Organosodium chemistry

Last updated

Organosodium chemistry is the chemistry of organometallic compounds containing a carbon to sodium chemical bond. [1] [2] The application of organosodium compounds in chemistry is limited in part due to competition from organolithium compounds, which are commercially available and exhibit more convenient reactivity.

Contents

The principal organosodium compound of commercial importance is sodium cyclopentadienide. Sodium tetraphenylborate can also be classified as an organosodium compound since in the solid state sodium is bound to the aryl groups.

Organometal bonds in group 1 are characterised by high polarity with corresponding high nucleophilicity on carbon. This polarity results from the disparate electronegativity of carbon (2.55) and that of lithium 0.98, sodium 0.93 potassium 0.82 rubidium 0.82 caesium 0.79). The carbanionic nature of organosodium compounds can be minimized by resonance stabilization, for example, Ph3CNa. One consequence of the highly polarized Na-C bond is that simple organosodium compounds often exist as polymers that are poorly soluble in solvents.

Synthesis

Transmetallation routes

In the original work the alkylsodium compound was accessed from the dialkylmercury compound by transmetallation. For example, diethylmercury in the Schorigin reaction or Shorygin reaction: [3] [4] [5]

(C2H5)2Hg + 2 Na → 2 C2H5Na + Hg

The high solubility of lithium alkoxides in hexane is the basis of a useful synthetic route: [6]

LiCH2SiMe3 + NaO–t–Bu → LiOt–Bu + NaCH2SiMe3

Deprotonation routes

For some acidic organic compounds, the corresponding organosodium compounds arise by deprotonation. Sodium cyclopentadienide is thus prepared by treating sodium metal and cyclopentadiene: [7]

2 Na+ 2 C5H6 → 2 Na+ C5H5 + H2

Sodium acetylides form similarly. Often strong sodium bases are employed in place of the metal. Sodium methylsulfinylmethylide is prepared by treating DMSO with sodium hydride: [8]

CH3SOCH3 + NaH → CH3SOCH
2
Na+ + H2

Metal-halogen exchange

Trityl sodium can be prepared by sodium-halogen exchange: [9]

Ph3CCl + 2 Na → Ph3C Na+ + NaCl

Electron transfer

Sodium also reacts with polycyclic aromatic hydrocarbons via one-electron reduction. With solutions of naphthalene, it forms the deeply coloured radical sodium naphthalene, which is used as a soluble reducing agent:

C10H8 + Na → Na+[C10H8]−•

Structural studies show however that sodium naphthalene has no Na-C bond, the sodium is invariably coordinated by ether or amine ligands. [10] The related anthracene as well as lithium derivatives are well known.

Structures

Structure of (C6H5)3CNa(thf)3 ("trityl sodium"), omitting all but the oxygen of the thf ligands. Selected distances: rNa-C(central)=256 pm, rNa-C(ipso) = 298 pm (avg of three). YICYOB.png
Structure of (C6H5)3CNa(thf)3 ("trityl sodium"), omitting all but the oxygen of the thf ligands. Selected distances: rNa-C(central)=256 pm, rNa-C(ipso) = 298 pm (avg of three).

Simple organosodium compounds such as the alkyl and aryl derivatives are generally insoluble polymers. Because of its large radius, Na prefers a higher coordination number than does lithium in organolithium compounds. Methyl sodium adopts a polymeric structure consisting of interconnected [NaCH3]4 clusters. [12] When the organic substituents are bulky and especially in the presence of chelating ligands like TMEDA, the derivatives are more soluble. For example, [NaCH2SiMe3]TMEDA is soluble in hexane. Crystals have been shown to consist of chains of alternating Na(TMEDA)+ and CH2SiMe
3
groups with Na–C distances ranging from 2.523(9) to 2.643(9) Å. [6]

Structure of the phenylsodium-PMDTA adduct, hydrogen atoms omitted for clarity. Phenylsodium-PMDTA adduct.svg
Structure of the phenylsodium-PMDTA adduct, hydrogen atoms omitted for clarity.

