Mesoionic carbene

Last updated

Mesoionic carbenes (MICs) are a type of reactive intermediate that are related to N-heterocyclic carbenes (NHCs) and are used in scientific research in chemistry. Unlike simple NHCs, the canonical resonance structures of these carbenes are mesoionic: an MIC cannot be drawn without adding additional charges to some of the atoms. MICs re also called abnormal N-heterocyclic carbenes (aNHC) or remote N-heterocyclic carbenes (rNHC). A variety of free carbenes can be isolated and are stable at room temperature. Other free carbenes are not stable and are susceptible to intermolecular decomposition pathways. MICs do not dimerize according to Wanzlick equilibrium as do normal NHCs. This results in relaxed steric requirements for mesoionic carbenes as compared to NHCs. [1] [2] [3] There are several mesoionic carbenes that cannot be generated as free compounds, but can be synthesized as a ligand in a transition metal complex. Most MIC transition metal complexes are less sensitive to air and moisture than phosphine or normal NHC complexes. They are also resistant to oxidation. The robust nature of MIC complexes is due to the ligand’s strong σ-donating ability. They are stronger σ-donors than phosphines, as well as normal N-heterocyclic carbenes due to decreased heteroatom stabilization. The strength of carbene ligands is attributed to the electropositive carbon center that forms strong bonds of a covalent nature with the metal. [1] [2] They have been shown to lower the frequency of CO stretching vibrations in metal complexes [4] [5] and exhibit large trans effects. [4] [6]

Contents

Classes

Imidazolin-4-ylidenes

The most studied mesoionic carbenes are based on imidazole and are referred to as imidazolin-4-ylidenes. These complexes were first reported by Crabtree in 2001. [7] The formation of imidazolin-4-ylidenes (MIC) instead of imidazolin-2-ylidenes (NHC) is typically a matter of blocking the C2 position. Most imidazolin-4-ylidenes are trisubstituted in the N1, C2, and N3 positions or tetrasubstituted. Electron-withdrawing groups in the N3 and C5 positions stabilize the carbenes more than electron-donating groups. [8] Free carbenes [8] [9] [10] as well as numerous transition metal complexes have been synthesized.

1,2,3-triazolylidenes

Also well studied are the mesoionic carbenes based on 1,2,3-triazole, referred to as 1,2,3-triazol-4(or 5)-ylidenes. The first triazolylidenes were reported by Albrecht in 2008. [11] These carbenes are typically trisubstituted with alkyl groups in the N1 and N3 positions and an aryl group in the C4 or C5 position. Free carbenes as well as numerous transition metal complexes have been reported. Free carbenes that are alkylated at N3 tend to undergo decomposition reactions in which the alkyl group participates in a nucleophilic attack at the carbene position. If N3 is substituted with a bulky alkyl group or an aryl group, the stability of the carbene increases significantly.

Pyrazolinylidenes

The first mesoionic carbenes based on pyrazole have been reported by Huynh in 2007. [12] These carbenes are referred to as pyrazolin-3(or 4)-ylidenes. Pyrazolin-4-ylidenes are often tetrasubstituted with alkyl or aryl groups; however, the C3 and C5 positions could be substituted with nitrogen- or oxygen-based groups. The electronic properties of the groups in the C3 and C5 positions affect the overall electron properties of the ligand and influence catalytic activity. Free carbene have been produced as well as transition metal complexes. [13] [14]

Others

Examples of tetrazol-5-ylidenes based on tetrazole have been prepared by Araki. [15] The N1 and N3 positions are substituted with alkyl or aryl groups. Transition metal complexes of these carbenes have been generated in situ. Mesoionic carbenes based on isoxazole and thiazole have been reported by Albrecht [16] and Bertrand [17] respectively. The isoxazol-4-ylidenes are trisubstituted in the N2, C3, and C5 positions with alkyl groups. The thiazol-5-ylidenes are trisubstituted in the C2, N3, and C4 positions with aryl groups. Transition metal complexes of both types of carbenes have been generated in situ. Bertrand also reported a 1,3-dithiol-5-ylidene based on 1,3-dithiolane, but it can only be isolated as a transition metal complex. [3]

Synthesis of free carbenes

Many free mesoionic carbenes are synthesized from their protonated salt form by deprotonation using strong potassium bases, such as potassium bis(trimethylsilyl)amide (KHMDS) or potassium tert-butoxide (KOt-Bu). Potassium bases are used because they do not form stable carbene-alkali metal adducts. [1] [8] [13] [14] [18]

