Abstract
A highly enantioselective formal hydroformylation of vinyl arenes enabled by copper hydride (CuH) catalysis is reported. Key to the success of the method was the use of the mild Lewis acid zinc triflate to promote the formation of oxocarbenium electrophiles through the activation of diethoxymethyl acetate. Using the newly developed protocol, a broad range of vinyl arene substrates underwent efficient hydroacetalization reactions to provide access to highly enantioenriched α-aryl acetal products in good yields with exclusively branched regioselectivity. The acetal products could be converted to the corresponding aldehydes, alcohols, and amines with full preservation of the enantiomeric purity. Density functional theory studies support that the key C–C bond-forming event between the alkyl copper intermediate and the oxocarbenium electrophile takes place with inversion of configuration of the Cu–C bond in a backside SE2-type mechanism.
Enantiomerically enriched aldehydes hold value as versatile synthetic intermediates, and as such, methods for their efficient construction are of great interest.1,2 Among numerous synthetic approaches for their synthesis, hydroformylation of olefins represents an atom economical and cost-effective strategy.3 Since the seminal discovery of “the oxo process” by Roelen in the 1930s, alkene hydroformylation has become one of the most widely employed homogeneous industrial catalytic processes, producing millions of tons of aldehyde or alcohol products from available feedstock olefins.4,5 The adoption of hydroformylation on industrial scales can be in part attributed to the large amount of work invested in the development of efficient transition metal catalysts. However, the intrinsic challenge of enantioselective hydroformylation involves—in addition to the formation of a new stereogenic center—addressing the issues of chemoselectivity (hydroformylation versus alkene hydrogenation or overreduction to form primary alcohols) and regioselectivity (branched versus linear aldehyde products).6 Accordingly, considerable effort has been exerted to develop chiral ligand-supported metal catalysts for enantioselective alkene hydroformylation. Since the first highly enantioselective vinyl arene hydroformylation reported by Stille,7 various successful methods using Pt- and Rh-based catalysts have emerged for many classes of alkenes.8 More recently, pioneering work from Nozaki,9 Zhang,11 Landis,12 Clarke,13 and others has led to the emergence of new chiral phosphine ligands, including Binaphos,9 Ph-BPE,10 Yanphos,11 bis(diazaphospholane)(BDP),12 Bobphos,13 and more,14 has enabled the development of efficient rhodium-catalyzed asymmetric hydroformylation reactions. In addition, a great deal of work has gone into inventing alternative hydroformylation methods that avoid the use of carbon monoxide gas.15–17 Catalytic methods other than hydroformylation for the synthesis of pharmaceutically relevant chiral α-aryl aldehydes have also been developed based on the enantioselective α-arylation of aldehydes,18 and photochemical deracemization.19
While significant progress has been made in precious metal-catalyzed enantioselective olefin hydroformylation, a protocol that employs an abundant and environmentally benign base metal catalyst along with a nontoxic carbonyl source in enantioselective hydroformylation, to the best of our knowledge, has yet to be reported.20 Nonetheless, in recent years base metal hydrides has been proven successful in various asymmetric olefin functionalizations.21 Herein, we describe the development of a highly enantioselective formal hydroformylation of vinylarene copper hydride-catalyzed process that operates efficiently using a simple, commercially available ortho ester as the precursor to the active electrophile (Figure 1).
Figure 1.
(A) Precious metal-catalyzed asymmetric hydroformylation. (B) CuH and Lewis acid mediated hydroacetalization (formylation) of vinyl arenes.
We hypothesized that a chiral alkylcopper species, generated from the highly enantioselective hydrocupration of vinyl arenes,21a could react with an oxocarbenium ion to afford an enantioenriched aldehyde synthon. However, given the high electrophilicity of oxocarbenium ions, we anticipated that the strongly reducing conditions required for copper hydride catalysis might be problematic. Based our prior work on hydroalkylation22 we considered the use of a protocol in which the slow in situ generation of an active electrophile might mitigate unproductive side reactions between the oxocarbenium and a silane. We envisioned that this could be accomplished through exposure of the commercially available ortho ester, diethoxy methyl acetate, to a suitable Lewis acid catalyst. The resultant oxocarbenium ion would be expected to rapidly react with alkyl copper intermediates to form hydrofunctionalized products. As these products are alkyl acetals, undesired carbonyl reduction or racemization, which plague conventional hydroformylation methods, would be precluded.6,8,19,23 The acetals thus formed could then be readily hydrolyzed to provide the corresponding aldehydes.