Reactions

Organosodium compounds are traditionally used as strong bases, [9] although this application has been supplanted by other reagents such as sodium bis(trimethylsilyl)amide.

The higher alkali metals are known to metalate even some unactivated hydrocarbons and are known to self-metalate:

2 NaC2H5 → C2H4Na2 + C2H6

In the Wanklyn reaction (1858) [14] [15] organosodium compounds react with carbon dioxide to give carboxylates:

C2H5Na + CO2 → C2H5CO2Na

Grignard reagents undergo a similar reaction.

Some organosodium compounds degrade by beta-elimination:

NaC2H5 → NaH + C2H4

Industrial applications

Although organosodium chemistry has been described to be of "little industrial importance", it once was central to the production of tetraethyllead. [16] A similar Wurtz coupling-like reaction is the basis of the industrial route to triphenylphosphine:

3 PhCl + PCl3 + 6 Na → PPh3 + 6 NaCl

The polymerization of butadiene and styrene is catalyzed by sodium metal. [3]

Organic derivatives of the heavier alkali metals

Organopotassium, organorubidium, and organocaesium compounds are less commonly encountered than organosodium compounds and are of limited utility. These compounds can be prepared by treatment of alkyl lithium compounds with the potassium, rubidium, and caesium alkoxides. Alternatively they arise from the organomercury compound, although this method is dated. The solid methyl derivatives adopt polymeric structures. Reminiscent of the nickel arsenide structure, MCH3 (M = K, Rb, Cs) has six alkali metal centers bound to each methyl group. The methyl groups are pyramidal, as expected. [12]

A notable reagent that is based on a heavier alkali metal alkyl is Schlosser's base, a mixture of n-butyllithium and potassium tert-butoxide. This reagent reacts with toluene to form the red-orange compound benzyl potassium (KCH2C6H5).

Evidence for the formation of heavy alkali metal-organic intermediates is provided by the equilibration of cis-but-2-ene and trans-but-2-ene catalysed by alkali metals. The isomerization is fast with lithium and sodium, but slow with the higher alkali metals. The higher alkali metals also favor the sterically congested conformation. [17] Several crystal structures of organopotassium compounds have been reported, establishing that they, like the sodium compounds, are polymeric. [6]

See also

Related Research Articles

<span class="mw-page-title-main">Alkali metal</span> Group of highly reactive chemical elements

The alkali metals consist of the chemical elements lithium (Li), sodium (Na), potassium (K), rubidium (Rb), caesium (Cs), and francium (Fr). Together with hydrogen they constitute group 1, which lies in the s-block of the periodic table. All alkali metals have their outermost electron in an s-orbital: this shared electron configuration results in their having very similar characteristic properties. Indeed, the alkali metals provide the best example of group trends in properties in the periodic table, with elements exhibiting well-characterised homologous behaviour. This family of elements is also known as the lithium family after its leading element.

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

A metallocene is a compound typically consisting of two cyclopentadienyl anions (C
5
H
5
, abbreviated Cp) bound to a metal center (M) in the oxidation state II, with the resulting general formula (C5H5)2M. Closely related to the metallocenes are the metallocene derivatives, e.g. titanocene dichloride or vanadocene dichloride. Certain metallocenes and their derivatives exhibit catalytic properties, although metallocenes are rarely used industrially. Cationic group 4 metallocene derivatives related to [Cp2ZrCH3]+ catalyze olefin polymerization.

