Skip to main content
Wiley Open Access Collection logoLink to Wiley Open Access Collection
. 2021 Oct 13;27(70):17559–17564. doi: 10.1002/chem.202103197

Synergistic Peptide and Gold Catalysis: Enantioselective Addition of Branched Aldehydes to Allenamides

Leo D M Nicholls 1, Helma Wennemers 1,
PMCID: PMC9293318  PMID: 34496089

Abstract

The combination of a peptide catalyst and a gold catalyst is presented for enantioselective addition reactions between branched aldehydes and allenamides. The two catalysts act in concert to provide γ,δ‐enamide aldehydes bearing a fully substituted, benzylic stereogenic center – a structural motif common in many natural products and therapeutically active compounds – with good yields and enantioselectivities. The reaction tolerates a variety of alkyl and alkoxy substituted aldehydes and the products can be elaborated into several chiral building blocks bearing either 1,4‐ or 1,5‐ functional group relationships. Mechanistic studies showed that the conformational features of the peptide are important for both the catalytic efficiency and stereochemistry, while a balance of acid/base additives is key for ensuring formation of the desired product over undesired side reactions.

Keywords: asymmetric catalysis, enamides, fully substituted stereogenic centers, gold, peptides


A synergistic combination of a tripeptide and a gold complex catalyzes the enantioselective addition between branched aldehydes and allenamides to furnish synthetically versatile γ,δ‐enamide aldehydes with a fully substituted stereogenic center. Mechanistic studies offered key insights into the role of the peptide and acid/base additives.

graphic file with name CHEM-27-17559-g006.jpg


Fully substituted, benzylic stereogenic centers are important building blocks of many natural products and biologically active compounds (Figure 1a). [1] These structural motifs have become the target of a variety of synthetic methodologies.[ 2 , 3 , 4 ] Among them, an attractive approach relies on amine catalyzed reactions between branched aldehydes and suitable carbon‐based electrophiles, which proceed through reactive enamine intermediates.[ 5 , 6 ] In recent years, the combination of a metal catalyst with an amine catalyst has expanded the scope to include otherwise unreactive electrophiles.[ 7 , 8 ] Interesting electrophiles to be used in these types of reactions are allenamides, [9] which can act as synthetically versatile C‐2 or C‐3 synthons through synthetic elaboration of the enamide moiety formed after C−C bond formation (Figure 1b). [10] Recently, two elegant studies by Mascareñas/López and González showcased the feasibility of reactions between branched aldehydes and allenamides catalyzed by prolinol silyl ethers and Au‐based catalysts. [11] These reports also pointed out that dual catalysis is challenging since the organocatalyst and the metal catalyst must act in concert and not interfere with each other.[ 11 , 12 ] We reasoned that amine‐based catalysts, in which the amino group is shielded and therefore less prone to interact with the metal catalyst, may offer advantages in this reaction.

Figure 1.

Figure 1

a) Selected therapeutically active compounds containing fully substituted benzylic stereogenic centers. b) The enantioselective addition of α‐branched aldehydes to allenamides. c) Tripeptide catalyst H‐dPro‐Pro‐Glu−NH2 and its conformation.

Our group developed the peptide H‐dPro‐Pro‐Glu‐NH2, which is a highly efficient and stereoselective organocatalyst for C−C bond formations that rely on the formation of enamine intermediates.[ 13 , 14 ] Detailed NMR spectroscopic analyses revealed that this peptide adopts a stable ground state conformation, in which the amine moiety forms a salt bridge with the glutamic acid side chain (Figure 1c). [15] We envisioned that this intramolecular coordination would disfavour non‐productive interactions between the peptide and the metal center and that as a result, peptides of the type H‐dPro‐Pro‐Xaa would be efficient catalysts for the targeted transformations.

Herein, we report the stereoselective addition of branched aldehydes to allenamides, catalyzed by a synergistic combination of peptide and gold catalysts. The reaction proceeds under mild conditions, tolerates a broad range of alkyl‐aryl and alkoxy‐aryl aldehydes and provides the addition products with good to excellent enantioselectivities. The aldehyde‐enamide products proved to be synthetically versatile for further elaboration into chiral building blocks including diols, lactones and piperidines. We also show that the conformational properties of the tripeptide catalyst and a fine balance between the reaction components (peptide, gold, acid, and base) are key for the efficiency of the reaction.