We began by examining the reaction of 4-phenyl styrene (1a), employing diethoxy methyl acetate as the precursor to the active electrophile and a catalyst derived from (S)-DTBM-SEGPHOS, CuOAc, and dimethoxy(methyl)silane. No formation of desired product 2a was observed unless zinc triflate was added as a Lewis acid cocatalyst. Using 6 mol % of Zn(OTf)2, 2a was formed in nearly quantitative yield (97% NMR yield) with a 98:2 er (Table 1, entry 2). Notably, the reaction mixture was found to exhibit complete selectivity for the formation of the branched product. Evaluating the performance of various Lewis acids revealed that Zn(OTf)2 was uniquely effective, among those we tested, in promoting the formation of 2a, with other common Lewis acids failing to provide any observable product (Table 1, entry 5). The ineffectiveness of ZnCl2 is particularly noteworthy and may be due to formation of ligated CuCl which cannot be converted to the copper hydride under the reaction conditions. Consistent with this hypothesis, no formation of product was observed when exogenous chloride was added (nBu4NCl or ZnCl2) along with Zn(OTf)2 (Table 1, entry 6). The previously described complex ((S)-DTBM-SEGPHOS)CuOAc, (P1)24 could also be used and afforded 2a in similar yield and selectivity as that obtained using a mixture of CuOAc and (S)-DTBM-SEGPHOS. A control experiment in the absence of CuOAc provided no product, supporting the notion that copper is responsible for the hydroacetalization reaction and ruling out a potential ligated zinc-catalyzed pathway. Finally, the use of other bidentate chiral (bis)phosphine ligands such as (S,S)-Ph-BPE resulted in a lower yield and enantioselectivity of 2a (Table 1, entry 7–8).
Table 1.
Optimization of the Enantioselective Hydroacetalization(formylation) of 4-Phenyl Styrene (1a) Employing Diethoxy Methyl Acetate as the Formyl Precursora
![]() | |||
---|---|---|---|
| |||
entry | deviation from above conditions | yield (2a) (%) | e.r. |
1 | no Zn(OTf)2 | nr | - |
2 | none | 97 | 98:2 |
3 | P1 (6 mol%) | 98 | 98:2 |
4 | no CuOAc | nr | - |
5 | other Lewis acids (ZnCl2, Mg(OTf)2, Sc(OTf)3,In(OTf)3) (6 mol%) | <1% | - |
6 | nBu4NCI or ZnCl2 (6 mol%) with Zn(OTf)2 (6 mol%) | <1% | - |
7 | L2 (7 mol%) | 87 | 35:65 |
8 | L3 (7 mol%) | 80 | 25:75 |
9 | L4 (7 mol%) | nr | - |
10 | L5 (7 mol%) | nr | - |
| |||
![]() |
Reaction conditions: 4-Phenyl styrene (1a, 0.20 mmol, 1.0 equiv), diethoxy methyl acetate (0.30 mmol, 1.5 equiv), (MeO)2MeSiH (DMMS) (0.80 mmol, 4.0 equiv), copper(I) acetate (6 mol %) and specified ligand (7 mol %), and THF (0.4 M), rt; reaction yields were estimated by 1H NMR analysis of the crude reaction mixture, using 1,1,2,2-tetrachloroethane as an internal standard (see SI for details). Enantiomeric ratio (e.r.) of 2a was determined by chiral supercritical fluid chromatography (SFC). nr = no reaction.
Having established useful conditions for enantioselective hydroacetalization, we next examined the scope of vinylarene (Table 2) that could be accommodated. A variety of electronically varied vinyl arenes—both electron-rich and -deficient—afforded the α-aryl acetals in good yields and enantioselectivities. The use of vinyl arenes with increased electron density led to products with slightly decreased enantioselectivity (2h, 2i, and 2k). Functional groups, such as aryl halides (2d, 2e, 2j), an alcohol (2m), ethers (2g, 2i, 2r, 2p), an amine (2k), an ester (2y), and a tertiary amide (2q), were all found to be compatible with the protocol. Vinyl arenes containing a pyridine (2t), carbazole (2n), thiophene (2v), piperidine (2k), pyrrole (2x), and benzofuran (2o) unit efficiently underwent the hydroacetalization reaction to provide the corresponding α-aryl acetals in good yield and excellent enantioselectivity. The scope of the hydroacetalization reaction was found to extend to β-Substituted styrenes as well, with compounds 2w, 2x, and 2y successfully produced with high levels of enantioselectivity.
Table 2.
Substrate Scope of the Enantioselective Hydroacetalization Reactiona
![]() |
Reported yields are the average of two isolated yields, using 0.50 mmol of vinyl arene substrate and 0.75 mmol of diethoxy methyl acetate. bReaction time 24 h. cCu(OAc) (10 mol %), (S)-DTBM-SEGPHOS (12 mol %), and reaction time 24 h. dVinyl arene substrate used as a 9:1 mixture of E/Z-isomers; enantiomeric ratio (e.r.) was determined by chiral SFC, 1enantiomeric ratio determined for aldehyde, 2enantiomeric ratio determined for alcohol.
One limitation of this methodology is the competitive formation of unusual dearomatized products when ortho substituted substrates were employed (Scheme SI1). Dearomatized products were also observed to form when trans-β-methylstyrene and several vinyl heteroarenes were used as substrates (Scheme SI1).25 Due to instability of the dearomatized product, we were unable to isolate and fully characterize these dearomatized products. Nonetheless, DFT calculations support a competing pathway for dearomative electrophilic addition for ortho-substituted styrenes (Scheme SI3).
We found that the α-aryl acetals produced by the above-described protocol could be converted to the corresponding aldehydes (3a–b) by treatment with a formic acid and pentane mixture (1:1) for 1 h at room temperature (Scheme 1).26 Importantly, this procedure was found to preserve enantiopurity, enabling the possibility of further derivatization to a variety of enantiomerically enriched products. Hydride reduction (3c–d) or reductive amination (3e–f) of aldehyde products could be efficiently carried out to provide alcohols or amines with high levels of enantiomeric purity.27 In addition, it has been reported that oxidation of α-aryl aldehydes to the corresponding carboxylic acids occurs without deterioration of enantiomeric purity.19 These resulting compounds have applications as nonsteroidal anti-inflammatory drugs. The absolute configuration of the products was assigned by comparing the specific optical rotation values of 3a, 3b, and 3d to those in the literature.19,28
Scheme 1. Derivatization of CuH-Catalyzed Asymmetric Hydroacetalization Products.
