Grubbs catalyst

Last updated

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 also been developed. [1] [2] Grubbs catalysts tolerate many functional groups in the alkene substrates, are air-tolerant, and are compatible with a wide range of solvents. [3] [4] For these reasons, Grubbs catalysts have become popular in synthetic organic chemistry. [5] 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.

Contents

First-generation Grubbs catalyst

First-generation Grubbs catalyst
Grubbs Catalyst 1st Generation.svg
Grubbs-1G-from-xtal-2010-3D-balls.png
Names
IUPAC name
Benzylidene-bis(tricyclohexylphosphino)-dichlororuthenium
Identifiers
3D model (JSmol)
ChemSpider
PubChem CID
UNII
  • InChI=1S/C43H72P2.2ClH.Ru/c1-8-22-36(23-9-1)43(44(37-24-10-2-11-25-37,38-26-12-3-13-27-38)39-28-14-4-15-29-39)45(40-30-16-5-17-31-40,41-32-18-6-19-33-41)42-34-20-7-21-35-42;;;/h1,8-9,22-23,37-43H,2-7,10-21,24-35H2;2*1H;/q+2;;;+2/p-2
    Key: NDDFAYQFCZRYDT-UHFFFAOYSA-L
  • Cl[Ru-2](Cl)([P+](C1CCCCC1)(C1CCCCC1)C1CCCCC1)([P+](C1CCCCC1)(C1CCCCC1)C1CCCCC1)=Cc1ccccc1
Properties
C43H72Cl2P2Ru
Molar mass 822.97 g·mol−1
AppearancePurple solid
Melting point 153 °C (307 °F; 426 K) (decomposition)
Except where otherwise noted, data are given for materials in their standard state (at 25 °C [77 °F], 100 kPa).
X mark.svgN  verify  (what is  Yes check.svgYX mark.svgN ?)

In the 1960s, ruthenium trichloride was found to catalyze olefin metathesis. Processes were commercialized based on these discoveries. These ill-defined but highly active homogeneous catalysts remain in industrial use. [6] The first well-defined ruthenium catalyst was reported in 1992. [7] It was prepared from RuCl2(PPh3)4 and diphenylcyclopropene.

First Grubbs-type catalyst MetathesisGrubbs1992.svg
First Grubbs-type catalyst

This initial ruthenium catalyst was followed in 1995 by what is now known as the first-generation Grubbs catalyst. It is synthesized from RuCl2(PPh3)3, phenyldiazomethane, and tricyclohexylphosphine in a one-pot synthesis. [8] [9]

Preparation of the first-generation Grubbs catalyst SynthesisofGrubbs1stGen.png
Preparation of the first-generation Grubbs catalyst

The first-generation Grubbs catalyst was the first well-defined Ru-based catalyst. It is also important as a precursor to all other Grubbs-type catalysts.

Second-generation Grubbs catalyst

Second-generation Grubbs catalyst
Grubbs catalyst Gen2.svg
Grubbs-2G-from-xtal-2005-3D-balls.png
Names
IUPAC name
[1,3-bis-(2,4,6-trimethylphenyl)-2-imidazolidinylidene]dichloro(phenylmethylene)(tricyclohexylphosphino)ruthenium
Identifiers
3D model (JSmol)
ChemSpider
PubChem CID
  • InChI=1S/C21H26N2.C18H33P.C7H6.2ClH.Ru/c1-14-9-16(3)20(17(4)10-14)22-7-8-23(13-22)21-18(5)11-15(2)12-19(21)6;1-4-10-16(11-5-1)19(17-12-6-2-7-13-17)18-14-8-3-9-15-18;1-7-5-3-2-4-6-7;;;/h9-12H,7-8H2,1-6H3;16-18H,1-15H2;1-6H;2*1H;/q;;;;;+2/p-2
    Key: FCDPQMAOJARMTG-UHFFFAOYSA-L
  • Cl[Ru-2](Cl)([c+]0n(-c1c(C)cc(C)cc1C)CCn0-c1c(C)cc(C)cc1C)([P+](C1CCCCC1)(C1CCCCC1)C1CCCCC1)=Cc1ccccc1
Properties
C46H65Cl2N2PRu
Molar mass 848.98 g·mol−1
AppearancePinkish brown solid
Melting point 143.5 to 148.5 °C (290.3 to 299.3 °F; 416.6 to 421.6 K)
Hazards
GHS labelling:
GHS-pictogram-flamme.svg
Warning
H228
P210, P240, P241, P280, P378
Except where otherwise noted, data are given for materials in their standard state (at 25 °C [77 °F], 100 kPa).

