Late transition metal-mediated C–H bond activation has become a popular method to generate metal-carbon σ bonds in metallacycle synthesis. These cyclometalation reactions are usually facilitated by a heteroatom (X)-based functional group nearby the target C–H bond.[1] Such directed C–H bond activation strategy has been widely used to generate cyclometalated late transition catalysts and catalyst precursors in various homogeneous catalytic processes.[2] Among reported metallacycles as catalyst precursors, a dominant majority feature one 5- or 6-member chelating ring with a η2-[C,X] ligand as both anionic carbon-donor and neutral heteroatom-donor. By contrast, bis-cyclometalated late transition metal complexes with two η2-[C,X] ligands [3] are rarely exploited as organometallic catalysts.[4-6] We herein report the synthesis, structural characterization, and catalytic applications of several bis-cyclometalated ruthenium(II) complexes with η2-[C,N] and η2-[C,O] ligands. The potential of these ruthenacycles as catalyst precursors is demonstrated by a room-temperature catalytic alkene-alkyne coupling to synthesize α,β,γ,δ-unsaturated esters and amides.
We have recently reported a bis-cyclometalated octahedral Ru(II) complex, {Ru(η4-cod)[η2-HN=C(C6H5)C6H4]2} (1a), as a catalyst precursor in Ru(II)-catalyzed [3+2] carbocyclization between N–H ketimines and alkynes using a N-heterocyclic carbene ligand IPr (Scheme 1).[7] The proposed mechanism involves carbon-carbon bond formation by alkyne insertion into the Ru–C σ bond of a ruthenacycle intermediate (A→B), presumably facilitated by the IPr ligand that has replaced cod ligand on Ru center. Thus, the substrate-derived η2-[C,N] imine ligands appear to play the dual role of actor ligand and spectator ligand, eventually incorporated into the cyclization product and replaced by incoming ketimine substrates via cyclometalation. It is noteworthy that 1a did not react with alkyne substrates without added IPr ligand, suggesting significant ligand effect on its stability and reactivity. We envision that bis-cyclometalated 1a and its structural analogs can be further explored as Ru(II) catalyst precursors with η2-[C,X] ligands solely as spectator ligands that occupy four of the six coordination sites and affect reactions occurring at the other two cis-coordination sites (Scheme 2, C). In particular, ancillary ligands (L) can be replaced by alkene/alkyne substrates via π-complexation (C→D), setting the stage for C–C bond formation by oxidative cyclization (D→E).[8] The latter transformation is a key step in a number of Ru-catalyzed C–C coupling reactions such as alkene-alkyne (enyne) couplings for diene synthesis,[9] [2+2] or [2+2+2] cycloadditions, [10] and [2+2+1] cycloadditions such as the Pauson-Khand reaction.[11] With easily accessible η2-[C,X] ligands via C–H activation, bis-cyclometalated Ru(II) complexes may serve as an attractive alternative to existing Ru catalysts,[12] allowing modular catalyst design and tunable ligand effects on catalyst efficiency and selectivity.
Scheme 1.

Proposed C-C bond formation by alkyne insertion into Ru-C bond of ruthenacycle intermediates for Ru(II)-catalyzed [3+2] imine/alkyne carbocyclization.[7]
Scheme 2.

Envisioned C-C bond formation by oxidative cyclization with bis-cyclometalated Ru(II) complexes having two η2-[C,X] ligands
As reported previously,[7] complex 1a was synthesized by room-temperature cyclometalation of benzophenone imine with a commercially available Ru(II) π-allyl complex [(cod)Ru(η3-methallyl)2] in a 2:1 ratio (Scheme 3). A bis-cyclometalated ketone analog {Ru(η4-cod)[η2-OC(C6H5)C6H4]2} (1b) was synthesized using benzophenone in a similar fashion, albeit with lower reactivity and requiring heating at 80 °C for complete conversion. The chelating cod ligand in both 1a and 1b could be replaced by two pyridine ligands via heating in neat pyridine at 60 °C to form bis(pyridine)-ligated 2a and 2b respectively. The solid-state structures of complexes 1a,[7] 1b, 2a and 2b were determined by single crystal X-ray diffraction (see Supporting Information for details). All four of these bis-cyclometalated Ru(II) complexes displayed a near-octahedral Ru(II) center with two cis η2-[C,X] ligands, where the two N or O atoms were trans to each other and the two Ru–C σ bonds were cis to each other.[7][13]
Scheme 3.