IReaction conditions: Pentane: Formic acid (1:1), rt, 1 h. iiReaction conditions: NaBH4 (5.0 equiv), methanol, rt, 1 h. iiiReaction conditions: NaBH(OAc)3 (2.0 equiv), amine (1.0 equiv), AcOH (1.0 equiv), DCE, rt, 8 h.
Based on our experimental findings, we hypothesize that the reaction proceeds through the catalytic cycle shown in Figure 2A. CuH complex (A) undergoes enantioselective hydrocupration of styrene substrate 1 to generate the Cu-alkyl intermediate (B), which goes on to react stereospecifically with oxocarbenium intermediate 5. This leads to the formation of acetal product 2, as well as ligated CuOAc (D). D can then be transformed to the active CuH catalyst through the reaction with DMMS ((MeO)2MeSiH).
Figure 2.
Computational studies of the mechanisms of electrophile activation and CuH-catalyzed hydroacetalization. Bond distances are in Å.
The proposed catalytic cycle was investigated using density functional theory (DFT), including both the Zn-promoted ionization and the Cu-catalyzed hydroacetalization steps.29 The DFT calculations were performed at the M06/6–311+G(d,p)–SDD(Cu, Zn)/SMD(THF)//B3LYP-D3/6–31G(d)–SDD(Cu, Zn) level of theory with the (S)-DTBM-SEGPHOS (L1)-supported Cu catalyst and p-methylstyrene 1b as the model substrate (see the Supporting Information for Computational Details). Calculations revealed that when the Zn(II) tetrahydrofuran complex was used as a catalyst, the generation of oxocarbenium ion 5 is thermodynamically favorable (ΔG = −10.9 kcal/mol) with a low activation barrier (ΔG‡ = 8.9 kcal/mol) via TS-1 (Figure 2B).
Upon the generation of (R)-Cu-alkyl intermediate 9 via the highly enantioselctive hydrocupration (TS-2) of substrate 1b by CuH catalyst 7,30 9 forms a weak van der Waals (vdW) complex with the oxocarbenium ion. From the vdW complex 10, our DFT calculations support that the electrophilic addition of 5 to 9 takes place through a stereoinvertive SE2-type mechanism,31 i.e., backside electrophilic substitution via TS-3 (Figure 2D). This SE2 pathway is calculated to occur readily with an activation barrier of 1.8 kcal/mol relative to Cu-alkyl intermediate 9. The stereochemical outcome of this most favorable pathway (shown in black in Figure 2C) is consistent with the absolute configuration of the hydroacetalization products observed experimentally. We located a transition state (TS-4) leading to the formation of the enantiomer of 2b via a stereoretentive SE2-type process. However, the activation barrier for TS-4 is 2.9 kcal/mol higher in energy than the stereoinvertive substitution TS-3.
Beyond the aforementioned SE2 mechanisms, several alternative pathways were considered. The oxocarbenium addition to the Cu center (TS-5) to form Cu(III) intermediate 11 and subsequent stereospecific C–C reductive elimination (TS-6) was predicted to occur with a significantly higher barrier relative to TS-3 (ΔΔG‡ = 15.8 kcal/mol).32 A dearomative electrophilic addition to the ortho position of 9 (TS-7) requires a barrier that is 3.6 kcal/mol higher relative to TS-3. Formation of dearomatized intermediate 12 via 1,3-Cu migration25,33 is unlikely because 12 is 13.2 kcal/mol less stable than TS-3. Taken together, the DFT calculations support that the acetalization of Cu-alkyl intermediate 9 takes place via a stereoinvertive SE2-type mechanism involving electrophile 5 generated from the facile Zn-catalyzed ionization of orthoester 4.
In summary, we have disclosed a CuH-catalyzed enantioselective formal hydroformylation of vinyl arenes, providing access to α-aryl acetal and aldehyde derivatives with high levels of enantioselectivity. This method was found to be tolerant of various functional groups and heterocycles. DFT computational modeling suggests a key role for the catalytic Lewis acid zinc(II) triflate in the activation of the orthoester electrophile to form a highly reactive oxocarbenium ion. Beyond this, calculations demonstrate that the stereoenriched alkyl copper intermediate generated from enantioselective alkene hydrocupration reacts with in situ formed oxocarbenium ion through a stereospecific invertive SE2-type mechanism to yield α-aryl acetals as product. We anticipate that the results presented in this study may have implications in the design of other base metal catalyzed hydroformylation processes.
Supplementary Material
ACKNOWLEDGMENTS
Research reported in this publication was supported by the National Institutes of Health (R35-GM122483 and R35GM128779). We thank the National Institutes of Health for a supplemental grant for the purchase of supercritical fluid chromatography (SFC) equipment (GM058160–17S1). DFT calculations were carried out at the University of Pittsburgh Center for Research Computing and the Advanced Cyberinfrastructure Coordination Ecosystem: Services & Support (ACCESS) program, supported by NSF award numbers OAC-2117681, OAC-1928147, and OAC-1928224. We are grateful to Drs. Dennis Kutateladze, Michael Strauss, James Law, and Christine Nguyen (MIT) for advice on the preparation of this manuscript.
