Abstract
Atom-efficient processes that occur via addition, redistribution or removal of hydrogen underlie many large volume industrial processes and pervade all segments of chemical industry. Although carbonyl addition is one of the oldest and most broadly utilized methods for C-C bond formation, the delivery of non-stabilized carbanions to carbonyl compounds has relied on premetalated reagents or metallic/organometallic reductants, which pose issues of safety and challenges vis-à-vis large volume implementation. Catalytic carbonyl reductive couplings promoted via hydrogenation, transfer hydrogenation and hydrogen auto-transfer allow abundant unsaturated hydrocarbons to serve as substitutes to organometallic reagents, enabling C-C bond formation in the absence of stoichiometric metals. This perspective (a) highlights past milestones in catalytic hydrogenation, hydrogen transfer and hydrogen auto-transfer, (b) summarizes current methods for catalytic enantioselective carbonyl reductive couplings, and (c) describes future opportunities based on the patterns of reactivity that animate transformations of this type.
Keywords: Hydrogenation, Transfer Hydrogenation, Hydrogen Auto-Transfer, Enantioselective, Carbonyl Addition, Ruthenium, Iridium
Graphical Abstract

I. Introduction and Historical Perspective
Beginning with Dobereiner’s lighter (1823),1 catalytic hydrogenation is both the Proteus and Prometheus of metal-catalysis (Figure 1). The hydrogenation of unsaturated organic compounds was first reported by Sabatier (1897).2,3 Sustaining half of all human life, the Haber-Bosch process (1905), the hydrogenation of atmospheric nitrogen to form ammonia, was developed shortly thereafter.4 The first homogenous hydrogenations were developed by Melvin Calvin (1938),5 who reduced 1,4-benzoquinone to 1,4-hydroquinone under an atmosphere of hydrogen using substoichiometric quantities of cuprous acetate in quinoline solvent. Mechanistic studies of homogenous hydrogenation were initiated by Halpern (1956),6 leading to the first homogenous hydrogenations of activated olefins (1961).7 This work, along with the observation of Vaska (1962) that IrCl(CO)(PPh3)2 and elemental hydrogen combine to form isolable dihydrides via “oxidative addition,”8 galvanized the conceptual foundations of catalytic hydrogenation. Finally, nearly 70 years after the discovery of heterogeneous hydrogenation by Sabatier, Wilkinson (1965) described the homogenous hydrogenations of unactivated alkenes using his eponymous rhodium catalyst, RhCl(PPh3)3.9 Empowered by advances in the synthesis and design of chiral phosphines,10 enantioselective rhodium-catalyzed hydrogenations were introduced by Knowles (1968).11 Taken together with parallel progress in transfer hydrogenation, especially ruthenium-catalyzed processes,12–14 a vast field of research was born. Catalytic hydrogenation is now applied across all segments of the chemical industry, with asymmetric hydrogenation superseding all other catalytic enantioselective methods for the manufacture of chiral compounds.14 For pharmaceutical ingredients, catalytic hydrogenation is estimated to encompass 14% of all GMP reactions,14e and patent analyses indicate hydrogenation catalysis has yet to reach its apex and will continue to grow.15
Figure 1.

Selected milestones in catalytic hydrogenation, hydrogen transfer and hydrogen auto-transfer.
As organic molecules are composed of carbon and hydrogen, atom-efficient C-C bond forming processes that occur via addition, redistribution or removal of hydrogen underlie many large volume industrial processes. The Fischer-Tropsch process (1922)16 and alkene hydroformylation (Co, 1938; Rh, 1968)17 represent two iconic hydrogen-mediated reductive C-C couplings. The principal source of hydrogen used in these reactions is derived from methane-steam reforming (1912),18 that is, hydrocarbon dehydrogenation. The Guerbet reaction (1899),19 the prototypical “borrowing hydrogen” process,20 converts lower alcohols to higher alcohols through rebounding hydrogen transfers that first generate carbonyl partners for aldol dehydration and then reduce the resulting enones (Figure 2, I). The ability to exploit alcohols as both carbonyl proelectrophiles and reducing agents is also exemplified by the Meerwein-Ponndorf-Verley reaction (1925),21 and the work of Adkins (1932)22 on heterogeneous nickel-catalyzed alcohol aminations, which occur through alcohol dehydrogenation-reductive amination (Figure 2, II). Homogenous metal catalysts for alcohol amination were subsequently reported by Grigg and Watanabe (1981).23 Borrowing hydrogen processes20 are finding ever-increasing use in commercial chemical synthesis,24,25 as formal substitution of hydroxyl groups through dehydrogenation-condensation-reduction enhances step-economy and minimizes waste generation by avoiding discrete oxidation-state changes.
Figure 2.

Carbonyl condensation via hydrogen-borrowing and carbonyl addition via hydrogenation, transfer hydrogenation and H2-auto-transfer.
Whereas borrowing hydrogen processes like the Guerbet reaction and alcohol amination operate through pathways involving carbonyl condensation,20 related carbonyl additions that occur via hydrogenation, transfer hydrogenation and hydrogen auto-transfer have been developed in the present authors laboratory (Figure 2).26 These carbonyl addition processes can occur by way of three general catalytic pathways (Figure 2, A-C).
Mechanism A: C=O/C=C Oxidative coupling followed by (transfer)hydrogenolysis of the resulting oxametalacycles.
Mechanism B: C-X Bond reductive cleavage promoted via (transfer) hydrogenation or hydrogen auto-transfer.
Mechanism C: C=C Hydrometalation promoted via (transfer) hydrogenation or hydrogen auto-transfer.
Such metal-catalyzed carbonyl reductive couplings complement classical methods for carbonyl addition as they exploit neutral π-unsaturated feedstocks as equivalents to premetalated C-nucleophiles and enable carbonyl addition from the alcohol oxidation level. In this perspective, carbonyl and imine additions that occur via metal-catalyzed hydrogenation, transfer hydrogenation and hydrogen auto-transfer are surveyed. Coverage is limited to enantioselective reactions involving formal addition of H2 across the π-unsaturated pronucleophile and the π-bond of the carbonyl (or imine) electrophile using low molecular weight feedstock reductants (hydrogen, 2-propanol or formic acid and its salts) or through hydrogen auto-transfer (processes in which alcohols serve dually as carbonyl proelectrophiles and hydrogen carriers). Transformations involving substrate/catalyst-directed asymmetric induction also are covered. Applications in natural product synthesis are not exhaustively described. As established by analysis of >9 million pharmaceutical industry patents, carbonyl addition (alongside cross-coupling) remains one of the most widely utilized methods for C-C bond formation.27 By combining the characteristics of catalytic hydrogenation and carbonyl addition, it is the authors’ hope to advance a broad, new class of environmentally benign catalytic enantioselective carbonyl addition for use in chemical synthesis.
II. Hydrogenation
The aldol reaction was discovered in late 18th century and is among the earliest examples of carbonyl additions.28 The development of catalytic enantioselective aldol additions29 remains a vibrant area of research as such methods are of great utility in type I polyketide total synthesis.30 Stereoselective aldol additions are typically conducted using preformed (metallo)enolates based on lithium,31 boron30a,32 and silicon.30f We have found that hydrogenation of vinyl ketones in the presence of aldehyde using cationic rhodium catalysts results in carbonyl addition, representing a novel, atom-efficient reductive variant of the aldol reaction (Scheme 1).33 Upon use of tri-2-furylphosphine34 as ligand, high levels of syn-diastereoselectivity are achieved,33e enabling asymmetric reductive aldol additions to N-Boc-α-aminoaldehydes.33g Notably, the enantiomeric purity of the α-aminoaldehyde is preserved under the neutral condition of hydrogen-mediated reductive coupling. True catalytic enantioselective variants were subsequently developed using a benzothiophene-substituted TADDOL-like phosphonite ligand.33h Remarkably, as illustrated in the total synthesis of the marine macrolide swinholide A, hydrogen-mediated reductive aldol addition occurs in a chemoselective manner in the presence of mono-olefins and conjugated dienes.35 The collective data suggest these reactions occur through pathways involving enone-carbonyl oxidative coupling (Figure 2, Mechanism A). For stereochemical models and a detailed discussion of the reaction mechanism, the reader is directed to the review literature.36 Cationic rhodium catalysts are required to suppress hydrogen oxidative addition37 and vacate a coordination site to permit simultaneous binding of enone and aldehyde at the square planar metal center.
Scheme 1.

Catalytic asymmetric reductive aldol reactions of methyl vinyl ketone via rhodium-catalyzed hydrogenation.
The C=O/C=C oxidative coupling-hydrogenolysis pathway for hydrogen-mediated reductive aldol addition was transferable to carbonyl additions of alkyne pronucleophiles (Scheme 2). Using cationic rhodium catalysts under an atmosphere of hydrogen gas, reductive cyclizations of acetylenic aldehydes occur efficiently with high levels of enantioselectivity (Scheme 2, eq. B).38 Intermolecular reductive coupling requires careful matching between the π-donicity of the metal catalyst and the π-acidity of the unsaturated reactant to achieve efficient C=X (X = O, NR)/C≡C oxidative coupling. For cationic rhodium catalysts, which are less strongly reducing, conjugated alkynes (1,3-enynes or 1,3-diynes) and vicinal dicarbonyl partners39 (or related imines40) are required (Scheme 2, eq. A, C, D). Isostructural iridium catalysts are more electron rich and are capable of promoting oxidative coupling between nonconjugated alkynes and unactivated carbonyl or imine partners (Scheme 2, eq. E).41 In hydrogen-mediated reductive couplings of acetylene, products of carbonyl or imine (Z)-dienylation are formed (Scheme 2, eq. F).42 Oxidative dimerization of acetylene is faster than C=O/C≡C oxidative coupling. As corroborated by detailed mechanistic studies,42d the resulting rhodacyclopentadiene inserts the carbonyl or imine partner to deliver oxa- or azarhodacycloheptadienes, which upon hydrogenolysis provide the products (Z)-dienylation. Remarkably, in all cases, conventional hydrogenation of the reaction products is not observed as the alkyne pronucleophiles are more π-acidic and, hence, better ligands for the metal than the olefin containing adducts. Total syntheses of several natural products were accomplished using C-C bond forming hydrogenations of this type. For example, the anti-tumor promoting marine macrolide bryostatin and several of its analogues were prepared via asymmetric hydrogen-mediated enyne-α-ketoaldehyde reductive coupling (Scheme 2, Bottom).43
Scheme 2.

Catalytic asymmetric alkyne-carbonyl reductive coupling via rhodium-catalyzed hydrogenation.
III.A. Transfer Hydrogenation and Hydrogen Auto-Transfer (Pathways Involving C-X Bond Cleavage)
Following Hoffmann’s seminal report of asymmetric carbonyl allylation using chiral diol-modified allylboron reagents (1978),44a the first catalytic enantioselective carbonyl allylations were reported by Yamamoto (1991).44b As documented in the review literature, the field of catalytic enantioselective carbonyl allylation has since experienced enormous growth.45,46 However, notwithstanding protocols developed in the present authors laboratory,26h all reported catalytic enantioselective carbonyl allylations rely on preformed allylmetal reagents45 or, as exemplified by the Nozaki-Hiyama-Kishi allylations,46,47 metallic reductants. Mechanistic pathways in which alcohol dehydrogenation is balanced by reductive C-X bond cleavage enable generation of transient allylmetal nucleophiles (Figure 2, Mechanism B). Based on this concept, diverse catalytic enantioselective carbonyl allylations were developed (Scheme 3, Top).48 These processes are catalyzed by π-allyliridium-C,O-benzoate complexes discovered in our laboratory and are unique in that asymmetric allylation is achieved from the alcohol or aldehyde oxidation level in the absence of stoichiometric metals. The indicated catalytic cycle and stereochemical model are based on DFT calculations (Scheme 3, Bottom).49 The iridium center of the π-allyliridium-C,O-benzoate is itself stereogenic, which provides an especially well-defined chiral environment for carbonyl addition (Scheme 3).
Scheme 3.

Enantioselective π-allyliridium-C,O-benzoate-catalyzed carbonyl allylations that occur via alcohol-mediated C-X reductive cleavage, reaction mechanism and stereochemical model. Levels of asymmetric induction typically exceed 90% ee.