The second-generation catalyst has the same uses in organic synthesis as the first generation catalyst, but generally with higher activity. This catalyst is stable toward moisture and air, thus is easier to handle in laboratories.

Shortly before the discovery of the second-generation Grubbs catalyst, a very similar catalyst based on an unsaturated N-heterocyclic carbene (1,3-bis(2,4,6-trimethylphenyl)imidazole) was reported independently by Nolan [10] and Grubbs [11] in March 1999, and by Fürstner [12] in June of the same year. Shortly thereafter, in August 1999, Grubbs reported the second-generation catalyst, based on a saturated N-heterocyclic carbene (1,3-bis(2,4,6-trimethylphenyl)dihydroimidazole): [13]

Synthesis of the second-generation Grubbs catalyst SynthesisofGrubbs2ndGen.svg
Synthesis of the second–generation Grubbs catalyst

In both the saturated and unsaturated cases a phosphine ligand is replaced with an N-heterocyclic carbene (NHC), which is characteristic of all second-generation-type catalysts. [3]

Both the first- and second-generation catalysts are commercially available, along with many derivatives of the second-generation catalyst.

Hoveyda–Grubbs catalysts

First-generation Hoveyda–Grubbs catalyst
Hoveyda-katalysator.svg
Hoveyda-Grubbs-catalyst-1st-gen 3D-balls.png
Names
IUPAC name
Dichloro(o-isopropoxyphenylmethylene)(tricyclohexylphosphine)ruthenium(II)
Identifiers
3D model (JSmol)
ChemSpider
PubChem CID
  • InChI=1S/C18H33P.C10H12O.2ClH.Ru/c1-4-10-16(11-5-1)19(17-12-6-2-7-13-17)18-14-8-3-9-15-18;1-8(2)11-10-7-5-4-6-9(10)3;;;/h16-18H,1-15H2;3-8H,1-2H3;2*1H;/q;;;;+2/p-2
    Key: KMKCJXPECJFQPQ-UHFFFAOYSA-L
  • Cl[Ru-2]2(Cl)([P+](C1CCCCC1)(C1CCCCC1)C1CCCCC1)=Cc1ccccc1[O+]2C(C)C
Properties
C28H45Cl2OPRu
Molar mass 600.61 g·mol−1
AppearanceBrown solid
Melting point 195 to 197 °C (383 to 387 °F; 468 to 470 K)
Hazards
GHS labelling:
GHS-pictogram-flamme.svg
Warning
H228
P210, P240, P241, P280, P378
Except where otherwise noted, data are given for materials in their standard state (at 25 °C [77 °F], 100 kPa).
Second-generation Hoveyda–Grubbs catalyst
Wikipedia-HoveydaGrubbsCatalysts.png
Hoveyda-Grubbs-catalyst-from-xtal-2007-3D-balls.png
Names
IUPAC name
[1,3-Bis-(2,4,6-trimethylphenyl)-2-imidazolidinylidene]dichloro(o-isopropoxyphenylmethylene)ruthenium
Identifiers
3D model (JSmol)
ChemSpider
EC Number
  • 608-446-3
PubChem CID
  • InChI=1S/C21H26N2.C10H12O.2ClH.Ru/c1-14-9-16(3)20(17(4)10-14)22-7-8-23(13-22)21-18(5)11-15(2)12-19(21)6;1-8(2)11-10-7-5-4-6-9(10)3;;;/h9-12H,7-8H2,1-6H3;3-8H,1-2H3;2*1H;/q;;;;+2/p-2
    Key: ZRPFJAPZDXQHSM-UHFFFAOYSA-L
  • Cl[Ru-2]2(Cl)([c+]0n(-c1c(C)cc(C)cc1C)CCn0-c1c(C)cc(C)cc1C)=Cc1ccccc1[O+]2C(C)C
Properties
C31H38Cl2N2ORu
Molar mass 626.63 g·mol−1
AppearanceGreen solid
Melting point 216 to 220 °C (421 to 428 °F; 489 to 493 K)
Except where otherwise noted, data are given for materials in their standard state (at 25 °C [77 °F], 100 kPa).