We started our investigations to evaluate the duality of peptide and metal catalysts using the reaction between 2‐phenylpropanal (1a) and allenesulfonamide 2a as a model reaction (Table 1). Firstly, we tested a range of different Cu, Au, Pd, Rh, and Ir complexes known to activate allenamides and/or allenes and used H‐dPro‐Pro‐Glu‐NH2 (4a) as peptide catalyst (Table 1, entries 1–5, Table S1). Many of the tested metal‐complexes did not allow access to the desired addition product or only in poor yields. Furthermore, dimerization of 2a to allenamide dimer 6 was observed.[ 16 , 17 ] These findings underscored the challenge of identifying metal complexes that are compatible with amine‐based catalysts. The best results were obtained with gold carbene complex 5a[ 18 , 19 ] that in combination with peptide 4a provided a promising conversion of 62 % to γ,δ‐enamide aldehyde 3a with an enantioselectivity of 47 % (Table 1, entry 2). N‐methyl morpholine (NMM) was included as a base since the trifluoroacetic acid (TFA) salt of the peptide was used.

Table 1.

Addition reaction between aldehydes 1a and 1b and allenesulfonamide 2a catalyzed by a combination of a peptide and a metal complex.

graphic file with name CHEM-27-17559-g008.jpg

Entry

4

5

1

Base

Solvent

Conv. [%][a]

ee (%)[b]

1

4a

1a

NMM

CH2Cl2

<5

2

4a

5a

1a

NMM

CH2Cl2

62

47

3

4a

5b

1a

NMM

CH2Cl2

22

45

4

4a

5c

1a

NMM

CH2Cl2

<5

5

4a

5d

1a

NMM

CH2Cl2

<5

6

4b

5a

1a

NMM

CH2Cl2

43

34

7

4c

5a

1a

NMM

CH2Cl2

16

25

8

4d

5a

1a

NMM

CH2Cl2

59

69

9

4d

5a

1a

NMM

CHCl3

64

80

10[c]

4d

5a

1a

NMM

CHCl3

81

79

11[c]

4d

5a

1a

DMAP

CHCl3

(80)

83

12

4d

5a

1b

NMM

CHCl3

(80)

88

13[d]

4e

5a

1b

iPr2NEt

CHCl3

(80)

95

Unless otherwise stated, reactions were carried out using an equal stoichiometry of aldehyde 1a/1b and allenesulfonamide 2a at a concentration of 0.2 M in CH2Cl2. [a] Conversion to 3a/3b estimated by 1H NMR spectroscopic analysis of the crude mixture using 1,3,5‐trimethoxybenzene as internal standard. Yields in parentheses. [b] Enantiomeric excess determined by chiral stationary phase SFC. [c] 2 equiv. aldehyde used. [d] 1.5 equiv. aldehyde used, at 0.05 M. NMM=N‐methylmorpholine, DMAP=4‐(dimethylamino)pyridine, Dipp=1,6‐diisopropylphenyl.

Variations of the peptide catalyst revealed that the stereochemistry, the CO2H group and its position as well as the length of the side chain of the C‐terminal amino acid are important for high yield and selectivity (Table 1, entries 2, 6, 7; Table S2). The DLL stereochemistry and the glutamic acid residue (peptide 4a), features that allow for the intramolecular salt bridge, provided the best results and the lowest amount of side products such as the allenamide dimer 6. Installation of an alkyl chain at the C‐terminus of the peptide (4d) improved the enantioselectivity of the reaction further (Table 1, entry 8). It is possible that the alkyl chain increases the local hydrophobicity and thereby stabilizes the salt bridge and improves the selectivity of the reaction. [20] Examination of other key reaction parameters revealed that the use of chloroform as solvent increased the enantioselectivity and an excess of the aldehyde improved the yield (Table 1, entries 9 and 10, Tables S5 and S6). TFA, the counterion of the peptide, proved to be important for the reactivity and enantioselectivity and better suited than other acids (e. g., benzoic acid, H2ClCCO2H, Ph2P(O)OH; Table S7). Finally, a screen of different bases, including DBU, DABCO, and Et3N, identified that DMAP is slightly superior to NMM (Table 1, entry 11, Table S8). Under these optimized conditions, γ,δ‐ensulfonamide aldehyde 3a was obtained in 80 % yield with an enantioselectivity of 83 %.