Among the unique capabilities associated with π-allyliridium-C,O-benzoate-catalyzed enantioselective allylations, the capacity to engage diols and higher polyols in site-selective asymmetric allylation is especially powerful, as it precludes the use of hydroxyl protecting groups (Scheme 4, Left).48f,g,v,w Such transformations are made possible by a pronounced kinetic (and contra-thermodynamic) preference for primary alcohol dehydrogenation. Additionally, chiral β-stereogenic alcohols are subject to asymmetric allylation with good levels of catalyst-directed diastereoselectivity, which avoids the handling of
Scheme 4.

Catalyst-directed diastereoselectivity in asymmetric allylations of 1,3-diols and two-directional allylation and crotylation of 1,3-diols and use of structurally complex nitrogen-rich allyl pronucleophiles.
configurationally labile chiral α-stereogenic aldehydes (Scheme 4, Middle).48e,f Finally, whereas malondialdehydes are highly intractable, 1,3-diols are abundant chemical feedstocks and are subject to asymmetric double allylation48c or double crotylation (Scheme 4, Top Right).48n The products of such two-directional chain syntheses are formed as single enantiomers due to the Horeau effect,50 and have been exploited as triketide building blocks in numerous total syntheses of polyketide natural products where they have been shown to significantly enhance step-economy.35,43b,51,52
The functional group compatibility of the π-allyliridium-C,O-benzoate-catalysts is illustrated in reactions of nitrogen-rich allylic acetates incorporating the top 10 N-heterocycles found in FDA-approved drugs (eq. 1).48b′ Allyl pronucleophiles of hitherto unattainable structural complexity are used as limiting reagents in C-C couplings of ethanol, the largest volume renewable small molecule feedstock (>85 million tons/year worldwide). The use of ethanol as a C2-feedstock in enantioselective C-C coupling was previously unknown.
![]() |
(eq. 1) |
Hydrogen auto-transfer mechanisms also can occur via internal redox processes wherein C-X bond reductive cleavage is balanced by oxidative esterification, as demonstrated in enantioselective allylations of o-phthalaldehydes (eq. 2).53 As the iridium catalyst is highly sensitive to rather modest steric and electronic perturbations, the indicated nonsymmetric o-phthalaldehyde can be engaged in site-selective internal redox allylation. Using this method, formal syntheses of ent-spirolaxine methyl ether and CJ-12,954 were achieved.
![]() |
(eq. 2) |
Enantioselective carbonyl propargylations comprise another major class of carbonyl additions that traditionally have relied on the use of premetalated reagents or metallic reductants.54 In a departure from prior art, it was found that iridium catalysts modified by SEGPHOS enable alcohol-mediated reductive C-Cl bond cleavage of trialkylsilyl substituted propargyl chlorides to furnish transient allenyliridium nucleophiles that participate in enantioselective carbonyl propargylation from alcohol or aldehyde oxidation level (Scheme 5).55a 3-Hexyne was added to suppress competing transfer hydrogenation of the alkyne moiety of the homopropargyl alcohol products. Using rhodium catalysts modified by (R)-BINAP, the parent propargyl chloride could be deployed in asymmetric carbonyl additions to chiral α-stereogenic amino alcohol proelectrophiles to furnish products of propargylation with good to complete levels of diastereocontrol (Scheme 5).55b
Scheme 5.

Catalytic enantioselective carbonyl propargylation via alcohol-mediated C-X reductive cleavage.
III.B. Transfer Hydrogenation and Hydrogen Auto-Transfer (Hydrometalative Pathways)
The activation of π-unsaturated pronucleophiles via alcohol-mediated hydrometalation represents another pathway for catalytic enantioselective carbonyl reductive coupling (Figure 2, Mechanism C). Neutral cyclometallated iridium(III) complexes have proven especially effective in reactions of this type. For example, as illustrated in reactions of allene gas (Scheme 6, eq. A),49 (N-phthalimido)allene (Scheme 6, eq. B)56 and dimethylallene (Scheme 6, eq. C),57 metal hydrides that arise in the course of π-allyliridium-C,O-benzoate-mediated alcohol dehydrogenation can be intercepted via allene hydrometalation to generate transient allyliridium nucleophiles, directly converting primary alcohol reactants to (branched) secondary homoallylic alcohols in the absence of stoichiometric byproducts. Remarkably, the cyclometallated iridium complex derived from PhanePhos catalyzes highly regio-, diastereo- and enantioselective carbonyl reductive couplings of methanol with dienes (Scheme 6, eq. D)58a and CF3-allenes (Scheme 6, eq. E)58b to form primary neopentyl alcohols bearing acyclic quaternary carbon stereocenters.59 The cyclometallated iridium-PhanePhos complex is also competent in 2-propanol-mediated reductive couplings of 1,1-disubstituted allene with fluoral,58c which are conducted in aqueous organic media (Scheme 6, eq. F).
Scheme 6.

Enantioselective carbonyl allylations via allene or diene hydrometalation catalyzed by cyclometallated iridium complexes. Levels of asymmetric induction typically exceed 90% ee.
Non-cyclometallated iridium and ruthenium complexes modified by (DM)-SEGPHOS or BINAP, respectively, are also competent catalysts for enantioselective carbonyl reductive coupling, as illustrated in enyne-mediated carbonyl propargylations (Scheme 7).60 In these processes, alcohol-mediated enyne hydrometalation delivers allenylmetal nucleophiles, which engage in carbonyl addition through six-centered transition structures. For the indicated iridium-catalyzed propargylation, axial chirality of the chiral ligand is relayed to the allenyliridium intermediate and the central chirality of the product. This method was recently used in the total synthesis and structural assignment of formosalides A and B.61
Scheme 7.

Enantioselective carbonyl propargylations that occur via enyne hydrometalation catalyzed by non-cyclometallated iridium and ruthenium complexes.
Hydrometalative activation of diene pronucleophiles via alcohol-mediated hydrogen (auto)transfer underlies novel ruthenium-catalyzed carbonyl crotylation protocols (Scheme 8).62 Trialkylsilyl substituents at the C2-position of the butadiene define the geometry of the transient crotylruthenium nucleophile, which, in turn, directs syn-diastereoselectivity.62a As initially observed in our laboratory in 2006,39c chiral phosphate counterions are effective stereocontrol elements metal-catalyzed C-C coupling. Using a ruthenium complex modified by the indicated BINOL-derived phosphate counterion, anti-diastereo- and enantioselective carbonyl crotylation could be achieved from the alcohol or aldehyde oxidation level using butadiene, a chemical feedstock (12 × 106 tons/year), as pronucleophile.62b Less Lewis basic BINOL-phosphate counterions form non-contact ion pairs with the ruthenium catalyst. In contrast, as borne out by DFT calculations62d and x-ray diffraction data, more Lewis basic TADDOL-phosphate counterions form contact ion pairs, permitting preservation of kinetic selectivity in the hydrometalation of the s-cis-diene conformer. The resulting (Z)-crotylruthenium intermediate engages in carbonyl addition faster than relaxation to the (E)-crotylruthenium isomer, enabling syn-diastereo- and enantioselective carbonyl crotylation to occur.62c,d
Scheme 8.

Enantioselective ruthenium- catalyzed carbonyl crotylations that occur via diene hydrometalation.
Alkyne pronucleophiles are subject to a dual catalytic process in which metal-catalyzed alkyne-to-allene isomerization is followed by a second catalytic cycle involving allene-carbonyl reductive coupling (Scheme 9, Top).63a,b A series of deuterium labelling experiments corroborate a mechanism involving allene hydrometalation to deliver a nucleophilic allylruthenium intermediate. Remarkably, propargyl ethers react through an entirely different and, to our knowledge, previously undescribed mechanism involving hydride shift enabled π-allyl formation (Scheme 9, Middle).63c Here, alkyne coordination triggers the generation of a vinylruthenium carbene species, which upon protonation forms an allylruthenium nucleophile. This mechanism also could be enacted using an iridium catalyst modified by H8-BINAP to deliver products of (siloxy)allylation with complete levels of alkene (Z)-stereoselectivity and high levels of enantioselectivity (Scheme 9, Bottom).63d Hydride shift enabled π-allyl formation does not involve hydrometalation and, mechanistically, represents a fourth distinct class of dehydrogenative carbonyl additions. These transformations demonstrate that tractable, commercially available alkynes can serve as synthetic equivalents to allylmetal nucleophiles.64
Scheme 9.

Enantioselective carbonyl allylations and crotylations via alkyne pronucleophiles catalyzed by (non-cyclometallated) iridium and ruthenium complexes.
IV. Future Opportunities and Objectives
This emerging class of catalytic C-C couplings presents numerous unmet challenges (Scheme 10). The development of dehydrogenative imine additions of π-unsaturated pronucleophiles via hydrogen auto-transfer of amine proelectrophiles (so-called “hydroaminoalkylations”) has largely focused on early transition metal catalysts.65 With only three exceptions,66,67 corresponding late transition metal-catalyzed variants have required non-native directing groups.68,69 Metal-catalyzed intermolecular reductive couplings of unactivated olefins with unactivated aldehydes using feedstock reductants have proven elusive, although promising strategies have emerged.26g,70 While intra- and intermolecular hydrogenative and transfer hydrogenative carbonyl reductive couplings of aryl halides have been accomplished,71 enantioselective variants, the use of alkyl halide pronucleophiles and hydrogen auto-transfer processes have not been described. Additionally, despite truly impressive progress on the reductive carboxylation of π-unsaturated reactants with CO2 mediated by silane,72,73 the use of inexpensive feedstock reductants in such processes has not been described (although photochemical promotion may unlock this challenge).73b
Scheme 10.

Unmet challenges in dehydrogenative carbonyl and imine addition of alcohol and amine proelectrophiles.
Beyond the transformations discussed in this perspective, there exist numerous other classical carbanionic and metal-mediated processes that could potentially be accomplished via hydrogenation, transfer hydrogenation or hydrogen auto-transfer, potentially availing more scalable, sustainable and cost-effective methods for chemical manufacture.74 Indeed, progress toward transfer hydrogenative pinacol reactions75 and cross-electrophile reductive couplings76 has been described. Recent advances in catalytic asymmetric photoreduction might offer new approaches to the development of more benign reductive C-C coupling protocols.77 Finally, the new patterns of reactivity described in this perspective can inform processes that have no counterpart in the lexicon of classical carbanion chemistry. For example, a broad, new family of ruthenium(0)-catalyzed cycloadditions of vicinal diols, ketols or diones with π-unsaturated partners has been developed based on transfer hydrogenative carbonyl addition (Scheme 11).78 It is the authors’ hope that this perspective will encourage further research into catalytic C-C bond formations that occur through the addition or redistribution of hydrogen.
Scheme 11.

Catalytic transformations beyond the classical lexicon of carbonyl addition chemistry.
ACKNOWLEDGMENT
The Robert A. Welch Foundation (F-0038) and the NIH-NIGMS (RO1-GM069445) are acknowledged for partial support of this research.
Footnotes
ASSOCIATED CONTENT
Supporting Information. For supporting information associated with this perspective article, please see the primary literature.
The authors declare no competing financial interest.
REFERENCES
- (1).(a) Hoffman R Dobereiner’s Lighter. American Scientist 1998, 86, 326–329. [Google Scholar]; (b) Williams WD Dobereiner’s Hydrogen Lighter. Bull. Hist. Chem 1999, 24, 66–68. [Google Scholar]
- (2).Sabatier P; Senderens J-BCR Action du Nickel sur l’Éthylène. Synthèse de l’Éthane. Hebd. Seances Acad. Sci 1897, 124, 1358–1361. [Google Scholar]
- (3). Prior to Sabatier’s seminal work, the catalytic hydrogenation of vegetable oils was supposedly reported by James F. Boyce. However, to our knowledge, the earliest written record of Boyce’s work appears in a patent from 1913: Boyce, J. Process of Producing an Edible Compound. U.S. Patent 1061254, May 6, 1913.