Synthesis of bis-cyclometalated Ru(II) complexes with η2-[C,N] benzophenone imine and η2-[C,O] benzophenone ligands.
The catalytic activity of bis-cyclometalated Ru(II) complexes was evaluated by intermolecular alkene-alkyne coupling between diphenylacetylene (3a) and methyl acrylate (4a) to form a (2E,4Z)-1,3-diene product 5a (Table 1).[9][14-16] Using 5 mol% [(cod)Ru(η3-methallyl)2] (6) as catalyst precursor and no added ligands led to only 12% conversion after heating at 80 °C for 24 hours in toluene (entry 1). By contrast, in situ generated ruthenacycle 1a by pre-activation of 6 with 2 equivalents of benzophenone imine effectively promoted formation of 5a in quantitative yield (entry 2). Compared to benzophenone imine ligand, much lower reactivity was observed when catalyst pre-activation was carried out using other aromatic compounds that are capable of generating bis-cyclometalated Ru(II) complexes.[13] For example, using 2-phenylpyridine and benzophenone ligands both result in lower than 10% conversion (entries 3,4; see Supporting Information for complete results).
Table 1.
Development of the catalytic reactions.[a]
| ||||||
|---|---|---|---|---|---|---|
| Entry | Ru Catalyst | Ligand | Solvent | Method[b] | Temp. | Yield (%)[c] |
| 1 | [Ru(cod)(C4H7)2] (6) | none | toluene | A | 80 °C | 12 |
| 2 | 6 | Ph2C=NH | toluene | B | 80 °C | >98 |
| 3 | 6 | 2-phenylpyridine | toluene | B | 80 °C | 8 |
| 4 | 6 | Ph2C=O | toluene | B | 80 °C | <5 |
| 5 | 6 | Ph2C=NH | toluene | B | r.t. | >98 |
| 6 | 6 | Ph2C=NH | toluene | A | r.t. | 15 |
| 7 | 1a | none | toluene | A | r.t. | >98 |
| 8 | 1a | none | THF | A | r.t. | 96 |
| 9 | 1a | none | DME | A | r.t. | 66 |
| 10 | 1a | none | DMF | A | r.t. | 93 |
| 11 | 1a | none | hexane | A | r.t. | 82 |
| 12 | 1b | none | toluene | A | r.t. | <2 |
| 13 | 2a | none | toluene | A | r.t. | >98 |
| 14 | 2b | none | toluene | A | r.t. | <2 |
Conditions: 3a (0.20 mmol, 1 equiv), 4a (2.0 equiv), Ru catalyst (0.05 equiv), ligand (0.10 equiv), solvent (0.5 mL), room temperature (20-22 °C) or 80 °C, 24 h.
Mixing methods: (A) All components were mixed without pre-activation; (B) Ru precursor 6 and the ligand were added to toluene and stirred at 80 °C for 30 min; the mixture was then cooled down to room temperature and added with substrates 3a and 4a.
GC yields with n-dodecane as internal standard.
Gratifyingly, the in situ formed 1a was sufficiently active at room temperature, promoting quantitative formation of 5a after 24 hours with 5 mol% catalyst loading (entry 5). To our best knowledge, this is the first example of room-temperature catalytic acrylate-alkyne coupling to form α,β,γ,δ-unsaturated esters and complements other catalyst systems for mild enyne couplings.[15][16] Notably, skipping the 80 °C pre-activation led to much lower catalyst reactivity (entry 6). Thus, isolated 1a was used as catalyst precursor to further evaluate the solvent effect, with highest reactivity observed in toluene and THF solvents (entries 7-11). Under optimized conditions, room-temperature coupling between 3a (1.0 equiv) and 4a (2.0 equiv)[17] proceeded smoothly in toluene solvent with 5.0 mol% 1a, giving 5a in quantitative yield by GC analysis (entry 7). The pyridine-ligated bis(imine) complex 2a was less stable than 1a in solution phase but displayed comparable catalytic activity (entry 13). By contrast, the bis(ketone) analogs 1b and 2b were virtually unreactive as catalyst precursors (entries 12 and 14). Such reactivity distinction is consistent with the mechanistic hypothesis that cod or pyridine ligands will be replaced by alkene/alkyne substrates (Scheme 2, C→D), thus having little effects on catalytic activity beyond the initial stage of catalyst pre-activation. On the other hand, the η2-[C,X] imine or ketone ligands are expected to stay on the Ru center throughout catalytic cycles and play a dominant role on catalyst activity.