Imidazolin-4-ylidenes (MIC) would form rather than imidazolin-2-ylidenes (NHC) due to blocking the C2 position. The C2 carbenes are thermodynamically more stable than their C4 counterparts due to resonance and inductive carbon-nitrogen interactions. Also, calculations show that the C4 hydrogen is less acidic than the C2 hydrogen of imidazole. This data suggests that the C2 position should be activated preferentially to the C4 position unless the C2 position is blocked. Aryl and bulky alkyl groups (such as isopropyl) are good at blocking the C2 position from being activated. [4] [18]

Carbene metal complexes

Many mesoionic carbenes may not be able to be isolated as a free carbene; however, these MICs can be generated as a ligand for transition metal complexes. Numerous mesoionic carbene transition metal complexes are known with metals including Fe, Os, Rh, Ir, Ni, Pd, Pt, Cu, and Ag. Metal complexes with Sm and Y are also known. MIC complexes are formed by a variety of mechanisms. [4] [18] [19]

Mesoionic carbenes may be generated in situ with addition of a strong base to their salt forms. The carbenes immediately form complexes with metals present in the reaction mixture through ligand exchange. [9] [17]

Direct metalation through C-H bond activation [1] [4] [11] [18] [19] [20] [21] [22] or C-H oxidative addition [4] [18] [23] is one method often utilized. Activation of a C‒H bond leads to oxidative addition of the carbene ligand to the metal center. Typically, direct metalation requires the blockage of sites that would lead to normal NHC complexes — phenyl and isopropyl groups are good blocking substituents, as discussed earlier. Smaller substituents may be cleaved. Direct metalation by silver(I) with imidazolium salts can cause cleavage at the C2 position if methyl is used as the blocking group. The result is formation of normal NHC carbenes. n-alkyl and benzyl groups may undergo the same fate as the methyl group. Steric bulk may also influence the formation of MIC complexes over NHC complexes. For imidazolium salts, the C2 position may not need to be blocked if the nitrogen substituents (N1 or N3) are sterically-demanding. Interactions between the nitrogen substituents and the metal center prevent normal NHC complexes from forming. If the carbene is part of a bidentate ligand with a forced geometry, the MIC complex may form preferentially as well. The counteranion of imidazolium salts participates in NHC vs. MIC formation. NHC formation typically occurs by heterolytic bond cleavage, so small, coordinating anions favor this pathway. MIC formation typically occurs by an oxidative addition pathway, so non-coordinating and apolar anions are preferred, such as BF4 or SbF6. [4] Other techniques focus on the activation of the desired carbon rather than blocking undesired carbons. A carbon may be activated by a halogen. A C-X bond (X = halide) is more favorable for activation than a C-H bond. This pathway results in the oxidative addition of the MIC carbene halide to a low valent metal center. [4] [18]

Transmetalation is another method commonly utilized. [4] [11] [18] [19] [22] [24] [25] Typically, a silver carbene complex is produced by direct metalation. This silver complex is reacted via transmetalation with a salt of the desired metal. The metal MIC complex is produced and silver salts generally precipitate.

Applications in catalysis

Since mesoionic carbene ligands are very strong σ-donors and make it easier for a metal center to undergo oxidative addition, MIC ligands have the potential to be useful in catalysis. [18] MIC transition metal complexes have been tested as catalysts in olefin metathesis, ring closure metathesis, and ring opening polymerization metathesis. [26] [27] The MIC complexes work very well, and in many cases, they outperform their NHC counterparts. MIC complexes have been successful as catalysts for Suzuki-Miyaura and Heck-Mizoroki cross-coupling reactions. [4] [9] [28] [29] [30] Again, in many cases, MIC catalysts are superior to their NHC counterparts. For example, in olefin metathesis, MIC catalysts are active at room temperature after simply addition of a Brønsted acid, such as hydrochloric acid or trifluoroacetic acid, compared to the large amount of thermal activation required for NHC catalysts. [27] MIC complexes have found use as catalysts in olefin hydrogenation. They have been shown to hydrogenate terminal and cis-alkenes. [4] [5] They work better than their NHC counterparts due to the MIC ligand’s stronger electron-donating properties. They are better able to provide electron density to promote hydrogen gas oxidative addition to the metal. MIC complexes have been used in transfer hydrogenation reactions. For example, they have been used to hydrogenate a diaryl ketone using isopropanol as a hydrogen source., [4] [21] MIC complexes are being considered as green chemistry catalysts. They act as catalysts for base- and oxidant-free oxidation of alcohols and amines. Some complexes have also been shown to synthesize certain aryl amides. [31] Other MIC complexes have been used in hydroarylation, involving the addition of an electron-rich aryl group and a hydrogen across a multiple bond. [32] The reactions that mesoionic carbene complexes catalyze will continue to expand as more research is done. [18] [33]

Related Research Articles

In chemistry, a carbene is a molecule containing a neutral carbon atom with a valence of two and two unshared valence electrons. The general formula is R-(C:)-R' or R=C: where the R represent substituents or hydrogen atoms.