<span class="mw-page-title-main">Organolithium reagent</span> Chemical compounds containing C–Li bonds

In organometallic chemistry, organolithium reagents are chemical compounds that contain carbon–lithium (C–Li) bonds. These reagents are important in organic synthesis, and are frequently used to transfer the organic group or the lithium atom to the substrates in synthetic steps, through nucleophilic addition or simple deprotonation. Organolithium reagents are used in industry as an initiator for anionic polymerization, which leads to the production of various elastomers. They have also been applied in asymmetric synthesis in the pharmaceutical industry. Due to the large difference in electronegativity between the carbon atom and the lithium atom, the C−Li bond is highly ionic. Owing to the polar nature of the C−Li bond, organolithium reagents are good nucleophiles and strong bases. For laboratory organic synthesis, many organolithium reagents are commercially available in solution form. These reagents are highly reactive, and are sometimes pyrophoric.

<span class="mw-page-title-main">Cyclopentadienyl complex</span> Coordination complex of a metal and Cp⁻ ions

A cyclopentadienyl complex is a coordination complex of a metal and cyclopentadienyl groups. Cyclopentadienyl ligands almost invariably bind to metals as a pentahapto (η5-) bonding mode. The metal–cyclopentadienyl interaction is typically drawn as a single line from the metal center to the center of the Cp ring.

In organometallic chemistry, acetylide refers to chemical compounds with the chemical formulas MC≡CH and MC≡CM, where M is a metal. The term is used loosely and can refer to substituted acetylides having the general structure RC≡CM. Acetylides are reagents in organic synthesis. The calcium acetylide commonly called calcium carbide is a major compound of commerce.

<span class="mw-page-title-main">Radical anion</span> 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 .

<span class="mw-page-title-main">Xanthate</span> Salt that is a metal-thioate/O-esters of dithiocarbonate

Xanthate usually refers to a salt of xanthic acid. The formula of the salt of xanthic acid is [R−O−CS2]M+ ,. Xanthate also refers to the anion [R−O−CS2]. Xanthate also may refer to an ester of xanthic acid. The formula of xanthic acid is R−O−C(=S)−S−H, while the formula of the esters of xanthic acid is R−O−C(=S)−S−R', where R and R' are organyl groups. The salts of xanthates are also called O-organyl dithioates. The esters of xanthic acid are also called O,S-diorganyl esters of dithiocarbonic acid. The name xanthate is derived from Ancient Greek ξανθός xanthos, meaning “yellowish, golden”, and indeed most xanthate salts are yellow. They were discovered and named in 1823 by Danish chemist William Christopher Zeise. These organosulfur compounds are important in two areas: the production of cellophane and related polymers from cellulose and for extraction of certain sulphide bearing ores. They are also versatile intermediates in organic synthesis.

<i>n</i>-Butyllithium Chemical compound

n-Butyllithium C4H9Li (abbreviated n-BuLi) is an organolithium reagent. It is widely used as a polymerization initiator in the production of elastomers such as polybutadiene or styrene-butadiene-styrene (SBS). Also, it is broadly employed as a strong base (superbase) in the synthesis of organic compounds as in the pharmaceutical industry.

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

Methyllithium is the simplest organolithium reagent with the empirical formula CH3Li. This s-block organometallic compound adopts an oligomeric structure both in solution and in the solid state. This highly reactive compound, invariably used in solution with an ether as the solvent, is a reagent in organic synthesis as well as organometallic chemistry. Operations involving methyllithium require anhydrous conditions, because the compound is highly reactive toward water. Oxygen and carbon dioxide are also incompatible with MeLi. Methyllithium is usually not prepared, but purchased as a solution in various ethers.

<span class="mw-page-title-main">Organomercury chemistry</span> Group of chemical compounds containing mercury

Organomercury chemistry refers to the study of organometallic compounds that contain mercury. Typically the Hg–C bond is stable toward air and moisture but sensitive to light. Important organomercury compounds are the methylmercury(II) cation, CH3Hg+; ethylmercury(II) cation, C2H5Hg+; dimethylmercury, (CH3)2Hg, diethylmercury and merbromin ("Mercurochrome"). Thiomersal is used as a preservative for vaccines and intravenous drugs.