- (4).Smil V Enriching the Earth: Fritz Haber, Carl Bosch, and the Transformation of World Food Production, MIT Press: Cambridge, MA, 2004; pp 68–107. [Google Scholar]
- (5).(a) Calvin M Homogeneous Catalytic Hydrogenation. Trans. Far. Soc 1938, 34, 1181–1191. [Google Scholar]; (b) Calvin M Homogeneous Catalytic Hydrogenation. J. Am. Chem. Soc 1939, 61, 2230–2234. [Google Scholar]
- (6).Halpern J Homogeneous Reactions of Molecular Hydrogen in Solution. Q. Rev. Chem. Soc 1956, 10, 463–479. [Google Scholar]
- (7).Halpern J; Harrod JF; James BR Homogeneous Catalytic Hydrogenation of Olefinic Compounds. J. Am. Chem. Soc 1961, 83, 753–754. [Google Scholar]
- (8).Vaska L; DiLuzio JW Activation of Hydrogen by a Transition Metal Complex at Normal Conditions Leading to a Stable Molecular Dihydride. J. Am. Chem. Soc 1962, 84, 679–680. [Google Scholar]
- (9).(a) Young JF; Osborn JA; Jardine FH; Wilkinson G Hydride Intermediates in Homogeneous Hydrogenation Reactions of Olefins and Acetylenes using Rhodium Catalysts. Chem. Commun 1965, 131–132.; (b) Jardine FH; Osborn JA; Wilkinson G; Young JF Homogeneous Catalytic Hydrogenation and Hydroformylation of Acetylenic Compounds. Chem. Ind 1965, 560–561.5825541
- (10).(a) Horner L; Winkler H; Rapp A; Mentrup A; Hoffmann H; Beck P Phosphororganische Verbindungen Optisch Aktive Tertiare Phosphine aus Optisch Aktiven Quartaren Phosphoniumsalzen. Tetrahedron Lett 1961, 2, 161–166. [Google Scholar]; (b) Korpiun O; Mislow K New Route to the Preparation and Configurational Correlation of Optically Active Phosphine Oxides. J. Am. Chem. Soc 1967, 89, 4784–4786. [Google Scholar]; (c) Dang TP; Kagan HB The Asymmetric Synthesis of Hydratropic Acid and Amino-Acids by Homogeneous Catalytic Hydrogenation. J. Chem. Soc. D 1971, 481.; (d) Miyashita A; Yasuda A; Takaya H; Toriumi K; Ito T; Souchi T; Noyori R Synthesis of 2,2’-bis(diphenylphosphino)-1,1’-binaphthyl (BINAP), an Atropisomeric Chiral Bis(triaryl)phosphine, and Its Use in The Rhodium(I)-Catalyzed Asymmetric Hydrogenation of α-(Acylamino)acrylic Acids. J. Am. Chem. Soc 1980, 102, 7932–7934. [Google Scholar]
- (11).Knowles WS; Sabacky MJ Catalytic Asymmetric Hydrogenation Employing a Soluble, Optically Active, Rhodium Complex Chem. Commun 1968, 1445–1446. [Google Scholar]
- (12). For the first ruthenium-catalyzed alcohol-mediated transfer hydrogenations of olefins and ketones, and the first highly enantioselective ruthenium-catalyzed hydrogenations and transfer hydrogenations of ketones, see:; (a) Sasson Y; Blum J Homogeneous Catalytic Transfer-Hydrogenation of α,β-Unsaturated Carbonyl Compounds by Dichlorotris(triphenylphosphine)ruthenium(II). Tetrahedron Lett 1971, 12, 2167–2170. [Google Scholar]; (b) Sasson Y; Blum J Dichlorotris(triphenylphosphine)ruthenium-Catalyzed Hydrogen Transfer from Alcohols to Saturated and α,β-Unsaturated Ketones. J. Org. Chem 1975, 40, 1887–1896. [Google Scholar]; (c) Noyori R; Ohta M; Hsiao Y; Kitamura M; Ohta T; Takaya H Asymmetric Synthesis of Isoquinoline Alkaloids by Homogeneous Catalysis. J. Am. Chem. Soc 1986, 108, 7117–7119. [Google Scholar]; (d) Hashiguchi S; Fujii A; Takehara J; Ikariya T; Noyori R Asymmetric Transfer Hydrogenation of Aromatic Ketones Catalyzed by Chiral Ruthenium(II) Complexes. J. Am. Chem. Soc 1995, 117, 7562–7563. [Google Scholar]
- (13). Perhaps the earliest example of metal-catalyzed hydrogen-transfer involves the heterogeneous palladium-catalyzed disproportionation of dimethyl 1,4-dihydroterephthalate:; (a) Knoevenagel E; Bergdolt B Ueber das Verhalten des Δ2.5-Dihydroterephtalsauredimethylesters bei Höheren Temperaturen und in Gegenwart von Palladiummohr. Chem. Ber 1903, 36, 2857–2860. [Google Scholar]; For early examples of alcohol-mediated transfer hydrogenations of ketones and olefins using iridium and platinum catalysts, see:; (b) Haddad YMY; Henbest HB; Husbands J; Mitchell TRB Reduction of Cyclohexanones to Axial Alcohols via Iridium-containing Catalysts. Proc. Chem. Soc 1964, 361.; (c) Trocha-Grimshaw J; Henbest HB Catalysis of the Transfer of Hydrogen from Propan-2-ol to α,β-Unsaturated Ketones by Organoiridium Compounds. A Carbon-Iridium Compound Containing a Chelate Keto-group. Chem. Comm 1967, 544.; (d) Bailar JC Jr.; Itatani H Homogeneous Catalysis in the Reactions of Olefinic Substances. VI. Selective Hydrogenation of Methyl Linoleate and Isomerization of Methyl Oleate by Homogeneous Catalysis with Platinum Complexes Containing Triphenylphosphine, -arsine, or –stibine. J. Am. Chem. Soc 1967, 89, 1592–1599. [Google Scholar]
- (14). For selected reviews on enantioselective hydrogenation and transfer hydrogenation in the synthesis of pharmaceutical ingredients, see:; (a) Hawkins JM; Watson TJN Asymmetric Catalysis in the Pharmaceutical Industry. Angew. Chem. Int. Ed 2004, 43, 3224–3228. [DOI] [PubMed] [Google Scholar]; (b) Thommen M. Homogeneous Asymmetric Hydrogenation: Mature and Fit for Early Stage Drug Development. Spec. Chem. Mag 2005, 25, 26–28. [Google Scholar]; (c) Thayer AM Chiral Catalysis. Chem. Eng. News 2005, 83, 40–58. [Google Scholar]; (d) Farina V; Reeves JT; Senanayake CH; Song JJ Asymmetric Synthesis of Active Pharmaceutical Ingredients. Chem. Rev 2006, 106, 2734–2793. [DOI] [PubMed] [Google Scholar]; (e) Carey JS; Laffan D; Thomson C; Williams MT Analysis of The Reactions Used for The Preparation of Drug Candidate Molecules. Org. Biomol. Chem 2006, 4, 2337–2347. [DOI] [PubMed] [Google Scholar]; (f) Ager DJ; de Vries AHM; de Vries JG Asymmetric Homogeneous Hydrogenations at Scale. Chem. Soc. Rev 2012, 41, 3340–3380. [DOI] [PubMed] [Google Scholar]; (g) Etayo P; Vidal-Ferran A Rhodium-Catalyzed Asymmetric Hydrogenation as a Valuable Synthetic Tool for The Preparation of Chiral Drugs. Chem. Soc. Rev 2013, 42, 728–754. [DOI] [PubMed] [Google Scholar]; (h) Hayler JD; Leahy DK; Simmons EM A Pharmaceutical Industry Perspective on Sustainable Metal Catalysis. Organometallics 2019, 38, 36–46. [Google Scholar]
- (15).Stoffels MA; Klauck FJR; Hamadi T; Glorius F; Leker J Technology Trends of Catalysts in Hydrogenation Reactions: A Patent Landscape Analysis. Adv. Synth. Catal 2020, 362, 1258–1274. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (16).(a) Fischer F; Tropsch H Process for the Synthesis of Alcohols and Other Oxygen-Compounds via Catalytic Reduction of Carbon Monoxide, DRP 411216, Nov 3, 1922 ; (b) Fischer F; Tropsch H The Preparation of Synthetic Oil Mixtures (Synthol) from Carbon Monoxide and Hydrogen. Brennstoff Chem 1923, 4, 276–285. [Google Scholar]; (c) Fischer F; Tropsch H Synthesis of Higher Members of the Aliphatic Series from Carbon Monoxide. Chem. Ber 1923, 56B, 2428–2443. [Google Scholar]
- (17).(a) Roelen O (to Chemische Verwertungsgesellschaft Oberhausen m.b.H.) German Patent DE 849548, 1938/1952; U.S. Patent 2327066, 1943; Chem. Abstr 1944, 38, 3631. [Google Scholar]; (b) Evans D; Osborn JA; Wilkinson G Hydroformylation of Alkenes by Use of Rhodium Complex Catalysts. J. Chem. Soc. A, 1968, 3133–3142.