With the standard reaction conditions established, various internal alkynes and acrylic esters or amides were studied for Ru-catalyzed room-temperature alkene-alkyne coupling (Table 2). Coupling between 3a and unsubstituted alkyl acrylates proceeded smoothly to form 1,3-diene products 5a-d and 5f in over 90% yields and with exclusive stereoselectivity for the (2E,4Z)-isomers. For phenyl acrylate coupling product 5e, the yield was improved from 58% to 87% by replacing 1a with bis(pyridine) ruthenacycle 2a as the catalyst precursor. Such reactivity enhancement is likely due to facile catalyst activation by substrate replacement of more labile pyridine ligands compared to the chelating diene ligand. When coupling between 3a and 4a was scaled up from 0.2 mmol to 20 mmol, the loading of 1a could be reduced to 1.0 mol% to acquire 5a in 90% isolated yield (4.8 gram purified product) over 48 hours. Coupling between 3a and N,N-dimethyl acrylamide gave product 5g in 72% yield, while heating was needed to improve the yield of N,N-diethyl product 5h to 85% at 60 °C. Compared to less reactive N,N-dialkylacrylamides, N-isopropyl- and N-t-butylacrylamide reacted with 3a in good reactivity to form products 5i and 5j, although the latter required 2a as catalyst precursor for satisfactory yield. The scope of alkyne substrates was studied by coupling reactions with methyl acrylate (4a) to give products 5k-v. High reactivity and regioselectivity was observed for phenylacetylene derivatives with alkyl substituents (products 5k-s), favoring the formation of 4-alkyl-5-aryl regioisomer in >10:1 selectivity. The mild reaction conditions allow good compatibility with functional groups such as acyl, formyl, and Br substituents (products 5p, 5r and 5s), providing synthetic handles for further functional group transformations. Aliphatic internal alkynes such as 3-hexyne and 4-octyne displayed lower reactivity than aromatic alkynes, and a 2:1 alkyne/acrylate stoichiometry was used to get coupling products 5t and 5u in moderate yields. Coupling between 4a and terminal alkynes generally suffered from low reactivity and gave a complex mixture of products.[18] Nevertheless, coupling between 4a and phenylacetylene was effectively catalyzed by 2a to form product 5v with (E,E)-stereoselectivity in 65% isolated yield.[14b]
Table 2.
Scope of the catalytic alkene-alkyne coupling.[a]
|
Conditions: 3 (0.20 mmol, 1.0 equiv), 4 (2.0 equiv), 1a (0.050 equiv), toluene (0.5 mL), 20-22 °C, 24h; averaged isolated yields of two runs.
Isolated yield under scale-up conditions: 3 (20 mmol, 1.0 equiv), 4 (2.0 equiv), 1a (0.010 equiv), toluene (6.0 mL), 22 °C, 48h.
Using 2a as catalyst precursor.
Reactions at 60 °C
Combined yield of two regioisomers (ratios determined by NMR analysis); structure of major isomer is shown.
Using 0.20 mmol methyl acrylate as limiting reagent and 2.0 equiv of alkyne.
Isolated yield for the major stereoisomer in 5:1 selectivity; minor isomer was not purified.
