Mechanism and Selectivity of Rhodium-Catalyzed 1:2 Coupling of Aldehydes and AllenesClick to copy article linkArticle link copied!
Abstract
The rhodium-catalyzed highly regioselective 1:2 coupling of aldehydes and allenes was investigated by means of density functional theory calculations. Full free energy profiles were calculated, and several possible reaction pathways were evaluated. It is shown that the energetically most plausible catalytic cycle is initiated by oxidative coupling of the two allenes, which was found to be the rate-determining step of the overall reaction. Importantly, Rh–allyl complexes that are able to adopt both η3 and η1 configurations were identified as key intermediates present throughout the catalytic cycle with profound implications for the selectivity of the reaction. The calculations reproduced and rationalized the experimentally observed selectivities and provided an explanation for the remarkable alteration in the product distribution when the catalyst precursor is changed from [RhCl(nbd)]2 (nbd = norbornadiene) to complexes containing noncoordinating counterions ([Rh(cod)2X]; X = OTf, BF4, PF6; cod = 1,5-cyclooctadiene). It turns out that the overall selectivity of the reaction is controlled by a combination of the inherent selectivities of several of the elementary steps and that both the mechanism and the nature of the selectivity-determining steps change when the catalyst is changed.
1 Introduction
2 Computational Details
3 Results and Discussion
3.1 Initial Step of the Catalytic Cycle: Oxidative Coupling versus Oxidative Addition
3.2 Formation of the Second C–C Bond
3.3 β-Hydride Elimination
3.4 Reductive Elimination
3.5 Overall Catalytic Cycle and Origins of the Regioselectivity in the [RhCl(dppe)]-Catalyzed Reaction
3.6 Regioselectivity in the [Rh(dppe)]+-Catalyzed Reaction
4 Conclusions
Supporting Information
Complete ref 13, additional results not shown in the text, and Cartesian coordinates of all optimized structures discussed in the paper. This material is available free of charge via the Internet at http://pubs.acs.org.
Terms & Conditions
Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.
Acknowledgment
We acknowledge financial support from the Swedish Research Council, the Göran Gustafsson Foundation, and the Knut and Alice Wallenberg Foundation. G.H. thanks the Carl Trygger Foundation for a postdoctoral fellowship. Computer time was generously provided by the Swedish National Infrastructure for Computing.
References
This article references 28 other publications.
- 1(a) Lautens, M.; Klute, W.; Tam, W. Chem. Rev. 1996, 96, 49– 92Google Scholar1ahttps://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK28XjsVKktw%253D%253D&md5=3776dd13a62c7dbef311a7cbf457dd7fTransition Metal-Mediated Cycloaddition ReactionsLautens, Mark; Klute, Wolfgang; Tam, WilliamChemical Reviews (Washington, D. C.) (1996), 96 (1), 49-92CODEN: CHREAY; ISSN:0009-2665. (American Chemical Society)A review with 226 refs. A discussion of established and interesting new cycloaddn. reactions promoted by metal species is presented.(b) Kotha, S.; Brahmachary, E.; Lahiri, K. Eur. J. Org. Chem. 2005, 4741– 4767Google Scholar1bhttps://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2MXht1KgurrM&md5=93220a4594f4dfa8e3cf02aac00ba43aTransition metal-catalyzed [2+2+2] cycloaddition and application in organic synthesisKotha, Sambasivarao; Brahmachary, Enugurthi; Lahiri, KakaliEuropean Journal of Organic Chemistry (2005), (22), 4741-4767CODEN: EJOCFK; ISSN:1434-193X. (Wiley-VCH Verlag GmbH & Co. KGaA)A review. The [2+2+2] cycloaddn. strategy is complementary to the well known Diels-Alder reaction for the generation of polycyclic compds. This [2+2+2] approach is atom-economical and, with the availability of new catalysts to effect the [2+2+2] cycloaddn. reaction, synthesis of a wide variety of highly functionalized polycycles is possible. This review deals with some recent advances relating to [2+2+2] cycloaddn. reactions involving the syntheses both of polycycles and of heterocycles. More specifically, syntheses of various biol. active mols., unusual amino acids, and theor. interesting mols. are described. An attempt has also been made to give an overview of recent advances in the achievement of chemo-, regio-, and stereoselectivity in cyclotrimerization reactions.(c) Murakami, M. Angew. Chem., Int. Ed. 2003, 42, 718– 720Google ScholarThere is no corresponding record for this reference.(d) Inglesby, P. A.; Evans, P. A. Chem. Soc. Rev. 2010, 39, 2791– 2805Google Scholar1dhttps://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3cXptFyhu7c%253D&md5=f8e29c1b1b3795588b1be0f7e9c11a97Stereoselective transition metal-catalysed higher-order carbocyclisation reactionsInglesby, Phillip A.; Evans, P. AndrewChemical Society Reviews (2010), 39 (8), 2791-2805CODEN: CSRVBR; ISSN:0306-0012. (Royal Society of Chemistry)A review. Transition metal-catalyzed higher-order carbocyclization reactions represent an important class of reactions due to their ability to construct complex polycyclic systems in a highly selective and atom-economical fashion. A key and striking feature with these transformations is the dichotomy in reactivity that a substrate displays with different transition metal complexes, which is akin to the manner enzymes direct terpene biosynthesis. This tutorial review details the historical development of higher-order carbocyclization reactions, specifically the variants of [m+2+2] that involve carbon-based π-systems, where m = 2, 3 and 4, in the context of crit. developments with various transition metal complexes.(e) Shibata, T.; Tsuchikama, K. Org. Biomol. Chem. 2008, 6, 1317– 1323Google ScholarThere is no corresponding record for this reference.(f) Tanaka, K. Chem.—Asian J. 2009, 4, 508– 518Google Scholar1fhttps://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD1MXks1Cht7c%253D&md5=941a1af8b94d4a5e67af5a6b0f552d64Transition-metal-catalyzed enantioselective [2+2+2] cycloadditions for the synthesis of axially chiral biarylsTanaka, KenChemistry - An Asian Journal (2009), 4 (4), 508-518CODEN: CAAJBI; ISSN:1861-4728. (Wiley-VCH Verlag GmbH & Co. KGaA)A review. Recent advances towards a development of novel transition-metal-catalyzed enantioselective [2+2+2] cycloaddn. for the synthesis of axially chiral biaryls are reviewed. A no. of efficient enantioselective biaryl syntheses were accomplished by chiral cobalt(I), iridium(I), and rhodium(I) complex catalyzed [2+2+2] cycloaddn. Furthermore, the enantioselective synthesis of axially chiral biaryls possessing non-biaryl axial chirality was also developed by using chiral rhodium(I) complexes as catalysts.(g) Yu, Z.-X.; Wang, Y.; Wang, Y. Chem.—Asian J. 2010, 5, 1072– 1088Google Scholar1ghttps://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3cXls1Wkur8%253D&md5=867a869887dd578c5621f6992c3b6319Transition-metal-catalyzed cycloadditions for the synthesis of eight-membered carbocyclesYu, Zhi-Xiang; Wang, Yi; Wang, YuanyuanChemistry - An Asian Journal (2010), 5 (5), 1072-1088CODEN: CAAJBI; ISSN:1861-4728. (Wiley-VCH Verlag GmbH & Co. KGaA)A review. Eight-membered carbocycles are in a wide variety of natural products that exhibit a broad range of biol. and medicinal activities (cf. the most potent anticancer drug, taxol). However, the synthesis of eight-membered carbocycles is quite challenging as traditional approaches are met with entropic and enthalpic penalties in the ring-forming transition states. These neg. effects can be totally or partially avoided with the implementation of transition-metal-catalyzed/mediated cycloaddns. In this focus review, examples of elegant and efficient metal-catalyzed and some metal-mediated cycloaddns. (including Ni- catalyzed [4+4] and Rh-catalyzed [4+2+2] and [5+2+1] reactions) are presented to illustrate this. Application of these cycloaddn. reactions in total synthesis was also presented to show the significance of these reactions in addressing challenges in natural product synthesis.
- 2(a) Willis, M. C. Chem. Rev. 2010, 110, 725– 748Google Scholar2ahttps://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD1MXhtleisL3N&md5=bd9108b8b2ab0b60865cdd64044736faTransition Metal Catalyzed Alkene and Alkyne HydroacylationWillis, Michael C.Chemical Reviews (Washington, DC, United States) (2010), 110 (2), 725-748CODEN: CHREAY; ISSN:0009-2665. (American Chemical Society)This review is a coordinated anal. of the intramol. or intermol. alkene hydroacylation using transition metal catalysts. The mechanism of such transformations is also presented in this systematic review.(b) Jun, C.-H.; Moon, C. W.; Lee, D.-Y. Chem.—Eur. J. 2002, 8, 2422– 2428Google Scholar2bhttps://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD38XksFWgs7s%253D&md5=9f3fa5e5d3340dfb3541509a40f2bcc3Chelation-assisted carbon-hydrogen and carbon-carbon bond activation by transition metal catalystsJun, Chul-Ho; Moon, Choong Woon; Lee, Dae-YonChemistry - A European Journal (2002), 8 (11), 2422-2428CODEN: CEUJED; ISSN:0947-6539. (Wiley-VCH Verlag GmbH)A review with refs. on the chelation-assisted C-H and C-C bond activation of carbonyl compds. by RhI catalysts. Hydroacylation of olefins was accomplished by utilizing 2-amino-3-picoline as a chelation auxiliary. The same strategy was employed for the C-C bond activation of unstrained ketones. Allylamine 24 was devised as a synthon of formaldehyde. Hydroiminoacylation of alkynes with allylamine 24 was applied to the alkyne cleavage by the aid of cyclohexylamine.(c) Leung, J. C.; Krische, M. J. Chem. Sci. 2012, 3, 2202– 2209Google Scholar2chttps://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC38XotFGku7g%253D&md5=551e827a85202d1ca97cfff97e5b5ec4Catalytic intermolecular hydroacylation of C-C π-bonds in the absence of chelation assistanceLeung, Joyce C.; Krische, Michael J.Chemical Science (2012), 3 (7), 2202-2209CODEN: CSHCCN; ISSN:2041-6520. (Royal Society of Chemistry)A review. Intermol. rhodium catalyzed hydroacylation in the absence of chelation assistance remains a largely unmet challenge due to competing decarbonylation of the acylrhodium intermediates to form catalytically inactive carbonyl complexes. Here, catalytic systems for intermol. hydroacylation in the absence of chelation assistance are reviewed, with an emphasis on recently described processes that operate through mechanistic pathways beyond aldehyde C-H oxidative addn.
- 3(a) Jang, H.-Y.; Krische, M. J. Acc. Chem. Res. 2004, 37, 653– 661Google ScholarThere is no corresponding record for this reference.(b) Ngai, M.-Y.; Kong, J.-R.; Krische, M. J. J. Org. Chem. 2007, 72, 1063– 1072Google ScholarThere is no corresponding record for this reference.(c) Iida, H.; Krische, M. J. Top. Curr. Chem. 2007, 279, 77– 104Google ScholarThere is no corresponding record for this reference.(d) Bower, J. F.; Kim, I. S.; Patman, R. L.; Krische, M. J. Angew. Chem., Int. Ed. 2009, 48, 34– 46Google ScholarThere is no corresponding record for this reference.(e) Skucas, E.; Ngai, M.-Y.; Komanduri, V.; Krische, M. J. Acc. Chem. Res. 2007, 40, 1394– 1401Google ScholarThere is no corresponding record for this reference.
- 4Kong, J. R.; Krische, M. J. J. Am. Chem. Soc. 2006, 128, 16040– 16041Google Scholar4https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD28Xht1GjsLvO&md5=4fed0d21910ffedc7657e0f145866142Catalytic Carbonyl Z-Dienylation via Multicomponent Reductive Coupling of Acetylene to Aldehydes and α-Ketoesters Mediated by Hydrogen: Carbonyl Insertion into Cationic RhodacyclopentadienesKong, Jong Rock; Krische, Michael J.Journal of the American Chemical Society (2006), 128 (50), 16040-16041CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)Exposure of aldehydes or α-ketoesters to equal vols. of acetylene and hydrogen gas at ambient temp. and pressure in the presence of cationic rhodium catalysts provides products of carbonyl Z-butadienylation, which arise via multicomponent coupling of four mols.: two mols. of acetylene, a mol. of vicinal dicarbonyl compd., and a mol. of elemental hydrogen. The collective data suggest a catalytic mechanism involving carbonyl insertion into a cationic rhodacyclopentadiene intermediate derived via oxidative dimerization of acetylene. Hydrogenolytic cleavage of the resulting oxarhodacycloheptadiene via formal σ-bond metathesis provides the product of carbonyl addn. and cationic rhodium(I) to close the catalytic cycle. Studies involving the hydrogenation of 1,6-diyne 14a (I) in the presence of α-ketoester 6a (II) corroborate the proposed catalytic mechanism. These multicomponent couplings represent the first use of acetylene gas, a basic chem. feedstock, in metal-catalyzed reductive C-C bond formation.