The Sonogashira reaction is a cross-coupling reaction used in organic synthesis to form carbon–carbon bonds. It employs a palladium catalyst as well as copper co-catalyst to form a carbon–carbon bond between a terminal alkyne and an aryl or vinyl halide.

Grubbs catalyst

Grubbs catalysts are a series of transition metal carbene complexes used as catalysts for olefin metathesis. They are named after Robert H. Grubbs, the chemist who supervised their synthesis. Several generations of the catalyst have been developed. Grubbs catalysts tolerate many functional groups in the alkene substrates, are air-tolerant, and are compatible with a wide range of solvents. For these reasons, Grubbs catalysts have become popular in synthetic organic chemistry. Grubbs, together with Richard R. Schrock and Yves Chauvin, won the Nobel Prize in Chemistry in recognition of their contributions to the development of olefin metathesis.

Organoboron chemistry

Organoborane or organoboron compounds are chemical compounds of boron and carbon that are organic derivatives of BH3, for example trialkyl boranes. Organoboron chemistry or organoborane chemistry is the chemistry of these compounds.

A transition metal carbene complex is an organometallic compound featuring a divalent organic ligand. The divalent organic ligand coordinated to the metal center is called a carbene. Carbene complexes for almost all transition metals have been reported. Many methods for synthesizing them and reactions utilizing them have been reported. The term carbene ligand is a formalism since many are not derived from carbenes and almost none exhibit the reactivity characteristic of carbenes. Described often as M=CR2, they represent a class of organic ligands intermediate between alkyls (−CR3) and carbynes (≡CR). They feature in some catalytic reactions, especially alkene metathesis, and are of value in the preparation of some fine chemicals.

Persistent carbene

A persistent carbene (also known as stable carbene) is a type of carbene demonstrating particular stability. The best-known examples and by far largest subgroup are the N-heterocyclic carbenes (NHC) (sometimes called Arduengo carbenes), for example diaminocarbenes with the general formula (R2N)2C:, where the 'R's are typically alkyl and aryl groups. The groups can be linked to give heterocyclic carbenes, such as those derived from imidazole, imidazoline, thiazole or triazole.

A migratory insertion is a type of reaction in organometallic chemistry wherein two ligands on a metal complex combine. It is a subset of reactions that very closely resembles the insertion reactions, and both are differentiated by the mechanism that leads to the resulting stereochemistry of the products. However, often the two are used interchangeably because the mechanism is sometimes unknown. Therefore, migratory insertion reactions or insertion reactions, for short, are defined not by the mechanism but by the overall regiochemistry wherein one chemical entity interposes itself into an existing bond of typically a second chemical entity e.g.:

Organoruthenium chemistry

Organoruthenium chemistry is the chemistry of organometallic compounds containing a carbon to ruthenium chemical bond. Several organoruthenium catalysts are of commercial interest and organoruthenium compounds have been considered for cancer therapy. The chemistry has some stoichiometric similarities with organoiron chemistry, as iron is directly above ruthenium in group 8 of the periodic table. The most important reagents for the introduction of ruthenium are ruthenium(III) chloride and triruthenium dodecacarbonyl.

Guy Bertrand, born on July 17, 1952 at Limoges is a chemistry professor at the University of California, San Diego.

An insertion reaction is a chemical reaction where one chemical entity interposes itself into an existing bond of typically a second chemical entity e.g.:

PEPPSI

PEPPSI is an abbreviation for pyridine-enhanced precatalyst preparation stabilization and initiation. It refers to a group of palladium catalysts developed around 2005 by Prof. Michael G. Organ and co-workers at York University, which can accelerate various aminations and cross-coupling reactions. In comparison to many alternative palladium catalysts, PEPPSI-type complexes are stable to air and moisture and are relatively easy to synthesize and handle.

