Methylidynetricobaltnonacarbonyl

Last updated
Methylidynetricobalt­nonacarbonyl
HCCo3(CO)9.png
HCCo3sample.jpg
Names
Other names
methylidyne-tris(tricarbonylcobalt)
Identifiers
3D model (JSmol)
PubChem CID
  • InChI=1S/9CO.CH.3Co/c9*1-2;;;;/h;;;;;;;;;1H;;;/q9*+1;;3*-3
    Key: WPOHNFRRRVAWSE-UHFFFAOYSA-N
  • [O+]#C[Co-3]12(C#[O+])(C#[O+])[Co-3]3(C#[O+])(C#[O+])(C#[O+])[Co-3]1(C#[O+])(C#[O+])(C#[O+])C23
Properties
C10HCo3O9
Molar mass 441.909 g·mol−1
Appearancepurple solid
Density 2.01 g/cm3
Melting point 105–107 °C (221–225 °F; 378–380 K)
Except where otherwise noted, data are given for materials in their standard state (at 25 °C [77 °F], 100 kPa).

Methylidynetricobaltnonacarbonyl is an organometallic cobalt cluster with the chemical formula Co3(CO)9CH that contains a metal carbonyl core with the methylidyne ligand, first discovered in the late 1950s. A variety of substituents can be added to the methylidyne group to form derivatives of the parent compound that have unique spectroscopic properties and reactivity. This page will explore the discovery and synthesis of methylidynetricobaltnonacarbonyl, the structure and bonding of the parent compound, as well as some examples reactivity and catalysis with the cluster.

Contents

Discovery and Synthesis

Methylidynetricobaltnonacarbonyl was first discovered in the late 1950s by Markby and Wender, though the absolute structure of the molecule could not be determined. [1] The group obtained Nuclear Magnetic Resonance (NMR), Infrared Spectroscopy (IR), and dipole moment data on a new trinuclear cobalt complex formed from reacting dicobaltoctacarbonyl with acetylenes, but were unable to determine the structure and hoped that X-ray crystallography would provide the evidence needed to determine the structure of the new compound. In 1962, independent groups from Italy and Germany reported the synthesis and properties of methylidynetricobaltnonacarbonyl as well as several derivatives. [2] The structure has been unambiguously determined via X-ray crystallography, demonstrating idealized C3v symmetry with three equivalent six-coordinate cobalt atoms. [3]

The synthetic procedure developed by Bor and coworkers relied on reacting the [Co(CO)4]- anion with chloroform, bromoform, or iodoform in a solution of acetone or THF to afford the desired product as dark violet crystals in ~20% yield with chloroform, ~18% yield with bromoform, and ~1-2% yield with iodoform. The chloride derivative Co3(CO)9CCl was obtained by reacting CCl4 with the [Co(CO)4]- anion to afford shiny, lilac crystals. Several other functional groups could be installed, including C6H5 and CO2CH3, all of which afforded dark violet crystals. The compounds demonstrated extraordinary stability in air and were dissolved without decomposition in organic solvents. Treatment of the parent compound with 100 atm CO at 160 °C resulted in the destruction of the complex and the formation of dicobaltoctacarbonyl: Co2(CO)8. [3]

Bor's synthesis of methylidynetricobaltnonacarbonyl from chloroform and the cobalttetracarbonyl anion Formation of methylidynetricobaltnonacarbonyl from chloroform.png
Bor's synthesis of methylidynetricobaltnonacarbonyl from chloroform and the cobalttetracarbonyl anion

Structure and Bonding

Crystal structure of methylidynetricobaltnonacarbonyl Crystal structure of methylidynetricobaltnonacarbonyl.png
Crystal structure of methylidynetricobaltnonacarbonyl

The Co-Co bond lengths, as determined by X-ray crystallography are ~2.47 Å and the Co-C bond distances are ~1.88 Å. [3] [4] These bond lengths do not change significantly depending on the substituent on the carbon group. The bond lengths confirm the presence of C-Co and Co-Co single bonds, confirming the structure drawn above. [4] The Co-C-Co angle is approximately 80°, indicating a significant bending of the carbon tetrahedral structure, and the ligands are arranged to provide maximum cobalt-carbon interaction. [5] The IR spectra demonstrate four absorption bands of varying intensities in the region of terminal C-O groups: between 2111-2101, 2066-2054, 2047-2038, and 2034-2018. The electronegative character of the chlorine atom in Co3(CO)9CCl causes lower C-O frequencies than Co3(CO)9CH, indicating that electron density is being pulled from the metal atoms to the chlorine, causing electrons to shift from the carbonyl groups to the metal atoms, giving them more double-bond character and decreasing their frequency. This effect can be seen with other electron withdrawing substituents, as well as the inverse with electron-donating substituents. [3] This suggests that the cluster behaves as an electron sink capable of interacting with the substituent on carbon via pi bonds, which is a theory that is also supported by the chemical shift of the carbon on the substituent in 13C NMR, as the chemical shift of the apical carbon increases with more electron withdrawing substituents. [6] The mass spectra demonstrate parent molecular ions (Co3(CO)9CH .+) in good abundance, which demonstrates the high stability of the Co3(CO)9 clusters towards oxidation. Initial loss of CO, which is characteristic of most metal-carbonyls, though rupture of the metal-metal bond also often occurs. [5]