- 5
The reaction scope was later expanded by inclusion of other aldehydes and imines as coupling partners. See:
(a) Skucas, E.; Kong, J. R.; Krische, M. J. J. Am. Chem. Soc. 2007, 129, 7242– 7243Google Scholar5ahttps://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2sXlsVKjtL4%253D&md5=a7e874a53cbb7c68ca588ab664b01ab6Enantioselective Reductive Coupling of Acetylene to N-Arylsulfonyl Imines via Rhodium Catalyzed C-C Bond-Forming Hydrogenation: (Z)-Dienyl Allylic AminesSkucas, Eduardas; Kong, Jong Rock; Krische, Michael J.Journal of the American Chemical Society (2007), 129 (23), 7242-7243CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)The first highly enantioselective catalytic vinylation of aldimines to furnish allylic amines is reported. Exposure of arom. and aliph. N-arylsulfonyl aldimines to equal vols. of acetylene and hydrogen gas at 45 °C and ambient pressure in the presence of chirally modified cationic rhodium catalysts provides the (Z)-dienyl allylic amines, e.g., I, in highly optically enriched form (93-98% ee) and as single geometrical isomers (>95:5, Z/E). The coupling products arise via multicomponent coupling of four mols.: two mols. of acetylene, a mol. of aldimine, and elemental hydrogen. Unlike other imine addns. involving nonstabilized carbanions, the present protocol circumvents use of preformed organometallic reagents.(b) Han, S. B.; Kong, J. R.; Krische, M. J. Org. Lett. 2008, 10, 4133– 4135Google ScholarThere is no corresponding record for this reference. - 6Toyoshima, T.; Miura, T.; Murakami, M. Angew. Chem., Int. Ed. 2011, 50, 10436– 10439Google Scholar6https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3MXhtFOlsLnF&md5=671aecf04716fd7642c2c48cb11cad4eSelective 1:2 Coupling of Aldehydes and Allenes with Control of RegiochemistryToyoshima, Takeharu; Miura, Tomoya; Murakami, MasahiroAngewandte Chemie, International Edition (2011), 50 (44), 10436-10439, S10436/1-S10436/72CODEN: ACIEF5; ISSN:1433-7851. (Wiley-VCH Verlag GmbH & Co. KGaA)A new rhodium-catalyzed coupling reaction of one mol. of aldehyde and two mol. of allene was developed and gave selectively either of two constitutional isomers of β,γ-dialkylidene ketones, e.g. I and II. Interestingly, the regioselectivity of the reaction depends on the counterion of a rhodium(I) complex.
- 7(a) Oonishi, Y.; Hosotani, A.; Sato, Y. J. Am. Chem. Soc. 2011, 133, 10386– 10389Google Scholar7ahttps://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3MXns1enu70%253D&md5=71bb3766dfa4cc0ebc76992057e5c9d9Rh(I)-Catalyzed Formal [6 + 2] Cycloaddition of 4-Allenals with Alkynes or Alkenes in a TetherOonishi, Yoshihiro; Hosotani, Akihito; Sato, YoshihiroJournal of the American Chemical Society (2011), 133 (27), 10386-10389CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)Rh(I)-catalyzed formal [6 + 2] cycloaddn. of allenal, e.g., I, having an alkyne or alkene in a tether proceeded smoothly, giving fused bicyclic ketone derivs., e.g., II, in good to excellent yields. It was also found that cyclization of enantiomerically enriched I (94% ee) gave cyclic ketone deriv. II in high yield with reasonable chirality transfer (86% ee). This result indicates that this cyclization proceeds through stereoselective formation of rhodacycle H' followed by insertion of a multiple bond.(b) Hojo, D.; Tanaka, K. Org. Lett. 2012, 14, 1492– 1495Google Scholar7bhttps://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC38XivVOht7o%253D&md5=b2c27512626dc31fb7185f1ced643811Rhodium-Catalyzed C-H Bond Activation/[4+2] Annulation/Aromatization Cascade To Produce Phenol, Naphthol, Phenanthrenol, and Triphenylenol DerivativesHojo, Daiki; Tanaka, KenOrganic Letters (2012), 14 (6), 1492-1495CODEN: ORLEF7; ISSN:1523-7052. (American Chemical Society)It has been established that a cationic rhodium(I)/dppp complex catalyzes the aldehyde C-H bond activation/[4 + 2] annulation/aromatization cascade to produce phenol, naphthol, phenanthrenol, and triphenylenol derivs. from readily available conjugated alkynyl aldehydes and alkynes.
- 8
The total of 36 refers only to the number of possible linear compounds containing the reactant molecules coupled in a aldehyde–allene–allene sequence. In fact, two of the above isomers contain two asymmetric carbon atoms and hence can exist as two diastereomers, raising the number of possible products with distinct energies to 38. If cyclic structures and those containing allene–aldehyde–allene sequence are also considered, the number of possible products is even greater.
There is no corresponding record for this reference. - 9Williams, V. M.; Kong, J. R.; Ko, B. J.; Mantri, Y.; Brodbelt, J. S.; Baik, M.-H.; Krische, M. J. J. Am. Chem. Soc. 2009, 131, 16054– 16062Google ScholarThere is no corresponding record for this reference.
- 10Miura, T.; Biyajima, T.; Toyoshima, T.; Murakami, M. Beilstein J. Org. Chem. 2011, 7, 578– 581Google Scholar10https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3MXmsV2itb4%253D&md5=6ed76f458f528c62436bb4c886c73d5dSynthesis of cross-conjugated trienes by rhodium-catalyzed dimerization of monosubstituted allenesMiura, Tomoya; Biyajima, Tsuneaki; Toyoshima, Takeharu; Murakami, MasahiroBeilstein Journal of Organic Chemistry (2011), 7 (), 578-581, No. 67CODEN: BJOCBH; ISSN:1860-5397. (Beilstein-Institut zur Foerderung der Chemischen Wissenschaften)A rhodium(I)/dppe catalyst promoted dimerization of monosubstituted allenes in a stereoselective manner to give cross-conjugated trienes, which are different from those obtained by a palladium catalyst. A palladium-catalyzed reaction of 1,2-undecadiene delivered the previously reported 11-methyl-10-methylene-8,11-eicosadiene and a new product [i.e., (8E)-10,11-bis(methylene)-8-eicosene]. [4+2] Cycloaddn. products of (8E)-10,11-bis(methylene)-8-eicosene with 1,1,2,2-ethenetetracarbonitrile and 4-phenyl-3H-1,2,4-triazole-3,5(4H)-dione were reported.
- 11
In fact, to the best of our knowledge, all of the transition-metal-catalyzed [n + m + o] cycloadditions involving allenes developed to date incorporate only one or two molecules of allene into the product. See:
(a) Murakami, M.; Ubukata, M.; Itami, K.; Ito, Y. Angew. Chem., Int. Ed. 1998, 37, 2248– 2250Google ScholarThere is no corresponding record for this reference.(b) Shanmugasundaram, M.; Wu, M.-S.; Cheng, C.-H. Org. Lett. 2001, 3, 4233– 4236Google ScholarThere is no corresponding record for this reference.(c) Shanmugasundaram, M.; Wu, M.-S.; Jeganmohan, M.; Huang, C.-W.; Cheng, C.-H. J. Org. Chem. 2002, 67, 7724– 7729Google ScholarThere is no corresponding record for this reference.(d) Miura, T.; Morimoto, M.; Murakami, M. J. Am. Chem. Soc. 2010, 132, 15836– 15838Google ScholarThere is no corresponding record for this reference.(e) Brusoe, A. T.; Alexanian, E. J. Angew. Chem., Int. Ed. 2011, 50, 6596– 6600Google ScholarThere is no corresponding record for this reference. - 12(a) McCarren, P. R.; Liu, P.; Cheong, P. H.-Y.; Jamison, T. F.; Houk, K. N. J. Am. Chem. Soc. 2009, 131, 6654– 6655Google ScholarThere is no corresponding record for this reference.(b) Liu, P.; McCarren, P.; Cheong, P. H.-Y.; Jamison, T. F.; Houk, K. N. J. Am. Chem. Soc. 2010, 132, 2050– 2057Google Scholar12bhttps://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3cXptlKjtQ%253D%253D&md5=85aa276762e5d3f13c963e993c1e1be9Origins of Regioselectivity and Alkene-Directing Effects in Nickel-Catalyzed Reductive Couplings of Alkynes and AldehydesLiu, Peng; McCarren, Patrick; Cheong, Paul Ha-Yeon; Jamison, Timothy F.; Houk, K. N.Journal of the American Chemical Society (2010), 132 (6), 2050-2057CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)The origins of reactivity and regioselectivity in nickel-catalyzed reductive coupling reactions of alkynes and aldehydes were investigated with d. functional calcns. The regioselectivities of reactions of simple alkynes are controlled by steric effects, while conjugated enynes and diynes are predicted to have increased reactivity and very high regioselectivities, placing alkenyl or alkynyl groups distal to the forming C-C bond. The reactions of enynes and diynes involve 1,4-attack of the Ni-carbonyl complex on the conjugated enyne or diyne. The consequences of these conclusions on reaction design are discussed.(c) Liu, P.; Krische, M. J.; Houk, K. N. Chem.—Eur. J. 2011, 17, 4021– 4029Google Scholar12chttps://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3MXjsVGgt7s%253D&md5=8c3bd7ae34ef3c5a3ef62ca1c4b6e573Mechanism and Origins of Regio- and Enantioselectivities in RhI-Catalyzed Hydrogenative Couplings of 1,3-Diynes and Activated Carbonyl Partners: Intervention of a Cumulene IntermediateLiu, Peng; Krische, Michael J.; Houk, Kendall N.Chemistry - A European Journal (2011), 17 (14), 4021-4029, S4021/1-S4021/24CODEN: CEUJED; ISSN:0947-6539. (Wiley-VCH Verlag GmbH & Co. KGaA)The mechanism of the rhodium-catalyzed reductive coupling of 1,3-diynes and vicinal dicarbonyl compds. employing H2 as reductant was studied by d. functional theory. Oxidative coupling through 1,4-addn. of the RhI-bound dicarbonyl to the conjugated diyne via a seven-membered cyclic cumulene transition state leads to exclusive formation of linear adducts. Diyne 1,4-addn. is much faster than the 1,2-addn. to simple alkynes. The 1,2-dicarbonyl compd. is bound to rhodium in a bidentate fashion during the oxidative coupling event. The chemo-, regio-, and enantioselectivities of this reaction were studied and are attributed to this unique 1,4-addn. pathway. The close proximity of the ligand and the alkyne substituent distal to the forming C-C bond controls the regio- and enantioselectivity: coupling occurs at the sterically more demanding alkyne terminus, which minimizes nonbonded interaction with the ligand. A stereochem. model is proposed that accounts for preferential formation of the (R)-configurated coupling product when (R)-biaryl phosphine ligands are used.
- 13Frisch, M. J.; Gaussian 03, revision D.01; Gaussian, Inc.: Wallingford, CT, 2004.Google ScholarThere is no corresponding record for this reference.
- 14(a) Becke, A. D. J. Chem. Phys. 1993, 98, 5648– 5652Google Scholar14ahttps://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK3sXisVWgtrw%253D&md5=291bbfc119095338bb1624f0c21c7ca8Density-functional thermochemistry. III. The role of exact exchangeBecke, Axel D.Journal of Chemical Physics (1993), 98 (7), 5648-52CODEN: JCPSA6; ISSN:0021-9606.Despite the remarkable thermochem. accuracy of Kohn-Sham d.-functional theories with gradient corrections for exchange-correlation, the author believes that further improvements are unlikely unless exact-exchange information is considered. Arguments to support this view are presented, and a semiempirical exchange-correlation functional (contg. local-spin-d., gradient, and exact-exchange terms) is tested for 56 atomization energies, 42 ionization potentials, 8 proton affinities, and 10 total at. energies of first- and second-row systems. This functional performs better than previous functionals with gradient corrections only, and fits expt. atomization energies with an impressively small av. abs. deviation of 2.4 kcal/mol.(b) Lee, C. T.; Yang, W. T.; Parr, R. G. Phys. Rev. B 1988, 37, 785– 789Google Scholar14bhttps://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaL1cXktFWrtbw%253D&md5=ee7b59267a2ff72e15171a481819ccf8Development of the Colle-Salvetti correlation-energy formula into a functional of the electron densityLee, Chengteh; Yang, Weitao; Parr, Robert G.Physical Review B: Condensed Matter and Materials Physics (1988), 37 (2), 785-9CODEN: PRBMDO; ISSN:0163-1829.A correlation-energy formula due to R. Colle and D. Salvetti (1975), in which the correlation energy d. is expressed in terms of the electron d. and a Laplacian of the 2nd-order Hartree-Fock d. matrix, is restated as a formula involving the d. and local kinetic-energy d. On insertion of gradient expansions for the local kinetic-energy d., d.-functional formulas for the correlation energy and correlation potential are then obtained. Through numerical calcns. on a no. of atoms, pos. ions, and mols., of both open- and closed-shell type, it is demonstrated that these formulas, like the original Colle-Salvetti formulas, give correlation energies within a few percent.
- 15Hay, P. J.; Wadt, W. R. J. Chem. Phys. 1985, 82, 270– 283Google Scholar15https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaL2MXhtlyju70%253D&md5=29271d2a54b5c81acd19762c570e64d7Ab initio effective core potentials for molecular calculations. Potentials for the transition metal atoms scandium to mercuryHay, P. Jeffrey; Wadt, Willard R.Journal of Chemical Physics (1985), 82 (1), 270-83CODEN: JCPSA6; ISSN:0021-9606.Ab initio effective core potentials (ECP's) were generated to replace the Coulomb, exchange, and core-orthogonality effects of the chem. inert core electron in the transition metal atoms Sc to Hg. For the second and third transition series relative ECP's were generated which also incorporate the mass-velocity and Darwin relativistic effects into the potential. The ab initio ECP's should facilitate valence electron calcns. on mols. contg. transition-metal atoms with accuracies approaching all-electron calcns. at a fraction of the computational cost. Analytic fits to the potentials are presented for use in multicenter integral evaluation. Gaussian orbital valence basis sets are developed for the (3d,4s,4p), (4d,5s,5p), and (5d,6s,6p) orbitals of the first, second, and third transition series atoms, resp. All-electron and valence-electron at. excitation energies are also compared for the low-lying states of Sc-Hg, and the valence-electron calcns. reproduce the all-electron excitation energies (typically within a few tenths of an eV).