In the Hoveyda–Grubbs catalysts, the benzylidene ligands have a chelating ortho-isopropoxy group attached to the benzene rings. The ortho-isopropoxybenzylidene moiety is sometimes referred to as a Hoveyda chelate. The chelating oxygen atom replaces a phosphine ligand, which in the case of the 2nd generation catalyst, gives a completely phosphine-free structure. The 1st generation Hoveyda–Grubbs catalyst was reported in 1999 by Amir H. Hoveyda's group, [14] and in the following year, the second-generation Hoveyda–Grubbs catalyst was described in nearly simultaneous publications by the Blechert [15] and Hoveyda [16] laboratories. Siegfried Blechert's name is not commonly included in the eponymous catalyst name. The Hoveyda–Grubbs catalysts, while more expensive and slower to initiate than the Grubbs catalyst from which they are derived, are popular because of their improved stability. [3] By changing the steric and electronic properties of the chelate, the initiation rate of the catalyst can be modulated, [17] [18] such as in the Zhan catalysts. Hoveyda–Grubbs catalysts are easily formed from the corresponding Grubbs catalyst by the addition of the chelating ligand and the use of a phosphine scavenger like copper(I) chloride: [16]

The second-generation Hoveyda–Grubbs catalysts can also be prepared from the 1st generation Hoveyda–Grubbs catalyst by the addition of the NHC: [15]

Preparation of the first-generation Hoveyda-Grubbs catalyst from the first-generation Grubbs catalyst SynthesisofGrubbsHoveyda2.png
Preparation of the first-generation Hoveyda–Grubbs catalyst from the first-generation Grubbs catalyst
Preparation of the second-generation Hoveyda-Grubbs catalyst from the second-generation Grubbs catalyst SynthesisofGrubbsHoveyda.png
Preparation of the second-generation Hoveyda–Grubbs catalyst from the second–generation Grubbs catalyst

In one study published by Grubbs and Hong in 2006, a water-soluble Grubbs catalyst was prepared by attaching a polyethylene glycol chain to the imidazolidine group. [19] This catalyst is used in the ring-closing metathesis reaction in water of a diene carrying an ammonium salt group making it water-soluble as well.

Ring closing metathesis reaction in water Grubbsreactioninwater.png
Ring closing metathesis reaction in water

Third-generation Grubbs catalyst (fast-initiating catalysts)

The rate of the Grubbs catalyst can be altered by replacing the phosphine ligand with more labile pyridine ligands. By using 3-bromopyridine the initiation rate is increased more than a millionfold. [20] Both pyridine and 3-bromopyridine are commonly used, with the bromo- version 4.8 times more labile resulting in even faster rates. [21] The catalyst is traditionally isolated as a two pyridine complex, however one pyridine is lost upon dissolving and reversibly inhibits the ruthenium center throughout any chemical reaction.

Third-generation Grubbs catalyst large.png

The principal application of the fast-initiating catalysts is as initiators for ring opening metathesis polymerisation (ROMP). Because of their usefulness in ROMP these catalysts are sometimes referred to as the 3rd generation Grubbs catalysts. [22] The high ratio of the rate of initiation to the rate of propagation makes these catalysts useful in living polymerization, yielding polymers with low polydispersity. [23]

Applications

Grubbs catalysts are of interest for olefin metathesis. It is mainly applied to fine chemical synthesis. Large-scale commercial applications of olefin metathesis almost always employ heterogeneous catalysts or ill-defined systems based on ruthenium trichloride. [6]

Related Research Articles

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

Olefin metathesis is an organic reaction that entails the redistribution of fragments of alkenes (olefins) by the scission and regeneration of carbon-carbon double bonds. Because of the relative simplicity of olefin metathesis, it often creates fewer undesired by-products and hazardous wastes than alternative organic reactions. For their elucidation of the reaction mechanism and their discovery of a variety of highly active catalysts, Yves Chauvin, Robert H. Grubbs, and Richard R. Schrock were collectively awarded the 2005 Nobel Prize in Chemistry.

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.

<span class="mw-page-title-main">Robert H. Grubbs</span> American chemist and Nobel Laureate (1942–2021)

Robert Howard GrubbsForMemRS was an American chemist and the Victor and Elizabeth Atkins Professor of Chemistry at the California Institute of Technology in Pasadena, California. He was a co-recipient of the 2005 Nobel Prize in Chemistry for his work on olefin metathesis.