Footnotes
ASSOCIATED CONTENT
Supporting Information
The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/jacs.4c04287.
Experimental procedures and characterization data for all new compounds, including NMR spectra, SFC traces, computational details, and Cartesian coordinates of all computed structures (PDF)
The authors declare no competing financial interest.
Contributor Information
Subhash Garhwal, Department of Chemistry, Massachusetts Institute of Technology, Cambridge, Massachusetts 02139, United States.
Yuyang Dong, Department of Chemistry, Massachusetts Institute of Technology, Cambridge, Massachusetts 02139, United States.
Binh Khanh Mai, Department of Chemistry, University of Pittsburgh, Pittsburgh, Pennsylvania 15260, United States.
Peng Liu, Department of Chemistry, University of Pittsburgh, Pittsburgh, Pennsylvania 15260, United States.
Stephen L. Buchwald, Department of Chemistry, Massachusetts Institute of Technology, Cambridge, Massachusetts 02139, United States
REFERENCES
- (1).(a) Botteghi C; Paganelli S; Schionato A; Marchetti M The Asymmetric Hydroformylation in the Synthesis of Pharmaceuticals. Chirality 1991, 3, 355–369. [DOI] [PubMed] [Google Scholar]; (b) Breit B Synthetic aspects of stereoselective hydroformylation. Acc. Chem. Res 2003, 36, 264–275. [DOI] [PubMed] [Google Scholar]
- (2).Stille JK In Comprehensive Organic Synthesis; Trost BM, Flemming I, Paquette LA, Eds.; Pergamon Press: Oxford, 1991; Vol. 4, p 913. [Google Scholar]
- (3).(a) Bai S-T; Wen J; Zhang X Asymmetric Carbonylation Reactions. In The Chemical Transformations of C1 Compounds, Wu X-F; Han B; Ding K; Liu Z, Eds.; Wiley-VCH: 2022; pp 611–666;. [Google Scholar]; (b) Franke R; Selent D; Börner A Applied Hydroformylation. Chem. Rev 2012, 112, 5675–5732. [DOI] [PubMed] [Google Scholar]
- (4).Bohnen HW; Cornils B Hydroformylation of alkenes: An industrial view of the status and importance. Adv. Catal 2002, 47, 1–64. [Google Scholar]
- (5).(a) Claver C; van Leeuwen PWNM; Claver C; van Leeuwen PWNM Rhodium Catalyzed Hydroformylation; Springer: Berlin, 2002. [Google Scholar]; (b) Hebrard F; Kalck P Cobalt-Catalyzed Hydroformylation of Alkenes: Generation and Recycling of the Carbonyl Species, and Catalytic Cycle. Chem. Rev 2009, 109, 4272–4282. [DOI] [PubMed] [Google Scholar]
- (6).(a) Agbossou F; Carpentier J-F; Mortreux A Asymmetric Hydroformylation. Chem. Rev 1995, 95, 2485–2506. [Google Scholar]; (b) Brezny AC; Landis CR Recent Developments in the Scope, Practicality, and Mechanistic Understanding of Enantioselective Hydroformylation. Acc. Chem. Res 2018, 51, 2344–2354. [DOI] [PubMed] [Google Scholar]
- (7).Parrinello G; Stille JK Asymmetric Hydroformylation Catalyzed by Homogeneous and Polymer-Supported Platinum Complexes Containing Chiral. J. Macromol. Sci., Chem 1987, 109, 7. [Google Scholar]
- (8).(a) Chakrabortty S; Almasalma AA; De Vries JG Recent Developments in Asymmetric Hydroformylation. Catal. Sci. Technol 2021, 11, 5388–5411. [Google Scholar]; (b) Chen C; Dong X-Q; Zhang X Chiral Ligands for Rhodium-Catalyzed Asymmetric Hydroformylation: A Personal Account. Chem. Rec 2016, 16, 2674–2686. [DOI] [PubMed] [Google Scholar]; (c) Deng Y; Wang H; Sun Y; Wang X Principles and Applications of Enantioselective Hydroformylation of Terminal Disubstituted Alkenes. ACS Catal 2015, 5, 6828–6837. [Google Scholar]; (d) Chikkali SH; van der Vlugt JI; Reek JNH Hybrid Diphosphorus Ligands in Rhodium Catalysed Asymmetric Hydroformylation. Coord. Chem. Rev 2014, 262, 1–15. [Google Scholar]; (e) Klosin J; Landis CR Ligands for Practical Rhodium-Catalyzed Asymmetric Hydroformylation. Acc. Chem. Res 2007, 40, 1251–1259. [DOI] [PubMed] [Google Scholar]
- (9).(a) Sakai N; Mano S; Nozaki K; Takaya H Highly Enantioselective Hydroformylation of Olefins Catalyzed by New Phosphinephosphite-Rh(I) Complexes. J. Am. Chem. Soc 1993, 115, 7033–7034. [Google Scholar]; (b) Nozaki K; Sakai N; Nanno T; Higashijima T; Mano S; Horiuchi T; Takaya H Highly Enantioselective Hydroformylation of Olefins Catalyzed by Rhodium(I) Complexes of New Chiral Phosphine-Phosphite Ligands. J. Am. Chem. Soc 1997, 119, 4413–4423. [Google Scholar]