Three types of reaction mechanism have been proposed for Ru-catalyzed alkene-alkyne couplings to form 1,3-dienes:[14][19] (1) C–C bond formation by alkene-alkyne oxidative cyclization (Scheme 2), followed by β-H elimination and C–H reductive elimination (Path 1); (2) alkyne insertion into a Ru hydride, followed by alkene insertion into the resulting Ru-alkenyl linkage and subsequent β-H elimination (Path 2); (3) sp2 C–H bond activation of alkene, followed by alkyne insertion into Ru alkenyl, and C–H bond formation by either reductive elimination or protonation of Ru–C bond (Path 3). Although the latter two pathways cannot be completely ruled out, the oxidative cyclization mechanism (Path 1) is most consistent with the observed regio- and stereochemistry in coupling products. In particular, high regioselectivity with non-symmetric alkyne substrates (products 5k-s) supports C–C bond formation by oxidative cyclization (Path 1) or alkyne insertion into Ru-alkenyl linkage (Path 3), not by alkyne insertion into Ru–H linkage (Path 2).[19] The complete lack of (2Z)-stereoisomers as coupling products also argues against the proposed alkene C–H activation stereochemistry in Path 3, which should favor (2Z)-isomers by ester- or amide-directed C–H activation/cyclometalation.
The proposed oxidative cyclization pathway has prompted us to extend our study to other mechanistically related C–C couplings using current catalyst system. Thus, 1a was found to catalyze room-temperature dimerization of methyl acrylate with high efficiency and exclusive tail-to-tail regioselectivity (Eq. 1).[17] In addition, a [2+2] norbornene/alkyne cycloaddition was effectively catalyzed by 1a at 120 °C (Eq. 2), which further supports the proposed Ru(II)/Ru(IV) catalytic cycle involving oxidative cyclization.[19][20]
In summary, we have developed a new class of bis-cyclometalated Ru(II) catalyst precursors with readily available η2-[C,X] ligands based on aromatic N–H ketimines and ketones. The catalytic activity of bis(imine) complex 1a was demonstrated with several catalytic C– C coupling reactions by proposed Ru(II)/Ru(IV) catalytic cycles involving oxidative cyclization. A room-temperature alkene-alkyne coupling was promoted to form α,β,γ,δ-unsaturated esters and amides with high regio- and stereoselectivities, good functional group tolerance, and very high catalyst efficiency in a representative gram-scale synthesis. The major limitation of current catalyst system is the limited scope of alkene substrates,[21][22] which we aim to improve by a more systematic future study on structure-reactivity correlations in bis-cyclometalated Ru(II) complexes with various η2-[C,X] ligands.
![]() |
(1) |
![]() |
(2) |
Supplementary Material
Acknowledgments
**Financial support for this work was provided by NSF (CHE-1301409 to PZ and CHE-1300912 to YZ) and NIH (Grant Number 2P20 RR015566) from the National Center for Research Resources.
Footnotes
Supporting information for this article is available on the WWW under http://www.angewandte.org or from the author.
Contributor Information
Ms. Jing Zhang, Department of Chemistry and Biochemistry, North Dakota State University, Fargo, ND 58108-6050, USA
Dr. Angel Ugrinov, Department of Chemistry and Biochemistry, North Dakota State University, Fargo, ND 58108-6050, USA
Prof. Dr. Yong Zhang, Email: yong.zhang@stevens.edu, Department of Chemistry, Chemical Biology, and Biomedical Engineering, Stevens Institute of Technology, Castle Point on Hudson Hoboken, NJ 07030, USA
Prof. Dr. Pinjing Zhao, Email: pinjing.zhao@ndsu.edu, Department of Chemistry and Biochemistry, North Dakota State University, Fargo, ND 58108-6050, USA.
References
- 1.Recent reviews on transition metal-mediated C–H bond activation with directing groups: Goldman AS, Goldberg KI, editors. Activation and Functionalization of C-H Bonds. Vol. 885. Washington DC: 2004. (ACS Symposium Series).Dyker G, editor. Handbook of C–H Transformations. Vol. 1. Wiley-VCH; 2005. Chen X, Engle KM, Wang DH, Yu JQ. Angew Chem. 2009;121:5196. doi: 10.1002/anie.200806273.Angew Chem Int Ed. 2009;48:5094.Colby DA, Bergman RG, Ellman JA. Chem Rev. 2010;110:624. doi: 10.1021/cr900005n.Lyons TW, Sanford MS. Chem Rev. 2010;110:1147. doi: 10.1021/cr900184e.Ackermann L. Chem Rev. 2011;111:1315. doi: 10.1021/cr100412j.Sun CL, Li BJ, Shi ZJ. Chem Rev. 2011;111:1293. doi: 10.1021/cr100198w.