<span class="mw-page-title-main">Dicarbonyltris(triphenylphosphine)ruthenium(0)</span> Chemical compound

Dicarbonyltris(triphenylphosphine)ruthenium(0) or Roper's complex is a ruthenium metal carbonyl. In it, two carbon monoxide ligands and three triphenylphosphine ligands are coordinated to a central ruthenium(0) center.

Ring-closing metathesis (RCM) is a widely used variation of olefin metathesis in organic chemistry for the synthesis of various unsaturated rings via the intramolecular metathesis of two terminal alkenes, which forms the cycloalkene as the E- or Z- isomers and volatile ethylene.

<span class="mw-page-title-main">Persistent carbene</span> Type of carbene demonstrating particular stability

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 four R moieties 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.

Ring-opening metathesis polymerization (ROMP) is a type of olefin metathesis chain-growth polymerization. The driving force of the reaction is relief of ring strain in cyclic olefins. A variety of heterogeneous and homogeneous catalysts have been developed. Most large-scale commercial processes rely on the former while some fine chemical syntheses rely on the homogeneous catalysts. Catalysts are based on transition metals such as W, Mo, Re, Ru, and Ti.

<span class="mw-page-title-main">Enyne metathesis</span> Organic reaction

An enyne metathesis is an organic reaction taking place between an alkyne and an alkene with a metal carbene catalyst forming a butadiene. This reaction is a variation of olefin metathesis.

<span class="mw-page-title-main">Amir H. Hoveyda</span>

Amir H. Hoveyda is an American organic chemist and professor of chemistry at Boston College, and held the position of department chair until 2018. In 2019, he embarked as researcher at the Institute of Science and Supramolecular Engineering at University of Strasbourg.

<span class="mw-page-title-main">2,4,6-Trimethylaniline</span> Chemical compound

2,4,6-Trimethylaniline is an organic compound with formula (CH3)3C6H2NH2. It is an aromatic amine that is of commercial interest as a precursor to dyes. It is prepared by selective mononitration of mesitylene, avoiding oxidation of the methyl groups. The resulting nitro compound is reduced to the aniline.

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

In organometallic chemistry, a metallacycle is a derivative of a carbocyclic compound wherein a metal has replaced at least one carbon center; this is to some extent similar to heterocycles. Metallacycles appear frequently as reactive intermediates in catalysis, e.g. olefin metathesis and alkyne trimerization. In organic synthesis, directed ortho metalation is widely used for the functionalization of arene rings via C-H activation. One main effect that metallic atom substitution on a cyclic carbon compound is distorting the geometry due to the large size of typical metals.

Asymmetric hydrogenation is a chemical reaction that adds two atoms of hydrogen to a target (substrate) molecule with three-dimensional spatial selectivity. Critically, this selectivity does not come from the target molecule itself, but from other reagents or catalysts present in the reaction. This allows spatial information to transfer from one molecule to the target, forming the product as a single enantiomer. The chiral information is most commonly contained in a catalyst and, in this case, the information in a single molecule of catalyst may be transferred to many substrate molecules, amplifying the amount of chiral information present. Similar processes occur in nature, where a chiral molecule like an enzyme can catalyse the introduction of a chiral centre to give a product as a single enantiomer, such as amino acids, that a cell needs to function. By imitating this process, chemists can generate many novel synthetic molecules that interact with biological systems in specific ways, leading to new pharmaceutical agents and agrochemicals. The importance of asymmetric hydrogenation in both academia and industry contributed to two of its pioneers — William Standish Knowles and Ryōji Noyori — being awarded one half of the 2001 Nobel Prize in Chemistry.

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

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.

<span class="mw-page-title-main">Dichlorotris(triphenylphosphine)ruthenium(II)</span> Chemical compound

Dichlorotris(triphenylphosphine)ruthenium(II) is a coordination complex of ruthenium. It is a chocolate brown solid that is soluble in organic solvents such as benzene. The compound is used as a precursor to other complexes including those used in homogeneous catalysis.

A metal carbido complex is a coordination complex that contains a carbon atom as a ligand. They are analogous to metal nitrido complexes. Carbido complexes are a molecular subclass of carbides, which are prevalent in organometallic and inorganic chemistry. Carbido complexes represent models for intermediates in Fischer–Tropsch synthesis, olefin metathesis, and related catalytic industrial processes. Ruthenium-based carbido complexes are by far the most synthesized and characterized to date. Although, complexes containing chromium, gold, iron, nickel, molybdenum, osmium, rhenium, and tungsten cores are also known. Mixed-metal carbides are also known.

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.