- 16(a) Barone, V.; Cossi, M. J. Phys. Chem. A 1998, 102, 1995– 2001Google Scholar16ahttps://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK1cXht1Cgt7o%253D&md5=7fe7f5f4627f26fd16a34e25219efaa6Quantum Calculation of Molecular Energies and Energy Gradients in Solution by a Conductor Solvent ModelBarone, Vincenzo; Cossi, MaurizioJournal of Physical Chemistry A (1998), 102 (11), 1995-2001CODEN: JPCAFH; ISSN:1089-5639. (American Chemical Society)A new implementation of the conductor-like screening solvation model (COSMO) in the GAUSSIAN94 package is presented. It allows Hartree-Fock (HF), d. functional (DF) and post-HF energy, and HF and DF gradient calcns.: the cavities are modeled on the mol. shape, using recently optimized parameters, and both electrostatic and nonelectrostatic contributions to energies and gradients are considered. The calcd. solvation energies for 19 neutral mols. in water are found in very good agreement with exptl. data; the solvent-induced geometry relaxation is studied for some closed and open shell mols., at HF and DF levels. The computational times are very satisfying: the self-consistent energy evaluation needs a time 15-30% longer than the corresponding procedure in vacuo, whereas the calcn. of energy gradients is only 25% longer than in vacuo for medium size mols.(b) Cossi, M.; Rega, N.; Scalmani, G.; Barone, V. J. Comput. Chem. 2003, 24, 669– 681Google Scholar16bhttps://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD3sXivFWqsbc%253D&md5=570ef9f44e925c9f78de6c7d97123229Energies, structures, and electronic properties of molecules in solution with the C-PCM solvation modelCossi, Maurizio; Rega, Nadia; Scalmani, Giovanni; Barone, VincenzoJournal of Computational Chemistry (2003), 24 (6), 669-681CODEN: JCCHDD; ISSN:0192-8651. (John Wiley & Sons, Inc.)The conductor-like solvation model, as developed in the framework of the polarizable continuum model (PCM), has been reformulated and newly implemented in order to compute energies, geometric structures, harmonic frequencies, and electronic properties in soln. for any chem. system that can be studied in vacuo. Particular attention is devoted to large systems requiring suitable iterative algorithms to compute the solvation charges: the fast multipole method (FMM) has been extensively used to ensure a linear scaling of the computational times with the size of the solute. A no. of test applications are presented to evaluate the performances of the method.
- 17Grimme, S. J. Comput. Chem. 2006, 27, 1787– 1799Google Scholar17https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD28XhtFenu7bO&md5=0b4aa16bebc3a0a2ec175d4b161ab0e4Semiempirical GGA-type density functional constructed with a long-range dispersion correctionGrimme, StefanJournal of Computational Chemistry (2006), 27 (15), 1787-1799CODEN: JCCHDD; ISSN:0192-8651. (John Wiley & Sons, Inc.)A new d. functional (DF) of the generalized gradient approxn. (GGA) type for general chem. applications termed B97-D is proposed. It is based on Becke's power-series ansatz from 1997 and is explicitly parameterized by including damped atom-pairwise dispersion corrections of the form C6·R-6. A general computational scheme for the parameters used in this correction has been established and parameters for elements up to xenon and a scaling factor for the dispersion part for several common d. functionals (BLYP, PBE, TPSS, B3LYP) are reported. The new functional is tested in comparison with other GGAs and the B3LYP hybrid functional on std. thermochem. benchmark sets, for 40 noncovalently bound complexes, including large stacked arom. mols. and group II element clusters, and for the computation of mol. geometries. Further cross-validation tests were performed for organometallic reactions and other difficult problems for std. functionals. In summary, it is found that B97-D belongs to one of the most accurate general purpose GGAs, reaching, for example for the G97/2 set of heat of formations, a mean abs. deviation of only 3.8 kcal mol-1. The performance for noncovalently bound systems including many pure van der Waals complexes is exceptionally good, reaching on the av. CCSD(T) accuracy. The basic strategy in the development to restrict the d. functional description to shorter electron correlation lengths scales and to describe situations with medium to large interat. distances by damped C6·R-6 terms seems to be very successful, as demonstrated for some notoriously difficult reactions. As an example, for the isomerization of larger branched to linear alkanes, B97-D is the only DF available that yields the right sign for the energy difference. From a practical point of view, the new functional seems to be quite robust and it is thus suggested as an efficient and accurate quantum chem. method for large systems where dispersion forces are of general importance.
- 18
For instance, see:
(a) Minenkov, Y.; Occhipinti, G.; Jensen, V. R. J. Phys. Chem. A 2009, 113, 11833– 11844Google ScholarThere is no corresponding record for this reference.(b) Siegbahn, P. E. M.; Blomberg, M. R. A.; Chen, S.-L. J. Chem. Theory Comput. 2010, 6, 2040– 2044Google Scholar18bhttps://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3cXntF2jtbc%253D&md5=5d6e81a233c545b03a68b0435229dc00Significant van der Waals Effects in Transition Metal ComplexesSiegbahn, Per E. M.; Blomberg, Margareta R. A.; Chen, Shi-LuJournal of Chemical Theory and Computation (2010), 6 (7), 2040-2044CODEN: JCTCCE; ISSN:1549-9618. (American Chemical Society)There is, in general, very good experience using hybrid DFT to study mechanisms of enzyme reactions contg. transition metals. For redox reactions, the B3LYP* functional, which has 15% exact exchange, has been shown to be particularly accurate. Still, there are some cases which have turned out to be quite difficult with large errors. In the present study, the effects of van der Waals interaction have been investigated for these cases, using the empirical formula of Grimme. The results are encouraging.(c) Harvey, J. N. Faraday Discuss. 2010, 145, 487– 505Google ScholarThere is no corresponding record for this reference.(d) McMullin, C. L.; Jover, J.; Harvey, J. N.; Fey, N. Dalton Trans. 2010, 39, 10833– 10836Google Scholar18dhttps://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3cXhtl2gt77E&md5=51d676bc1051a68e20a6228a58dc0a7bAccurate modelling of Pd(0) + PhX oxidative addition kineticsMcMullin, Claire L.; Jover, Jesus; Harvey, Jeremy N.; Fey, NatalieDalton Transactions (2010), 39 (45), 10833-10836CODEN: DTARAF; ISSN:1477-9226. (Royal Society of Chemistry)We have used dispersion-cor. DFT (DFT-D) together with solvation to examine possible mechanisms for reaction of PhX (X = Cl, Br, I) with Pd(PtBu3)2 and compare our results to recently published kinetic data (F. Barrios-Landeros, B. P. Carrow and J. F. Hartwig, J. Am. Chem. Soc., 2009, 131, 8141-8154). The calcd. activation free energies agree near-quant. with exptl. obsd. rate consts.(e) Lonsdale, R.; Harvey, J. N.; Mulholland, A. J. J. Phys. Chem. Lett. 2010, 1, 3232– 3237Google Scholar18ehttps://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3cXhtleitbvK&md5=9bc8d74cf3976d4da46535ec7d926182Inclusion of Dispersion Effects Significantly Improves Accuracy of Calculated Reaction Barriers for Cytochrome P450 Catalyzed ReactionsLonsdale, Richard; Harvey, Jeremy N.; Mulholland, Adrian J.Journal of Physical Chemistry Letters (2010), 1 (21), 3232-3237CODEN: JPCLCD; ISSN:1948-7185. (American Chemical Society)Prediction of cytochrome P 450 reactivity is of great importance to the development of new medicinal compds. D. functional theory (DFT) has proven itself as a useful tool in the characterization of the elusive reactive species, compd. I, and of the mechanisms of substrate oxidn. B3LYP is the most widely used d. functional in the study of P 450s; however, a major drawback of B3LYP is its inaccurate treatment of dispersion, leading to discrepancies between expt. and theory in some systems. Recent work has shown that an added empirical dispersion correction to B3LYP (B3LYP-D) yields more promising results for similar systems. In the present work, two previously studied systems, camphor hydroxylation and alkene oxidn., have been recalcd. using B3LYP-D. Our work shows that inclusion of dispersion has a significant effect on the energies and geometries of transition states and encounter complexes; furthermore, an improved agreement with exptl. data is obsd.(f) Osuna, S.; Swart, M.; Solà, M. J. Phys. Chem. A 2011, 115, 3491– 3496Google ScholarThere is no corresponding record for this reference.(g) Santoro, S.; Liao, R.-Z.; Himo, F. J. Org. Chem. 2011, 76, 9246– 9252Google ScholarThere is no corresponding record for this reference.(h) Nordin, M.; Liao, R.-Z.; Ahlford, K.; Adolfsson, H.; Himo, F. ChemCatChem 2012, 4, 1095– 1104Google ScholarThere is no corresponding record for this reference.(i) Xu, X.; Liu, P.; Lesser, A.; Sirois, L. E.; Wender, P. A.; Houk, K. N. J. Am. Chem. Soc. 2012, 134, 11012– 11025Google Scholar18ihttps://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC38XnvF2mtr4%253D&md5=92a384dcb6fe790b625d62dec27b40e5Ligand Effects on Rates and Regioselectivities of Rh(I)-Catalyzed (5 + 2) Cycloadditions: A Computational Study of Cyclooctadiene and Dinaphthocyclooctatetraene as LigandsXu, Xiufang; Liu, Peng; Lesser, Adam; Sirois, Lauren E.; Wender, Paul A.; Houk, K. N.Journal of the American Chemical Society (2012), 134 (26), 11012-11025CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)The first theor. study on the effects of ligands on the mechanism, reactivities, and regioselectivities of Rh(I)-catalyzed (5 + 2) cycloaddns. of vinylcyclopropanes (VCPs) and alkynes has been performed using d. functional theory (DFT) calcns. Highly efficient and selective intermol. (5 + 2) cycloaddns. of VCPs and alkynes have been achieved recently using two novel rhodium catalysts, [Rh(dnCOT)]+SbF6- and [Rh(COD)]+SbF6-, which provide superior reactivities and regioselectivities relative to that of the previously reported [Rh(CO)2Cl]2 catalyst. Computationally, the high reactivities of the dnCOT and COD ligands are attributed to the steric repulsions that destabilize the Rh-product complex, the catalyst resting state in the catalytic cycle. The regioselectivities of reactions with various alkynes and different Rh catalysts are investigated, and a predictive model is provided that describes substrate-substrate and ligand-substrate steric repulsions, electronic effects, and noncovalent π/π and C-H/π interactions. In the reactions with dnCOT or COD ligands, the first new C-C bond is formed proximal to the bulky substituent on the alkyne to avoid ligand-substrate steric repulsions. This regioselectivity is reversed either by employing the smaller [Rh(CO)2Cl]2 catalyst to diminish the ligand-substrate repulsions or by using aryl alkynes, for which the ligand-substrate interactions become stabilizing due to π/π and C-H/π dispersion interactions. Electron-withdrawing groups on the alkyne prefer to be proximal to the first new C-C bond to maximize metal-substrate back-bonding interactions. These steric, electronic, and dispersion effects can all be utilized in designing new ligands to provide regiochem. control over product formation with high selectivities. The computational studies reveal the potential of employing the dnCOT family of ligands to achieve unique regiochem. control due to the steric influences and dispersion interactions assocd. with the rigid aryl substituents on the ligand.(j) Jiménez-Halla, J. O. C.; Kalek, M.; Stawinski, J.; Himo, F. Chem.—Eur. J. 2012, 18, 12424– 12436Google Scholar18jhttps://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC38XhtFOmtrbL&md5=d3ea852f2a6ceef323d0426288ace547Computational Study of the Mechanism and Selectivity of Palladium-Catalyzed Propargylic Substitution with Phosphorus NucleophilesJimenez-Halla, J. Oscar C.; Kalek, Marcin; Stawinski, Jacek; Himo, FahmiChemistry - A European Journal (2012), 18 (39), 12424-12436, S12424/1-S12424/155CODEN: CEUJED; ISSN:0947-6539. (Wiley-VCH Verlag GmbH & Co. KGaA)The mechanism and sources of selectivity in the palladium-catalyzed propargylic substitution reaction that involves phosphorus nucleophiles, and which yields predominantly allenylphosphonates and related compds., have been studied computationally by d. functional theory. Full free-energy profiles are computed for both H-phosphonate and H-phosphonothioate substrates. The calcns. show that the special behavior of H-phosphonates among other heteroatom nucleophiles is indeed reflected in higher energy barriers for the attack on the central carbon atom of the allenyl/propargyl ligand relative to the ligand-exchange pathway, which leads to the exptl. obsd. products. It is argued that, to explain the preference of allenyl- vs. propargyl-phosphonate/phosphonothioate formation in reactions that involve H-phosphonates and H-phosphonothioates, anal. of the complete free-energy surfaces is necessary, because the product ratio is detd. by different transition states in the resp. branches of the catalytic cycle. In addn., these transition states change in going from a H-phosphonate to a H-phosphonothioate nucleophile.(k) Kalek, M.; Himo, F. J. Am. Chem. Soc. 2012, 134, 19159– 19169Google ScholarThere is no corresponding record for this reference.(l) Huang, G.; Xia, Y.; Sun, C.; Li, J.; Lee, D. J. Org. Chem. 2013, 78, 988– 995Google ScholarThere is no corresponding record for this reference. - 19