Tolman electronic parameter

The Tolman electronic parameter (TEP) is a measure of the electron donating or withdrawing ability of a ligand. It is determined by measuring the frequency of the A1 C-O vibrational mode (ν(CO)) of a (pseudo)-C3v symmetric complex, [LNi(CO)3] by infrared spectroscopy, where L is the ligand of interest. [LNi(CO)3] was chosen as the model compound because such complexes are readily prepared from tetracarbonylnickel(0). The shift in ν(CO) is used to infer the electronic properties of a ligand, which can aid in understanding its behavior in other complexes. The analysis was introduced by Chadwick A. Tolman.

Diiminopyridines are a class of diimine ligands. They featuring a pyridine nucleus with imine sidearms appended to the 2,6–positions. The three nitrogen centres bind metals in a tridentate fashion, forming pincer complexes. Diiminopyridines are notable as non-innocent ligand that can assume more than one oxidation state. Complexes of DIPs participate in a range of chemical reactions, including ethylene polymerization, hydrosilylation, and hydrogenation.

Palladium–NHC complex

In organometallic chemistry, palladium-NHC complexes are a family of organopalladium compounds in which palladium forms a coordination complex with N-Heterocyclic carbenes (NHCs). They have been investigated for applications in homogeneous catalysis, particularly cross-coupling reactions.

Transition metal NHC complex

In coordination chemistry, a transition metal NHC complex is a metal complex containing one or more N-heterocyclic carbene ligands. Such compounds are the subject of much research, in part because of prospective applications in homogeneous catalysis. One such success is the second generation Grubbs catalyst.

Zhan catalyst

A Zhan catalyst is a type of ruthenium-based organometallic complex used in olefin metathesis. This class of chemicals is named after the chemist who first synthesized them, Zheng-Yun J. Zhan.

Janis Louie is a Chemistry professor and Henry Eyring Fellow at The University of Utah. Louie contributes to the chemistry world with her research in inorganic, organic, and polymer chemistry.

Cyclic alkyl amino carbenes

In chemistry, cyclic(alkyl)(amino)carbenes (CAACs) are a family of stable singlet carbene ligands developed by Prof. Guy Bertrand and his group in 2005 at UC Riverside. In marked contrast with the popular N-heterocyclic carbenes (NHC) which possess two "amino" substituents adjacent to the "carbene" center, CAACs possess one "amino" substituent and an sp3 carbon atom "alkyl". This specific configuration makes the CAACs very good σ-donors and π-acceptors when compared to NHCs. Moreover the reduced heteroatom stabilization of the carbene center in CAACs versus NHCs also gives rise to a smaller ΔEST.

Borylene

A borylene is the boron analogue of a carbene. The general structure is R-B: with R an organic residue and B a boron atom with two unshared electrons. Borylenes are of academic interest in organoboron chemistry. A singlet ground state is predominant with boron having two vacant sp2 orbitals and one doubly occupied one. With just one additional substituent the boron is more electron deficient than the carbon atom in a carbene. For this reason stable borylenes are more uncommon than stable carbenes. Some borylenes such as boron monofluoride (BF) and boron monohydride (BH) the parent compound also known simply as borylene, have been detected in microwave spectroscopy and may exist in stars. Other borylenes exist as reactive intermediates and can only be inferred by chemical trapping.

Transition metal isocyanide complexes

Transition metal isocyanide complexes are coordination compounds containing isocyanide ligands. Because isocyanide are relatively basic, but also good pi-acceptors, a wide range of complexes are known. Some isocyanide complexes are used in medical imaging.