We then allowed aldehyde 1b, substituted with a methoxy instead of a methyl group, to react with allenesulfonamide 2a and obtained the product with an enantioselectivity of 88 % ee (Table 1, entry 12). This stereoselectivity could be further increased to 95 % by using tripeptide 4e that has an aspartic acid in place of a glutamic acid residue as well as through fine‐tuning of the reaction conditions (aldehyde:allenamide=1.5 : 1, 0.05 M, iPr2NEt as base) (Table 1, entry 13; Tables S10‐S16).

With optimized reaction conditions in hand, we proceeded to evaluate a scope of the reactions between allenamides and alkyl‐aryl aldehydes (Scheme 1) and alkoxy‐aryl aldehydes (Scheme 2). Pleasingly, the products from reactions with branched alkyl‐aryl aldehydes were obtained with good yields and enantioselectivities (3a, 3c3f, 60–80 % yield, 78–83 % ee; Scheme 1). Even carbamate protecting groups (Boc and Alloc) at the nitrogen in the allenamine were tolerated and provided the enamide products 3g and 3h with comparable stereoselectivity. Branched allensulfonamides resulted in only traces of product (3i), most likely due to the increased steric requirement.

Scheme 1.

Scheme 1

Scope of the peptide and gold catalyzed addition of branched alkyl‐aryl aldehydes to allenamides. Isolated yields, enantiomeric excess determined by chiral stationary phase SFC.

Scheme 2.

Scheme 2

Scope of the peptide and gold‐catalyzed addition of branched alkoxy‐aryl aldehydes to allenesulfonamides. Isolated yields, enantiomeric excesses determined by chiral stationary phase SFC.

Also aldehydes bearing a methoxy group and different electron rich and electron poor aryl moieties at Cα formed γ,δ‐ensulfonamide aldehydes with good yields (58‐80 %) and excellent enantioselectivities (90‐95 % ee) (3b, 3j3l; Scheme 2). Reactions with aldehydes bearing sterically more demanding alkoxy substituent proceeded with moderate to good yields but the high enantioselectivities were retained (3m3o; Scheme 2). [21] Single crystals of the major enantiomer of 3b suitable for X‐Ray diffraction could be grown and revealed the absolute stereochemistry at the fully substituted stereogenic center to be (S).[ 22 , 23 ]

We envisioned that the combination of enamide and aldehyde moieties would provide versatile sites for derivatisation of the addition products into a range of different chiral building blocks (Scheme 3). Indeed, ozonolysis of 3a, followed by reduction afforded 1,4‐diol 7, and acid catalyzed hydrolysis, followed by reduction afforded 1,5‐diol 8. Here, the enamide functional group serves as a C‐2 or C‐3 synthon, respectively. Reductive amination of 3b under neutral conditions proceeded to afford amino enamide 9. Reduction of 3b also proceeded smoothly to enamide alcohol 10, which upon ozonolysis and oxidation afforded lactone 11 through the corresponding intermediate lactol. Finally, piperidine 12 was accessed through acid promoted hydrolysis of 3a in the presence of benzyl amine, followed by reductive cyclisation using Na(CN)BH3. Such 3,3‐disubstituted piperidines have been previously studied as opioid agonists but synthetic access towards them has been limited. [2] In all reactions, only minor changes in the enantioselectivity were observed. These results highlight the versatility of γ,δ‐enamide aldehydes as building blocks to access a broad range of different chiral compounds.

Scheme 3.

Scheme 3

Derivatisation of products 3a and 3b: (a) O3, CH2Cl2/MeOH, −78 °C, then NaBH4. (b) p‐TSA, CH2Cl2, 0 °C, then NaBH4. (c) i) BnNH2, Na2SO4, CH2Cl2, rt, ii) NaBH4, MeOH. (d) NaBH4, MeOH, 0 °C. (e) i) O3, CH2Cl2, −78 °C, then SMe2, ii) CrO3/H2SO4, Me2CO, rt. (f) BnNH2, p‐TSA, CH2Cl2, 0 °C, then NaCN(BH3), MeOH, rt. Enantiomeric excesses determined by SFC or HPLC.