For selected reports of η3-allyl Rh species, see:
(a) Evans, P. A.; Lawler, M. J. J. Am. Chem. Soc. 2004, 126, 8642– 8643Google ScholarThere is no corresponding record for this reference.(b) Evans, P. A.; Leahy, D. K. Chemtracts 2003, 16, 567– 578Google ScholarThere is no corresponding record for this reference.(c) Evans, P. A.; Leahy, D. K.; Slieker, L. M. Tetrahedron: Asymmetry 2003, 14, 3613– 3618Google ScholarThere is no corresponding record for this reference.(d) Evans, P. A.; Robinson, J. E.; Moffett, K. K. Org. Lett. 2001, 3, 3269– 3271Google ScholarThere is no corresponding record for this reference.(e) Arnold, J. S.; Cizio, G. T.; Nguyen, H. M. Org. Lett. 2011, 13, 5576– 5579Google ScholarThere is no corresponding record for this reference.(f) Arnold, J. S.; Cizio, G. T.; Heitz, D. R.; Nguyen, H. M. Chem. Commun. 2012, 48, 11531– 11533Google ScholarThere is no corresponding record for this reference.(g) Arnold, J. S.; Stone, R. F.; Nguyen, H. M. Org. Lett. 2010, 12, 4580– 4583Google ScholarThere is no corresponding record for this reference.(h) Arnold, J. S.; Nguyen, H. M. J. Am. Chem. Soc. 2012, 134, 8380– 8383Google ScholarThere is no corresponding record for this reference.(i) Hayashi, T.; Okada, A.; Suzuka, T.; Kawatsura, M. Org. Lett. 2003, 5, 1713– 1715Google ScholarThere is no corresponding record for this reference.(j) Vrieze, D. C.; Hoge, G. S.; Hoerter, P. Z.; Van Haitsma, J. T.; Samas, B. M. Org. Lett. 2009, 11, 3140– 3142Google ScholarThere is no corresponding record for this reference.(k) Koschker, P.; Lumbroso, A.; Breit, B. J. Am. Chem. Soc. 2011, 133, 20746– 20749Google ScholarThere is no corresponding record for this reference.(l) Lumbroso, A.; Koschker, P.; Vautravers, N. R.; Breit, B. J. Am. Chem. Soc. 2011, 133, 2386– 2389Google Scholar19lhttps://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3MXhsVOnsLc%253D&md5=d71b0c3b438ed69fac23c09a8bc16177Redox-neutral atom-economic rhodium-catalyzed coupling of terminal alkynes with carboxylic acids toward branched allylic estersLumbroso, Alexandre; Koschker, Philipp; Vautravers, Nicolas R.; Breit, BernhardJournal of the American Chemical Society (2011), 133 (8), 2386-2389CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)A new method for the prepn. of a wide range of branched allylic esters from terminal alkynes that proceeds via a redox-neutral propargylic CH activation employing a rhodium(I)/DPEphos catalyst is reported.(m) Nishimura, T.; Hirabayashi, S.; Yasuhara, Y.; Hayashi, T. J. Am. Chem. Soc. 2006, 128, 2556– 2557Google ScholarThere is no corresponding record for this reference.(n) Choi, J.-c.; Osakada, K.; Yamamoto, T. Organometallics 1998, 17, 3044– 3050Google ScholarThere is no corresponding record for this reference.(o) Barros, H. J. V.; Guimarães, C. C.; dos Santos, E. N.; Gusevskaya, E. V. Organometallics 2007, 26, 2211– 2218Google ScholarThere is no corresponding record for this reference.(p) Jiao, L.; Lin, M.; Zhuo, L.-G.; Yu, Z.-X. Org. Lett. 2010, 12, 2528– 2531Google ScholarThere is no corresponding record for this reference.(q) Jiao, L.; Lin, M.; Yu, Z.-X. J. Am. Chem. Soc. 2011, 133, 447– 461Google ScholarThere is no corresponding record for this reference.(r) Lin, M.; Li, F.; Jiao, L.; Yu, Z.-X. J. Am. Chem. Soc. 2011, 133, 1690– 1693Google ScholarThere is no corresponding record for this reference.(s) Lin, M.; Kang, G.-Y.; Guo, Y.-A.; Yu, Z.-X. J. Am. Chem. Soc. 2012, 134, 398– 405Google ScholarThere is no corresponding record for this reference.(t) Li, Q.; Yu, Z.-X. Organometallics 2012, 31, 5185– 5195Google ScholarThere is no corresponding record for this reference. - 20(a) Szabó, K. J. Chem.—Eur. J. 2004, 10, 5268– 5275Google Scholar20ahttps://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2cXpsF2rtLk%253D&md5=d619dbf6a865b2ab693c7d7950c6ca76Palladium-catalyzed electrophilic allylation reactions via bis(allyl)palladium complexes and related intermediatesSzabo, Kalman J.Chemistry - A European Journal (2004), 10 (21), 5268-5275CODEN: CEUJED; ISSN:0947-6539. (Wiley-VCH Verlag GmbH & Co. KGaA)A review. The synthetic scope of the allyl-Pd chem. can be extended to involve electrophilic reagents. The greatest challenge in these reactions is the catalytic generation of an allyl-Pd intermediate incorporating a nucleophilic allyl moiety. A vast majority of the published reactions that involve Pd-catalyzed allylation of electrophiles proceed via bis(allyl)palladium intermediates. The η1-moiety of the bis(allyl)palladium intermediates reacts with electrophiles, including aldehydes, imines, or Michael acceptors. Recently, catalytic electrophilic allylations via monoallylpalladium complexes have also been presented by employment of so-called pincer complex catalysts.(b) Szabó, K. J. Chem.—Eur. J. 2000, 6, 4413– 4421Google Scholar20bhttps://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD3cXovVyqs70%253D&md5=430af920ef49e31d8963558650e0a3aaUmpolung of the allylpalladium reactivity: mechanism and regioselectivity of the electrophilic attack on bis-allylpalladium complexes formed in palladium-catalyzed transformationsSzabo, Kalman J.Chemistry - A European Journal (2000), 6 (23), 4413-4421CODEN: CEUJED; ISSN:0947-6539. (Wiley-VCH Verlag GmbH)The structure and reactivity of various bis-allylpalladium complexes occurring as catalytic intermediates in important synthetic transformations were studied by applying d. functional theory at the B3PW91 (DA + P) level. η1,η3 Coordinated bis-allylpalladium complexes are readily formed from the corresponding η3,η3 complexes, esp. in the presence of π-acceptor phosphine ligands. The theor. calcns. indicate dσ → π* type hyperconjugative interactions occurring in the η1-coordinated allyl moiety of the η1,η3 coordinated complexes. These hyperconjugative interactions influence the structure of the complexes and dramatically increase the reactivity of the double bond in the η1-moiety. The DFT results indicate a remarkably low activation barrier for the electrophilic attack on the η1-allyl functionality. In bridged η1,η3 complexes, the electrophilic attack occurs with a very high regioselectivity, which can be explained from d-π type hyperconjugative interactions.
- 21
There is an additional possibility of a migratory insertion of a third molecule of allene instead of the aldehyde. Such a pathway would eventually lead to [2 + 2 + 2] trimerization of the allene. However, since experimentally allene 2 only dimerizes (see ref 10), this possibility was not considered computationally.
There is no corresponding record for this reference. - 22
For selected reviews of allylation of carbonyl compounds, see:
(a) Denmark, S. E.; Fu, J. Chem. Rev. 2003, 103, 2763– 2794Google Scholar22ahttps://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD3sXltVGjs78%253D&md5=b9f7c57ce56327693b6c8be57a50cda6Catalytic enantioselective addition of allylic organometallic reagents to aldehydes and ketonesDenmark, Scott E.; Fu, JipingChemical Reviews (Washington, DC, United States) (2003), 103 (8), 2763-2793CODEN: CHREAY; ISSN:0009-2665. (American Chemical Society)A review discusses the enantioselective allylation of aldehydes and ketones with a variety of allylmetal reagents to provide nonracemic homoallylic alcs. The mechanism of and chiral Lewis acid catalysts for the addn. of allylic silanes, stannanes and boranes to aldehydes and ketones, catalytic enantioselective allylation reactions with allylic halides, Lewis base-catalyzed enantioselective allylation with allylic trichlorosilanes, and allenylation and propargylation of aldehydes are discussed in the review; a table of conditions and selectivities for the enantioselective allylation of a variety of substrates is also provided.(b) Kennedy, J. W. J.; Hall, D. G. Angew. Chem., Int. Ed. 2003, 42, 4732– 4739Google Scholar22bhttps://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD3sXosFKntLg%253D&md5=9661db0988639ab2c9ef7ab6b04e91b6Recent advances in the activation of boron and silicon reagents for stereocontrolled allylation reactionsKennedy, Jason W. J.; Hall, Dennis G.Angewandte Chemie, International Edition (2003), 42 (39), 4732-4739CODEN: ACIEF5; ISSN:1433-7851. (Wiley-VCH Verlag GmbH & Co. KGaA)A review. Despite the popularity of boron and silicon allylation reagents in stereocontrolled synthesis, they suffer from a no. of inherent limitations that have slowed down their development as synthetic tools for nucleophilic addns. to carbonyl compds. and imine derivs. These limitations are the low reactivity and diastereoselectivity of allyl trialkylsilane reagents, and the lack of catalytic systems for the activation and substoichiometric control of enantioselectivity in the addns. of allyl boron reagents. To develop more efficient and general methods for the control of abs. stereochem. in the resulting homoallylic alcs., new approaches aimed at solving the problem of activation of allylic boron and silicon reagents are needed. This Minireview describes a no. of recent approaches that have been devised to address this problem.(c) Yu, C.-M.; Youn, J.; Jung, H.-K. Bull. Korean Chem. Soc. 2006, 27, 463– 472Google Scholar22chttps://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD28XkvVektr8%253D&md5=e940b908b85abcd16616c1858cef0085Regulation of stereoselectivity and reactivity in inter- and intramolecular allylic transfer reactionsYu, Chan-Mo; Youn, Jinsoup; Jung, Hee-KeumBulletin of the Korean Chemical Society (2006), 27 (4), 463-472CODEN: BKCSDE; ISSN:0253-2964. (Korean Chemical Society)A review. The prepn. of enantiomerically enriched homoallylic alcs. through asym. addn. of chiral allylic transfer reagents and allylating reagents with chiral catalysts to carbonyl functionalities represents an important chem. transformation. Excellent progress has been made over the past decade in the development and application of catalytic asym. allylic transfer reactions. In this account, our efforts for the various intermol. allylic transfer reactions such as allylation, propargylation, allenylation, and dienylation utilizing an accelerating strategy and sequential allylic transfer reactions to achieve multiple stereoselection, mainly using transition metal catalysts, are described.(d) Marek, I.; Sklute, G. Chem. Commun. 2007, 1683– 1691Google Scholar22dhttps://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2sXksFSit70%253D&md5=6f75de6048df9ee00d842748c108b870Creation of quaternary stereocenters in carbonyl allylation reactionsMarek, Ilan; Sklute, GeniaChemical Communications (Cambridge, United Kingdom) (2007), (17), 1683-1691CODEN: CHCOFS; ISSN:1359-7345. (Royal Society of Chemistry)A review. Despite the advances in stereoselective carbonyl allylation reactions, the creation of quaternary stereocenters by the addn. of 3,3'-disubstituted allylmetals to aldehydes is still a challenging issue. This feature article describes the most powerful approaches that have been devised to address this problem. The application of allyl boronates, allyl(trichloro)silanes and allylzinc reagents are discussed.(e) Hall, D. G. Synlett 2007, 1644– 1655Google ScholarThere is no corresponding record for this reference. - 23
For examples of [2 + 2 + 2] cyclization involving carbonyl compounds, see:
(a) Tsuchikama, K.; Yoshinami, Y.; Shibata, T. Synlett 2007, 1395– 1398Google ScholarThere is no corresponding record for this reference.(b) Tanaka, K.; Otake, Y.; Wada, A.; Noguchi, K.; Hirano, M. Org. Lett. 2007, 9, 2203– 2206Google ScholarThere is no corresponding record for this reference. - 24
Transition states connecting INT6 to INT7 and INT6′ to INT7′ could not be optimized. However, energy scans showed that the barriers are quite low, much lower than those for the direct reductive elimination (see the Supporting Information).
There is no corresponding record for this reference. - 25
The methyl group in the η3-allyl complexes INT7 and INT7′, contrary to the case of INT2 discussed above, can undergo a relocation from the anti position to the syn position. This could potentially lead to the formation of isomers of products 4 and 6 containing different configurations at one of the double bonds. However, the barriers for the syn/anti substituent exchange in INT7 and INT7′ were calculated to be much higher than those for the reductive elimination. Therefore, INT7 and INT7′ undergo the reductive elimination before they can isomerize. See the Supporting Information for details.
There is no corresponding record for this reference. - 26
The ratios were calculated using the Eyring equation. For example, the ratio of INT6 to INT6′ was calculated as
There is no corresponding record for this reference. - 27Wade, L. G., Jr. Organic Chemistry, 6th ed.; Prentice-Hall: Upper Saddle River, NJ, 2006.Google ScholarThere is no corresponding record for this reference.
- 28Nielsen, R. J.; Goddard, W. A., III. J. Am. Chem. Soc. 2006, 128, 9651– 9660Google ScholarThere is no corresponding record for this reference.
Cited By
This article is cited by 22 publications.
