencia . [11]

Oxoniumionlaurencia2.png

Several members of these elusive species have been prepared explicitly by total synthesis, demonstrating the possibility of their existence. [11] The key to their successful generation was the use of a weakly coordinating anion (Krossing's anion, [Al(pftb)4], pftb = perfluoro-tert-butoxy) as the counteranion. [12] As shown in the example below, this was executed by a transannular halide abstraction strategy through the reaction of the oxonium ion precursor (an organic halide) with the silver salt of the Krossing's anion Ag[Al(pftb)4]•CH2Cl2, generating the desired oxonium ion with simultaneous precipitation of inorganic silver halides. The resulting oxonium ions were characterized comprehensively by nuclear magnetic resonance spectroscopy at low temperature (−78 °C) with support from density functional theory computation.

Wiki fig 2.png

These oxonium ions were also demonstrated to directly give rise to multiple related natural products by reacting with various nucleophiles, such as water, bromide, chloride, and acetate. [13] [14] [15]

Wiki fig 3.png

See also

Related Research Articles

Williamson ether synthesis

The Williamson ether synthesis is an organic reaction, forming an ether from an organohalide and a deprotonated alcohol (alkoxide). This reaction was developed by Alexander Williamson in 1850. Typically it involves the reaction of an alkoxide ion with a primary alkyl halide via an SN2 reaction. This reaction is important in the history of organic chemistry because it helped prove the structure of ethers.

<span class="mw-page-title-main">Zinc chloride</span> Chemical compound

Zinc chloride is the name of inorganic chemical compounds with the formula ZnCl2 and its hydrates. Zinc chlorides, of which nine crystalline forms are known, are colorless or white, and are highly soluble in water. This salt is hygroscopic and even deliquescent. Zinc chloride finds wide application in textile processing, metallurgical fluxes, and chemical synthesis. No mineral with this chemical composition is known aside from the very rare mineral simonkolleite, Zn5(OH)8Cl2·H2O.

<span class="mw-page-title-main">Aluminium chloride</span> Chemical compound

Aluminium chloride, also known as aluminium trichloride, is an inorganic compound with the formula AlCl3. It forms hexahydrate with the formula [Al(H2O)6]Cl3, containing six water molecules of hydration. Both are colourless crystals, but samples are often contaminated with iron(III) chloride, giving a yellow color.

In chemistry, the phosphonium cation describes polyatomic cations with the chemical formula PR+
4
. These cations have tetrahedral structures. The salts are generally colorless or take the color of the anions.

Anions that interact weakly with cations are termed non-coordinating anions, although a more accurate term is weakly coordinating anion. Non-coordinating anions are useful in studying the reactivity of electrophilic cations. They are commonly found as counterions for cationic metal complexes with an unsaturated coordination sphere. These special anions are essential components of homogeneous alkene polymerisation catalysts, where the active catalyst is a coordinatively unsaturated, cationic transition metal complex. For example, they are employed as counterions for the 14 valence electron cations [(C5H5)2ZrR]+ (R = methyl or a growing polyethylene chain). Complexes derived from non-coordinating anions have been used to catalyze hydrogenation, hydrosilylation, oligomerization, and the living polymerization of alkenes. The popularization of non-coordinating anions has contributed to increased understanding of agostic complexes wherein hydrocarbons and hydrogen serve as ligands. Non-coordinating anions are important components of many superacids, which result from the combination of Brønsted acids and Lewis acids.

Radical anion Free radical species

In organic chemistry, a radical anion is a free radical species that carries a negative charge. Radical anions are encountered in organic chemistry as reduced derivatives of polycyclic aromatic compounds, e.g. sodium naphthenide. An example of a non-carbon radical anion is the superoxide anion, formed by transfer of one electron to an oxygen molecule. Radical anions are typically indicated by .

In the Ullmann condensation or Ullmann-type reaction is the copper-promoted conversion of aryl halides to aryl ethers, aryl thioethers, aryl nitriles, and aryl amines. These reactions are examples of cross-coupling reactions.

<span class="mw-page-title-main">Sulfonium</span> Cation of the form [SR3]+

In organic chemistry, a sulfonium ion, also known as sulphonium ion or sulfanium ion, is a positively-charged ion featuring three organic substituents attached to sulfur. These organosulfur compounds have the formula [SR3]+. Together with a negatively-charged counterion, they give sulfonium salts. They are typically colorless solids that are soluble in organic solvent.

<span class="mw-page-title-main">Diazonium compound</span> Diazonium salts of formula R-N≡N+

Diazonium compounds or diazonium salts are a group of organic compounds sharing a common functional group [R−N+≡N]X where R can be any organic group, such as an alkyl or an aryl, and X is an inorganic or organic anion, such as a halide.