- (10).(a) Axtell AT; Cobley CJ; Klosin J; Whiteker GT; Zanotti-Gerosa A; Abboud KA Highly Regio- and Enantioselective Asymmetric Hydroformylation of Olefins Mediated by 2,5-Disubstituted Phospholane Ligands. Angew. Chem., Int. Ed 2005, 44, 5834–5838. [DOI] [PubMed] [Google Scholar]; (b) Zheng X; Cao B; Liu TL; Zhang X Rhodium-Catalyzed Asymmetric Hydroformylation of 1,1-Disubstituted Allylphthalimides: A Catalytic Route to β3-Amino Acids. Adv. Synth. Catal 2013, 355, 679–684. [Google Scholar]
- (11).(a) Yan Y; Zhang X A Hybrid Phosphorus Ligand for Highly Enantioselective Asymmetric Hydroformylation. J. Am. Chem. Soc 2006, 128, 7198–7202. [DOI] [PubMed] [Google Scholar]; (b) Zhang X; Cao B; Yan Y; Yu S; Ji B; Zhang X Synthesis and Application of Modular Phosphine–Phosphoramidite Ligands in Asymmetric Hydroformylation: Structure-Selectivity Relationship. Chem. - Eur. J 2010, 16, 871–877. [DOI] [PubMed] [Google Scholar]; (c) Zhang X; Cao B; Yu S; Zhang X Rhodium-Catalyzed Asymmetric Hydroformylation of N-Allylamides: Highly Enantioselective Approach to β2-Amino Aldehydes. Angew. Chem., Int. Ed 2010, 49, 4047–4050. [DOI] [PubMed] [Google Scholar]; (d) You C; Wei B; Li X; Yang Y; Liu Y; Lv H; Zhang X Rhodium-Catalyzed Desymmetrization by Hydroformylation of Cyclopentenes: Synthesis of Chiral Carbocyclic Nucleosides. Angew. Chem., Int. Ed 2016, 55, 6511–6514. [DOI] [PubMed] [Google Scholar]; (e) Chen C; Jin S; Zhang Z; Wei B; Wang H; Zhang K; Lv H; Dong X-Q; Zhang X Rhodium/Yanphos-Catalyzed Asymmetric Interrupted Intramolecular Hydroaminomethylation of trans-1,2-Disubstituted Alkenes. J. Am. Chem. Soc 2016, 138, 9017–9020. [DOI] [PubMed] [Google Scholar]; (f) Dangat Y; Popli S; Sunoj RB Unraveling the Importance of Noncovalent Interactions in Asymmetric Hydroformylation Reactions. J. Am. Chem. Soc 2020, 142, 17079–17092. [DOI] [PubMed] [Google Scholar]
- (12).(a) Abrams ML; Foarta F; Landis CR Asymmetric Hydroformylation of Z-Enamides and Enol Esters with Rhodium-Bisdiazaphos Catalysts. J. Am. Chem. Soc 2014, 136, 14583–14588. [DOI] [PubMed] [Google Scholar]; (b) Risi RM; Maza AM; Burke SD Asymmetric Hydroformylation-Initiated Tandem Sequences for Syntheses of (+)-Patulolide C, (–)-Pyrenophorol, (+)-Decarestrictine L, and (+)-Prelog Djerassi Lactone. J. Org. Chem 2015, 80, 204–216. [DOI] [PubMed] [Google Scholar]; (c) Clark TP; Landis CR; Freed SL; Klosin J; Abboud KA Highly Active, Regioselective, and Enantioselective Hydroformylation with Rh Catalysts Ligated by Bis-3,4-diazaphospholanes. J. Am. Chem. Soc 2005, 127, 5040–5042. [DOI] [PubMed] [Google Scholar]; (d) Watkins AL; Hashiguchi BG; Landis CR Highly Enantioselective Hydroformylation of Aryl Alkenes with Diazaphospholane Ligands. Org. Lett 2008, 10, 4553. [DOI] [PubMed] [Google Scholar]
- (13).(a) Fuentes JA; Janka ME; Rodgers J; Fontenot KJ; Bühl M; Slawin AMZ; Clarke ML Effect of ligand backbone on the selectivity and stability of rhodium hydroformylation catalysts derived from phospholane-phosphites. Organometallics 2021, 40, 3966–3978. [Google Scholar]; (b) Noonan GM; Fuentes JA; Cobley CJ; Clarke ML An asymmetric hydroformylation catalyst that delivers branched aldehydes from alkyl alkenes. Angew. Chem., Int. Ed 2012, 51, 2477–2480. [DOI] [PubMed] [Google Scholar]; (c) Noonan GM; Cobley CJ; Mahoney T; Clarke ML Rhodium/phospholane–phosphite catalysts give unusually high regioselectivity in the enantioselective hydroformylation of vinyl arenes. Chem. Commun 2014, 50, 1475–1477. [DOI] [PubMed] [Google Scholar]