- Ping Wu, Fanshu Cao, Yu Zhou, Zuqian Xue, Ni Zhang, Lei Shi, Gen Luo. Substrate Facilitating Roles in Rare-Earth-Catalyzed C–H Alkenylation of Pyridines with Allenes: Mechanism and Origins of Regio- and Stereoselectivity. Inorganic Chemistry 2022, 61
(43)
, 17330-17341. https://doi.org/10.1021/acs.inorgchem.2c02953
- Shihan Liu, Tao Zhang, Lei Zhu, Fenru Liu, Ruopeng Bai, Yu Lan. Layered Chirality Relay Model in Rh(I)-Mediated Enantioselective C–Si Bond Activation: A Theoretical Study. Organic Letters 2020, 22
(6)
, 2124-2128. https://doi.org/10.1021/acs.orglett.9b04636
- Wei Ding, Yan King Terence Ho, Yasuhiro Okuda, Christopher Kevin Wijaya, Zheng Hao Tan, Naohiko Yoshikai. Cobalt-Catalyzed Hydroacylative Dimerization of Allenes Leading to Skipped Dienes. Organic Letters 2019, 21
(15)
, 6173-6178. https://doi.org/10.1021/acs.orglett.9b02465
- Mei Zhang, Lingfei Hu, Yanmin Lang, Yang Cao, and Genping Huang . Mechanism and Origins of Regio- and Enantioselectivities of Iridium-Catalyzed Hydroarylation of Alkenyl Ethers. The Journal of Organic Chemistry 2018, 83
(5)
, 2937-2947. https://doi.org/10.1021/acs.joc.8b00377
- Tao Yang and Masahiro Ehara . Computational Studies on Reaction Mechanism and Origins of Selectivities in Nickel-Catalyzed (2 + 2 + 2) Cycloadditions and Alkenylative Cyclizations of 1,6-Ene–Allenes and Alkenes. The Journal of Organic Chemistry 2017, 82
(4)
, 2150-2159. https://doi.org/10.1021/acs.joc.6b02957
- Stefano Santoro, Marcin Kalek, Genping Huang, and Fahmi Himo . Elucidation of Mechanisms and Selectivities of Metal-Catalyzed Reactions using Quantum Chemical Methodology. Accounts of Chemical Research 2016, 49
(5)
, 1006-1018. https://doi.org/10.1021/acs.accounts.6b00050
- Jiajia Chen, Wei Guo, and Yuanzhi Xia . Computational Revisit to the β-Carbon Elimination Step in Rh(III)-Catalyzed C–H Activation/Cycloaddition Reactions of N-Phenoxyacetamide and Cyclopropenes. The Journal of Organic Chemistry 2016, 81
(6)
, 2635-2638. https://doi.org/10.1021/acs.joc.6b00003
- Genping Huang and Peng Liu . Mechanism and Origins of Ligand-Controlled Linear Versus Branched Selectivity of Iridium-Catalyzed Hydroarylation of Alkenes. ACS Catalysis 2016, 6
(2)
, 809-820. https://doi.org/10.1021/acscatal.5b02201
- Theresa Sperger, Italo A. Sanhueza, Indrek Kalvet, and Franziska Schoenebeck . Computational Studies of Synthetically Relevant Homogeneous Organometallic Catalysis Involving Ni, Pd, Ir, and Rh: An Overview of Commonly Employed DFT Methods and Mechanistic Insights. Chemical Reviews 2015, 115
(17)
, 9532-9586. https://doi.org/10.1021/acs.chemrev.5b00163
- Genping Huang . Mechanism and Selectivity in Rhodium-Catalyzed [7 + 2] Cycloaddition and Cyclopropanation/Cyclization of Allenylcyclopentane-alkynes: Metallacycle-Directed C(sp3)-C(sp3) vs C(sp3)-H Activation. The Journal of Organic Chemistry 2015, 80
(15)
, 7564-7571. https://doi.org/10.1021/acs.joc.5b01148
- Tao Liu, Lingli Han, Shumin Han, and Siwei Bi . Theoretical Investigation on Rhodium(I)-Catalyzed Cycloisomerizations of 4-Allenal Species with Linked Alkyne: Ketone vs Alcohol Products. Organometallics 2015, 34
(1)
, 280-288. https://doi.org/10.1021/om501118e
- Xin Hong, Matthew C. Stevens, Peng Liu, Paul A. Wender, and K. N. Houk . Reactivity and Chemoselectivity of Allenes in Rh(I)-Catalyzed Intermolecular (5 + 2) Cycloadditions with Vinylcyclopropanes: Allene-Mediated Rhodacycle Formation Can Poison Rh(I)-Catalyzed Cycloadditions. Journal of the American Chemical Society 2014, 136
(49)
, 17273-17283. https://doi.org/10.1021/ja5098308
- Hongli Wu, Xiaojie Li, Xiangyang Tang, Genping Huang. Mechanism and origins of chemo- and regioselectivities of (NHC)NiH-catalyzed cross-hydroalkenylation of vinyl ethers with α-olefins: a computational study. Organic Chemistry Frontiers 2018, 5
(23)
, 3410-3420. https://doi.org/10.1039/C8QO01020J
- Xiaojie Li, Hongli Wu, Yanmin Lang, Genping Huang. Mechanism, selectivity, and reactivity of iridium- and rhodium-catalyzed intermolecular ketone α-alkylation with unactivated olefins
via
an enamide directing strategy. Catalysis Science & Technology 2018, 8
(9)
, 2417-2426. https://doi.org/10.1039/C8CY00290H
- Hongyan Zou, Zhong‐Liang Wang, Genping Huang. Mechanism and Origins of the Chemo‐ and Regioselectivities in Nickel‐Catalyzed Intermolecular Cycloadditions of Benzocyclobutenones with 1,3‐Dienes. Chemistry – A European Journal 2017, 23
(51)
, 12593-12603. https://doi.org/10.1002/chem.201702316
- Sérgio Filipe Sousa, António J. M. Ribeiro, Rui P. P. Neves, Natércia F. Brás, Nuno M. F. S. A. Cerqueira, Pedro A. Fernandes, Maria João Ramos. Application of quantum mechanics/molecular mechanics methods in the study of enzymatic reaction mechanisms. WIREs Computational Molecular Science 2017, 7
(2)
https://doi.org/10.1002/wcms.1281
- Li Wang, Hongyan Zou, Xinwen Zhang, Genping Huang. Mechanism and origins of selectivity in rhodium-catalyzed intermolecular [3 + 2] cycloadditions of vinylaziridines with allenes. Organic Chemistry Frontiers 2017, 4
(4)
, 587-596. https://doi.org/10.1039/C6QO00671J
- B. A. Murray. Reactions of Aldehydes and Ketones and Their Derivatives. 2016, 1-66. https://doi.org/10.1002/9781118707838.ch1
- Xinwen Zhang, Hongyan Zou, Genping Huang. Mechanism and Origins of Ligand‐Controlled Selectivity of Rhodium‐Catalyzed Intermolecular Cycloadditions of Vinylaziridines with Alkynes. ChemCatChem 2016, 8
(15)
, 2549-2556. https://doi.org/10.1002/cctc.201600349
- Mei Zhang, Genping Huang. Mechanism and Selectivity of Ru
II
‐ and Rh
III
‐Catalyzed Oxidative Spiroannulation of Naphthols and Phenols with Alkynes through a C−H Activation/Dearomatization Strategy. Chemistry – A European Journal 2016, 22
(27)
, 9356-9365. https://doi.org/10.1002/chem.201600884
- Mei Zhang, Genping Huang. Mechanism of iridium-catalysed branched-selective hydroarylation of vinyl ethers: a computational study. Dalton Transactions 2016, 45
(8)
, 3552-3557. https://doi.org/10.1039/C5DT04973C
- Elena Soriano, Israel Fernández. Allenes and computational chemistry: from bonding situations to reaction mechanisms. Chemical Society Reviews 2014, 43
(9)
, 3041. https://doi.org/10.1039/c3cs60457h
Article Views are the COUNTER-compliant sum of full text article downloads since November 2008 (both PDF and HTML) across all institutions and individuals. These metrics are regularly updated to reflect usage leading up to the last few days.
Citations are the number of other articles citing this article, calculated by Crossref and updated daily. Find more information about Crossref citation counts.
The Altmetric Attention Score is a quantitative measure of the attention that a research article has received online. Clicking on the donut icon will load a page at altmetric.com with additional details about the score and the social media presence for the given article. Find more information on the Altmetric Attention Score and how the score is calculated.
Recommended Articles
References
This article references 28 other publications.
- 1(a) Lautens, M.; Klute, W.; Tam, W. Chem. Rev. 1996, 96, 49– 921ahttps://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK28XjsVKktw%253D%253D&md5=3776dd13a62c7dbef311a7cbf457dd7fTransition Metal-Mediated Cycloaddition ReactionsLautens, Mark; Klute, Wolfgang; Tam, WilliamChemical Reviews (Washington, D. C.) (1996), 96 (1), 49-92CODEN: CHREAY; ISSN:0009-2665. (American Chemical Society)A review with 226 refs. A discussion of established and interesting new cycloaddn. reactions promoted by metal species is presented.(b) Kotha, S.; Brahmachary, E.; Lahiri, K. Eur. J. Org. Chem. 2005, 4741– 47671bhttps://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2MXht1KgurrM&md5=93220a4594f4dfa8e3cf02aac00ba43aTransition metal-catalyzed [2+2+2] cycloaddition and application in organic synthesisKotha, Sambasivarao; Brahmachary, Enugurthi; Lahiri, KakaliEuropean Journal of Organic Chemistry (2005), (22), 4741-4767CODEN: EJOCFK; ISSN:1434-193X. (Wiley-VCH Verlag GmbH & Co. KGaA)A review. The [2+2+2] cycloaddn. strategy is complementary to the well known Diels-Alder reaction for the generation of polycyclic compds. This [2+2+2] approach is atom-economical and, with the availability of new catalysts to effect the [2+2+2] cycloaddn. reaction, synthesis of a wide variety of highly functionalized polycycles is possible. This review deals with some recent advances relating to [2+2+2] cycloaddn. reactions involving the syntheses both of polycycles and of heterocycles. More specifically, syntheses of various biol. active mols., unusual amino acids, and theor. interesting mols. are described. An attempt has also been made to give an overview of recent advances in the achievement of chemo-, regio-, and stereoselectivity in cyclotrimerization reactions.(c) Murakami, M. Angew. Chem., Int. Ed. 2003, 42, 718– 720There is no corresponding record for this reference.(d) Inglesby, P. A.; Evans, P. A. Chem. Soc. Rev. 2010, 39, 2791– 28051dhttps://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3cXptFyhu7c%253D&md5=f8e29c1b1b3795588b1be0f7e9c11a97Stereoselective transition metal-catalysed higher-order carbocyclisation reactionsInglesby, Phillip A.; Evans, P. AndrewChemical Society Reviews (2010), 39 (8), 2791-2805CODEN: CSRVBR; ISSN:0306-0012. (Royal Society of Chemistry)A review. Transition metal-catalyzed higher-order carbocyclization reactions represent an important class of reactions due to their ability to construct complex polycyclic systems in a highly selective and atom-economical fashion. A key and striking feature with these transformations is the dichotomy in reactivity that a substrate displays with different transition metal complexes, which is akin to the manner enzymes direct terpene biosynthesis. This tutorial review details the historical development of higher-order carbocyclization reactions, specifically the variants of [m+2+2] that involve carbon-based π-systems, where m = 2, 3 and 4, in the context of crit. developments with various transition metal complexes.(e) Shibata, T.; Tsuchikama, K. Org. Biomol. Chem. 2008, 6, 1317– 1323There is no corresponding record for this reference.(f) Tanaka, K. Chem.—Asian J. 2009, 4, 508– 5181fhttps://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD1MXks1Cht7c%253D&md5=941a1af8b94d4a5e67af5a6b0f552d64Transition-metal-catalyzed enantioselective [2+2+2] cycloadditions for the synthesis of axially chiral biarylsTanaka, KenChemistry - An Asian Journal (2009), 4 (4), 508-518CODEN: CAAJBI; ISSN:1861-4728. (Wiley-VCH Verlag GmbH & Co. KGaA)A review. Recent advances towards a development of novel transition-metal-catalyzed enantioselective [2+2+2] cycloaddn. for the synthesis of axially chiral biaryls are reviewed. A no. of efficient enantioselective biaryl syntheses were accomplished by chiral cobalt(I), iridium(I), and rhodium(I) complex catalyzed [2+2+2] cycloaddn. Furthermore, the enantioselective synthesis of axially chiral biaryls possessing non-biaryl axial chirality was also developed by using chiral rhodium(I) complexes as catalysts.(g) Yu, Z.-X.; Wang, Y.; Wang, Y. Chem.—Asian J. 2010, 5, 1072– 10881ghttps://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3cXls1Wkur8%253D&md5=867a869887dd578c5621f6992c3b6319Transition-metal-catalyzed cycloadditions for the synthesis of eight-membered carbocyclesYu, Zhi-Xiang; Wang, Yi; Wang, YuanyuanChemistry - An Asian Journal (2010), 5 (5), 1072-1088CODEN: CAAJBI; ISSN:1861-4728. (Wiley-VCH Verlag GmbH & Co. KGaA)A review. Eight-membered carbocycles are in a wide variety of natural products that exhibit a broad range of biol. and medicinal activities (cf. the most potent anticancer drug, taxol). However, the synthesis of eight-membered carbocycles is quite challenging as traditional approaches are met with entropic and enthalpic penalties in the ring-forming transition states. These neg. effects can be totally or partially avoided with the implementation of transition-metal-catalyzed/mediated cycloaddns. In this focus review, examples of elegant and efficient metal-catalyzed and some metal-mediated cycloaddns. (including Ni- catalyzed [4+4] and Rh-catalyzed [4+2+2] and [5+2+1] reactions) are presented to illustrate this. Application of these cycloaddn. reactions in total synthesis was also presented to show the significance of these reactions in addressing challenges in natural product synthesis.
- 2(a) Willis, M. C. Chem. Rev. 2010, 110, 725– 7482ahttps://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD1MXhtleisL3N&md5=bd9108b8b2ab0b60865cdd64044736faTransition Metal Catalyzed Alkene and Alkyne HydroacylationWillis, Michael C.Chemical Reviews (Washington, DC, United States) (2010), 110 (2), 725-748CODEN: CHREAY; ISSN:0009-2665. (American Chemical Society)This review is a coordinated anal. of the intramol. or intermol. alkene hydroacylation using transition metal catalysts. The mechanism of such transformations is also presented in this systematic review.(b) Jun, C.-H.; Moon, C. W.; Lee, D.-Y. Chem.—Eur. J. 2002, 8, 2422– 24282bhttps://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD38XksFWgs7s%253D&md5=9f3fa5e5d3340dfb3541509a40f2bcc3Chelation-assisted carbon-hydrogen and carbon-carbon bond activation by transition metal catalystsJun, Chul-Ho; Moon, Choong Woon; Lee, Dae-YonChemistry - A European Journal (2002), 8 (11), 2422-2428CODEN: CEUJED; ISSN:0947-6539. (Wiley-VCH Verlag GmbH)A review with refs. on the chelation-assisted C-H and C-C bond activation of carbonyl compds. by RhI catalysts. Hydroacylation of olefins was accomplished by utilizing 2-amino-3-picoline as a chelation auxiliary. The same strategy was employed for the C-C bond activation of unstrained ketones. Allylamine 24 was devised as a synthon of formaldehyde. Hydroiminoacylation of alkynes with allylamine 24 was applied to the alkyne cleavage by the aid of cyclohexylamine.(c) Leung, J. C.; Krische, M. J. Chem. Sci. 2012, 3, 2202– 22092chttps://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC38XotFGku7g%253D&md5=551e827a85202d1ca97cfff97e5b5ec4Catalytic intermolecular hydroacylation of C-C π-bonds in the absence of chelation assistanceLeung, Joyce C.; Krische, Michael J.Chemical Science (2012), 3 (7), 2202-2209CODEN: CSHCCN; ISSN:2041-6520. (Royal Society of Chemistry)A review. Intermol. rhodium catalyzed hydroacylation in the absence of chelation assistance remains a largely unmet challenge due to competing decarbonylation of the acylrhodium intermediates to form catalytically inactive carbonyl complexes. Here, catalytic systems for intermol. hydroacylation in the absence of chelation assistance are reviewed, with an emphasis on recently described processes that operate through mechanistic pathways beyond aldehyde C-H oxidative addn.