Series of mixed organic and inorganic tetrahedrane structures resulting from the isolobal nature of Co(CO)3 and methylidyne ClusterScheme.png
Series of mixed organic and inorganic tetrahedrane structures resulting from the isolobal nature of Co(CO)3 and methylidyne

The d9-ML3 fragment Co(CO)3 fragment is isolobal with methylidyne, CH. Though there are differences in the a1 and e energies of the two fragments, there are a series of organometallic tetrahedranes of the C4, CoC3, Co2C2, Co3C, and Co4 variety, where a variety of ligands and functional groups can be installed on the cobalt and carbon groups, respectively. [8] Hoffmann demonstrated that a substituent with an acceptor orbital of appropriate symmetry could interact with the highest energy orbital to withdraw electron density from the filled e set on the apex carbon. [8] Seyferth demonstrated that the carbonium ion had the correct symmetry to accept electron density from the apex carbon via s-p stabilization. They explained that apex carbon stabilize the carbonium ion via s-p conjugation. [9]

Visual representation of methylidynetricobaltnonacarbonyl isolobal analogy with the CH fragment Methylidynetricobaltnonacarbonyl isolobal analogy.png
Visual representation of methylidynetricobaltnonacarbonyl isolobal analogy with the CH fragment

Early reactivity studies on methylidynetricobaltnoncarbonyl were conducted by Cetini et al. in the early 1960s by conducting 14CO labelling studies on the kinetics of carbonyl ligand exchange. [10] The group concluded the exchange rate of the carbonyl ligands decreases according to the order F > Cl > Br > H, which they attributed to the electronegativity of the atom bound to the methylidyne carbon. Additionally, they demonstrated that the nine carbonyl groups are not kinetically equivalent, and that only 3 CO groups exchange within the 35 – 55 °C temperature range studied, which agrees with the calculated frontier molecular orbitals discussed above. In the case with the COOCH3 group, the exchange occurs for only one CO group which they attributed to the more electron deficient nature of the complex. [10]

Reactivity

MIT’s Dietmar Seyferth explored the reactivity of methylidynetricobaltnonacarbonyls beginning the early 1970s, and his work uncovered many of the interesting properties of the metal carbonyl cluster. In 1970, his group demonstrated that a variety of organomercury compounds of the R2Hg and RHgX type are capable of alkylating methylidynetricobaltnonacarbonyl with good yields and facile synthetic procedures. [11] The group observed the side product Hg[Co(CO)4]2 was forming during the reaction and hypothesized that if the decomposition of the starting material was due to the reversible loss of CO ligand, then conducting these reactions under a CO atmosphere could improve yields. This was observed, and yields of the reaction increased to >90%. [11]

Methylidynetricobaltnonacarbonyl reactivity with alkylated mercury compounds to form derivatives of the parent compound Methylidynetricobaltnonacarbonyl reactivity with mercury.png
Methylidynetricobaltnonacarbonyl reactivity with alkylated mercury compounds to form derivatives of the parent compound

In 1973, Seyferth and coworkers demonstrated that adding azobisisobutryonitrile (AIBN), a common reagent for the initiation of radical reactions, to a solution of Co3(CO)9CH and allyl acetate afforded a red crystalline solid that they believed formed via a radical mechanism. [12] Around the same time, Czájlik and coworkers were exploring how metal carbonyls initiate radical polymerization, as they hoped that the tricobalt complex would be more active than Co2(CO)8 and more stable than Mo(CO)6. The derivatives of Co3(CO)9CX demonstrated different activity depending on the structure of X, which aligns with their hypothesis that the first step of initiation would be loss of a CO ligand. The order of initiator activity was found to be X = Cl > H > Br > Ph > F > i-Pr > C2F5, which aligns with the trend seen for CO ligand exchange studies. [13] In 1974, Seyferth and coworkers published their mechanistic insights into the reaction described above. The group highlighted that though the reaction may proceed via the homolytic decomposition of some Co3(CO)9CH molecules, the presence of a radical catalyst or a radical inhibitor did not change the rate of the reaction. [14] They also proposed that the electrophilic cleavage of the mercury-carbon bond could be initiated by Co3(CO)9CH, but the group couldn’t provide evidence for this theory either, so no conclusions could be drawn about the absolute mechanism for this reaction. [14]