Sodium atoms have 11 electrons, one more than the stable configuration of the noble gas neon. As a result, sodium usually forms ionic compounds involving the Na+ cation. Sodium is a reactive alkali metal and is much more stable in ionic compounds. It can also form intermetallic compounds and organosodium compounds. Sodium compounds are often soluble in water.

<span class="mw-page-title-main">Sodium tetraphenylborate</span> Chemical compound

Sodium tetraphenylborate is the organic compound with the formula NaB(C6H5)4. It is a salt, wherein the anion consists of four phenyl rings bonded to boron. This white crystalline solid is used to prepare other tetraphenylborate salts, which are often highly soluble in organic solvents. The compound is used in inorganic and organometallic chemistry as a precipitating agent for potassium, ammonium, rubidium, and cesium ions, and some organic nitrogen compounds.

<span class="mw-page-title-main">Sodium cyclopentadienide</span> Chemical compound

Sodium cyclopentadienide is an organosodium compound with the formula C5H5Na. The compound is often abbreviated as NaCp, where Cp is the cyclopentadienide anion. Sodium cyclopentadienide is a colorless solid, although samples often are pink owing to traces of oxidized impurities.

Organovanadium chemistry is the chemistry of organometallic compounds containing a carbon (C) to vanadium (V) chemical bond. Organovanadium compounds find only minor use as reagents in organic synthesis but are significant for polymer chemistry as catalysts.

Plumbocene is an organometallic compound of lead. It is a member of the class of metallocenes. It is soluble in benzene, acetone, ether, and petroleum ether, and insoluble in water. Plumbocene is stable in cold water.

<span class="mw-page-title-main">Metal bis(trimethylsilyl)amides</span>

Metal bis(trimethylsilyl)amides are coordination complexes composed of a cationic metal with anionic bis(trimethylsilyl)amide ligands and are part of a broader category of metal amides.

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

Phenylsodium C6H5Na is an organosodium compound. Solid phenylsodium was first isolated by Nef in 1903. Although the behavior of phenylsodium and phenyl magnesium bromide are similar, the organosodium compound is very rarely used.

<span class="mw-page-title-main">Lithium naphthalene</span> Chemical compound

Lithium naphthalene is an organic salt with the chemical formula Li+C
10
H
8
. In the research laboratory, it is used as a reductant in the synthesis of organic, organometallic, and inorganic chemistry. It is usually generated in situ. Lithium naphthalene crystallizes with ligands bound to Li+.

An arsinide, arsanide, dihydridoarsenate(1−) or arsanyl compound is a chemical derivative of arsine, where one hydrogen atom is replaced with a metal or cation. The arsinide ion has formula AsH−2. It can be considered as a ligand with name arsenido or arsanido. Researchers are unenthusiastic about studying arsanyl compounds, because of the toxic chemicals, and their instability. The IUPAC names are arsanide and dihydridoarsenate(1−). For the ligand the name is arsanido. The neutral −AsH2 group is termed arsanyl.

Germyl, trihydridogermanate(1-), trihydrogermanide, trihydridogermyl or according to IUPAC Red Book: germanide is an anion containing germanium bounded with three hydrogens, with formula GeH−3. Germyl is the IUPAC term for the –GeH3 group. For less electropositive elements the bond can be considered covalent rather than ionic as "germanide" indicates. Germanide is the base for germane when it loses a proton.