- 3(a) Jang, H.-Y.; Krische, M. J. Acc. Chem. Res. 2004, 37, 653– 661There is no corresponding record for this reference.(b) Ngai, M.-Y.; Kong, J.-R.; Krische, M. J. J. Org. Chem. 2007, 72, 1063– 1072There is no corresponding record for this reference.(c) Iida, H.; Krische, M. J. Top. Curr. Chem. 2007, 279, 77– 104There is no corresponding record for this reference.(d) Bower, J. F.; Kim, I. S.; Patman, R. L.; Krische, M. J. Angew. Chem., Int. Ed. 2009, 48, 34– 46There is no corresponding record for this reference.(e) Skucas, E.; Ngai, M.-Y.; Komanduri, V.; Krische, M. J. Acc. Chem. Res. 2007, 40, 1394– 1401There is no corresponding record for this reference.
- 4Kong, J. R.; Krische, M. J. J. Am. Chem. Soc. 2006, 128, 16040– 160414https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD28Xht1GjsLvO&md5=4fed0d21910ffedc7657e0f145866142Catalytic Carbonyl Z-Dienylation via Multicomponent Reductive Coupling of Acetylene to Aldehydes and α-Ketoesters Mediated by Hydrogen: Carbonyl Insertion into Cationic RhodacyclopentadienesKong, Jong Rock; Krische, Michael J.Journal of the American Chemical Society (2006), 128 (50), 16040-16041CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)Exposure of aldehydes or α-ketoesters to equal vols. of acetylene and hydrogen gas at ambient temp. and pressure in the presence of cationic rhodium catalysts provides products of carbonyl Z-butadienylation, which arise via multicomponent coupling of four mols.: two mols. of acetylene, a mol. of vicinal dicarbonyl compd., and a mol. of elemental hydrogen. The collective data suggest a catalytic mechanism involving carbonyl insertion into a cationic rhodacyclopentadiene intermediate derived via oxidative dimerization of acetylene. Hydrogenolytic cleavage of the resulting oxarhodacycloheptadiene via formal σ-bond metathesis provides the product of carbonyl addn. and cationic rhodium(I) to close the catalytic cycle. Studies involving the hydrogenation of 1,6-diyne 14a (I) in the presence of α-ketoester 6a (II) corroborate the proposed catalytic mechanism. These multicomponent couplings represent the first use of acetylene gas, a basic chem. feedstock, in metal-catalyzed reductive C-C bond formation.
- 5
The reaction scope was later expanded by inclusion of other aldehydes and imines as coupling partners. See:
(a) Skucas, E.; Kong, J. R.; Krische, M. J. J. Am. Chem. Soc. 2007, 129, 7242– 72435ahttps://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2sXlsVKjtL4%253D&md5=a7e874a53cbb7c68ca588ab664b01ab6Enantioselective Reductive Coupling of Acetylene to N-Arylsulfonyl Imines via Rhodium Catalyzed C-C Bond-Forming Hydrogenation: (Z)-Dienyl Allylic AminesSkucas, Eduardas; Kong, Jong Rock; Krische, Michael J.Journal of the American Chemical Society (2007), 129 (23), 7242-7243CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)The first highly enantioselective catalytic vinylation of aldimines to furnish allylic amines is reported. Exposure of arom. and aliph. N-arylsulfonyl aldimines to equal vols. of acetylene and hydrogen gas at 45 °C and ambient pressure in the presence of chirally modified cationic rhodium catalysts provides the (Z)-dienyl allylic amines, e.g., I, in highly optically enriched form (93-98% ee) and as single geometrical isomers (>95:5, Z/E). The coupling products arise via multicomponent coupling of four mols.: two mols. of acetylene, a mol. of aldimine, and elemental hydrogen. Unlike other imine addns. involving nonstabilized carbanions, the present protocol circumvents use of preformed organometallic reagents.(b) Han, S. B.; Kong, J. R.; Krische, M. J. Org. Lett. 2008, 10, 4133– 4135There is no corresponding record for this reference. - 6Toyoshima, T.; Miura, T.; Murakami, M. Angew. Chem., Int. Ed. 2011, 50, 10436– 104396https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3MXhtFOlsLnF&md5=671aecf04716fd7642c2c48cb11cad4eSelective 1:2 Coupling of Aldehydes and Allenes with Control of RegiochemistryToyoshima, Takeharu; Miura, Tomoya; Murakami, MasahiroAngewandte Chemie, International Edition (2011), 50 (44), 10436-10439, S10436/1-S10436/72CODEN: ACIEF5; ISSN:1433-7851. (Wiley-VCH Verlag GmbH & Co. KGaA)A new rhodium-catalyzed coupling reaction of one mol. of aldehyde and two mol. of allene was developed and gave selectively either of two constitutional isomers of β,γ-dialkylidene ketones, e.g. I and II. Interestingly, the regioselectivity of the reaction depends on the counterion of a rhodium(I) complex.
- 7(a) Oonishi, Y.; Hosotani, A.; Sato, Y. J. Am. Chem. Soc. 2011, 133, 10386– 103897ahttps://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3MXns1enu70%253D&md5=71bb3766dfa4cc0ebc76992057e5c9d9Rh(I)-Catalyzed Formal [6 + 2] Cycloaddition of 4-Allenals with Alkynes or Alkenes in a TetherOonishi, Yoshihiro; Hosotani, Akihito; Sato, YoshihiroJournal of the American Chemical Society (2011), 133 (27), 10386-10389CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)Rh(I)-catalyzed formal [6 + 2] cycloaddn. of allenal, e.g., I, having an alkyne or alkene in a tether proceeded smoothly, giving fused bicyclic ketone derivs., e.g., II, in good to excellent yields. It was also found that cyclization of enantiomerically enriched I (94% ee) gave cyclic ketone deriv. II in high yield with reasonable chirality transfer (86% ee). This result indicates that this cyclization proceeds through stereoselective formation of rhodacycle H' followed by insertion of a multiple bond.(b) Hojo, D.; Tanaka, K. Org. Lett. 2012, 14, 1492– 14957bhttps://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC38XivVOht7o%253D&md5=b2c27512626dc31fb7185f1ced643811Rhodium-Catalyzed C-H Bond Activation/[4+2] Annulation/Aromatization Cascade To Produce Phenol, Naphthol, Phenanthrenol, and Triphenylenol DerivativesHojo, Daiki; Tanaka, KenOrganic Letters (2012), 14 (6), 1492-1495CODEN: ORLEF7; ISSN:1523-7052. (American Chemical Society)It has been established that a cationic rhodium(I)/dppp complex catalyzes the aldehyde C-H bond activation/[4 + 2] annulation/aromatization cascade to produce phenol, naphthol, phenanthrenol, and triphenylenol derivs. from readily available conjugated alkynyl aldehydes and alkynes.
- 8
The total of 36 refers only to the number of possible linear compounds containing the reactant molecules coupled in a aldehyde–allene–allene sequence. In fact, two of the above isomers contain two asymmetric carbon atoms and hence can exist as two diastereomers, raising the number of possible products with distinct energies to 38. If cyclic structures and those containing allene–aldehyde–allene sequence are also considered, the number of possible products is even greater.
There is no corresponding record for this reference. - 9Williams, V. M.; Kong, J. R.; Ko, B. J.; Mantri, Y.; Brodbelt, J. S.; Baik, M.-H.; Krische, M. J. J. Am. Chem. Soc. 2009, 131, 16054– 16062There is no corresponding record for this reference.
- 10Miura, T.; Biyajima, T.; Toyoshima, T.; Murakami, M. Beilstein J. Org. Chem. 2011, 7, 578– 58110https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3MXmsV2itb4%253D&md5=6ed76f458f528c62436bb4c886c73d5dSynthesis of cross-conjugated trienes by rhodium-catalyzed dimerization of monosubstituted allenesMiura, Tomoya; Biyajima, Tsuneaki; Toyoshima, Takeharu; Murakami, MasahiroBeilstein Journal of Organic Chemistry (2011), 7 (), 578-581, No. 67CODEN: BJOCBH; ISSN:1860-5397. (Beilstein-Institut zur Foerderung der Chemischen Wissenschaften)A rhodium(I)/dppe catalyst promoted dimerization of monosubstituted allenes in a stereoselective manner to give cross-conjugated trienes, which are different from those obtained by a palladium catalyst. A palladium-catalyzed reaction of 1,2-undecadiene delivered the previously reported 11-methyl-10-methylene-8,11-eicosadiene and a new product [i.e., (8E)-10,11-bis(methylene)-8-eicosene]. [4+2] Cycloaddn. products of (8E)-10,11-bis(methylene)-8-eicosene with 1,1,2,2-ethenetetracarbonitrile and 4-phenyl-3H-1,2,4-triazole-3,5(4H)-dione were reported.
- 11
In fact, to the best of our knowledge, all of the transition-metal-catalyzed [n + m + o] cycloadditions involving allenes developed to date incorporate only one or two molecules of allene into the product. See:
(a) Murakami, M.; Ubukata, M.; Itami, K.; Ito, Y. Angew. Chem., Int. Ed. 1998, 37, 2248– 2250There is no corresponding record for this reference.(b) Shanmugasundaram, M.; Wu, M.-S.; Cheng, C.-H. Org. Lett. 2001, 3, 4233– 4236There is no corresponding record for this reference.(c) Shanmugasundaram, M.; Wu, M.-S.; Jeganmohan, M.; Huang, C.-W.; Cheng, C.-H. J. Org. Chem. 2002, 67, 7724– 7729There is no corresponding record for this reference.(d) Miura, T.; Morimoto, M.; Murakami, M. J. Am. Chem. Soc. 2010, 132, 15836– 15838There is no corresponding record for this reference.(e) Brusoe, A. T.; Alexanian, E. J. Angew. Chem., Int. Ed. 2011, 50, 6596– 6600There is no corresponding record for this reference. - 12(a) McCarren, P. R.; Liu, P.; Cheong, P. H.-Y.; Jamison, T. F.; Houk, K. N. J. Am. Chem. Soc. 2009, 131, 6654– 6655There is no corresponding record for this reference.(b) Liu, P.; McCarren, P.; Cheong, P. H.-Y.; Jamison, T. F.; Houk, K. N. J. Am. Chem. Soc. 2010, 132, 2050– 205712bhttps://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3cXptlKjtQ%253D%253D&md5=85aa276762e5d3f13c963e993c1e1be9Origins of Regioselectivity and Alkene-Directing Effects in Nickel-Catalyzed Reductive Couplings of Alkynes and AldehydesLiu, Peng; McCarren, Patrick; Cheong, Paul Ha-Yeon; Jamison, Timothy F.; Houk, K. N.Journal of the American Chemical Society (2010), 132 (6), 2050-2057CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)The origins of reactivity and regioselectivity in nickel-catalyzed reductive coupling reactions of alkynes and aldehydes were investigated with d. functional calcns. The regioselectivities of reactions of simple alkynes are controlled by steric effects, while conjugated enynes and diynes are predicted to have increased reactivity and very high regioselectivities, placing alkenyl or alkynyl groups distal to the forming C-C bond. The reactions of enynes and diynes involve 1,4-attack of the Ni-carbonyl complex on the conjugated enyne or diyne. The consequences of these conclusions on reaction design are discussed.(c) Liu, P.; Krische, M. J.; Houk, K. N. Chem.—Eur. J. 2011, 17, 4021– 402912chttps://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3MXjsVGgt7s%253D&md5=8c3bd7ae34ef3c5a3ef62ca1c4b6e573Mechanism and Origins of Regio- and Enantioselectivities in RhI-Catalyzed Hydrogenative Couplings of 1,3-Diynes and Activated Carbonyl Partners: Intervention of a Cumulene IntermediateLiu, Peng; Krische, Michael J.; Houk, Kendall N.Chemistry - A European Journal (2011), 17 (14), 4021-4029, S4021/1-S4021/24CODEN: CEUJED; ISSN:0947-6539. (Wiley-VCH Verlag GmbH & Co. KGaA)The mechanism of the rhodium-catalyzed reductive coupling of 1,3-diynes and vicinal dicarbonyl compds. employing H2 as reductant was studied by d. functional theory. Oxidative coupling through 1,4-addn. of the RhI-bound dicarbonyl to the conjugated diyne via a seven-membered cyclic cumulene transition state leads to exclusive formation of linear adducts. Diyne 1,4-addn. is much faster than the 1,2-addn. to simple alkynes. The 1,2-dicarbonyl compd. is bound to rhodium in a bidentate fashion during the oxidative coupling event. The chemo-, regio-, and enantioselectivities of this reaction were studied and are attributed to this unique 1,4-addn. pathway. The close proximity of the ligand and the alkyne substituent distal to the forming C-C bond controls the regio- and enantioselectivity: coupling occurs at the sterically more demanding alkyne terminus, which minimizes nonbonded interaction with the ligand. A stereochem. model is proposed that accounts for preferential formation of the (R)-configurated coupling product when (R)-biaryl phosphine ligands are used.
- 13Frisch, M. J.; Gaussian 03, revision D.01; Gaussian, Inc.: Wallingford, CT, 2004.There is no corresponding record for this reference.
- 14(a) Becke, A. D. J. Chem. Phys. 1993, 98, 5648– 565214ahttps://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK3sXisVWgtrw%253D&md5=291bbfc119095338bb1624f0c21c7ca8Density-functional thermochemistry. III. The role of exact exchangeBecke, Axel D.Journal of Chemical Physics (1993), 98 (7), 5648-52CODEN: JCPSA6; ISSN:0021-9606.Despite the remarkable thermochem. accuracy of Kohn-Sham d.-functional theories with gradient corrections for exchange-correlation, the author believes that further improvements are unlikely unless exact-exchange information is considered. Arguments to support this view are presented, and a semiempirical exchange-correlation functional (contg. local-spin-d., gradient, and exact-exchange terms) is tested for 56 atomization energies, 42 ionization potentials, 8 proton affinities, and 10 total at. energies of first- and second-row systems. This functional performs better than previous functionals with gradient corrections only, and fits expt. atomization energies with an impressively small av. abs. deviation of 2.4 kcal/mol.(b) Lee, C. T.; Yang, W. T.; Parr, R. G. Phys. Rev. B 1988, 37, 785– 78914bhttps://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaL1cXktFWrtbw%253D&md5=ee7b59267a2ff72e15171a481819ccf8Development of the Colle-Salvetti correlation-energy formula into a functional of the electron densityLee, Chengteh; Yang, Weitao; Parr, Robert G.Physical Review B: Condensed Matter and Materials Physics (1988), 37 (2), 785-9CODEN: PRBMDO; ISSN:0163-1829.A correlation-energy formula due to R. Colle and D. Salvetti (1975), in which the correlation energy d. is expressed in terms of the electron d. and a Laplacian of the 2nd-order Hartree-Fock d. matrix, is restated as a formula involving the d. and local kinetic-energy d. On insertion of gradient expansions for the local kinetic-energy d., d.-functional formulas for the correlation energy and correlation potential are then obtained. Through numerical calcns. on a no. of atoms, pos. ions, and mols., of both open- and closed-shell type, it is demonstrated that these formulas, like the original Colle-Salvetti formulas, give correlation energies within a few percent.