- (18).Mittasch A; Schneider C Verfahren zur Herstellung von Wasserstoff aus Kohlenwasserstoffen und Wasserdampf. German DRP Patent 296866, December 1, 1912. [Google Scholar]
- (19).Guerbet M Action de l’Alcool Amylique de Fermentation sur Son Dérivé Sodé. C. R. Acad. Sci. Paris 1899, 128, 1002–1004. [Google Scholar]
- (20). For selected reviews on hydrogen auto-transfer (borrowing hydrogen) processes, see:; (a) Hamid MHSA; Slatford PA; Williams JMJ Borrowing Hydrogen in The Activation of Alcohols. Adv. Synth. Catal 2007, 349, 1555–1575. [Google Scholar]; (b) Guillena G; Ramón DJ; Yus M Alcohols as Electrophiles in C-C Bond-Forming Reactions: The Hydrogen Autotransfer Process. Angew. Chem. Int. Ed 2007, 46, 2358–2364. [DOI] [PubMed] [Google Scholar]; (c) Dobereiner GE; Crabtree RH Dehydrogenation as a Substrate-Activating Strategy in Homogeneous Transition Metal Catalysis. Chem. Rev 2010, 110, 681–703. [DOI] [PubMed] [Google Scholar]; (d) Bähn S; Imm S; Neubert L; Zhang M; Neumann H; Beller M The Catalytic Amination of Alcohols. ChemCatChem 2011, 3, 1853–1864. [Google Scholar]; (e) Yang Q; Wang Q; Yu Z Substitution of Alcohols by N-Nucleophiles via Transition Metal-Catalyzed Dehydrogenation. Chem. Soc. Rev 2015, 44, 2305–2329. [DOI] [PubMed] [Google Scholar]; (f) Aitchison H; Wingad RL; Wass DF Homogeneous Ethanol to Butanol Catalysis - Guerbet Renewed. ACS Catal 2016, 6, 7125–7132. [Google Scholar]; (g) Quintard A; Rodriguez J Catalytic Enantioselective OFF ↔ ON Activation Processes Initiated by Hydrogen Transfer: Concepts and Challenges. Chem. Commun 2016, 52, 10456–10473. [DOI] [PubMed] [Google Scholar]; (h) Reed-Berendt BG; Polidano K; Morrill LC Recent Advances in Homogeneous Borrowing Hydrogen Catalysis Using Earth-Abundant First Row Transition Metals. Org. Biomol. Chem 2019, 17, 1595–1607. [DOI] [PubMed] [Google Scholar]; (i) Kwok T; Hoff O; Armstrong RJ; Donohoe TJ Control of Absolute Stereochemistry in Transition-Metal-Catalysed Hydrogen-Borrowing Reactions. Chem. Eur. J 2020, 26, 12912–12926. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (21).(a) Meerwein H; Schmidt R Ein Neues Verfahren zur Reduktion von Aldehyden und Ketonen. Justus Liebigs Ann. Chem 1925, 444, 221–238. [Google Scholar]; (b) Verley A Exchange of Functional Groups between Two Molecules. Exchange of Alcohol and Aldehyde Groups. Bull. Soc. Chim. Fr 1925, 37, 537–542. [Google Scholar]; (c) Ponndorf W Der Reversible Austausch der Oxydationsstufen zwischen Aldehyden oder Ketonen einerseits und Primären oder Sekundären Alkoholen anderseits. Angew. Chem 1926, 39, 138–143. [Google Scholar]
- (22).Winans CF; Adkins H The Alkylation of Amines as Catalyzed by Nickel. J. Am. Chem. Soc 1932, 54, 306–312. [Google Scholar]
- (23).(a) Grigg R; Mitchell TRB; Sutthivaiyakit S; Tongpenyai N Transition Metal-Catalysed N-Alkylation of Amines by Alcohols. J. Chem. Soc., Chem. Comm 1981, 611–612.; (b) Watanabe Y; Tsuji Y; Ohsugi Y The Ruthenium Catalyzed N-Alkylation and N-Heterocyclization of Aniline Using Alcohols and Aldehydes. Tetrahedron Lett 1981, 22, 2667–2670. [Google Scholar]
- (24).For notable industrial applications of borrowing hydrogen type reactions, see: Berliner MA; Dubant SPA; Makowski T; Ng K; Sitter B; Wager C; Zhang Y Use of an Iridium-Catalyzed Redox-Neutral Alcohol-Amine Coupling on Kilogram Scale for the Synthesis of a GlyT1 Inhibitor. Org. Process Res. Dev 2011, 15, 1052–1062. [Google Scholar]; (b) Shimizu H; Maeda H; Nara H Highly Productive α―Alkylation of Ketones with Alcohols Mediated by an Ir-Oxalamidato/Solid Base Catalyst System. Org. Process Res. Dev 2020, 24, 2772–2779. [Google Scholar]
- (25).For a review on industrial applications of borrowing hydrogen type reactions, see: Leonard J; Blacker AJ; Marsden SP; Jones MF; Mulholland KR; Newton R A Survey of the Borrowing Hydrogen Approach to the Synthesis of some Pharmaceutically Relevant Intermediates. Org. Process Res. Dev 2015, 19, 1400–1410. [Google Scholar]
- (26). For selected reviews on carbonyl addition via hydrogenation, transfer hydrogenation and hydrogen auto-transfer, see:; (a) Ngai M-Y; Kong J-R; Krische MJ Hydrogen-Mediated C-C Bond Formation: A Broad New Concept in Catalytic C-C Coupling. J. Org. Chem 2007, 72, 1063–1072. [DOI] [PubMed] [Google Scholar]; (b) Iida H; Krische MJ Catalytic Reductive Coupling of Alkenes and Alkynes to Carbonyl Compounds and Imines Mediated by Hydrogen. Top. Curr. Chem 2007, 279, 77–104. [Google Scholar]; (c) Hassan A; Krische MJ Unlocking Hydrogenation for C-C Bond Formation: A Brief Overview of Enantioselective Methods. Org. Proc. Res. Devel 2011, 15, 1236–1242. [DOI] [PMC free article] [PubMed] [Google Scholar]; (d) Moran J; Krische MJ Formation of C-C Bonds via Ruthenium-Catalyzed Transfer Hydrogenation. Pure Appl. Chem 2012, 84, 1729–1739. [DOI] [PMC free article] [PubMed] [Google Scholar]; (e) Ketcham JM; Shin I; Montgomery TP; Krische MJ Catalytic Enantioselective C-H Functionalization of Alcohols by Redox-Triggered Carbonyl Addition: Borrowing Hydrogen, Returning Carbon. Angew. Chem. Int. Ed 2014, 53, 9142–9150. [DOI] [PMC free article] [PubMed] [Google Scholar]; (f) Perez F; Oda S; Geary LM; Krische MJ Ruthenium-Catalyzed Transfer Hydrogenation for C-C Bond Formation: Hydrohydroxyalkylation and Hydroaminoalkylation via Reactant Redox Pairs. Top. Curr. Chem 2016, 374, 365–387. [DOI] [PMC free article] [PubMed] [Google Scholar]; (g) Nguyen KD; Park BY; Luong T; Sato H; Garza VJ; Krische MJ Metal-Catalyzed Reductive Coupling of Olefin-Derived Nucleophiles: Reinventing Carbonyl Addition. Science 2016, 354, aah5133. [DOI] [PMC free article] [PubMed] [Google Scholar]; (h) Kim SW; Zhang W; Krische MJ Catalytic Enantioselective Carbonyl Allylation and Propargylation via Alcohol-Mediated Hydrogen Transfer: Merging the Chemistry of Grignard and Sabatier. Acc. Chem. Res 2017, 50, 2371–2380. [DOI] [PMC free article] [PubMed] [Google Scholar]; (i) Doerksen RS; Meyer CC; Krische MJ Feedstock Reagents in Metal-Catalyzed Carbonyl Reductive Coupling: Minimizing Preactivation for Efficiency in Target-Oriented Synthesis. Angew. Chem. Int. Ed 2019, 58, 14055–14064. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (27).Schneider N; Lowe DM; Sayle RA; Tarselli MA; Landrum GA Big Data from Pharmaceutical Patents: A Computational Analysis of Medicinal Chemists’ Bread and Butter. J. Med. Chem 2016, 59, 4385–4402. [DOI] [PubMed] [Google Scholar]
- (28). The discovery of the aldol reaction is attributed to Borodin and Wurtz, however, there exist earlier observations by Kane:; (a) Kane R Ueber eine aus dem Essiggeist Entspringende Reihe von Verbindungen. Ann. Phys. Chem., Ser. 2 1838, 44, 473–494. [Google Scholar]; (b) Kane R Ueber den Essiggeist und einige davon Abgeleitete Verbindungen. J. Prakt. Chem 1838, 15, 129–155. [Google Scholar]; (c) Von Richter V aus St. Petersburg am 17. October 1869. Ber. Dtsch. Chem. Ges 1869, 2, 552–554. [Google Scholar]; (d) Borodin A Ueber einen Neuen Abkömmling des Valerals. Ber. Dtsch. Chem. Ges 1873, 6, 982–985. [Google Scholar]; (e) Würtz CA Sur un Aldéhyde-Alcool. Bull. Soc. Chim. Fr 1872, 17, 436–442. [Google Scholar]; (f) Würtz CA Ueber einen Aldehyd-Alkohol. J. Prakt. Chem 1872, 5, 457–464. [Google Scholar]
- (29). For selected reviews on catalytic enantioselective aldol additions, see:; (a) Trost BM; Brindle CS The Direct Catalytic Asymmetric Aldol Reaction. Chem. Soc. Rev 2010, 39, 1600–1632. [DOI] [PMC free article] [PubMed] [Google Scholar]; (b) Dias LC; de Lucca EC Jr.; Ferreira MAB; Polo EC Metal-Catalyzed Asymmetric Aldol Reactions. J. Braz. Chem. Soc 2012, 23, 2137–2158. [Google Scholar]; (c) Yamashita Y; Yasukawa T; Yoo W-J; Kitanosono T; Kobayashi S Catalytic Enantioselective Aldol Reactions. Chem. Soc. Rev 2018, 47, 4388–4480. [DOI] [PubMed] [Google Scholar]
- (30). For selected reviews on the use of asymmetric aldol additions in natural product total synthesis, see:; (a) Paterson I; Doughty VA; Florence G; Gerlach K; McLeod MD; Scott JP; Trieselmann T Asymmetric Aldol Reactions Using Boron Enolates: Applications to Polyketide Synthesis. ACS Symposium Series 2001, 783, 195–206. [Google Scholar]; (b) Schetter B; Mahrwald R Modern Aldol Methods for the Total Synthesis of Polyketides. Angew. Chem. Int. Ed 2006, 45, 7506–7525. [DOI] [PubMed] [Google Scholar]; (c) Paterson I Total Synthesis of Polyketides Using Asymmetric Aldol Reactions. In Asymmetric Synthesis, 2nd ed.; Christmann M, Bräse S, Eds.; Wiley-VCH Verlag GmbH & Co.: Weinheim, Germany, 2008; pp 293–298. [Google Scholar]; (d) Brodmann T; Lorenz M; Schäckel R; Simsek S; Kalesse M Highly Stereoselective Aldol Reactions in the Total Syntheses of Complex Natural Products. Synlett 2009, 174–192.; (e) Hosokawa S Asymmetric Aldol Reactions in the Total Syntheses of Natural Products. In Stereoselective Synthesis of Drugs and Natural Products; Andrushko V, Andrushko N, Eds.; John Wiley & Sons: New York, NY, 2013; Vol. 1; pp 215–247. [Google Scholar]; (f) Kan SBJ; Ng KK-H; Paterson I The Impact of the Mukaiyama Aldol Reaction in Total Synthesis. Angew. Chem. Int. Ed 2013, 52, 9097–9108. [DOI] [PubMed] [Google Scholar]
- (31).Heathcock CH The Aldol Reaction: Group I and Group II Enolates. In Comprehensive Organic Synthesis; Trost BM, Fleming I, Eds.; Pergamon Press: New York, NY, 1991; Vol. 2; Chapter 1.6; pp 181–238. [Google Scholar]
- (32). Kim BM; Williams SF; Masamune S The Aldol Reaction: Group III Enolates. In Comprehensive Organic Synthesis; Trost BM, Fleming I, Eds.; Pergamon Press: New York, NY, 1991; Vol. 2; Chapter 1.7; pp 239–275. [Google Scholar]