- (14).(a) Kuil M; Goudriaan PE; van Leeuwen PWNM; Reek JNH Template-induced formation of heterobidentate ligands and their application in the asymmetric hydroformylation of styrene. Chem. Commun 2006, 4679–4681. [DOI] [PubMed] [Google Scholar]; (b) Lightburn TE; Dombrowski MT; Tan KL Catalytic Scaffolding Ligands: An Efficient Strategy for Directing Reactions. J. Am. Chem. Soc 2008, 130, 9210–9211. [DOI] [PubMed] [Google Scholar]; (c) Jouffroy M; Gramage-Doria R; Armspach D; Semeril D; Oberhauser W; Matt D; Toupet L Confining Phosphanes Derived from Cyclodextrins for Efficient Regio- and Enantioselective Hydroformylation. Angew. Chem., Int. Ed 2014, 53, 3937–3940. [DOI] [PubMed] [Google Scholar]; (d) Bellini R; Reek JNH Application of Supramolecular Bidentate Hybrid Ligands in Asymmetric Hydroformylation. Chem.—Eur. J 2012, 18, 13510–13519. [DOI] [PubMed] [Google Scholar]; (e) Allmendinger S; Kinuta H; Breit B Easily Accessible TADDOL-Derived Bisphosphonite Ligands: Synthesis and Application in the Asymmetric Hydroformylation of Vinylarenes. Adv. Synth. Catal 2015, 357 (1), 41–45. [Google Scholar]
- (15).(a) Morimoto T; Kakiuchi K Evolution of Carbonylation Catalysis: No Need for Carbon Monoxide. Angew. Chem., Int. Ed 2004, 43, 5580–5588. [DOI] [PubMed] [Google Scholar]; (b) Wu L; Liu Q; Jackstell R; Beller M Carbonylations of Alkenes with CO Surrogates. Angew. Chem., Int. Ed 2014, 53, 6310–6320. [DOI] [PubMed] [Google Scholar]; (c) Cao J; Zheng Z-J; Xu Z; Xu L-W Transition-Metal-Catalyzed Transfer Carbonylation with HCOOH or HCHO as non-Gaseous C1 Source. Coord. Chem. Rev 2017, 336, 43–53. [Google Scholar]; (d) Hussain N; Chhalodia AK; Ahmed A; Mukherjee D Recent Advances in Metal-Catalyzed Carbonylation Reactions by Using Formic Acid as CO Surrogate. ChemistrySelect 2020, 5, 11272–11290. [Google Scholar]; (e) Ren W; Chang W; Dai J; Shi Y; Li J; Shi Y An Effective Pd-Catalyzed Regioselective Hydroformylation of Olefins with Formic Acid. J. Am. Chem. Soc 2016, 138, 14864–14867. [DOI] [PubMed] [Google Scholar]
- (16).(a) Morimoto T; Fujii T; Miyoshi K; Makado G; Tanimoto H; Nishiyama Y; Kakiuchi K Accessible Protocol for Asymmetric Hydroformylation of Vinylarenes Using Formaldehyde. Org. Biomol Chem 2015, 13, 4632–4636. [DOI] [PubMed] [Google Scholar]; (b) Makado G; Morimoto T; Sugimoto Y; Tsutsumi K; Kagawa N; Kakiuchi K Highly Linear-Selective Hydroformylation of 1-Alkenes Using Formaldehyde as a Syngas Substitute. Adv. Synth. Catal 2010, 352, 299–304. [Google Scholar]
- (17).(a) Pittaway R; Dingwall P; Fuentes JA; Clarke ML CO-Free Enantioselective Hydroformylation of Functionalised Alkenes: Using a Dual Catalyst System to Give Improved Selectivity and Yield. Adv. Synth. Catal 2019, 361, 4334–4341. [Google Scholar]; (b) Fuentes JA; Pittaway R; Clarke ML Rapid Asymmetric Transfer Hydroformylation (ATHF) of Disubstituted Alkenes Using Paraformaldehyde as a Syngas Surrogate. Chem.—Eur. J 2015, 21, 10645–10649. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (18).(a) Alemán J; Cabrera S; Maerten E; Overgaard J; Jørgensen KA Asymmetric Organocatalytic α-Arylation of Aldehydes. Angew. Chem., Int. Ed 2007, 46, 5520–5523. [DOI] [PubMed] [Google Scholar]; (b) Conrad JC; Kong J; Laforteza BN; MacMillan DWC Enantioselective α-Arylation of Aldehydes via Organo-SOMO Catalysis. An Ortho-Selective Arylation Reaction Based on an Open-Shell Pathway. J. Am. Chem. Soc 2009, 131, 11640–11641. [DOI] [PMC free article] [PubMed] [Google Scholar]; (c) Nicolaou KC; Reingruber R; Sarlah D; Bräse S Enantioselective Intramolecular Friedel-Crafts-Type α-Arylation of Aldehydes. J. Am. Chem. Soc 2009, 131, 2086–2087. [DOI] [PubMed] [Google Scholar]; (d) García-Fortanet J; Buchwald SL Asymmetric Palladium-Catalyzed Intramolecular α-Arylation of Aldehydes. Angew. Chem., Int. Ed 2008, 47, 8108–8111. [DOI] [PMC free article] [PubMed] [Google Scholar]; (e) Nareddy P; Mantilli L; Guénée L; Mazet C Atropoisomeric (P,N) Ligands for the Highly Enantioselective Pd-Catalyzed Intramolecular Asymmetric α-Arylation of α-Branched Aldehydes. Angew. Chem 2012, 124, 3892–3897. [DOI] [PubMed] [Google Scholar]; (f) Allen AE; MacMillan DWC Enantioselective α-Arylation of Aldehydes via the Productive Merger of Iodonium Salts and Organocatalysis. J. Am. Chem. Soc 2011, 133, 4260–4263. [DOI] [PMC free article] [PubMed] [Google Scholar]; (g) Bertuzzi G; Pecorari D; Bernardi L; Fochi M; Li R; Chemcomm, /; Communication, C.. An Organocatalytic Enantioselective Direct α-Heteroarylation of Aldehydes with Isoquinoline N-Oxides. Chem. Commun 2018, 54, 3977–3980. [DOI] [PubMed] [Google Scholar]; (h) Xu MM; Wang HQ; Mao YJ; Mei GJ; Wang SL; Shi F Cooperative Catalysis-Enabled Asymmetric α-Arylation of Aldehydes Using 2-Indolylmethanols as Arylation Reagents. J. Org. Chem 2018, 83, 5027–5034. [DOI] [PubMed] [Google Scholar]