Geminal halide hydrolysis is an organic reaction. The reactants are geminal dihalides with a water molecule or a hydroxide ion. The reaction yields ketones from secondary halides or aldehydes from primary halides.

<span class="mw-page-title-main">Enol ether</span>

In organic chemistry an enol ether is an alkene with an alkoxy substituent. The general structure is R2C=CR-OR where R = H, alkyl or aryl. A common subfamily of enol ethers are vinyl ethers, with the formula ROCH=CH2. Important enol ethers include the reagent 3,4-dihydropyran and the monomers methyl vinyl ether and ethyl vinyl ether.

18-Crown-6 Chemical compound

18-Crown-6 is an organic compound with the formula [C2H4O]6 and the IUPAC name of 1,4,7,10,13,16-hexaoxacyclooctadecane. It is a white, hygroscopic crystalline solid with a low melting point. Like other crown ethers, 18-crown-6 functions as a ligand for some metal cations with a particular affinity for potassium cations (binding constant in methanol: 106 M−1). The point group of 18-crown-6 is S6. The dipole moment of 18-crown-6 varies in different solvent and under different temperature. Under 25 °C, the dipole moment of 18-crown-6 is 2.76 ± 0.06 D in cyclohexane and 2.73 ± 0.02 in benzene. The synthesis of the crown ethers led to the awarding of the Nobel Prize in Chemistry to Charles J. Pedersen.

Grignard reagent Organometallic compounds used in organic synthesis

A Grignard reagent or Grignard compound is a chemical compound with the generic formula R−Mg−X, where X is a halogen and R is an organic group, normally an alkyl or aryl. Two typical examples are methylmagnesium chloride Cl−Mg−CH3 and phenylmagnesium bromide (C6H5)−Mg−Br. They are a subclass of the organomagnesium compounds.

In chemistry, an onium ion is a cation formally obtained by the protonation of mononuclear parent hydride of a pnictogen, chalcogen, or halogen. The oldest-known onium ion, and the namesake for the class, is ammonium, NH+4, the protonated derivative of ammonia, NH3.

<span class="mw-page-title-main">Chloro(triphenylphosphine)gold(I)</span> Chemical compound

Chloro(triphenylphosphine)gold(I) or triphenylphosphinegold(I) chloride is a coordination complex with the formula (Ph3P)AuCl. This colorless solid is a common reagent for research on gold compounds.

Disodium tetracarbonylferrate Chemical compound

Disodium tetracarbonylferrate is the organoiron compound with the formula Na2[Fe(CO)4]. It is always used as a solvate, e.g., with tetrahydrofuran or dimethoxyethane, which bind to the sodium cation. An oxygen-sensitive colourless solid, it is a reagent in organometallic and organic chemical research. The dioxane solvated sodium salt is known as Collman's reagent, in recognition of James P. Collman, an early popularizer of its use.

Organotellurium chemistry describes the synthesis and properties of chemical compounds containing a carbon-tellurium chemical bond. Organotellurium chemistry is a lightly studied area, in part because of the few applications.

Oxatriquinane (oxoniaperhydrotriquinacene) is an alkyl oxonium ion with formula (CH2CH2CH)3O+. It has a cyclononane backbone, with a trivalent oxygen connected to carbon 1, 4, and 7, forming three fused pentagonal rings. In contrast to most trialkyloxonium ions, oxatriquinane hydrolyzes slowly.

Organogold chemistry is the study of compounds containing gold–carbon bonds. They are studied in academic research, but have not received widespread use otherwise. The dominant oxidation states for organogold compounds are I with coordination number 2 and a linear molecular geometry and III with CN = 4 and a square planar molecular geometry. The first organogold compound discovered was gold(I) carbide Au2C2, which was first prepared in 1900.

Tetrakis(3,5-bis(trifluoromethyl)phenyl)borate Chemical compound

Tetrakis[3,5-bis(trifluoromethyl)phenyl]borate is an anion with chemical formula [{3,5-(CF3)2C6H3}4B], which is commonly abbreviated as [BArF4], indicating the presence of fluorinated aryl (ArF) groups. It is sometimes referred to as Kobayashi's anion in honour of Hiroshi Kobayashi who led the team that first synthesised it. More commonly it is affectionately nicknamed "BARF." The BARF ion is also abbreviated BArF24, to distinguish it from the closely related BArF
20
, [(C6F5)4B].