In 1976, Reed and coworkers sought to explore the electrochemical properties of these cobalt carbonyl clusters to discover how the nature of the X group in Co3(CO)9CX changes the anodic and cathodic properties of the cluster. [15] The group observed a reversible, one-electron reduction in the range of -0.7 to -0.9V versus the saturated calomel electrode. They observed that clusters with more electronegative substituents are reduced around -0.8V while substituents with more electron donating groups are reduced around -0.9V, which is almost identical to the pattern observed for the reduction of ferrocene with the same substituents, demonstrating that substituents have the same effect on the HOMO and LUMO of ferrocene and the cobalt cluster. Subsequent oxidation of the reduced species occurred irreversibly around +1.5V, which was then irreversibly reduced at -1.0V, leading to another decomposition product that oxidized irreversibly at 0.4V. [15]

Several years later, in 1998, Sugihara, et al. demonstrated that the Pauson-Khand reaction could be catalyzed by the methylidynetricobaltnonacarbonyl cluster. The Pauson-Khand reaction allows for the cyclic cotrimerization of an alkyne, an alkene, and carbon monoxide [16] via a [2+2+1] cycloaddition [17] and has been widely utilized for the synthesis of cyclopentenones for application in natural products. [18] Methylidynetricobaltnonacarbonyl is more air-stable than the parent dicobalt octacarbonyl, making it a more attractive catalyst. The clusters demonstrated no need for additives such as trimethylphosphite, as is necessary with the dicobalt octacarbonyl, and the best results were obtained with the parent methylidyne cluster. [18]

Pauson-Khand reaction using methlylidynetricobaltnonacarbonyl as a catalyst PausonKhand.png
Pauson-Khand reaction using methlylidynetricobaltnonacarbonyl as a catalyst

More recently, Nordlander et al. have been exploring using the methylidynetricobaltnonacarbonyl derivatives as a precatalyst for the asymmetric intramolecular Pauson-Khand reaction. [19] The group treated the parent Co3(CO)9CH with the chiral Josiphos diphosphines to form the desired asymmetric precatalyst. The group determined that the clusters acted as pre-catalysts only as no clusteres were recovered from the reaction mixture after catalysis. Despite the unique approach, the clusters were found to be only moderately effective for this reaction, and are inferior compared to other metal clusters with NORPHOS or Me-DuPHOS ligands, which gave higher yields and fewer side products. [19]

Related Research Articles

Ferrocene is an organometallic compound with the formula Fe(C5H5)2. The molecule is a complex consisting of two cyclopentadienyl rings sandwiching a central iron atom. It is an orange solid with a camphor-like odor that sublimes above room temperature, and is soluble in most organic solvents. It is remarkable for its stability: it is unaffected by air, water, strong bases, and can be heated to 400 °C without decomposition. In oxidizing conditions it can reversibly react with strong acids to form the ferrocenium cation Fe(C5H5)+2. Ferrocene and the ferrocenium cation are sometimes abbreviated as Fc and Fc+ respectively.

<span class="mw-page-title-main">Carborane</span> Class of chemical compounds

Carboranes are electron-delocalized clusters composed of boron, carbon and hydrogen atoms. Like many of the related boron hydrides, these clusters are polyhedra or fragments of polyhedra. Carboranes are one class of heteroboranes.

The Stille reaction is a chemical reaction widely used in organic synthesis. The reaction involves the coupling of two organic groups, one of which is carried as an organotin compound (also known as organostannanes). A variety of organic electrophiles provide the other coupling partner. The Stille reaction is one of many palladium-catalyzed coupling reactions.

A transition metal carbene complex is an organometallic compound featuring a divalent carbon ligand, itself also called a carbene. Carbene complexes have been synthesized from most transition metals and f-block metals, using many different synthetic routes such as nucleophilic addition and alpha-hydrogen abstraction. The term carbene ligand is a formalism since many are not directly derived from carbenes and most are much less reactive than lone carbenes. Described often as =CR2, carbene ligands are intermediate between alkyls (−CR3) and carbynes (≡CR). Many different carbene-based reagents such as Tebbe's reagent are used in synthesis. They also feature in catalytic reactions, especially alkene metathesis, and are of value in both industrial heterogeneous and in homogeneous catalysis for laboratory- and industrial-scale preparation of fine chemicals.