- (19).Huang M; Zhang L; Pan T; Luo S Deracemization through Photochemical E/Z Isomerization of Enamines. Science 2022, 375, 869–874. [DOI] [PubMed] [Google Scholar]
- (20).For selected examples of base metal-catalyzed hydroformylation, see: (a) Hebrard F; Kalck P Cobalt-Catalyzed Hydroformylation of Alkenes: Generation and Recycling of the Carbonyl Species, and Catalytic Cycle. Chem. Rev 2009, 109, 4272–4282. [DOI] [PubMed] [Google Scholar]; (b) Kumar R; Chikkali SH Hydroformylation of Olefins by Metals Other than Rhodium. J. Organomet. Chem 2022, 960, No. 122231. [Google Scholar]; (c) Pospech J; Fleischer I; Franke R; Buchholz S; Beller M Alternative metals for homogeneous catalyzed hydroformylation reactions. Angew. Chem., Int. Ed 2013, 52, 2852–2872. [DOI] [PubMed] [Google Scholar]; (d) Geng H-Q; Meyer T; Franke R; Wu X-F Copper-Catalyzed Hydroformylation and Hydroxymethylation of Styrenes. Chem. Sci 2021, 12, 14937–14943. [DOI] [PMC free article] [PubMed] [Google Scholar]; (e) Hood DM; Johnson RA; Carpenter AE; Younker JM; Vinyard DJ; Stanley GG Highly Active Cationic Cobalt(II) Hydroformylation Catalysts. Science 2020, 367 (6477), 542–548. [DOI] [PubMed] [Google Scholar]; (f) Pandey S; Raj KV; Shinde DR; Vanka K; Kashyap V; Kurungot S; Vinod CP; Chikkali SH Iron Catalyzed Hydroformylation of Alkenes under Mild Conditions: Evidence of an Fe(II) Catalyzed Process. J. Am. Chem. Soc 2018, 140 (12), 4430–4439. [DOI] [PubMed] [Google Scholar]
- (21).(a) Liu RY; Buchwald SL CuH-Catalyzed Olefin Functionalization: From Hydroamination to Carbonyl Addition. Acc. Chem. Res 2020, 53, 1229–1243. [DOI] [PMC free article] [PubMed] [Google Scholar]; (b) Hirano K; Miura M Hydroamination, Aminoboration, and Carboamination with Electrophilic Amination Reagents: Umpolung-Enabled Regio- and Stereo-selective Synthesis of N-Containing Molecules from Alkenes and Alkynes. J. Am. Chem. Soc 2022, 144, 648–661. [DOI] [PubMed] [Google Scholar]; (c) Wang Y; He Y; Zhu S NiH-Catalyzed Functionalization of Remote and Proximal Olefins: New Reactions and Innovative Strategies. Acc. Chem. Res 2022, 55, 3519–3536. [DOI] [PubMed] [Google Scholar]; (d) Sun X-Y; Yao B-Y; Xuan B; Xiao L-J; Zhou Q-L Recent advances in nickel-catalyzed asymmetric hydrofunctionalization of alkenes. Chem. Catal 2022, 2, 3140–3162. [Google Scholar]; (e) Wang Z; Yin H; Fu GC Catalytic enantioconvergent coupling of secondary and tertiary electrophiles with olefins. Nature 2018, 563, 379–383. [DOI] [PMC free article] [PubMed] [Google Scholar]; (f) Li Y; Nie W; Chang Z; Wang J-W; Lu X; Fu Y Cobalt-catalysed enantioselective C(sp3)–C(sp3) coupling. Nat. Catal 2021, 4, 901–911. [Google Scholar]; (g) Yang ZP; Fu GC Convergent Catalytic Asymmetric Synthesis of Esters of Chiral Dialkyl Carbinols. J. Am. Chem. Soc 2020, 142, 5870–5875. [DOI] [PMC free article] [PubMed] [Google Scholar]; (h) Choi J; Fu GC Transition metal-catalyzed alkyl-alkyl bond formation: Another dimension in cross-coupling chemistry. Science 2017, 356, No. eaaf7230. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (22).(a) Dong Y; Shin K; Mai BK; Liu P; Buchwald SL Copper Hydride-Catalyzed Enantioselective Olefin Hydromethylation. J. Am. Chem. Soc 2022, 144, 16303–16309. [DOI] [PMC free article] [PubMed] [Google Scholar]; (b) Kutateladze DA; Mai BK; Dong Y; Zhang Y; Liu P; Buchwald SL Stereoselective Synthesis of Trisubstituted Alkenes via Copper Hydride-Catalyzed Alkyne Hydroalkylation. J. Am. Chem. Soc 2023, 145, 17557–17563. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (23).Gui Y-Y; Hu N; Chen X-W; Liao LL; 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]