- 15Hay, P. J.; Wadt, W. R. J. Chem. Phys. 1985, 82, 270– 28315https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaL2MXhtlyju70%253D&md5=29271d2a54b5c81acd19762c570e64d7Ab initio effective core potentials for molecular calculations. Potentials for the transition metal atoms scandium to mercuryHay, P. Jeffrey; Wadt, Willard R.Journal of Chemical Physics (1985), 82 (1), 270-83CODEN: JCPSA6; ISSN:0021-9606.Ab initio effective core potentials (ECP's) were generated to replace the Coulomb, exchange, and core-orthogonality effects of the chem. inert core electron in the transition metal atoms Sc to Hg. For the second and third transition series relative ECP's were generated which also incorporate the mass-velocity and Darwin relativistic effects into the potential. The ab initio ECP's should facilitate valence electron calcns. on mols. contg. transition-metal atoms with accuracies approaching all-electron calcns. at a fraction of the computational cost. Analytic fits to the potentials are presented for use in multicenter integral evaluation. Gaussian orbital valence basis sets are developed for the (3d,4s,4p), (4d,5s,5p), and (5d,6s,6p) orbitals of the first, second, and third transition series atoms, resp. All-electron and valence-electron at. excitation energies are also compared for the low-lying states of Sc-Hg, and the valence-electron calcns. reproduce the all-electron excitation energies (typically within a few tenths of an eV).
- 16(a) Barone, V.; Cossi, M. J. Phys. Chem. A 1998, 102, 1995– 200116ahttps://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK1cXht1Cgt7o%253D&md5=7fe7f5f4627f26fd16a34e25219efaa6Quantum Calculation of Molecular Energies and Energy Gradients in Solution by a Conductor Solvent ModelBarone, Vincenzo; Cossi, MaurizioJournal of Physical Chemistry A (1998), 102 (11), 1995-2001CODEN: JPCAFH; ISSN:1089-5639. (American Chemical Society)A new implementation of the conductor-like screening solvation model (COSMO) in the GAUSSIAN94 package is presented. It allows Hartree-Fock (HF), d. functional (DF) and post-HF energy, and HF and DF gradient calcns.: the cavities are modeled on the mol. shape, using recently optimized parameters, and both electrostatic and nonelectrostatic contributions to energies and gradients are considered. The calcd. solvation energies for 19 neutral mols. in water are found in very good agreement with exptl. data; the solvent-induced geometry relaxation is studied for some closed and open shell mols., at HF and DF levels. The computational times are very satisfying: the self-consistent energy evaluation needs a time 15-30% longer than the corresponding procedure in vacuo, whereas the calcn. of energy gradients is only 25% longer than in vacuo for medium size mols.(b) Cossi, M.; Rega, N.; Scalmani, G.; Barone, V. J. Comput. Chem. 2003, 24, 669– 68116bhttps://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD3sXivFWqsbc%253D&md5=570ef9f44e925c9f78de6c7d97123229Energies, structures, and electronic properties of molecules in solution with the C-PCM solvation modelCossi, Maurizio; Rega, Nadia; Scalmani, Giovanni; Barone, VincenzoJournal of Computational Chemistry (2003), 24 (6), 669-681CODEN: JCCHDD; ISSN:0192-8651. (John Wiley & Sons, Inc.)The conductor-like solvation model, as developed in the framework of the polarizable continuum model (PCM), has been reformulated and newly implemented in order to compute energies, geometric structures, harmonic frequencies, and electronic properties in soln. for any chem. system that can be studied in vacuo. Particular attention is devoted to large systems requiring suitable iterative algorithms to compute the solvation charges: the fast multipole method (FMM) has been extensively used to ensure a linear scaling of the computational times with the size of the solute. A no. of test applications are presented to evaluate the performances of the method.
- 17Grimme, S. J. Comput. Chem. 2006, 27, 1787– 179917https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD28XhtFenu7bO&md5=0b4aa16bebc3a0a2ec175d4b161ab0e4Semiempirical GGA-type density functional constructed with a long-range dispersion correctionGrimme, StefanJournal of Computational Chemistry (2006), 27 (15), 1787-1799CODEN: JCCHDD; ISSN:0192-8651. (John Wiley & Sons, Inc.)A new d. functional (DF) of the generalized gradient approxn. (GGA) type for general chem. applications termed B97-D is proposed. It is based on Becke's power-series ansatz from 1997 and is explicitly parameterized by including damped atom-pairwise dispersion corrections of the form C6·R-6. A general computational scheme for the parameters used in this correction has been established and parameters for elements up to xenon and a scaling factor for the dispersion part for several common d. functionals (BLYP, PBE, TPSS, B3LYP) are reported. The new functional is tested in comparison with other GGAs and the B3LYP hybrid functional on std. thermochem. benchmark sets, for 40 noncovalently bound complexes, including large stacked arom. mols. and group II element clusters, and for the computation of mol. geometries. Further cross-validation tests were performed for organometallic reactions and other difficult problems for std. functionals. In summary, it is found that B97-D belongs to one of the most accurate general purpose GGAs, reaching, for example for the G97/2 set of heat of formations, a mean abs. deviation of only 3.8 kcal mol-1. The performance for noncovalently bound systems including many pure van der Waals complexes is exceptionally good, reaching on the av. CCSD(T) accuracy. The basic strategy in the development to restrict the d. functional description to shorter electron correlation lengths scales and to describe situations with medium to large interat. distances by damped C6·R-6 terms seems to be very successful, as demonstrated for some notoriously difficult reactions. As an example, for the isomerization of larger branched to linear alkanes, B97-D is the only DF available that yields the right sign for the energy difference. From a practical point of view, the new functional seems to be quite robust and it is thus suggested as an efficient and accurate quantum chem. method for large systems where dispersion forces are of general importance.
- 18
For instance, see:
(a) Minenkov, Y.; Occhipinti, G.; Jensen, V. R. J. Phys. Chem. A 2009, 113, 11833– 11844There is no corresponding record for this reference.(b) Siegbahn, P. E. M.; Blomberg, M. R. A.; Chen, S.-L. J. Chem. Theory Comput. 2010, 6, 2040– 204418bhttps://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3cXntF2jtbc%253D&md5=5d6e81a233c545b03a68b0435229dc00Significant van der Waals Effects in Transition Metal ComplexesSiegbahn, Per E. M.; Blomberg, Margareta R. A.; Chen, Shi-LuJournal of Chemical Theory and Computation (2010), 6 (7), 2040-2044CODEN: JCTCCE; ISSN:1549-9618. (American Chemical Society)There is, in general, very good experience using hybrid DFT to study mechanisms of enzyme reactions contg. transition metals. For redox reactions, the B3LYP* functional, which has 15% exact exchange, has been shown to be particularly accurate. Still, there are some cases which have turned out to be quite difficult with large errors. In the present study, the effects of van der Waals interaction have been investigated for these cases, using the empirical formula of Grimme. The results are encouraging.(c) Harvey, J. N. Faraday Discuss. 2010, 145, 487– 505There is no corresponding record for this reference.(d) McMullin, C. L.; Jover, J.; Harvey, J. N.; Fey, N. Dalton Trans. 2010, 39, 10833– 1083618dhttps://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3cXhtl2gt77E&md5=51d676bc1051a68e20a6228a58dc0a7bAccurate modelling of Pd(0) + PhX oxidative addition kineticsMcMullin, Claire L.; Jover, Jesus; Harvey, Jeremy N.; Fey, NatalieDalton Transactions (2010), 39 (45), 10833-10836CODEN: DTARAF; ISSN:1477-9226. (Royal Society of Chemistry)We have used dispersion-cor. DFT (DFT-D) together with solvation to examine possible mechanisms for reaction of PhX (X = Cl, Br, I) with Pd(PtBu3)2 and compare our results to recently published kinetic data (F. Barrios-Landeros, B. P. Carrow and J. F. Hartwig, J. Am. Chem. Soc., 2009, 131, 8141-8154). The calcd. activation free energies agree near-quant. with exptl. obsd. rate consts.(e) Lonsdale, R.; Harvey, J. N.; Mulholland, A. J. J. Phys. Chem. Lett. 2010, 1, 3232– 323718ehttps://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3cXhtleitbvK&md5=9bc8d74cf3976d4da46535ec7d926182Inclusion of Dispersion Effects Significantly Improves Accuracy of Calculated Reaction Barriers for Cytochrome P450 Catalyzed ReactionsLonsdale, Richard; Harvey, Jeremy N.; Mulholland, Adrian J.Journal of Physical Chemistry Letters (2010), 1 (21), 3232-3237CODEN: JPCLCD; ISSN:1948-7185. (American Chemical Society)Prediction of cytochrome P 450 reactivity is of great importance to the development of new medicinal compds. D. functional theory (DFT) has proven itself as a useful tool in the characterization of the elusive reactive species, compd. I, and of the mechanisms of substrate oxidn. B3LYP is the most widely used d. functional in the study of P 450s; however, a major drawback of B3LYP is its inaccurate treatment of dispersion, leading to discrepancies between expt. and theory in some systems. Recent work has shown that an added empirical dispersion correction to B3LYP (B3LYP-D) yields more promising results for similar systems. In the present work, two previously studied systems, camphor hydroxylation and alkene oxidn., have been recalcd. using B3LYP-D. Our work shows that inclusion of dispersion has a significant effect on the energies and geometries of transition states and encounter complexes; furthermore, an improved agreement with exptl. data is obsd.(f) Osuna, S.; Swart, M.; Solà, M. J. Phys. Chem. A 2011, 115, 3491– 3496There is no corresponding record for this reference.(g) Santoro, S.; Liao, R.-Z.; Himo, F. J. Org. Chem. 2011, 76, 9246– 9252There is no corresponding record for this reference.(h) Nordin, M.; Liao, R.-Z.; Ahlford, K.; Adolfsson, H.; Himo, F. ChemCatChem 2012, 4, 1095– 1104There is no corresponding record for this reference.(i) Xu, X.; Liu, P.; Lesser, A.; Sirois, L. E.; Wender, P. A.; Houk, K. N. J. Am. Chem. Soc. 2012, 134, 11012– 1102518ihttps://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC38XnvF2mtr4%253D&md5=92a384dcb6fe790b625d62dec27b40e5Ligand Effects on Rates and Regioselectivities of Rh(I)-Catalyzed (5 + 2) Cycloadditions: A Computational Study of Cyclooctadiene and Dinaphthocyclooctatetraene as LigandsXu, Xiufang; Liu, Peng; Lesser, Adam; Sirois, Lauren E.; Wender, Paul A.; Houk, K. N.Journal of the American Chemical Society (2012), 134 (26), 11012-11025CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)The first theor. study on the effects of ligands on the mechanism, reactivities, and regioselectivities of Rh(I)-catalyzed (5 + 2) cycloaddns. of vinylcyclopropanes (VCPs) and alkynes has been performed using d. functional theory (DFT) calcns. Highly efficient and selective intermol. (5 + 2) cycloaddns. of VCPs and alkynes have been achieved recently using two novel rhodium catalysts, [Rh(dnCOT)]+SbF6- and [Rh(COD)]+SbF6-, which provide superior reactivities and regioselectivities relative to that of the previously reported [Rh(CO)2Cl]2 catalyst. Computationally, the high reactivities of the dnCOT and COD ligands are attributed to the steric repulsions that destabilize the Rh-product complex, the catalyst resting state in the catalytic cycle. The regioselectivities of reactions with various alkynes and different Rh catalysts are investigated, and a predictive model is provided that describes substrate-substrate and ligand-substrate steric repulsions, electronic effects, and noncovalent π/π and C-H/π interactions. In the reactions with dnCOT or COD ligands, the first new C-C bond is formed proximal to the bulky substituent on the alkyne to avoid ligand-substrate steric repulsions. This regioselectivity is reversed either by employing the smaller [Rh(CO)2Cl]2 catalyst to diminish the ligand-substrate repulsions or by using aryl alkynes, for which the ligand-substrate interactions become stabilizing due to π/π and C-H/π dispersion interactions. Electron-withdrawing groups on the alkyne prefer to be proximal to the first new C-C bond to maximize metal-substrate back-bonding interactions. These steric, electronic, and dispersion effects can all be utilized in designing new ligands to provide regiochem. control over product formation with high selectivities. The computational studies reveal the potential of employing the dnCOT family of ligands to achieve unique regiochem. control due to the steric influences and dispersion interactions assocd. with the rigid aryl substituents on the ligand.(j) Jiménez-Halla, J. O. C.; Kalek, M.; Stawinski, J.; Himo, F. Chem.—Eur. J. 2012, 18, 12424– 1243618jhttps://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC38XhtFOmtrbL&md5=d3ea852f2a6ceef323d0426288ace547Computational Study of the Mechanism and Selectivity of Palladium-Catalyzed Propargylic Substitution with Phosphorus NucleophilesJimenez-Halla, J. Oscar C.; Kalek, Marcin; Stawinski, Jacek; Himo, FahmiChemistry - A European Journal (2012), 18 (39), 12424-12436, S12424/1-S12424/155CODEN: CEUJED; ISSN:0947-6539. (Wiley-VCH Verlag GmbH & Co. KGaA)The mechanism and sources of selectivity in the palladium-catalyzed propargylic substitution reaction that involves phosphorus nucleophiles, and which yields predominantly allenylphosphonates and related compds., have been studied computationally by d. functional theory. Full free-energy profiles are computed for both H-phosphonate and H-phosphonothioate substrates. The calcns. show that the special behavior of H-phosphonates among other heteroatom nucleophiles is indeed reflected in higher energy barriers for the attack on the central carbon atom of the allenyl/propargyl ligand relative to the ligand-exchange pathway, which leads to the exptl. obsd. products. It is argued that, to explain the preference of allenyl- vs. propargyl-phosphonate/phosphonothioate formation in reactions that involve H-phosphonates and H-phosphonothioates, anal. of the complete free-energy surfaces is necessary, because the product ratio is detd. by different transition states in the resp. branches of the catalytic cycle. In addn., these transition states change in going from a H-phosphonate to a H-phosphonothioate nucleophile.(k) Kalek, M.; Himo, F. J. Am. Chem. Soc. 2012, 134, 19159– 19169There is no corresponding record for this reference.(l) Huang, G.; Xia, Y.; Sun, C.; Li, J.; Lee, D. J. Org. Chem. 2013, 78, 988– 995There is no corresponding record for this reference. - 19