- 2.For a recent review on cyclometalation with d-block transition metals, see: Albrecht M. Chem Rev. 2010;110:576. doi: 10.1021/cr900279a.
- 3.There is no official classification of bis-cyclometalated complexes. Here we limit our discussions to metallacycles with two chelating η2-[C,X] ligands that bind to four coordination sites. For metallacycles with pincer-type chelating ligands that bind to three adjacent planar coordination sites, see: Morales-Morales D, Jensen C, editors. The Chemistry of Pincer Compounds. Elsevier Science; Amsterdam: 2007.
- 4.Representative studies on bis-cyclometalated Ru(II) and Pd(IV) complexes for stoichiometric organometallic transformations: Basu A, Bhaduri S, Sharma K, Jones PG. J Chem Soc, Chem Commun. 1987:1126.Guari Y, Castellanos A, Sabo-Etienne S, Chaudret B. J Mol Catal A: Chem. 2004;212:77.Dick AR, Kampf JW, Sanford MS. J Am Chem Soc. 2005;127:12790. doi: 10.1021/ja0541940.Whitfield SR, Sanford MS. J Am Chem Soc. 2007;129:15142. doi: 10.1021/ja077866q.
- 5.For recent studies on “chiral-at-metal” bis-cyclometalated Ir(III) complexes for enantioselective organocatalysis and Lewis-acid catalysis, see the following paper and references therein: Huo H, Fu C, Harms K, Meggers E. J Am Chem Soc. 2014;136:2990. doi: 10.1021/ja4132505.
- 6.Review on bis-cyclometalated Ir(III) catalysts with two η2-[C,N] arylpyridine ligands for electron-transfer photoredox reactions: Narayanam JMR, Stephenson CRJ. Chem Soc Rev. 2011;40:102. doi: 10.1039/b913880n.
- 7.Zhang J, Ugrinov A, Zhao P. Angew Chem. 2013;125:6813. doi: 10.1002/anie.201209031. [DOI] [PMC free article] [PubMed] [Google Scholar]; Angew Chem Int Ed. 2013;52:6681. [Google Scholar]
- 8.Both the terms “oxidative cyclization” and “reductive coupling” have been used to describe such processes where the organic π systems are formally reduced and the metal is oxidized upon cyclization. See: Hartwig JF. Organotransition Metal Chemistry: From Bonding to Catalysis. University Science Books; 2010. Section 22.9.3.1.
- 9.Review on Ru-catalyzed C–C couplings by oxidative cyclization processes: Trost BM, Frederiksen MU, Rudd MT. Angew Chem. 2005;117:6788. doi: 10.1002/anie.200500136.Angew Chem Int Ed. 2005;44:6630.
- 10.(a) Derien S, Monnier F, Dixneuf PH. Top Organomet Chem. 2004;11:1. [Google Scholar]; (b) Vovard-Le Bray C, Derien S, Dixneuf PH. C R Chim. 2010;13:292. [Google Scholar]
- 11.a) Morimoto T, Chatani N, Fukumoto Y, Murai S. J Org Chem. 1997;62:3762. [Google Scholar]; b) Kondo T, Suzuki N, Okada T, Mitsudo T. J Am Chem Soc. 1997;119:6187. [Google Scholar]
- 12.Reported Ru catalyst systems for C–C coupling reactions by oxidative cyclization most commonly use the catalyst precursors of Ru3(CO)12, Ru(η4-cod)(η6-cot), and Ru(II) chlorides supported by η5-Cp/Cp* or η6-cymene ligands. See refs 9-11 for more details.