<span class="mw-page-title-main">Pauson–Khand reaction</span> Chemical reaction

The Pauson–Khand (PK) reaction is a chemical reaction, described as a [2+2+1] cycloaddition. In it, an alkyne, an alkene and carbon monoxide combine into a α,β-cyclopentenone in the presence of a metal-carbonyl catalyst.

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

Organozinc chemistry is the study of the physical properties, synthesis, and reactions of organozinc compounds, which are organometallic compounds that contain carbon (C) to zinc (Zn) chemical bonds.

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

Dicobalt octacarbonyl is an organocobalt compound with composition Co2(CO)8. This metal carbonyl is used as a reagent and catalyst in organometallic chemistry and organic synthesis, and is central to much known organocobalt chemistry. It is the parent member of a family of hydroformylation catalysts. Each molecule consists of two cobalt atoms bound to eight carbon monoxide ligands, although multiple structural isomers are known. Some of the carbonyl ligands are labile.

The Nicholas reaction is an organic reaction where a dicobalt octacarbonyl-stabilized propargylic cation is reacted with a nucleophile. Oxidative demetallation gives the desired alkylated alkyne. It is named after Kenneth M. Nicholas.

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

Decamethyldizincocene is an organozinc compound with the formula [Zn25–C5Me5)2]. It is the first and an unusual example of a compound with a Zn-Zn bond. Decamethyldizincocene is a colorless crystalline solid that burns spontaneously in the presence of oxygen and reacts with water. It is stable at room temperature and especially soluble in diethyl ether, benzene, pentane, or tetrahydrofuran.

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

Hexadecacarbonylhexarhodium is a metal carbonyl cluster with the formula Rh6(CO)16. It exists as purple-brown crystals that are slightly soluble in dichloromethane and chloroform. It is the principal binary carbonyl of rhodium.

<span class="mw-page-title-main">Iron tetracarbonyl dihydride</span> Chemical compound

Iron tetracarbonyl dihydride is the organometallic compound with the formula H2Fe(CO)4. This compound was the first transition metal hydride discovered. The complex is stable at low temperatures but decomposes rapidly at temperatures above –20 °C.

Benzylic activation and stereocontrol in tricarbonyl(arene)chromium complexes refers to the enhanced rates and stereoselectivities of reactions at the benzylic position of aromatic rings complexed to chromium(0) relative to uncomplexed arenes. Complexation of an aromatic ring to chromium stabilizes both anions and cations at the benzylic position and provides a steric blocking element for diastereoselective functionalization of the benzylic position. A large number of stereoselective methods for benzylic and homobenzylic functionalization have been developed based on this property.

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

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

Transition metal carbyne complexes are organometallic compounds with a triple bond between carbon and the transition metal. This triple bond consists of a σ-bond and two π-bonds. The HOMO of the carbyne ligand interacts with the LUMO of the metal to create the σ-bond. The two π-bonds are formed when the two HOMO orbitals of the metal back-donate to the LUMO of the carbyne. They are also called metal alkylidynes—the carbon is a carbyne ligand. Such compounds are useful in organic synthesis of alkynes and nitriles. They have been the focus on much fundamental research.

<span class="mw-page-title-main">Metal-phosphine complex</span>

A metal-phosphine complex is a coordination complex containing one or more phosphine ligands. Almost always, the phosphine is an organophosphine of the type R3P (R = alkyl, aryl). Metal phosphine complexes are useful in homogeneous catalysis. Prominent examples of metal phosphine complexes include Wilkinson's catalyst (Rh(PPh3)3Cl), Grubbs' catalyst, and tetrakis(triphenylphosphine)palladium(0).

In organometallic chemistry, a transition metal alkyne complex is a coordination compound containing one or more alkyne ligands. Such compounds are intermediates in many catalytic reactions that convert alkynes to other organic products, e.g. hydrogenation and trimerization.

Metal arene complexes are organometallic compounds of the formula (C6R6)xMLy. Common classes are of the type (C6R6)ML3 and (C6R6)2M. These compounds are reagents in inorganic and organic synthesis. The principles that describe arene complexes extend to related organic ligands such as many heterocycles (e.g. thiophene) and polycyclic aromatic compounds (e.g. naphthalene).