For selected reports of η3-allyl Rh species, see:
(a) Evans, P. A.; Lawler, M. J. J. Am. Chem. Soc. 2004, 126, 8642– 8643There is no corresponding record for this reference.(b) Evans, P. A.; Leahy, D. K. Chemtracts 2003, 16, 567– 578There is no corresponding record for this reference.(c) Evans, P. A.; Leahy, D. K.; Slieker, L. M. Tetrahedron: Asymmetry 2003, 14, 3613– 3618There is no corresponding record for this reference.(d) Evans, P. A.; Robinson, J. E.; Moffett, K. K. Org. Lett. 2001, 3, 3269– 3271There is no corresponding record for this reference.(e) Arnold, J. S.; Cizio, G. T.; Nguyen, H. M. Org. Lett. 2011, 13, 5576– 5579There is no corresponding record for this reference.(f) Arnold, J. S.; Cizio, G. T.; Heitz, D. R.; Nguyen, H. M. Chem. Commun. 2012, 48, 11531– 11533There is no corresponding record for this reference.(g) Arnold, J. S.; Stone, R. F.; Nguyen, H. M. Org. Lett. 2010, 12, 4580– 4583There is no corresponding record for this reference.(h) Arnold, J. S.; Nguyen, H. M. J. Am. Chem. Soc. 2012, 134, 8380– 8383There is no corresponding record for this reference.(i) Hayashi, T.; Okada, A.; Suzuka, T.; Kawatsura, M. Org. Lett. 2003, 5, 1713– 1715There is no corresponding record for this reference.(j) Vrieze, D. C.; Hoge, G. S.; Hoerter, P. Z.; Van Haitsma, J. T.; Samas, B. M. Org. Lett. 2009, 11, 3140– 3142There is no corresponding record for this reference.(k) Koschker, P.; Lumbroso, A.; Breit, B. J. Am. Chem. Soc. 2011, 133, 20746– 20749There is no corresponding record for this reference.(l) Lumbroso, A.; Koschker, P.; Vautravers, N. R.; Breit, B. J. Am. Chem. Soc. 2011, 133, 2386– 238919lhttps://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3MXhsVOnsLc%253D&md5=d71b0c3b438ed69fac23c09a8bc16177Redox-neutral atom-economic rhodium-catalyzed coupling of terminal alkynes with carboxylic acids toward branched allylic estersLumbroso, Alexandre; Koschker, Philipp; Vautravers, Nicolas R.; Breit, BernhardJournal of the American Chemical Society (2011), 133 (8), 2386-2389CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)A new method for the prepn. of a wide range of branched allylic esters from terminal alkynes that proceeds via a redox-neutral propargylic CH activation employing a rhodium(I)/DPEphos catalyst is reported.(m) Nishimura, T.; Hirabayashi, S.; Yasuhara, Y.; Hayashi, T. J. Am. Chem. Soc. 2006, 128, 2556– 2557There is no corresponding record for this reference.(n) Choi, J.-c.; Osakada, K.; Yamamoto, T. Organometallics 1998, 17, 3044– 3050There is no corresponding record for this reference.(o) Barros, H. J. V.; Guimarães, C. C.; dos Santos, E. N.; Gusevskaya, E. V. Organometallics 2007, 26, 2211– 2218There is no corresponding record for this reference.(p) Jiao, L.; Lin, M.; Zhuo, L.-G.; Yu, Z.-X. Org. Lett. 2010, 12, 2528– 2531There is no corresponding record for this reference.(q) Jiao, L.; Lin, M.; Yu, Z.-X. J. Am. Chem. Soc. 2011, 133, 447– 461There is no corresponding record for this reference.(r) Lin, M.; Li, F.; Jiao, L.; Yu, Z.-X. J. Am. Chem. Soc. 2011, 133, 1690– 1693There is no corresponding record for this reference.(s) Lin, M.; Kang, G.-Y.; Guo, Y.-A.; Yu, Z.-X. J. Am. Chem. Soc. 2012, 134, 398– 405There is no corresponding record for this reference.(t) Li, Q.; Yu, Z.-X. Organometallics 2012, 31, 5185– 5195There is no corresponding record for this reference. - 20(a) Szabó, K. J. Chem.—Eur. J. 2004, 10, 5268– 527520ahttps://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2cXpsF2rtLk%253D&md5=d619dbf6a865b2ab693c7d7950c6ca76Palladium-catalyzed electrophilic allylation reactions via bis(allyl)palladium complexes and related intermediatesSzabo, Kalman J.Chemistry - A European Journal (2004), 10 (21), 5268-5275CODEN: CEUJED; ISSN:0947-6539. (Wiley-VCH Verlag GmbH & Co. KGaA)A review. The synthetic scope of the allyl-Pd chem. can be extended to involve electrophilic reagents. The greatest challenge in these reactions is the catalytic generation of an allyl-Pd intermediate incorporating a nucleophilic allyl moiety. A vast majority of the published reactions that involve Pd-catalyzed allylation of electrophiles proceed via bis(allyl)palladium intermediates. The η1-moiety of the bis(allyl)palladium intermediates reacts with electrophiles, including aldehydes, imines, or Michael acceptors. Recently, catalytic electrophilic allylations via monoallylpalladium complexes have also been presented by employment of so-called pincer complex catalysts.(b) Szabó, K. J. Chem.—Eur. J. 2000, 6, 4413– 442120bhttps://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD3cXovVyqs70%253D&md5=430af920ef49e31d8963558650e0a3aaUmpolung of the allylpalladium reactivity: mechanism and regioselectivity of the electrophilic attack on bis-allylpalladium complexes formed in palladium-catalyzed transformationsSzabo, Kalman J.Chemistry - A European Journal (2000), 6 (23), 4413-4421CODEN: CEUJED; ISSN:0947-6539. (Wiley-VCH Verlag GmbH)The structure and reactivity of various bis-allylpalladium complexes occurring as catalytic intermediates in important synthetic transformations were studied by applying d. functional theory at the B3PW91 (DA + P) level. η1,η3 Coordinated bis-allylpalladium complexes are readily formed from the corresponding η3,η3 complexes, esp. in the presence of π-acceptor phosphine ligands. The theor. calcns. indicate dσ → π* type hyperconjugative interactions occurring in the η1-coordinated allyl moiety of the η1,η3 coordinated complexes. These hyperconjugative interactions influence the structure of the complexes and dramatically increase the reactivity of the double bond in the η1-moiety. The DFT results indicate a remarkably low activation barrier for the electrophilic attack on the η1-allyl functionality. In bridged η1,η3 complexes, the electrophilic attack occurs with a very high regioselectivity, which can be explained from d-π type hyperconjugative interactions.
- 21
There is an additional possibility of a migratory insertion of a third molecule of allene instead of the aldehyde. Such a pathway would eventually lead to [2 + 2 + 2] trimerization of the allene. However, since experimentally allene 2 only dimerizes (see ref 10), this possibility was not considered computationally.
There is no corresponding record for this reference. - 22
For selected reviews of allylation of carbonyl compounds, see:
(a) Denmark, S. E.; Fu, J. Chem. Rev. 2003, 103, 2763– 279422ahttps://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD3sXltVGjs78%253D&md5=b9f7c57ce56327693b6c8be57a50cda6Catalytic enantioselective addition of allylic organometallic reagents to aldehydes and ketonesDenmark, Scott E.; Fu, JipingChemical Reviews (Washington, DC, United States) (2003), 103 (8), 2763-2793CODEN: CHREAY; ISSN:0009-2665. (American Chemical Society)A review discusses the enantioselective allylation of aldehydes and ketones with a variety of allylmetal reagents to provide nonracemic homoallylic alcs. The mechanism of and chiral Lewis acid catalysts for the addn. of allylic silanes, stannanes and boranes to aldehydes and ketones, catalytic enantioselective allylation reactions with allylic halides, Lewis base-catalyzed enantioselective allylation with allylic trichlorosilanes, and allenylation and propargylation of aldehydes are discussed in the review; a table of conditions and selectivities for the enantioselective allylation of a variety of substrates is also provided.(b) Kennedy, J. W. J.; Hall, D. G. Angew. Chem., Int. Ed. 2003, 42, 4732– 473922bhttps://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD3sXosFKntLg%253D&md5=9661db0988639ab2c9ef7ab6b04e91b6Recent advances in the activation of boron and silicon reagents for stereocontrolled allylation reactionsKennedy, Jason W. J.; Hall, Dennis G.Angewandte Chemie, International Edition (2003), 42 (39), 4732-4739CODEN: ACIEF5; ISSN:1433-7851. (Wiley-VCH Verlag GmbH & Co. KGaA)A review. Despite the popularity of boron and silicon allylation reagents in stereocontrolled synthesis, they suffer from a no. of inherent limitations that have slowed down their development as synthetic tools for nucleophilic addns. to carbonyl compds. and imine derivs. These limitations are the low reactivity and diastereoselectivity of allyl trialkylsilane reagents, and the lack of catalytic systems for the activation and substoichiometric control of enantioselectivity in the addns. of allyl boron reagents. To develop more efficient and general methods for the control of abs. stereochem. in the resulting homoallylic alcs., new approaches aimed at solving the problem of activation of allylic boron and silicon reagents are needed. This Minireview describes a no. of recent approaches that have been devised to address this problem.(c) Yu, C.-M.; Youn, J.; Jung, H.-K. Bull. Korean Chem. Soc. 2006, 27, 463– 47222chttps://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD28XkvVektr8%253D&md5=e940b908b85abcd16616c1858cef0085Regulation of stereoselectivity and reactivity in inter- and intramolecular allylic transfer reactionsYu, Chan-Mo; Youn, Jinsoup; Jung, Hee-KeumBulletin of the Korean Chemical Society (2006), 27 (4), 463-472CODEN: BKCSDE; ISSN:0253-2964. (Korean Chemical Society)A review. The prepn. of enantiomerically enriched homoallylic alcs. through asym. addn. of chiral allylic transfer reagents and allylating reagents with chiral catalysts to carbonyl functionalities represents an important chem. transformation. Excellent progress has been made over the past decade in the development and application of catalytic asym. allylic transfer reactions. In this account, our efforts for the various intermol. allylic transfer reactions such as allylation, propargylation, allenylation, and dienylation utilizing an accelerating strategy and sequential allylic transfer reactions to achieve multiple stereoselection, mainly using transition metal catalysts, are described.(d) Marek, I.; Sklute, G. Chem. Commun. 2007, 1683– 169122dhttps://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2sXksFSit70%253D&md5=6f75de6048df9ee00d842748c108b870Creation of quaternary stereocenters in carbonyl allylation reactionsMarek, Ilan; Sklute, GeniaChemical Communications (Cambridge, United Kingdom) (2007), (17), 1683-1691CODEN: CHCOFS; ISSN:1359-7345. (Royal Society of Chemistry)A review. Despite the advances in stereoselective carbonyl allylation reactions, the creation of quaternary stereocenters by the addn. of 3,3'-disubstituted allylmetals to aldehydes is still a challenging issue. This feature article describes the most powerful approaches that have been devised to address this problem. The application of allyl boronates, allyl(trichloro)silanes and allylzinc reagents are discussed.(e) Hall, D. G. Synlett 2007, 1644– 1655There is no corresponding record for this reference. - 23
For examples of [2 + 2 + 2] cyclization involving carbonyl compounds, see:
(a) Tsuchikama, K.; Yoshinami, Y.; Shibata, T. Synlett 2007, 1395– 1398There is no corresponding record for this reference.(b) Tanaka, K.; Otake, Y.; Wada, A.; Noguchi, K.; Hirano, M. Org. Lett. 2007, 9, 2203– 2206There is no corresponding record for this reference. - 24
Transition states connecting INT6 to INT7 and INT6′ to INT7′ could not be optimized. However, energy scans showed that the barriers are quite low, much lower than those for the direct reductive elimination (see the Supporting Information).
There is no corresponding record for this reference. - 25
The methyl group in the η3-allyl complexes INT7 and INT7′, contrary to the case of INT2 discussed above, can undergo a relocation from the anti position to the syn position. This could potentially lead to the formation of isomers of products 4 and 6 containing different configurations at one of the double bonds. However, the barriers for the syn/anti substituent exchange in INT7 and INT7′ were calculated to be much higher than those for the reductive elimination. Therefore, INT7 and INT7′ undergo the reductive elimination before they can isomerize. See the Supporting Information for details.
There is no corresponding record for this reference. - 26
The ratios were calculated using the Eyring equation. For example, the ratio of INT6 to INT6′ was calculated as
There is no corresponding record for this reference. - 27Wade, L. G., Jr. Organic Chemistry, 6th ed.; Prentice-Hall: Upper Saddle River, NJ, 2006.There is no corresponding record for this reference.
- 28Nielsen, R. J.; Goddard, W. A., III. J. Am. Chem. Soc. 2006, 128, 9651– 9660There is no corresponding record for this reference.
Supporting Information
Supporting Information
Complete ref 13, additional results not shown in the text, and Cartesian coordinates of all optimized structures discussed in the paper. This material is available free of charge via the Internet at http://pubs.acs.org.
Terms & Conditions
Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.