NMR spectroscopic studies with peptides 4d and 4e in deuterated chloroform, the solvent used for the reactions, revealed characteristic Nuclear Overhauser effects (NOEs) that are closely related to those previously observed for 4a. These signals are typical for a type I β‐turn (Figure 1c and 2a). Most importantly, these studies support that the conformation of 4d and 4e is similar to that of 4a. [15] The stereochemical outcome of the addition reaction is therefore consistent with reaction of the Au‐activated electrophile at the upper, less sterically hindered face of the enamine, whereby the aromatic residue is in conjugation with the proline nitrogen (Figure 2b).

Figure 2.

Figure 2

a) NOEs observed in the NMR spectra of peptides 4d and 4e. b) Stereochemical model of C−C bond formation.

We next aimed to further understand the driving forces and mechanism behind the catalytic efficiency of the reaction. Here, we focused on the addition of alkyl‐aryl aldehydes to allenamides and its catalytic mixture of peptide 4d⋅TFA, gold carbene complex 5a, and DMAP. We started by mixing these three components in a 1 : 1 : 1 ratio and analysed the mixture by 1H NMR spectroscopy. The spectrum shows one major set of signals relating to the peptide catalyst and two sets of signals in a ratio of ∼10 : 1 relating to the gold catalyst (Figure 3). The chemical shifts of the signals relating to the peptide indicate that peptide 4d exists in an equilibrium between the TFA‐salt and the “desalted” form, a finding that was corroborated by a titration of the peptide‐TFA salt with DMAP (Figure S5). The major set of signals for the gold species relates to Au‐DMAP complex 13 and the minor set of signals to Au‐TFA complex 14. In addition, a small amount of the salt formed by DMAP and TFA is visible. These assignments were supported by comparison with authentic samples by 1H and 19F NMR spectroscopy and, in case of 13, further confirmed by crystal structure analysis (Figures S2‐S4).[ 17 , 22 ] These experiments demonstrate that an equilibrium exists at the gold center, between the DMAP and TFA ligands. No complexation is, however, observed between the peptide and the gold center in the catalytic mixture. In contrast, a complex mixture was observed between 4d and 5a in the absence of DMAP and TFA (Figure S3). Together with the intramolecular salt bridge within the structure of the peptide, these additives thus help to separate the peptide and gold catalysts and disfavour non‐productive interactions. These findings suggest that acid and base additives are important for the efficiency of the catalytic reaction.

Figure 3.

Figure 3

1H NMR spectrum of a 1 : 1 : 1 mixture of 4d⋅TFA, 5a and DMAP.

A series of experiments in which TFA and/or DMAP were excluded or added in excess to the reaction mixture further supported this conclusion (Table 2). In the absence of TFA hardly any product formed and in the absence of DMAP or both additives only small amounts of γ,δ‐enamide aldehyde 3a formed along with considerable amounts of allenamide dimer 6 (Table 2, entries 1–3). Addition of a two‐fold excess of DMAP with respect to the Au‐complex and the peptide prevented product formation (Table 2, entry 4) Thus, the absence of DMAP favours dimerization of 2a to form 6 over reaction to form 3a. An excess DMAP or the absence of TFA has an inhibitory effect.

Table 2.

Addition reaction between aldehydes 1a and 1b and allenamide 2a catalyzed by a combination of a peptide and a metal complex.

graphic file with name CHEM-27-17559-g003.jpg

Entry

Deviation from conditions

3a [%][a]

6 [%][a]

1

No TFA, no DMAP

10

21

2

No DMAP

9

29

3

No TFA

6

4

Additional DMAP (5 mol%)

Reaction conditions: aldehyde 1a: allenamide 2a=2 : 1, concentration=0.2 M. [a] Yield estimated by NMR using 1,3,5‐trimethoxybenzene as internal standard.

These findings support the following catalytic cycle (Scheme 4): The peptide catalyst forms an enamine intermediate with the aldehyde that then reacts in the stereochemistry determining step with an electrophilic π‐complex (I) formed through ligand exchange of the Au‐DMAP (or Au‐TFA) complex with the allenamide. Upon C−C bond formation, vinyl gold‐iminium intermediate II is formed, which after protodeauration and hydrolysis furnishes the γ,δ‐enamide aldehyde (3) and regenerates the gold and peptide catalysts, respectively.