References

  1. 1 2 3 4 G. Guisado-Barrios, J. Bouffard, B. Donnadieu, G. Bertrand. Angew. Chem., Int. Ed. 2010, 49, 4759-4762.
  2. 1 2 D. Martin, M. Melaimi, M. Soleilhavoup, G. Bertrand. Organometallics. 2011, 30, 5304-5313.
  3. 1 2 G. Ung, D. Mendoza-Espinosa, J. Bouffard, G. Bertrand. Angew. Chem., Int. Ed. 2011, 50, 4215-4218.
  4. 1 2 3 4 5 6 7 8 9 10 11 12 M. Albrecht. Chem. Commun. 2008, 3601-3610.
  5. 1 2 M. Heckenroth, E. Kluser, A. Neels, M. Albrecht. Angew. Chem., Int. Ed. 2007, 46, 6293-6296.
  6. M. Heckenroth, A. Neels, M. Garnier, P. Aebi, A. Ehlers, M. Albrecht. Chem. Eur. J. 2009, 15, 9375-9386.
  7. S. Gründemann, A. Kovacevic, M. Albrecht, J. Faller, R. H. Crabtree. Chem. Commun. 2001, 2274-2275.
  8. 1 2 3 G. Ung, G. Bertrand. Chem. Eur. J. 2011, 17, 8269-8272.
  9. 1 2 3 E. Aldeco-Perez, A. Rosenthal, B. Donnadieu, P. Parameswaran, G. Frenking, G. Bertrand. Science. 2009, 326, 556-559.
  10. D. Mendoza-Espinosa, B. Donnadieu, G. Bertrand. J. Am. Chem. Soc. 2010, 132, 7264-7265.
  11. 1 2 3 P. Mathew, A. Neels, M. Albrecht. J. Am. Chem. Soc. 2008, 130, 13534-13535.
  12. Y. Han, H. V. Huynh, G. K. Tan. Organometallics. 2007, 26, 6581-6585.
  13. 1 2 V. Lavallo, C. A. Dyker, B. Donnadieu, G. Bertrand. Angew. Chem., Int. Ed. 2008, 47, 5411-5414.
  14. 1 2 I. Fernández, C. A. Dyker, A. DeHope, B. Donnadieu, G. Frenking, G. Bertrand. J. Am. Chem. Soc. 2009, 131, 11875-11881.
  15. S. Araki, K. Yokoi, R. Sato, T. Hirashita, J. Setsune. J. Heterocyclic Chem. 2009, 46, 164-171.
  16. M. Iglesias, M. Albrecht. Dalton Trans. 2010, 39, 5213-5215.
  17. 1 2 D. Mendoza-Espinosa, G. Ung, B. Donnadieu, G. Bertrand. Chem. Commun. 2011, 47, 10614-10616.
  18. 1 2 3 4 5 6 7 8 9 Schuster, Oliver; Yang, Liangru; Raubenheimer, Helgard G.; Albrecht, Martin (2009-08-12). "Beyond Conventional N-Heterocyclic Carbenes: Abnormal, Remote, and Other Classes of NHC Ligands with Reduced Heteroatom Stabilization" (PDF). Chemical Reviews. 109 (8): 3445–3478. doi:10.1021/cr8005087. ISSN   0009-2665. PMID   19331408.
  19. 1 2 3 Arnold, Polly L.; Pearson, Stephen (2007-03-01). "Abnormal N-heterocyclic carbenes". Coordination Chemistry Reviews. Recent developments in the organometallic chemistry of N-heterocyclic carbenes. 251 (5): 596–609. doi:10.1016/j.ccr.2006.08.006. ISSN   0010-8545.
  20. G. Song. Y. Zhang. X. Li. Organometallics. 2008, 27, 1936-1943.
  21. 1 2 L. Yang, A. Krüger, A. Neels, M. Albrecht. Organometallics. 2008, 27, 3161-3171.
  22. 1 2 A. Poulain, D. Canseco-Gonzalez, R. Hynes-Roche, H. Müller-Bunz, O. Schuster, H. Stoeckli-Evans, A. Neels, M. Albrecht. Organometallics. 2011, 30, 1021-1029.
  23. D. Bacciu, K. J. Cavell, I. A. Fallis, L. Ooi. Angew. Chem., Int. Ed. 2005, 44, 5282-5284.
  24. M. Alcarazo, S. Roseblade, A. R. Cowley, R. Fernández, J. M. Brown, J. M. Lassaletta. J. Am. Chem. Soc. 2005, 127, 3290-3291.
  25. A. Chianese, A. Kovacevic, B. Zeglis, J. W. Faller, R. H. Crabtree. Organometallics. 2004, 23, 2461-2468.
  26. J. Bouffard, B. K. Keitz, R. Tonner, G. Guisado-Barrios, G. Frenking, R. H. Grubbs, G. Bertrand. Organometallics. 2011, 30, 2617-2627.
  27. 1 2 B. K. Keitz, J. Bouffard, G. Bertrand, R. H. Grubbs. J. Am. Chem. Soc. 2011, 133, 8498-8501.
  28. H. Lebel, M. K. Janes, A. B. Charette, S. P. Nolan. J. Am. Chem. Soc. 2004, 126, 5046-5047.
  29. Y. Han, L. J. Lee, H. V. Huynh. Organometallics. 2009, 28, 2778-2786.
  30. T. Nakamura, K. Ogata, S. Fukuzawa. Chem. Lett. 2010, 39, 920-922.
  31. A. Prades, E. Peris, M. Albrecht. Organometallics. 2011, 30, 1162-1167.
  32. R. Saravanakumar, V. Ramkumar, S. Sankararaman. Organometallics. 2011, 30, 1689-1694.
  33. A. Prades, M. Viciano, M. Sanaú, E. Peris. Organometallics. 2008, 27, 4254-4259.