Functionalized polyolefins are olefin polymers with polar and nonpolar functionalities attached onto the polymer backbone. There has been an increased interest in functionalizing polyolefins due to their increased usage in everyday life. Polyolefins are virtually ubiquitous in everyday life, from consumer food packaging to biomedical applications; therefore, efforts must be made to study catalytic pathways towards the attachment of various functional groups onto polyolefins in order to affect the material's physical properties.

<span class="mw-page-title-main">Transition metal NHC complex</span>

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.

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

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.

<span class="mw-page-title-main">Cyclic alkyl amino carbenes</span> Family of chemical compounds

In chemistry, cyclic(alkyl)(amino)carbenes (CAACs) are a family of stable singlet carbene ligands developed by the research group of Guy Bertrand in 2005 at UC Riverside. In marked contrast with the popular N-heterocyclic carbenes (NHCs) 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.

References

  1. Grubbs, Robert H. (2003). Handbook of Metathesis (1st ed.). Weinheim: Wiley-VCH. ISBN   978-3-527-30616-9.
  2. Grubbs, R. H.; Trnka, T. M. (2004). "Ruthenium-Catalyzed Olefin Metathesis". In Murahashi, S. (ed.). Ruthenium‐Catalyzed Olefin Metathesis. Ruthenium in Organic Synthesis. Weinheim: Wiley-VCH. pp. 153–177. doi:10.1002/3527603832.ch6. ISBN   978-3-527-60383-1.
  3. 1 2 3 Vougioukalakis, G. C.; Grubbs, R. H. (2010). "Ruthenium-Based Heterocyclic Carbene-Coordinated Olefin Metathesis Catalysts". Chemical Reviews . 110 (3): 1746–1787. doi:10.1021/cr9002424. PMID   20000700.
  4. Trnka, T. M.; Grubbs, R. H. (2001). "The Development of L2X2Ru=CHR Olefin Metathesis Catalysts: An Organometallic Success Story". Accounts of Chemical Research . 34 (1): 18–29. doi:10.1021/ar000114f. PMID   11170353.
  5. Cossy, Janine; Arseniyadis, Stellios; Meyer, Christophe (2010). Metathesis in Natural Product Synthesis: Strategies, Substrates and Catalysts (1st ed.). Weinheim: Wiley-VCH. ISBN   978-3-527-32440-8.
  6. 1 2 Lionel Delaude; Alfred F. Noels (2005). "Metathesis". Kirk-Othmer Encyclopedia of Chemical Technology. Weinheim: Wiley-VCH. doi:10.1002/0471238961.metanoel.a01. ISBN   978-0-471-23896-6.
  7. Nguyen, S. T.; Johnson, L. K.; Grubbs, R. H.; Ziller, J. W. (1992). "Ring-opening metathesis polymerization (ROMP) of norbornene by a Group VIII carbene complex in protic media" (PDF). Journal of the American Chemical Society . 114 (10): 3974–3975. doi:10.1021/ja00036a053.
  8. Schwab, P.; France, M. B.; Ziller, J. W.; Grubbs, R. H. (1995). "A Series of Well-Defined Metathesis Catalysts – Synthesis of [RuCl2(=CHR′)(PR3)2] and Its Reactions". Angew. Chem. Int. Ed. 34 (18): 2039–2041. doi:10.1002/anie.199520391.
  9. Schwab, P.; Grubbs, R. H.; Ziller, J. W. (1996). "Synthesis and Applications of RuCl2(=CHR′)(PR3)2: The Influence of the Alkylidene Moiety on Metathesis Activity". J. Am. Chem. Soc. 118 (1): 100–110. doi:10.1021/ja952676d.
  10. Huang, J.-K.; Stevens, E. D.; Nolan, S. P.; Petersen, J. L. (1999). "Olefin Metathesis-Active Ruthenium Complexes Bearing a Nucleophilic Carbene Ligand". J. Am. Chem. Soc. 121 (12): 2674–2678. doi:10.1021/ja9831352.