<span class="mw-page-title-main">Metal cluster compound</span> Cluster of three or more metals

Metal cluster compounds are a molecular ion or neutral compound composed of three or more metals and featuring significant metal-metal interactions.

A Fischer carbene is a type of transition metal carbene complex, which is an organometallic compound containing a divalent organic ligand. In a Fischer carbene, the carbene ligand is a σ-donor π-acceptor ligand. Because π-backdonation from the metal centre is generally weak, the carbene carbon is electrophilic.

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

References

  1. Markby, R.; Wender, I.; Friedel, R. A.; Cotton, F. A.; Sternberg, H. W. J. Am. Chem. Soc. 1958, 80 (24), 6529-6533. DOI: 10.1021/ja01557a018.
  2. Ercoli, R. La Chimica e l’industria 1962, 44, 1344-1349.; Bor, G.; Markó, L.; Markó, B. Chem. Ber. 1962, 95 (2), 333-340. DOI: https://doi.org/10.1002/cber.19620950207.
  3. 1 2 3 4 Bor, G.; Markó, L.; Markó, B. Chem. Ber. 1962, 95 (2), 333-340. DOI: https://doi.org/10.1002/cber.19620950207.
  4. 1 2 Castellani, M. P.; Smith, W. G.; Patel, N. G.; Bott, S. G.; Richmond, M. G. J. Chem. Crystallogr. 1999, 29 (5), 609-617. DOI: 10.1023/A:1009561205829.
  5. 1 2 Robinson, B. H.; Tham, W. S. Journal of the Chemical Society A: Inorganic, Physical, Theoretical 1968,  (0), 1784-1787, 10.1039/J19680001784. DOI: 10.1039/J19680001784.
  6. Aime, S.; Milone, L.; Valle, M. Inorg. Chim. Acta 1976, 18, 9-11. DOI: https://doi.org/10.1016/S0020-1693(00)95577-4.
  7. Schilling, B. E. R.; Hoffmann, R. Journal fo the American Chemical Society 1979, 101 (13), 3456-3467.
  8. 1 2 Schilling, B. E. R.; Hoffmann, R. Journal fo the American Chemical Society 1979, 101 (13), 3456-3467.
  9. Seyferth, D.; Williams, G. H.; Hallgren, J. E. J. Am. Chem. Soc. 1973, 95 (1), 266-267. DOI: 10.1021/ja00782a060.
  10. 1 2 Cetini, N.; Ercoli, R.; Gambino, O.; Vaglio, G. Atti della Accademia delle Scienze di Torino, Clase di Scienze Fisiche, Matematiche e Naturali 1965, 99, 1123-1126.
  11. 1 2 Seyferth, D.; Hallgren, J. E.; Spohn, R. J. J. Organomet. Chem. 1970, 23 (2), C55-C57. DOI: https://doi.org/10.1016/S0022-328X(00)92937-1.
  12. Seyferth, D.; Hallgren, J. E. J. Organomet. Chem. 1973, 49 (1), C41-C42. DOI: https://doi.org/10.1016/S0022-328X(00)84929-3.
  13. Pályi, G.; Baumgartner, F.; Czájlik, I. J. Organomet. Chem. 1973, 49 (2), C85-C87. DOI: https://doi.org/10.1016/S0022-328X(00)84216-3.
  14. 1 2 Seyferth, D.; Hallgren, J. E.; Spohn, R. J.; Williams, G. H.; Nestle, M. O.; Hung, P. L. K. J. Organomet. Chem. 1974, 65 (1), 99-118. DOI: https://doi.org/10.1016/S0022-328X(00)83893-0.
  15. 1 2 Kotz, J. C.; Petersen, J. V.; Reed, R. C. J. Organomet. Chem. 1976, 120 (3), 433-437. DOI: https://doi.org/10.1016/S0022-328X(00)98054-9.
  16. Khand, I. U.; Knox, G. R.; Pauson, P. L.; Watts, W. E. Journal of the Chemical Society D: Chemical Communications 1971,  (1), 36a-36a, 10.1039/C2971000036A. DOI: 10.1039/C2971000036A.
  17. Schore, N. E. Chem. Rev. 1988, 88 (7), 1081-1119. DOI: 10.1021/cr00089a006.
  18. 1 2 Sugihara, T.; Yamaguchi, M. J. Am. Chem. Soc. 1998, 120 (41), 10782-10783. DOI: 10.1021/ja982635s.
  19. 1 2 Mottalib, M. A.; Haukka, M.; Nordlander, E. Polyhedron 2016, 103, 275-282. DOI: https://doi.org/10.1016/j.poly.2015.04.021.