- (33).(a) Jang H-Y; Huddleston RR; Krische MJ Reductive Generation of Enolates from Enones Using Elemental Hydrogen: Catalytic C−C Bond Formation under Hydrogenative Conditions. J. Am. Chem. Soc 2002, 124, 15156–15157. [DOI] [PubMed] [Google Scholar]; (b) Huddleston RR; Krische MJ Enolate Generation under Hydrogenation Conditions: Catalytic Aldol Cycloreduction of Keto-Enones. Org. Lett 2003, 5, 1143–1146. [DOI] [PubMed] [Google Scholar]; (c) Koech PK; Krische MJ Catalytic Addition of Metallo-Aldehyde Enolates to Ketones: A New C−C Bond-Forming Hydrogenation. Org. Lett 2004, 6, 691–694. [DOI] [PubMed] [Google Scholar]; (d) Marriner GA; Garner SA; Jang H-Y; Krische MJ Metallo-Aldehyde Enolates via Enal Hydrogenation: Catalytic Cross Aldolization with Glyoxal Partners As Applied to the Synthesis of 3,5-Disubstituted Pyridazines. J. Org. Chem 2004, 69, 1380–1382. [DOI] [PubMed] [Google Scholar]; (e) Jung C-K; Garner SA; Krische MJ Hydrogen-Mediated Aldol Reductive Coupling of Vinyl Ketones Catalyzed by Rhodium: High syn-Selectivity through the Effect of Tri-2-furylphosphine. Org. Lett 2006, 8, 519–522. [DOI] [PubMed] [Google Scholar]; (f) Han SB; Krische MJ Reductive Aldol Coupling of Divinyl Ketones via Rhodium-Catalyzed Hydrogenation: syn-Diastereoselective Construction of β-Hydroxyenones. Org. Lett 2006, 8, 5657–5660. [DOI] [PubMed] [Google Scholar]; (g) Jung C-K; Krische MJ Asymmetric Induction in Hydrogen-Mediated Reductive Aldol Additions to α-Amino Aldehydes Catalyzed by Rhodium: Selective Formation of syn-Stereotriads Directed by Intramolecular Hydrogen-Bonding. J. Am. Chem. Soc 2006, 128, 17051–17056. [DOI] [PubMed] [Google Scholar]; (h) Bee C; Han SB; Hassan A; Iida H; Krische MJ Diastereo- and Enantioselective Hydrogenative Aldol Coupling of Vinyl Ketones: Design of Effective Monodentate TADDOL-Like Phosphonite Ligands. J. Am. Chem. Soc 2008, 130, 2746–2747. [DOI] [PMC free article] [PubMed] [Google Scholar]; (i) Komanduri V; Grant CD; Krische MJ Branch-Selective Reductive Coupling of 2-Vinyl Pyridines and Imines via Rhodium Catalyzed C-C Bond Forming Hydrogenation. J. Am. Chem. Soc 2008, 130, 12592–12593. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (34).Anderson NG; Keay BA 2-Furyl Phosphines as Ligands for Transition-Metal-Mediated Organic Synthesis. Chem. Rev 2001, 101, 997–1030. [DOI] [PubMed] [Google Scholar]
- (35).Shin I; Hong S; Krische MJ Total Synthesis of Swinholide A: An Exposition in Hydrogen-Mediated C-C Bond Formation. J. Am. Chem. Soc 2016, 138, 14246–14249. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (36).Meyer CC; Ortiz E; Krische MJ Catalytic Reductive Aldol and Mannich Reactions of Enone, Acrylate and Vinyl Heteroaromatic Pronucleophiles. Chem. Rev 2020, 120, 3721–3748. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (37).Halpern J Mechanism and Stereoselectivity of Asymmetric Hydrogenation. Science 1982, 217, 401–407. [DOI] [PubMed] [Google Scholar]
- (38).Rhee JU; Krische MJ Highly Enantioselective Reductive Cyclization of Acetylenic Aldehydes via Rhodium Catalyzed Asymmetric Hydrogenation. J. Am. Chem. Soc 2006, 128, 10674–10675. [DOI] [PubMed] [Google Scholar]
- (39).(a) Kong J-R; Ngai M-Y; Krische MJ Highly Enantioselective Direct Reductive Coupling of Conjugated Alkynes and α-Ketoesters via Rhodium-Catalyzed Asymmetric Hydrogenation. J. Am. Chem. Soc 2006, 128, 718–719. [DOI] [PubMed] [Google Scholar]; (b) Cho C-W; Krische MJ α-Hydroxy Esters via Enantioselective Hydrogen-Mediated C-C Coupling: Regiocontrolled Reactions of Silyl-Substituted 1,3-Diynes. Org. Lett 2006, 8, 3873–3876. [DOI] [PubMed] [Google Scholar]; (c) Komanduri V; Krische MJ Enantioselective Reductive Coupling of 1,3-Enynes to Heterocyclic Aromatic Aldehydes and Ketones via Rhodium Catalyzed Asymmetric Hydrogenation: Mechanistic Insight into the Role of Brønsted Acid Additives. J. Am. Chem. Soc 2006, 128, 16448–16449. [DOI] [PubMed] [Google Scholar]; (d) Hong Y-T; Cho C-W; Skucas E; Krische MJ Enantioselective Reductive Coupling of 1,3-Enynes to Glyoxalates Mediated by Hydrogen: Asymmetric Synthesis of β,γ-Unsaturated α-Hydroxy Esters. Org. Lett 2007, 9, 3745–3748. [DOI] [PubMed] [Google Scholar]
- (40).Kong J-R; Cho C-W; Krische MJ Hydrogen-Mediated Reductive Coupling of Conjugated Alkynes with Ethyl (N-Sulfinyl)iminoacetates: Synthesis of Unnatural α-Amino Acids via Rhodium-Catalyzed C-C Bond Forming Hydrogenation. J. Am. Chem. Soc 2005, 127, 11269–11276. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (41).(a) Barchuk A; Ngai M-Y; Krische MJ Allylic Amines via Iridium Catalyzed C-C Bond Forming Hydrogenation: Imine Vinylation in the Absence of Stoichiometric Byproducts or Metallic Reagents. J. Am. Chem. Soc 2007, 129, 8432–8433. [DOI] [PubMed] [Google Scholar]; (b) Ngai M-Y; Barchuk A; Krische MJ Enantioselective Iridium Catalyzed Imine Vinylation: Optically Enriched Allylic Amines via Alkyne-Imine Reductive Coupling Mediated by Hydrogen. J. Am. Chem. Soc 2007, 129, 12644–12645. [DOI] [PubMed] [Google Scholar]
- (42).(a) Kong J-R; Krische MJ Catalytic Carbonyl (Z)-Dienylation via Multicomponent Reductive Coupling of Acetylene to Aldehydes and α-Ketoesters Mediated by Hydrogen: Carbonyl Insertion into Cationic Rhodacyclopentadienes. J. Am. Chem. Soc 2006, 128, 16040–16041. [DOI] [PubMed] [Google Scholar]; (b) Skucas E; Kong JR; Krische MJ Enantioselective Reductive Coupling of Acetylene to N-Arylsulfonyl Imines via Rhodium Catalyzed C-C Bond Forming Hydrogenation: (Z)-Dienyl Allylic Amines. J. Am. Chem. Soc 2007, 129, 7242–7243. [DOI] [PubMed] [Google Scholar]; (c) Han SB; Kong J-R; Krische MJ Catalyst-Directed Diastereoselectivity in Hydrogenative Couplings of Acetylene to α-Chiral Aldehydes: Formal Synthesis of All Eight L-Hexoses. Org. Lett 2008, 10, 4133–4135. [DOI] [PMC free article] [PubMed] [Google Scholar]; (d) Williams VM; Kong JR; Ko BJ; Mantri Y; Brodbelt JS; Baik M-H; Krische MJ ESI-MS, DFT and Synthetic Studies on the H2-Mediated Coupling of Acetylene: Insertion of C=X Bonds into Rhodacyclopentadienes and Brønsted Acid Cocatalyzed Hydrogenolysis of Organorhodium Intermediates. J. Am. Chem. Soc 2009, 131, 16054–16062. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (43).(a) Cho C-W; Krische MJ Enantioselective Reductive Coupling of Alkynes and α-Keto Aldehydes via Rhodium-Catalyzed Hydrogenation: An Approach to Bryostatin Substructures. Org. Lett 2006, 8, 891–894. [DOI] [PubMed] [Google Scholar]; (b) Lu Y; Woo SK; Krische MJ Total Synthesis of Bryostatin 7 via C-C Bond Forming Hydrogenation. J. Am. Chem. Soc 2011, 133, 13876–13879. [DOI] [PMC free article] [PubMed] [Google Scholar]; (c) Andrews IP; Ketcham JM; Blumberg PM; Kedei N; Lewin NE; Peach MK; Krische MJ Synthesis of seco-B-Ring Bryostatin Analogue WN-1 via C-C Bond Forming Hydrogenation: Critical Contribution of the B-Ring in Determining Bryostatin- like and Phorbol 12-Myristate 13-Acetate-like Properties. J. Am. Chem. Soc 2014, 136, 13209–13216. [DOI] [PMC free article] [PubMed] [Google Scholar]; (d) Ketcham JM; Volchkov I; Chen T-Y; Blumberg PM; Kedei N; Lewin NE; Krische MJ Evaluation of Chromane-Based Bryostatin Analogues Prepared via Hydrogen-Mediated C-C Bond Formation: Potency Does Not Confer Bryostatin-Like Biology J. Am. Chem. Soc 2016, 138, 13415–13423. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (44).(a) Herold T; Hoffmann RW Enantioselective Synthesis of Homoallyl Alcohols via Chiral Allylboronic Esters. Angew. Chem. Int. Ed. Eng 1978, 17, 768–769. [Google Scholar]; (b) Furuta K; Mouri M; Yamamoto H Chiral (Acyloxy)borane Catalyzed Asymmetric Allylation of Aldehydes. Synlett 1991, 561–562.
- (45). For selected reviews on catalytic enantioselective carbonyl allylation, see:; (a) Denmark SE; Fu J Catalytic Enantioselective Addition of Allylic Organometallic Reagents to Aldehydes and Ketones. Chem. Rev 2003, 103, 2763–2794. [DOI] [PubMed] [Google Scholar]; (b) Hall DG Lewis and Brønsted Acid Catalyzed Allylboration of Carbonyl Compounds: From Discovery to Mechanism and Applications. Synlett 2007, 1644–1655.; (c) Yus M; González-Gómez JC; Foubelo F Catalytic Enantioselective Allylation of Carbonyl Compounds and Imines. Chem. Rev 2011, 111, 7774–7854. [DOI] [PubMed] [Google Scholar]; (d) Huo H-X; Duvall JR; Huang M-Y; Hong R Catalytic Asymmetric Allylation of Carbonyl Compounds and Imines with Allylic Boronates. Org. Chem. Front 2014, 1, 303–320. [Google Scholar]; (e) Spielmann K; Niel G; de Figueiredo RM; Campagne J-M Catalytic Nucleophilic ‘Umpoled’ π-Allyl Reagents. Chem. Soc. Rev 2018, 47, 1159–1173. [DOI] [PubMed] [Google Scholar]
- (46). For selected reviews on the catalytic enantioselective NHK allylation, see:; (a) Hargaden GC; Guiry PJ The development of the asymmetric Nozaki-Hiyama-Kishi Reaction Adv. Synth. Catal 2007, 349, 2407–2424. [Google Scholar]; (b) Tian Q; Zhang G Recent Advances in the Asymmetric Nozaki-Hiyama-Kishi Reaction. Synthesis 2016, 48, 4038–4049. [Google Scholar]
- (47).(a) Okude Y; Hirano S; Hiyama T; Nozaki H Grignard-type Carbonyl Addition of Allyl Halides by Means of Chromous salt. A Chemospecific Synthesis of Homoallyl Alcohols. J. Am. Chem. Soc 1977, 99, 3179–3181. [Google Scholar]; (b) Takai K; Tagashira M; Kuroda T; Oshima K; Utimoto K; Nozaki H Reactions of Alkenylchromium Reagents Prepared from Alkenyl Trifluoromethanesulfonates (Triflates) with Chromium(II) Chloride under Nickel Catalysis. J. Am. Chem. Soc 1986, 108, 6048–6050. [DOI] [PubMed] [Google Scholar]; (c) Jin H; Uenishi J; Christ WJ; Kishi Y Catalytic Effect of Nickel(II) Chloride and Palladium(II) Acetate on Chromium(II)-Mediated Coupling Reaction of Iodo Olefins with Aldehydes. J. Am. Chem. Soc 1986, 108, 5644–5646. [Google Scholar]