References

  1. March, Jerry (2007), Advanced Organic Chemistry: Reactions, Mechanisms, and Structure (4th ed.), New York: Wiley, p. 497
  2. Olah, George A. (1998). Onium Ions. John Wiley & Sons. p. 509. ISBN   9780471148777.
  3. Olah, George A. (1993). "Superelectrophiles". Angew. Chem. Int. Ed. Engl. 32 (6): 767–788. doi:10.1002/anie.199307673.
  4. Raber, Douglas J.; Gariano Jr, Patrick; Brod, Albert O.; Gariano, Anne L.; Guida, Wayne C. (1977). "Esterification Of Carboxylic Acids With Trialkyloxonium Salts: Ethyl And Methyl 4-acetoxybenzoates". Org. Synth. 56: 59. doi:10.15227/orgsyn.056.0059.
  5. Struble, Justin R.; Bode, Jeffrey W. (2010). "Synthesis Of A N-mesityl Substituted Aminoindanol-derived Triazolium Salt". Org. Synth. 87: 362. doi: 10.15227/orgsyn.087.0362 .
  6. Hegedus, Lous S.; Mcguire, Michael A.; Schultze, Lisa M. (1987). "1,3-Dimethyl-3-methoxy-4-phenylazetidinone". Org. Synth. 65: 140. doi:10.15227/orgsyn.065.0140.
  7. Schmidbaur, Hubert (2000). "The Aurophilicity Phenomenon: A Decade of Experimental Findings, Theoretical Concepts and Emerging Application". Gold Bulletin . 33 (1): 3–10. doi: 10.1007/BF03215477 .
  8. Schmidbaur, Hubert (1995). "Ludwig Mond Lecture: High-Carat Gold Compounds". Chem. Soc. Rev. 24 (6): 391–400. doi:10.1039/CS9952400391.
  9. Bruce, M. I.; Nicholson, B. K.; Bin Shawkataly, O.; Shapley, J. R.; Henly, T. (1989). "Synthesis of Gold-Containing Mixed-Metal Cluster Complexes". In Kaesz, Herbert D. (ed.). Inorganic Syntheses. Vol. 26. John Wiley & Sons, Inc. pp. 324–328. doi:10.1002/9780470132579.ch59. ISBN   9780470132579.
  10. Sherry, Benjamin D.; Toste, F. Dean (2004). "Gold(I)-Catalyzed Propargyl Claisen Rearrangement" (PDF). Journal of the American Chemical Society . 126 (49): 15978–15979. CiteSeerX   10.1.1.604.7272 . doi:10.1021/ja044602k. ISSN   0002-7863. PMID   15584728.
  11. 1 2 Sam Chan, Hau Sun; Nguyen, Q. Nhu N.; Paton, Robert S.; Burton, Jonathan W. (2019-10-09). A full list of references encompassing the contributions from Braddock, Snyder, Murai, Suzuki, Fukuzawa, Burton, Kim, and Fox are available inside. "Synthesis, Characterization, and Reactivity of Complex Tricyclic Oxonium Ions, Proposed Intermediates in Natural Product Biosynthesis". Journal of the American Chemical Society. 141 (40): 15951–15962. doi:10.1021/jacs.9b07438. ISSN   0002-7863. PMID   31560524. S2CID   203580092.
  12. Krossing, Ingo (2001). "The Facile Preparation of Weakly Coordinating Anions: Structure and Characterisation of Silverpolyfluoroalkoxyaluminates AgAl(ORF)4, Calculation of the Alkoxide Ion Affinity". Chemistry – A European Journal. 7 (2): 490–502. doi:10.1002/1521-3765(20010119)7:2<490::aid-chem490>3.0.co;2-i. ISSN   1521-3765. PMID   11271536.
  13. Wang, Bin-Gui; Gloer, James B.; Ji, Nai-Yun; Zhao, Jian-Chun (March 2013). "Halogenated Organic Molecules of Rhodomelaceae Origin: Chemistry and Biology". Chemical Reviews. 113 (5): 3632–3685. doi:10.1021/cr9002215. ISSN   0009-2665. PMID   23448097.
  14. Zhou, Zhen-Fang; Menna, Marialuisa; Cai, You-Sheng; Guo, Yue-Wei (2015-02-11). "Polyacetylenes of Marine Origin: Chemistry and Bioactivity". Chemical Reviews. 115 (3): 1543–1596. doi:10.1021/cr4006507. ISSN   0009-2665. PMID   25525670.
  15. Wanke, Tauana; Philippus, Ana Cláudia; Zatelli, Gabriele Andressa; Vieira, Lucas Felipe Oliveira; Lhullier, Cintia; Falkenberg, Miriam (2015-11-01). "C15 acetogenins from the Laurencia complex: 50 years of research – an overview". Revista Brasileira de Farmacognosia. 25 (6): 569–587. doi: 10.1016/j.bjp.2015.07.027 . ISSN   0102-695X.