  11. Scholl, M.; Trnka, T. M.; Morgan, J. P.; Grubbs, R. H. (1999). "Increased Ring Closing Metathesis Activity of Ruthenium-Based Olefin Metathesis Catalysts Coordinated with Imidazolin-2-ylidene Ligands". Tetrahedron Letters . 40 (12): 2247–2250. doi:10.1016/S0040-4039(99)00217-8.
  12. Ackermann, L.; Fürstner, A.; Weskamp, T.; Kohl, F. J.; Herrmann, W. A. (1999). "Ruthenium Carbene Complexes with Imidazolin-2-ylidene Ligands Allow the Formation of Tetrasubstituted Cycloalkenes by RCM". Tetrahedron Lett. 40 (26): 4787–4790. doi:10.1016/S0040-4039(99)00919-3.
  13. Scholl, M.; Ding, S.; Lee, C. W.; Grubbs, R. H. (1999). "Synthesis and Activity of a New Generation of Ruthenium-Based Olefin Metathesis Catalysts Coordinated with 1,3-Dimesityl-4,5-dihydroimidazol-2-ylidene Ligands". Org. Lett. 1 (6): 953–956. doi:10.1021/ol990909q. PMID   10823227.
  14. Kingsbury, Jason S.; Harrity, Joseph P. A.; Bonitatebus, Peter J.; Hoveyda, Amir H. (1999). "A Recyclable Ru-Based Metathesis Catalyst". Journal of the American Chemical Society . 121 (4): 791–799. doi:10.1021/ja983222u.
  15. 1 2 Gessler, S.; Randl, S.; Blechert, S. (2000). "Synthesis and metathesis reactions of phosphine-free dihydroimidazole carbene ruthenium complex". Tetrahedron Letters. 41 (51): 9973–9976. doi:10.1016/S0040-4039(00)01808-6.
  16. 1 2 Garber, S. B.; Kingsbury, J. S.; Gray, B. L.; Hoveyda, A. H. (2000). "Efficient and Recyclable Monomeric and Dendritic Ru-Based Metathesis Catalysts". Journal of the American Chemical Society. 122 (34): 8168–8179. doi:10.1021/ja001179g.
  17. Engle, Keary M.; Lu, Gang; Luo, Shao-Xiong; Henling, Lawrence M.; Takase, Michael K.; Liu, Peng; Houk, K. N.; Grubbs, Robert H. (2015). "Origins of Initiation Rate Differences in Ruthenium Olefin Metathesis Catalysts Containing Chelating Benzylidenes". Journal of the American Chemical Society. 137 (17): 5782–5792. doi:10.1021/jacs.5b01144. PMID   25897653.
  18. Luo, Shao-Xiong; Engle, Keary M.; Deng, Xiaofei; Hejl, Andrew; Takase, Michael K.; Henling, Lawrence M.; Liu, Peng; Houk, K. N.; Grubbs, Robert H. (2018). "An Initiation Kinetics Prediction Model Enables Rational Design of Ruthenium Olefin Metathesis Catalysts Bearing Modified Chelating Benzylidenes". ACS Catalysis . 8 (5): 4600–4611. doi:10.1021/acscatal.8b00843. PMC   7289044 . PMID   32528741.
  19. Grubbs, Robert H.; Hong, Soon Hyeok (2006). "Highly Active Water-Soluble Olefin Metathesis Catalyst" (PDF). Journal of the American Chemical Society. 128 (11): 3508–3509. doi:10.1021/ja058451c. PMID   16536510.
  20. Love, J. A.; Morgan, J. P.; Trnka, T. M.; Grubbs, R. H. (2002). "A Practical and Highly Active Ruthenium-Based Catalyst that Effects the Cross Metathesis of Acrylonitrile". Angew. Chem. Int. Ed. Engl. 41 (21): 4035–4037. doi:10.1002/1521-3773(20021104)41:21<4035::AID-ANIE4035>3.0.CO;2-I. PMID   12412073.
  21. Walsh, Dylan J.; Lau, Sii Hong; Hyatt, Michael G.; Guironnet, Damien (2017-09-25). "Kinetic Study of Living Ring-Opening Metathesis Polymerization with Third-Generation Grubbs Catalysts". Journal of the American Chemical Society. 139 (39): 13644–13647. doi:10.1021/jacs.7b08010. ISSN   0002-7863. PMID   28944665.
  22. Leitgeb, Anita; Wappel, Julia; Slugovc, Christian (2010). "The ROMP toolbox upgraded". Polymer. 51 (14): 2927–2946. doi: 10.1016/j.polymer.2010.05.002 .
  23. Choi, T.-L.; Grubbs, R. H. (2003). "Controlled Living Ring-Opening-Metathesis Polymerization by a Fast-Initiating Ruthenium Catalyst". Angewandte Chemie International Edition. 42 (15): 1743–1746. doi:10.1002/anie.200250632. PMID   12707895.