- (48).(a) Kim IS; Ngai M-Y; Krische MJ Enantioselective Iridium Catalyzed Carbonyl Allylation from the Alcohol or Aldehyde Oxidation Level Using Allyl Acetate as an Allyl Metal Surrogate. J. Am. Chem. Soc 2008, 130, 6340–6341. [DOI] [PMC free article] [PubMed] [Google Scholar]; (b) Kim IS; Ngai M-Y; Krische MJ Enantioselective Iridium Catalyzed Carbonyl Allylation from the Alcohol or Aldehyde Oxidation Level via Transfer Hydrogenative Coupling of Allyl Acetate: Departure from Chirally Modified Allyl Metal Reagents in Carbonyl Addition. J. Am. Chem. Soc 2008, 130, 14891–14899. [DOI] [PMC free article] [PubMed] [Google Scholar]; (c) Lu Y; Kim IS; Hassan A; Del Valle DJ; Krische MJ 1,n-Glycols as Dialdehyde Equivalents in Iridium Catalyzed Enantioselective Carbonyl Allylation and Iterative Two-Directional Assembly of 1,3-Polyols. Angew. Chem. Int. Ed 2009, 48, 5018–5021. [DOI] [PMC free article] [PubMed] [Google Scholar]; (d) Hassan A; Lu Y; Krische MJ Elongation of 1,3-Polyols via Iterative Catalyst-Directed Carbonyl Allylation from the Alcohol Oxidation Level. Org. Lett 2009, 11, 3112–3115. [DOI] [PMC free article] [PubMed] [Google Scholar]; (e) Schmitt DC; Dechert-Schmitt A-MR; Krische MJ Iridium-Catalyzed Allylation of Chiral β-Stereogenic Alcohols: Bypassing Discrete Formation of Epimerizable Aldehydes. Org. Lett 2012, 14, 6302–6305. [DOI] [PMC free article] [PubMed] [Google Scholar]; (f) Dechert-Schmitt A-MR; Schmitt DC; Krische MJ Protecting-Group-Free Diastereoselective C-C Coupling of 1,3-Glycols and Allyl Acetate through Site-Selective Primary Alcohol Dehydrogenation. Angew. Chem. Int. Ed 2013, 52, 3195–3198. [DOI] [PMC free article] [PubMed] [Google Scholar]; (g) Shin I; Wang G; Krische MJ Catalyst-Directed Diastereo- and Site-Selectivity in Successive Nucleophilic and Electrophilic Allylations of Chiral 1,3-Diols: Protecting-Group-Free Synthesis of Substituted Pyrans. Chem. Eur. J 2014, 20, 13382–13389. [DOI] [PMC free article] [PubMed] [Google Scholar]; (h) Kim SW; Lee W; Krische MJ Asymmetric Allylation of Glycidols Mediated by Allyl Acetate via Iridium-Catalyzed Hydrogen Transfer. Org. Lett 2017, 19, 1252–1254. [DOI] [PMC free article] [PubMed] [Google Scholar]; (i) Hassan A; Townsend IA; Krische MJ Catalytic Enantioselective Grignard Nozaki-Hiyama Methallylation from the Alcohol Oxidation Level: Chloride Compensates for π-Complex Instability. Chem. Comm 2011, 47, 10028–10030. [DOI] [PMC free article] [PubMed] [Google Scholar]; (j) Hassan A; Montgomery TP; Krische MJ Consecutive Iridium Catalyzed C-C and C-H Bond Forming Hydrogenations for the Diastereo- and Enantioselective Synthesis of syn-3-Fluoro-1-Alcohols: C-H (2-Fluoro)allylation of Primary Alcohols. Chem. Comm 2012, 48, 4692–4694. [DOI] [PMC free article] [PubMed] [Google Scholar]; (k) Montgomery TP; Hassan A; Park BY; Krische MJ Enantioselective Conversion of Primary Alcohols to α-exo-Methylene Butyrolactones via Iridium Catalyzed C-C Bond Forming Transfer Hydrogenation: 2-(Alkoxycarbonyl)allylation. J. Am. Chem. Soc 2012, 134, 11100–11103. [DOI] [PMC free article] [PubMed] [Google Scholar]; (l) Kim IS; Han SB; Krische MJ anti-Diastereo- and Enantioselective Carbonyl Crotylation from the Alcohol or Aldehyde Oxidation Level Employing a Cyclometallated Iridium Catalyst: α-Methyl Allyl Acetate as a Surrogate to Preformed Crotylmetal Reagents. J. Am. Chem. Soc 2009, 131, 2514–2520. [DOI] [PMC free article] [PubMed] [Google Scholar]; (m) Gao X; Townsend IA; Krische MJ Enhanced anti-Diastereo- and Enantioselectivity in Alcohol-Mediated Carbonyl Crotylation Using an Isolable Single Component Iridium Catalyst. J. Org. Chem 2011, 76, 2350–2354. [DOI] [PMC free article] [PubMed] [Google Scholar]; (n) Gao X; Han H; Krische MJ Direct Generation of Acyclic Polypropionate Stereopolyads via Double Diastereo- and Enantioselective Iridium-Catalyzed Crotylation of 1,3-Diols: Beyond Stepwise Carbonyl Addition in Polyketide Construction. J. Am. Chem. Soc 2011, 133, 12795–12800. [DOI] [PMC free article] [PubMed] [Google Scholar]; (o) Gao X; Zhang YJ; Krische MJ Iridium-Catalyzed anti-Diastereo- and Enantioselective Carbonyl (α-Trifluoromethyl)allylation from the Alcohol or Aldehyde Oxidation Level. Angew. Chem. Int. Ed 2011, 50, 4173–4175. [DOI] [PMC free article] [PubMed] [Google Scholar]; (p) Han SB; Gao X; Krische MJ Iridium-Catalyzed anti-Diastereo- and Enantioselective Carbonyl (Trimethylsilyl)allylation from the Alcohol or Aldehyde Oxidation Level. J. Am. Chem. Soc 2010, 132, 9153–9156. [DOI] [PMC free article] [PubMed] [Google Scholar]; (q) Han SB; Han H; Krische MJ Diastereo- and Enantioselective anti-Alkoxyallylation Employing Allylic gem-Dicarboxylates as Ally Donors via Iridium Catalyzed Transfer Hydrogenation. J. Am. Chem. Soc 2010, 132, 1760–1761. [DOI] [PMC free article] [PubMed] [Google Scholar]; (r) Tsutsumi R; Hong S; Krische MJ Diastereo- and Enantioselective Iridium Catalyzed Carbonyl (α-Cyclopropyl)allylation via Transfer Hydrogenation. Chem. Eur. J 2015, 21, 12903–12907. [DOI] [PMC free article] [PubMed] [Google Scholar]; (s) Xiang M; Luo G; Wang Y; Krische MJ Enantioselective Iridium-Catalyzed Carbonyl Isoprenylation via Alcohol-Mediated Hydrogen Transfer. Chem. Comm 2019, 55, 981–984. [DOI] [PMC free article] [PubMed] [Google Scholar]; (t) Luo G; Xiang M; Krische MJ Successive Nucleophilic and Electrophilic Allylation for The Catalytic Enantioselective Synthesis of 2,4-Disubstituted Pyrrolidines. Org. Lett 2019, 21, 2493–2497. [DOI] [PMC free article] [PubMed] [Google Scholar]; (u) Zhang YJ; Yang JH; Kim SH; Krische MJ anti-Diastereo- and Enantioselective Carbonyl (Hydroxymethyl)allylation from The Alcohol or Aldehyde Oxidation Level: Allyl Carbonates as Allylmetal Surrogates J. Am. Chem. Soc 2010, 132, 4562–4563. [DOI] [PMC free article] [PubMed] [Google Scholar]; (v) Feng J; Garza VJ; Krische MJ Redox-Triggered C-C Coupling of Alcohols and Vinyl Epoxides: Diastereo- and Enantioselective Formation of All-Carbon Quaternary Centers via tert-(Hydroxy)-Prenylation. J. Am. Chem. Soc 2014, 136, 8911–8914. [DOI] [PMC free article] [PubMed] [Google Scholar]; (w) Wang G; Franke J; Ngo CQ; Krische MJ Diastereo- and Enantioselective Iridium Catalyzed Coupling of Vinyl Aziridines and Alcohols: Site-Selective Modification of Unprotected Diols and Synthesis of Substituted Piperidines. J. Am. Chem. Soc 2015, 137, 7915–7920. [DOI] [PMC free article] [PubMed] [Google Scholar]; (x) Moran J; Smith AG; Carris RM; Johnson JS; Krische MJ Polarity Inversion of Donor-Acceptor Cyclopropanes: Disubstituted δ-Lactones via Enantioselective Iridium Catalysis. J. Am. Chem. Soc 2011, 133, 18618–18621. [DOI] [PMC free article] [PubMed] [Google Scholar]; (y) Hassan A; Zbieg JR; Krische MJ Enantioselective Iridium Catalyzed Vinylogous Reformatsky-Aldol Reaction from the Alcohol Oxidation Level: Linear Regioselectivity by Way of Carbon-Bound Enolates. Angew. Chem. Int. Ed 2011, 50, 3493–3496. [DOI] [PMC free article] [PubMed] [Google Scholar]; (z) Garza VJ; Krische MJ Hydroxymethylation beyond Carbonylation: Enantioselective Iridium Catalyzed Reductive Coupling of Formaldehyde with Allylic Acetates via Enantiotopic π-Facial Discrimination. J. Am. Chem. Soc 2016, 138, 3655–3658. [DOI] [PMC free article] [PubMed] [Google Scholar]; (a′) Cabrera JM; Tauber J; Zhang W; Xiang M; Krische MJ Selection between Diastereomeric Kinetic vs Thermodynamic Carbonyl Binding Modes Enables Enantioselective Iridium-Catalyzed anti-(α-Aryl)allylation of Aqueous Fluoral Hydrate and Difluoroacetaldehyde Ethyl Hemiacetal. J. Am. Chem. Soc 2018, 140, 9392–9395. [DOI] [PMC free article] [PubMed] [Google Scholar]; (b′) Meyer CC; Stafford NP; Cheng MJ; Krische MJ Ethanol: Unlocking an Abundant Renewable C2-Feedstock for Catalytic Enantioselective C-C Coupling. Angew. Chem. Int. Ed 2021, 60, 10.1002/anie.202102694. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (49).Kim SW; Meyer CC; Mai BK; Liu P; Krische MJ Inversion of Enantioselectivity in Allene Gas versus Allyl Acetate Reductive Aldehyde Allylation Guided by Metal-Centered Stereogenicity: An Experimental and Computational Study. ACS Catal 2019, 9, 9158–9163. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (50).(a) Vigneron JP; Dhaenens M; Horeau A Nouvelle Methode pour Porter au Maximum la Purete Optique d’un Produit Partiellement Dedouble sans L’Aide d’Aucune Substance Chirale. Tetrahedron 1973, 29, 1055–1059. [Google Scholar]; (b) Harned AM From Determination of Enantiopurity to The Construction of Complex Molecules: The Horeau Principle and Its Application in Synthesis. Tetrahedron 2018, 74, 3797–3841. [Google Scholar]
- (51). For catalytic enantioselective double allylation and double crotylation of 1,3-diols in polyketide total synthesis, see:; (a) Han SB; Hassan A; Kim I-S; Krische MJ Total Synthesis of (+)-Roxaticin via C-C Bond Forming Transfer Hydrogenation: A Departure from Stoichiometric Chiral Reagents, Auxiliaries, and Premetalated Nucleophiles in Polyketide Construction. J. Am. Chem. Soc 2010, 132, 15559–15561. [DOI] [PMC free article] [PubMed] [Google Scholar]; (b) Yang Z; Zhang B; Zhao G; Yang J; Xie X; She X Concise Formal Synthesis of (+)-Neopeltolide. Org. Lett 2011, 13, 5916–5919. [DOI] [PubMed] [Google Scholar]; (c) Feng Y; Jiang X; De Brabander JK Studies toward the Unique Pederin Family Member Psymberin: Full Structure Elucidation, Two Alternative Total Syntheses, and Analogs. J. Am. Chem. Soc 2012, 134, 17083–17093. [DOI] [PMC free article] [PubMed] [Google Scholar]; (d) Gao X; Woo SK; Krische MJ Total Synthesis of 6-Deoxyerythronolide B via C-C Bond-Forming Transfer Hydrogenation. J. Am. Chem. Soc 2013, 135, 4223–4226. [DOI] [PMC free article] [PubMed] [Google Scholar]; (e) Waldeck AR; Krische MJ Total Synthesis of Cyanolide A in the Absence of Protecting Groups, Chiral Auxiliaries or Premetallated Carbon Nucleophiles. Angew. Chem. Int. Ed 2013, 52, 4470–4473. [DOI] [PMC free article] [PubMed] [Google Scholar]; (f) Kasun ZA; Gao X; Lipinski RM; Krische MJ Direct Generation of Triketide Stereopolyads via Merged Redox-Construction Events: Total Synthesis of (+)-Zincophorin Methyl Ester. J. Am. Chem. Soc 2015, 137, 8900–8903. [DOI] [PMC free article] [PubMed] [Google Scholar]; (g) Willwacher J; Heggen B; Wirtz C; Thiel W; Fürstner A Total Synthesis, Stereochemical Revision, and Biological Reassessment of Mandelalide A: Chemical Mimicry of Intrafamily Relationships. Chem. Eur. J 2015, 21, 10416–10430. [DOI] [PubMed] [Google Scholar]; Novaes LFT; Sarotti AM; Pilli RA Total Synthesis and Stereochemical Assignment of Cryptolatifolione. RSC Adv 2015, 5, 53471–53476. [Google Scholar]; (i) Cabrera JM; Krische MJ Total Synthesis of Clavosolide A via Asymmetric Alcohol-Mediated Carbonyl Allylation: Beyond Protecting Groups or Chiral Auxiliaries in Polyketide Construction. Angew. Chem. Int. Ed 2019, 58, 10718–10722. [DOI] [PMC free article] [PubMed] [Google Scholar]; (j) Oka K; Fuchi S; Komine K; Fukuda H; Hatakeyama S; Ishihara J Catalytic Asymmetric Total Synthesis of Exiguolide. Chem. Eur. J 2020, 26, 12862–12867. [DOI] [PubMed] [Google Scholar]