- (24).Schuppe AW; Borrajo-Calleja GM; Buchwald SL Enantioselective Olefin Hydrocyanation without Cyanide. J. Am. Chem. Soc 2019, 141, 18668–18672. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (25).Dong Y; Schuppe AW; Mai BK; Liu P; Buchwald SL Confronting the Challenging Asymmetric Carbonyl 1,2-Addition Using Vinyl Heteroarene Pronucleophiles: Ligand-Controlled Regiodivergent Processes through a Dearomatized Allyl-Cu Species. J. Am. Chem. Soc 2022, 144, 5985–5995. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (26).Barbot F; Miginiac P Preparation of 3-alkenals and 3-alkynals by hydrolysis of the corresponding acetals. Synthesis 1983, 1983, 651–654. [Google Scholar]
- (27).Abdel-Magid AF; Carson KG; Harris BD; Maryanoff CA; Shah RD Reductive Amination of Aldehydes and Ketones with Sodium Triacetoxyborohydride. Studies on Direct and Indirect Reductive Amination Procedures. J. Org. Chem 1996, 61, 3849–3862. [DOI] [PubMed] [Google Scholar]
- (28).Friest JA; Maezato Y; Broussy S; Blum P; Berkowitz DB Use of a robust dehydrogenase from an archael hyperthermophile in asymmetric catalysis-dynamic reductive kinetic resolution entry into (S)-profens. J. Am. Chem. Soc 2010, 132, 5930–5931. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (29).For computational studies of transition metal-catalyzed hydroformylation, see: (a) Carbo JJ; Maseras F; Bo C; van Leeuwen PWNM Unraveling the origin of regioselectivity in rhodium diphosphine catalyzed hydroformylation. A DFT QM/MM study. J. Am. Chem. Soc 2001, 123, 7630–7637. [DOI] [PubMed] [Google Scholar]; (b) Decker SA; Cundari TR DFT Study of the Ethylene Hydroformylation Catalytic Cycle Employing a HRh(PH3)2(CO) Model Catalyst. Organometallics 2001, 20, 2827–2841. [Google Scholar]; (c) Rush LE; Pringle PG; Harvey JN Computational kinetics of cobalt-catalyzed alkene hydroformylation. Angew. Chem., Int. Ed 2014, 53, 8672–8676. [DOI] [PubMed] [Google Scholar]; (d) Lei M; Wang Z; Du X; Zhang X; Tang Y Asymmetric hydroformylation catalyzed by RhH(CO)2[(R,S)-Yanphos]: mechanism and origin of enantioselectivity. J. Phys. Chem. A 2014, 118, 8960–8970. [DOI] [PubMed] [Google Scholar]; (e) Jacobs I; de Bruin B; Reek JNH Comparison of the Full Catalytic Cycle of Hydroformylation Mediated by Mono- and Bis-Ligated Triphenylphosphine-Rhodium Complexes by Using DFT Calculations. ChemCatChem 2015, 7 (11), 1708–1718. [Google Scholar]; (f) Kegl T Computational Aspects of Hydroformylation. RSC Adv 2015, 5, 4304–4327. [Google Scholar]; (g) You C; Li X; Yang Y; Yang Y-S; Tan X; Li S; Wei B; Lv H; Chung L-W; Zhang X Silicon-Oriented Regio- and Enantioselective Rhodium-Catalyzed Hydroformylation. Nat. Commun 2018, 9, 2045. [DOI] [PMC free article] [PubMed] [Google Scholar]; (h) Liu J; Wei Z; Yang J; Ge Y; Wei D; Jackstell R; Jiao H; Beller M Tuning the Selectivity of Palladium Catalysts for Hydroformylation and Semihydrogenation of Alkynes: Experimental and Mechanistic Studies. ACS Catal 2020, 10, 12167–12181. [Google Scholar]
- (30).Xi Y; Hartwig JF Mechanistic studies of copper-catalyzed asymmetric hydroboration of alkenes. J. Am. Chem. Soc 2017, 139, 12758–12772. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (31).Baughman NN; Akhmedov NG; Petersen JL; Popp BV Experimental and Computational Analysis of CO2 Addition Reactions Relevant to Copper-Catalyzed Boracarboxylation of Vinyl Arenes: Evidence for a Phosphine-Promoted Mechanism. Organometallics 2021, 40, 23–37. [Google Scholar]
- (32).Luo Y; Li Y; Wu J; Xue XS; Hartwig JF; Shen Q Oxidative addition of an alkyl halide to form a stable Cu(III) product. Science 2023, 381, 1072–1079. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (33).Li C; Liu RY; Jesikiewicz LT; Yang Y; Liu P; Buchwald SL CuH-Catalyzed Enantioselective Ketone Allylationwith 1,3-Dienes: Scope, Mechanism, and Applications. J. Am. Chem. Soc 2019, 141, 5062–5070. [DOI] [PMC free article] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.