References

  1. Synthesis of Organometallic Compounds: A Practical Guide Sanshiro Komiya Ed. 1997
  2. C. Elschenbroich, A. Salzer Organometallics : A Concise Introduction (2nd Ed) (1992) from Wiley-VCH: Weinheim. ISBN   3-527-28165-7
  3. 1 2 Dietmar Seyferth "Alkyl and Aryl Derivatives of the Alkali Metals: Strong Bases and Reactive Nucleophiles. 2. Wilhelm Schlenk's Organoalkali-Metal Chemistry. The Metal Displacement and the Transmetalation Reactions. Metalation of Weakly Acidic Hydrocarbons. Superbases" Organometallics 2009, volume 28, pp 2–33. doi : 10.1021/om801047n
  4. P. Schorigin, "Beiträge zur Kenntnis der Kondensationen, welche durch Natrium bewirkt werden" Chem. Berichte 1907, vol. 40, 3111. doi : 10.1002/cber.19070400371
  5. P. Schorigin, "Synthesen mittels Natrium und Halogenalkylen" Ber. 41, 2711. doi : 10.1002/cber.190804102208. P. Schorigin, "Über die Natriumalkyl-Verbindungen und über die mit ihrer Hilfe ausgeführten Synthesen" Berichte, 1908, vol. 41, 2717. doi : 10.1002/cber.190804102209. P. Schorigin, "Synthesen mittels Natrium und Halogenalkylen" Chem. Berichte, 1908, volume 41, 2711. doi : 10.1002/cber.190804102208. P. Schorigin, "Neue Synthese aromatischer Carbonsäuren aus den Kohlenwasserstoffen" Chem. Berichte 1908, vol. 41, 2723. doi : 10.1002/cber.190804102210.
  6. 1 2 3 William Clegg, Ben Conway, Alan R. Kennedy, Jan Klett, Robert E. Mulvey, Luca Russo "Synthesis and Structures of [(Trimethylsilyl)methyl]sodium and -potassium with Bi- and Tridentate N-Donor Ligands" Eur. J. Inorg. Chem. 2011, pp. 721–726. doi : 10.1002/ejic.201000983
  7. Tarun K. Panda, Michael T. Gamer, and Peter W. Roesky "An Improved Synthesis of Sodium and Potassium Cyclopentadienide" Organometallics 2003, 22, 877-878. doi : 10.1021/om0207865
  8. Iwai, I.; Ide, J. (1988). "2,3-Diphenyl-1,3-Butadiene". Organic Syntheses .; Collective Volume, vol. 6, p. 531
  9. 1 2 W. B. Renfrow Jr & C. R. Hauser (1939). "Triphenylmethylsodium". Organic Syntheses . 19: 83. doi:10.15227/orgsyn.019.0083.
  10. Bock, Hans; Arad, Claudia; Näther, Christian; Havlas, Zdenek (1995). "The Structures of Solvent-Separated Naphthalene and Anthracene Radical Anions". J. Chem. Soc., Chem. Commun. (23): 2393–2394. doi:10.1039/C39950002393.
  11. Vrána, Jan; Jambor, Roman; Růžička, Aleš; Dostál, Libor (2018). "New Synthetic Strategies Leading to [RNPNR]− Anions and the Isolation of the [P(Nt-Bu)3]3− Trianion". Dalton Transactions. 47 (25): 8434–8441. doi:10.1039/C8DT02141D. PMID   29897357.
  12. 1 2 E. Weiss, "Structures of Organo Alkali Metal Complexes and Related Compounds" Angewandte Chemie International Edition in English, 1993, volume 32, pages 1501–1523. doi : 10.1002/anie.199315013
  13. Schümann, Uwe; Behrens, Ulrich; Weiss, Erwin (April 1989). "Synthese und Struktur von Bis[μ-phenyl(pentamethyldiethylentriamin)natrium], einem Phenylnatrium-Solvat". Angewandte Chemie. 101 (4): 481–482. Bibcode:1989AngCh.101..481S. doi:10.1002/ange.19891010420.
  14. J. A. Wanklyn, Ann. 107, 125 (1858)
  15. The Merck index of chemicals and drugs: an encyclopedia for chemists, Paul G. Stecher
  16. Rolf Ackermann, Ludwig Lange "Sodium Compounds, Organic" in Ullmann's Encyclopedia of Industrial Chemistry, 2005 Wiley-VCH, Weinheim. doi : 10.1002/14356007.a24_341
  17. Manfred Schlosser (1988). "Superbases for organic synthesis". Pure Appl. Chem. 60 (11): 1627–1634. doi: 10.1351/pac198860111627 .