- (52).(a) Dechert-Schmitt A-MR; Schmitt DC; Gao X; Itoh T; Krische MJ Polyketide Construction via Hydrohydroxyalkylation and Related Alcohol C-H Functionalizations: Reinventing the Chemistry of Carbonyl Addition. Nat. Prod. Rep 2014, 31, 504–513. [DOI] [PMC free article] [PubMed] [Google Scholar]; (b) Feng J; Kasun ZA; Krische MJ Enantioselective Alcohol C-H Functionalization for Polyketide Construction: Unlocking Redox-Economy and Site-Selectivity for Ideal Chemical Synthesis. J. Am. Chem. Soc 2016, 138, 5467–5478. [DOI] [PMC free article] [PubMed] [Google Scholar]; (c) Liu H; Lin S; Jacobsen KM; Poulsen TB Chemical Syntheses and Chemical Biology of Carboxyl Polyether Ionophores: Recent Highlights. Angew.Chem. Int. Ed 2019, 58, 13630–13642. [DOI] [PubMed] [Google Scholar]
- (53).Cabrera JM; Tauber J; Krische MJ Enantioselective Iridium-Catalyzed Phthalide Formation through Internal Redox Allylation of Phthalaldehydes. Angew. Chem. Int. Ed 2018, 57, 1390–1393. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (54). For selected reviews on catalytic enantioselective carbonyl propargylation, see:; (a) Marshall JA Chiral Allylic and Allenic Stannanes as Reagents for Asymmetric Synthesis. Chem. Rev 1996, 96, 31–48. [DOI] [PubMed] [Google Scholar]; (b) Gung BW Additions of Allyl, Allenyl, and Propargylstannanes to Aldehydes and Imines. Org. React 2004, 64, 1–113. [Google Scholar]; (c) Ding C-H; Hou X-L Catalytic Asymmetric Propargylation. Chem. Rev 2011, 111, 1914–1937. [DOI] [PubMed] [Google Scholar]; (d) Thaima T; Zamani F; Hyland CJT; Pyne SG Allenylation and Propargylation Reactions of Ketones, Aldehydes, Imines, and Iminium Ions Using Organoboronates and Related Derivatives. Synthesis 2017, 49, 1461–1480. [Google Scholar]; (e) Ambler BR; Woo SK; Krische MJ Catalytic Enantioselective Carbonyl Propargylation Beyond Preformed Carbanions: Reductive Coupling and Hydrogen Auto-Transfer. ChemCatChem 2019, 11, 324–332. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (55).(a) Woo SK; Geary LM; Krische MJ Enantioselective Carbonyl Propargylation by Iridium-Catalyzed Transfer Hydrogenative Coupling of Alcohols and Propargyl Chlorides. Angew. Chem. Int. Ed 2012, 51, 7830–7834. [DOI] [PubMed] [Google Scholar]; (b) Liang T; Woo SK; Krische MJ C-Propargylation Overrides O-Propargylation in Reactions of Propargyl Chloride with Primary Alcohols: Rhodium Catalyzed Transfer Hydrogenation. Angew. Chem. Int. Ed 2016, 55, 9207–9211. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (56).Spielmann K; Xiang M; Schwartz LA; Krische MJ Direct Conversion of Primary Alcohols to 1,2-Amino Alcohols: Enantioselective Iridium-Catalyzed Carbonyl Reductive Coupling of Phthalimido-Allene via Hydrogen Auto-Transfer. J. Am. Chem. Soc 2019, 141, 14136–14141. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (57).(a) Han SB; Kim IS; Han H; Krische MJ Enantioselective Carbonyl Reverse Prenylation from the Alcohol or Aldehyde Oxidation Level Employing 1,1-Dimethylallene as the Prenyl Donor. J. Am. Chem. Soc 2009, 131, 6916–6917. [DOI] [PMC free article] [PubMed] [Google Scholar]; (b) Itoh J; Han SB; Krische MJ Enantioselective Allylation, Crotylation and Reverse Prenylation of Substituted Isatins: Iridium-Catalyzed C-C Bond-Forming Transfer Hydrogenation. Angew. Chem. Int. Ed 2009, 48, 6313–6316. [DOI] [PMC free article] [PubMed] [Google Scholar]; (c) Bechem B; Patman RL; Hashmi ASK; Krische MJ Enantioselective Carbonyl Allylation, Crotylation and tert-Prenylation of Furan Methanols and Furfurals via Iridium-Catalyzed Transfer Hydrogenation. J. Org. Chem 2010, 75, 1795–1798. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (58).(a) Nguyen KD; Herkommer D; Krische MJ Enantioselective Formation of All-Carbon Quaternary Centers via C-H Functionalization of Methanol: Iridium-Catalyzed Diene Hydrohydroxymethylation. J. Am. Chem. Soc 2016, 138, 14210–14213. [DOI] [PMC free article] [PubMed] [Google Scholar]; (b) Holmes M; Nguyen KD; Schwartz LA; Luong T; Krische MJ Enantioselective Formation of CF3-Bearing All-Carbon Quaternary Stereocenters via C-H Functionalization of Methanol: Iridium Catalyzed Allene Hydrohydroxymethylation. J. Am. Chem. Soc 2017, 139, 8114–8117. [DOI] [PMC free article] [PubMed] [Google Scholar]; (c) Schwartz LA; Holmes M; Brito GA; Gonç TP; Richardson J; Ruble JC; Huang K-W; Krische MJ Cyclometallated Iridium-PhanePhos Complexes Are Active Catalysts in Enantioselective Allene-Fluoral Reductive Coupling and Related Alcohol-Mediated Carbonyl Additions that Form Acyclic Quaternary Carbon Stereocenters. J. Am. Chem. Soc 2019, 141, 2087–2096. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (59). For selected reviews on catalytic enantioselective formation of acyclic quaternary carbon stereocenters, see:; (a) Marek I; Sklute G Creation of Quaternary Stereocenters in Carbonyl Allylation Reactions. Chem. Commun 2007, 1683–1691. [DOI] [PubMed] [Google Scholar]; (b) Das JP; Marek I Enantioselective Synthesis of All-Carbon Quaternary Stereogenic Centers in Acyclic Systems. Chem. Commun 2011, 47, 4593–4623. [DOI] [PubMed] [Google Scholar]; (c) Marek I; Minko Y; Pasco M; Mejuch T; Gilboa N; Chechik H; Das JP All-Carbon Quaternary Stereogenic Centers in Acyclic Systems through the Creation of Several C–C Bonds per Chemical Step. J. Am. Chem. Soc 2014, 136, 2682–2694. [DOI] [PubMed] [Google Scholar]; (d) Holmes M; Schwartz LA; Krische MJ Intermolecular Metal-Catalyzed Reductive Coupling of Dienes, Allenes and Enynes with Carbonyl Compounds and Imines. Chem. Rev 2018, 118, 6026–6052. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (60).(a) Geary LM; Woo SK; Leung JC; Krische MJ Diastereo- and Enantioselective Iridium Catalyzed Carbonyl Propargylation from the Alcohol or Aldehyde Oxidation Level: 1,3-Enynes as Allenylmetal Equivalents. Angew. Chem. Int. Ed 2012, 51, 2972–2976. [DOI] [PubMed] [Google Scholar]; (b) Nguyen KD; Herkommer D; Krische MJ Ruthenium-BINAP Catalyzed Alcohol C–H tert-Prenylation via 1,3-Enyne Transfer Hydrogenation: Beyond Stoichiometric Carbanions in Enantioselective Carbonyl Propargylation. J. Am. Chem. Soc 2016, 138, 5238–5241. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (61).Schulthoff S; Hamilton JY; Heinrich M; Kwon Y; Wirtz C Fürstner, A. The Formosalides: Structure Determination by Total Synthesis. Angew. Chem. Int. Ed 2021, 60, 446–454. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (62).(a) Zbieg JR; Moran J; Krische MJ Diastereo- and Enantioselective Ruthenium Catalyzed Hydrohydroxyalkylation of 2-Silyl-Butadienes: Carbonyl syn-Crotylation from the Alcohol Oxidation Level. J. Am. Chem. Soc 2011, 133, 10582–10586. [DOI] [PMC free article] [PubMed] [Google Scholar]; (b) Zbieg JR; Yamaguchi E; McInturff EL; Krische MJ Enantioselective C-H Crotylation of Primary Alcohols via Hydrohydroxyalkylation of Butadiene. Science 2012, 336, 324–327. [DOI] [PMC free article] [PubMed] [Google Scholar]; (c) McInturff EL; Yamaguchi E; Krische MJ Chiral-Anion-Dependent Inversion of Diastereo- and Enantioselectivity in Carbonyl Crotylation via Ruthenium-Catalyzed Butadiene Hydrohydroxyalkylation. J. Am. Chem. Soc 2012, 134, 20628–20631. [DOI] [PMC free article] [PubMed] [Google Scholar]; (d) Grayson MN; Krische MJ; Houk KN Ruthenium-Catalyzed Asymmetric Hydrohydroxyalkylation of Butadiene: The Role of the Formyl Hydrogen Bond in Stereochemical Control. J. Am. Chem. Soc 2015, 137, 8838–8850. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (63).(a) Liang T; Nguyen KD; Zhang W; Krische MJ Enantioselective Ruthenium-Catalyzed Carbonyl Allylation via Alkyne-Alcohol C-C Bond-Forming Transfer Hydrogenation: Allene Hydrometalation vs Oxidative Coupling. J. Am. Chem. Soc 2015, 137, 3161–3164. [DOI] [PMC free article] [PubMed] [Google Scholar]; (b) Xiang M; Ghosh A; Krische MJ Diastereo- and Enantioselective Ruthenium-Catalyzed C-C Coupling of 1-Arylpropynes and Alcohols: Alkynes as Chiral Allylmetal Precursors in Carbonyl anti-(α-Aryl)allylation. J. Am. Chem. Soc 2021, 143, 2838–2845. [DOI] [PMC free article] [PubMed] [Google Scholar]; (c) Liang T; Zhang W; Chen T-Y; Nguyen KD; Krische MJ Ruthenium Catalyzed Diastereo- and Enantioselective Coupling of Propargyl Ethers with Alcohols: Siloxy-Crotylation via Hydride Shift Enabled Conversion of Alkynes to π-Allyls. J. Am. Chem. Soc 2015, 137, 13066–13071. [DOI] [PMC free article] [PubMed] [Google Scholar]; (d) Liang T; Zhang W; Krische MJ Iridium-Catalyzed C-C Coupling of a Simple Propargyl Ether with Primary Alcohols: Enantioselective Homoaldol Addition via Redox-Triggered (Z)-Siloxyallylation. J. Am. Chem. Soc 2015, 137, 16024–16027. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (64).For a review, see: Haydl AM; Breit B; Liang T; Krische MJ Alkynes as Electrophilic or Nucleophilic Allylmetal Precursors in Transition-Metal Catalysis. Angew. Chem. Int. Ed 2017, 56, 11312–11325. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (65). For selected reviews on hydroaminoalkylation, see:; (a) Roesky PW Catalytic Hydroaminoalkylation. Angew. Chem. Int. Ed 2009, 48, 4892–4894. [DOI] [PubMed] [Google Scholar]; (b) Eisenberger P; Schafer LL Catalytic Synthesis of Amines and N-Containing Heterocycles: Amidate Complexes for Selective C-N and C-C Bond-Forming Reactions. Pure Appl. Chem 2010, 82, 1503–1515. [Google Scholar]; (c) Chong E; Garcia P; Schafer LL Hydroaminoalkylation: Early-Transition-Metal-Catalyzed α-Alkylation of Amines. Synthesis 2014, 2884–2896. [Google Scholar]; (d) Edwards PM; Schafer LL Early Transition Metal-Catalyzed C-H Alkylation: Hydroaminoalkylation for Csp3-Csp3 Bond Formation in the Synthesis of Selectively Substituted Amines. Chem. Comm 2018, 54, 12543–12560. [DOI] [PubMed] [Google Scholar]; (e) Hannedouche J; Schulz E Hydroamination and Hydroaminoalkylation of Alkenes by Group 3–5 Elements: Recent Developments and Comparison with Late Transition Metals. Organometallics 2018, 37, 4313–4326. [Google Scholar]; (f) Gonnard L; Guérinot A; Cossy J Transition Metal-Catalyzed α-Alkylation of Amines by C(sp3)‒H Bond Activation. Tetrahedron 2019, 75, 145–163. [Google Scholar]