- 13.For recently reported structures of other bis-cyclometalated Ru(II) analogs, see: Flower KR, Howard VJ, Pritchard RG, Warren JE. Organometallics. 2002;21:1184.Ma J, Xu J, Su WP. Acta Crystallogr Sect E: Struct Rep Online. 2007;63:m1421.Scherl P, Wadepohl H, Gade LH. Organometallics. 2013;32:4409.
- 14.Ru-catalyzed alkene-alkyne coupling to form 1,3-dienes upon heating at 60-120 °C: Mitsudo Ta, Zhang SW, Nagao M, Watanabe Y. J Chem Soc, Chem Commun. 1991:598.Nishimura T, Washitake Y, Uemura S. Adv Synth Catal. 2007;349:2563.Neisius NM, Plietker B. Angew Chem. 2009;121:5863. doi: 10.1002/anie.200901928.Angew Chem Int Ed. 2009;48:5752.Miura H, Shimura S, Hosokawa S, Yamazoe S, Wada K, Inoue M. Adv Synth Catal. 2011;353:2837.
- 15.Relevant reports on catalytic room-temperature enyne coupling to form 1,3-dienes with different substrates: a) (Pd) N-t-butyl-acrylamide and diphenylacetylene: Lindhardt (neé Hansen) AT, Mantel MLH, Skrydstrup T. Angew Chem. 2008;120:2708.Angew Chem Int Ed. 2008;47:2668. b) (Ru) allyl acetates and alkynoates: Trost BM, Martos-Redrejo A. Org Lett. 2009;11:1071. doi: 10.1021/ol8028324. c) (Rh) pyrrolidinyl acryl amides and alkynes: Shibata Y, Otake Y, Hirano M, Tanaka K. Org Lett. 2009;11:689. doi: 10.1021/ol802767s. d) (Co) vinyl- or silylethylene and alkynes: Mannathan S, Cheng CH. Chem Commun. 2010;46:1923. doi: 10.1039/b920071a. e) (Ru) ethylene and ynamides Saito N, Saito K, Shiro M, Sato Y. Org Lett. 2011;13:2718. doi: 10.1021/ol200812y.
- 16.Representative examples of other catalysts for enyne coupling to form 1,3-dienes under thermal conditions: (a) (Rh) Shibata Y, Hirano M, Tanaka K. Org Lett. 2008;10:2829. doi: 10.1021/ol800995r. (b) (Ni) Horie H, Koyama I, Kurahashi T, Matsubara H. Chem Commun. 2011;47:2658. doi: 10.1039/c0cc04061d.
- 17.The use of excess 4a was necessary to compensate the formation of a minor by-product by dimerization of 4a. See following reports on Ru-catalyzed acrylate dimerization with tail-to-tail regioselectivity: Sustmann R, Hornung HJ, Schupp T, Patzke B. J Mol Catal. 1993;85:149.Hirano M, Sakate Y, Komine N, Komiya S, Bennett MA. Organometallics. 2009;28:4902.
- 18.Terminal alkynes may form non-productive Ru(II) alkylidene or dienyl complexes by alkyne oligomerization: Chen DF, Zhang CY, Xu SS, Song HB, Wang BQ. Organometallics. 2011;30:676.
- 19.See Scheme S1 in Supporting Information for a detailed description of the three possible mechanisms and discussion on regioselectivity.
- 20.Examples of early studies on Ru-catalyzed [2+2] cycloaddition with norbornene and internal alkynes: Butenschoen H. Angew Chem. 1994;106:664.Angew Chem Int Ed. 1994;33:1636.Mitsudo Ta, Naruse H, Kondo T, Ozaki Y, Watanabe Y. Angew Chem. 1994;106:595.Angew Chem Int Ed. 1994;33:580.
- 21.The current catalyst system did not work with α- or β-substituted acrylates or less electron-deficient alkenes such as vinylarenes.
- 22.Preliminary results from DFT calculations suggest involvement of hydrogen-bonding interactions between cyclometalated imine NH moieties and carbonyl groups from acrylate substrates, which could contribute to the reactivity dependence on alkene substrates and classes of η2-[C,X] ligands. Results of this computational study on proposed catalytic mechanisms will be reported separately.
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.