- (66).(a) Schmitt DC; Lee J; Dechert-Schmitt A-MR Yamaguchi E; Krische MJ Ruthenium Catalyzed Hydroaminoalkylation of Isoprene via Transfer Hydrogenation: Byproduct-free Prenylation of Hydantoins. Chem. Comm 2013, 49, 6096–6098. [DOI] [PMC free article] [PubMed] [Google Scholar]; (b) Chen T-Y; Tsutsumi R; Montgomery TP; Volchkov I; Krische MJ Ruthenium Catalyzed C-C Coupling of Amino Alcohols with Dienes via Transfer Hydrogenation: Redox-Triggered Imine Addition and Related Hydroaminoalkylations. J. Am. Chem. Soc 2015, 137, 1798–1801. [DOI] [PubMed] [Google Scholar]
- (67).Exploiting concepts pioneered in our laboratory, a nickel-catalyzed alkyne hydroaminoalkylation via hydrogen auto-transfer was recently reported: Li L; Liu Y-C; Shi H Nickel-Catalyzed Enantioselective α-Alkenylation of N-Sulfonyl Amines: Modular Access to Chiral α-Branched Amines. J. Am. Chem. Soc 2021, 143, 4154–4161. Corresponding H2-mediated reductive couplings to form essentially identical products (ref. 41) were not cited. [DOI] [PubMed] [Google Scholar]
- (68). For ruthenium-catalyzed hydroaminoalkylation, see:; (a) Jun C-H; Hwang D-C; Na S −.J. Chelation-Assisted Alkylation of Benzylamine Derivatives by Ru0 Catalyst. Chem. Commun 1998, 1405–1406.; (b) Chatani N; Asaumi T; Yorimitsu S; Ikeda T; Kakiuchi F; Murai S Ru3(CO)12-Catalyzed Coupling Reaction of sp3 C-H Bonds Adjacent to a Nitrogen Atom in Alkylamines with Alkenes. J. Am. Chem. Soc 2001, 123, 10935–10941. [DOI] [PubMed] [Google Scholar]; (c) Schinkel M; Wang L; Bielefeld K; Ackermann L Ruthenium(II)-Catalyzed C(sp3)−H α-Alkylation of Pyrrolidines. Org. Lett 2014, 16, 1876–1879. [DOI] [PubMed] [Google Scholar]; (d) Kulago AA; Van Steijvoort BF; Mitchell EA; Meerpoel L; Maes BUW Directed Ruthenium-Catalyzed C(sp3)-H α-Alkylation of Cyclic Amines Using Dioxolane-Protected Alkenones. Adv. Synth. Catal 2014, 356, 1610–1618. [Google Scholar]; (e) Bergman SD; Storr TE; Prokopcová H; Aelvoet K; Diels G; Meerpoel L; Maes BUW The Role of the Alcohol and Carboxylic Acid in Directed Ruthenium-Catalyzed C(sp3)-H α-Alkylation of Cyclic Amines. Chem. Eur. J 2012, 18, 10393–10398. [DOI] [PubMed] [Google Scholar]
- (69). For iridium-catalyzed hydroaminoalkylation, see:; (a) Tsuchikama K; Kasagawa M; Endo K; Shibata T Cationic Ir(I)-Catalyzed sp3 C-H Bond Alkenylation of Amides with Alkynes. Org. Lett 2009, 11, 1821–1823. [DOI] [PubMed] [Google Scholar]; (b) Pan S; Endo K; Shibata T Ir(I)-Catalyzed Enantioselective Secondary sp3 C-H Bond Activation of 2-(Alkylamino)pyridines with Alkenes. Org. Lett 2011, 13, 4692–4695. [DOI] [PubMed] [Google Scholar]; (c) Pan S; Matsuo Y; Endo K; Shibata T Cationic Iridium-Catalyzed Enantioselective Activation of Secondary sp3 C-H Bond Adjacent to Nitrogen Atom. Tetrahedron 2012, 68, 9009–9015. [Google Scholar]; (d) Lahm G; Opatz T Unique Regioselectivity in the C(sp3)−H α‑Alkylation of Amines: The Benzoxazole Moiety as a Removable Directing Group. Org. Lett 2014, 16, 4201–4203. [DOI] [PubMed] [Google Scholar]; (e) Tahara Y.-k.; Michino M; Ito M; Kanyiva KS; Shibata T Enantioselective sp3 C-H Alkylation of γ-Butyrolactam by a Chiral Ir(I) Catalyst for the Synthesis of 4-Substituted γ-Amino Acids. Chem. Commun 2015, 51, 16660–16663. [DOI] [PubMed] [Google Scholar]; (f) Nagai M; Nagamoto M; Nishimura T; Yorimitsu H Iridium-Catalyzed sp3 CH Alkylation of 3-Carbonyl-2-(alkylamino)pyridines with Alkenes. Chem. Lett 2017, 46, 1176–1178. [Google Scholar]; (g) Yamauchi D; Nishimura T; Yorimitsu H Hydroxoiridium-Catalyzed Hydroalkylation of Terminal Alkenes with Ureas by C(sp3)-H Bond Activation. Angew. Chem. Int. Ed 2017, 56, 7200–7204. [DOI] [PubMed] [Google Scholar]; (h) Tran AT; Yu J-Q Practical Alkoxythiocarbonyl Auxiliaries for Iridium(I)-Catalyzed C-H Alkylation of Azacycles. Angew. Chem. Int. Ed 2017, 56, 10530–10534. [DOI] [PMC free article] [PubMed] [Google Scholar]; (i) Shirai T; Okamoto T; Yamamoto Y Iridium-Catalyzed Direct Asymmetric Alkylation of Aniline Derivatives using 2-Norbornene. Asian J. Org. Chem 2018, 7, 1054–1056. [Google Scholar]; (j) Nakamura I; Yamauchi D; Nishimura T Hydroxoiridium-Catalyzed sp3 C-H Alkylation of Indoline Derivatives with Terminal Alkenes. Asian J. Org. Chem 2018, 7, 1347–1350. [Google Scholar]; (k) Hattori H; Nishimura T Iridium-Catalyzed Sequential sp3 C-H Alkylation of an N-Methyl Group with Alkenes Towards the Synthesis of α-Substituted Amines. Adv. Synth. Catal 2018, 360, 4827–4831. [Google Scholar]
- (70). While promising advances in intermolecular catalytic carbonyl reductive couplings of olefins have been made, present approaches require activated carbonyl compounds (ref. a, b), aryl-substituted olefins (ref. c, d) or stoichiometric metals in combination with silane reductant (ref. e, f):; (a) Park BY; Luong T; Sato H; Krische MJ Osmium(0) Catalyzed C-C Coupling of Ethylene and α-Olefins with Diols, Ketols or Hydroxy Esters via Transfer Hydrogenation. J. Org. Chem 2016, 81, 8585–8594. [DOI] [PMC free article] [PubMed] [Google Scholar]; (b) Yamaguchi E; Mowat J; Luong T; Krische MJ Regio‐ and Diastereoselective C-C Coupling of α‐Olefins and Styrenes to 3‐Hydroxy‐2‐oxindoles by Ru‐Catalyzed Hydrohydroxyalkylation. Angew. Chem. Int. Ed 2013, 52, 8428–8431. [DOI] [PMC free article] [PubMed] [Google Scholar]; (c) Zheng Y-L; Liu Y-Y; Wu Y-M; Wang Y-X; Lin Y-T; Ye M Iron-Catalyzed Regioselective Transfer Hydrogenative Couplings of Unactivated Aldehydes with Simple Alkenes. Angew. Chem. Int. Ed 2016, 55, 6315–6318. [DOI] [PubMed] [Google Scholar]; (d) Xiao H; Wang G Krische, M. J. Regioselective Hydrohydroxyalkylation of Styrene with Primary Alcohols or Aldehydes via Ruthenium Catalyzed C-C Bond Forming Transfer Hydrogenation. Angew. Chem. Int. Ed 2016, 55, 16119–16122. [DOI] [PMC free article] [PubMed] [Google Scholar]; (e) Matos JLM; Vásquez-Céspedes S; Gu J; Oguma T; Shenvi RA Branch-Selective Addition of Unactivated Olefins into Imines and Aldehydes. J. Am. Chem. Soc 2018, 140, 49, 16976–16981. [DOI] [PMC free article] [PubMed] [Google Scholar]; (f) Saladrigas M; Puig J; Bonjoch J; Bradshaw B Iron-Catalyzed Radical Intermolecular Addition of Unbiased Alkenes to Aldehydes. Org. Lett 2020, 22, 8111–8115. [DOI] [PubMed] [Google Scholar]
- (71).(a) Shin I; Ramgren SD; Krische MJ Reductive Cyclization of Halo-Ketones to Form 3-Hydroxy-2-Oxindoles via Palladium Catalyzed Hydrogenation: A Hydrogen-Mediated Grignard Addition. Tetrahedron 2015, 71, 5776–5780. [DOI] [PMC free article] [PubMed] [Google Scholar]; (b) Swyka RA; Zhang W; Richardson J; Ruble JC; Krische MJ Rhodium-Catalyzed Aldehyde Arylation via Formate-Mediated Transfer Hydrogenation: Beyond Metallic Reductants in Grignard-Nozaki-Hiyami-Kishi-Type Addition. J. Am. Chem. Soc 2019, 141, 1828–1832. [DOI] [PMC free article] [PubMed] [Google Scholar]; (c) Shuler WG; Swyka RA; Schempp TT; Spinello BJ; Krische MJ Vinyl Triflate-Aldehyde Reductive Coupling-Redox Isomerization Mediated by Formate: Rhodium-Catalyzed Ketone Synthesis in the Absence of Stoichiometric Metals. Chem. Eur. J 2019, 25, 12517–12520. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (72). For selected examples of the catalytic enantioselective reductive coupling of CO2 with styrenes, dienes and allenes mediated by silane, see:; (a) Gui Y-Y; Hu N; Chen X-W; Liao L-L; Ju T; Ye J-H; Zhang Z; Li J; Yu D-G Highly Regio- and Enantioselective Copper-Catalyzed Reductive Hydroxymethylation of Styrenes and 1,3-Dienes with CO2. J. Am. Chem. Soc 2017, 139, 17011–17014. [DOI] [PubMed] [Google Scholar]; (b) Chen X-W; Zhu L; Gui Y-Y; Jing K; Bo Z-Y; Lan Y; Li J; Yu D-G Highly Selective and Catalytic Generation of Acyclic Quaternary Carbon Stereocenters via Functionalization of 1,3-Dienes with CO2. J. Am. Chem. Soc 2019, 141, 18825–18835. [DOI] [PubMed] [Google Scholar]; (c) Qiu J; Gao S; Li C; Zhang L; Wang Z; Wang X; Ding K Construction of All-Carbon Chiral Quaternary Centers through CuI-Catalyzed Enantioselective Reductive Hydroxymethylation of 1,1-Disubstituted Allenes with CO2. Chem. Eur. J 2019, 25, 13874–13878. [DOI] [PubMed] [Google Scholar]
- (73). For recent reviews encompassing reductive couplings to CO2, see:; (a) Ran C-K; Chen X-W; Gui Y-Y; Liu J; Song L; Ren K; Yu D-G Recent Advances in Asymmetric Synthesis with CO2. Sci. China Chem 2020, 63, 1336–1351. [Google Scholar]; (b) Zhang Z; Ye J-H; Ju T; Liao L-L; Huang H; Gui Y-Y; Zhou W-J; Yu D-G Visible-Light-Driven Catalytic Reductive Carboxylation with CO2. ACS Catal 2020, 10, 10871–10885. [Google Scholar]
- (74).Rossen K Greening Organic Chemistry with Process Chemistry. J. Org. Chem 2019, 84, 4580–4582. [DOI] [PubMed] [Google Scholar]
- (75).Enantioselective Hantzsch ester-mediated reductive pinacol reaction auger well for related processes employing feedstock reductants: Rono LJ; Yayla HG; Wang DY; Armstrong MF; Knowles RR Enantioselective Photoredox Catalysis Enabled by Proton-Coupled Electron Transfer: Development of an Asymmetric Aza-Pinacol Cyclization. J. Am. Chem. Soc 2013, 135, 17735–17738. [DOI] [PubMed] [Google Scholar]
- (76).Promising results on formate-mediated cross-electrophile reductive coupling were recently disclosed: Schwartz LA; Spielmann K; Swyka RA; Xiang M; Krische MJ Formate-Mediated Cross-Electrophile Reductive Coupling of Aryl Iodides and Bromopyridines. Isr. J. Chem 2020, 60, 1–5. 10.1002/ijch.202000069. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (77).Lin L; Bai X; Ye X; Zhao X; Tan C-H; Jiang Z Organocatalytic Enantioselective Protonation for Photoreduction of Activated Ketones and Ketimines Induced by Visible Light. Angew. Chem. Int. Ed 2017, 56, 13842–13846. [DOI] [PubMed] [Google Scholar]
- (78).For a review on transfer hydrogenative cycloadditions, see: Sato H; Turnbull BWH; Fukaya K; Krische MJ Ruthenium(0) Catalyzed Cycloaddition of 1,2-Diols, Ketols or Diones via Alcohol-Mediated Hydrogen Transfer. Angew. Chem. Int. Ed 2018, 57, 3012–3021. [DOI] [PMC free article] [PubMed] [Google Scholar]