Scheme 4.

Scheme 4

Proposed mechanistic cycle. [Au]=(IPr)Au+.

This mechanistic rationale is in agreement with the observed effects of TFA and DMAP on the reaction. The reduced reaction speed upon increasing the concentration of DMAP likely stems from the equilibrium between the electrophilic π‐complex (I) and Au‐DMAP complex 13. The strong coordinating ability of DMAP likely favours gold‐complex 13 over intermediate I. Increasing the amount of DMAP therefore shifts the equilibrium towards 13 and thus decreases the amount of intermediate I and therefore the overall reaction efficiency. Lowering the concentration of DMAP has the opposite effect, shifting the equilibrium to favour intermediate I. Upon formation of I, two reaction pathways can take place (A (green) and B (red) in Scheme 4). In pathway A, I reacts with the enamine, leading to product 3. Alternatively, in pathway B, reaction of I with a second equivalent of allenamide 2 takes place, leading to allenamide dimer 6 or higher oligomers. The DMAP thus plays a crucial role in modulating the amount of I, acting to perfectly balance the production of intermediate I, so that pathway A is favoured over other undesired reaction pathways. TFA has a similar effect as increasing its amount should facilitate decomplexation of the DMAP ligand from the gold center by protonation of the pyridine nitrogen. A lower amount of DMAP or a higher amount of TFA should therefore increase the reaction rate. This expectation proved to be correct when performing the reaction with either 3 mol% of DMAP or 7.5 mol% of TFA. In both cases the reaction proceeded faster without sacrificing the enantioselectivity, although with reduced conversion to the γ,δ‐enamide aldehyde (3a) (Figure S7).

In summary, we have disclosed the enantioselective addition of branched aldehydes to allenamides, catalyzed by a synergistic combination of a peptide and a gold catalyst. The reaction furnishes α‐alkylated aldehydes bearing either a quaternary or a fully substituted carbon stereogenic center in good to very good enantioselectivities and proceeds under mild conditions. The aldehyde‐enamide products can be elaborated into a range of different chiral building blocks including diols, lactones and piperidines. Acid and base additives proved important to control the formation of the electrophilic gold intermediate and its reaction with the enamine intermediate rather than undesired formation of allenamide dimers. Structural modifications as well as in‐solution NMR spectroscopic studies showed that the conformational properties of the peptidic catalyst are important for the catalytic efficiency and the selectivity of the reaction. These data demonstrate the robustness and chemoselectivity of peptide catalysts of the type H‐Pro‐Pro‐Xaa, even in increasingly complex environments with metal catalysts. They lay the foundation for the use of peptides in even more complex synergistic catalytic processes, reminiscent of catalytic cascades used by nature.

Conflict of interest

The authors declare no conflict of interest.

Supporting information

As a service to our authors and readers, this journal provides supporting information supplied by the authors. Such materials are peer reviewed and may be re‐organized for online delivery, but are not copy‐edited or typeset. Technical support issues arising from supporting information (other than missing files) should be addressed to the authors.

Supporting Information

Acknowledgements

We thank the European Union's Horizon 2020 research and innovation programme (Grant No 862081) as well as the Swiss National Science Foundation (Grant No 200020_188729) for financial support and the NMR Service at the Laboratory of Organic Chemistry at ETH Zürich for assistance with the kinetic measurements. Open access funding provided by Eidgenossische Technische Hochschule Zurich.

L. D. M. Nicholls, H. Wennemers, Chem. Eur. J. 2021, 27, 17559.

References

  • 1. 
  • 1a. Hjorth S., Waters S., Waters N., Tedroff J., Svensson P., Fagerberg A., Edling M., Svanberg B., Ljung E., Gunnergren J., McLean S. L., Grayson B., Idris N. F., Neill J. C., Sonesson C., J. Pharmacol. Exp. Ther. 2020, JPET-AR-2020-000037; [DOI] [PubMed] [Google Scholar]
  • 1b. Rasmussen K., Calligaro D. O., Czachura J. F., Dreshfield-Ahmad L. J., Evans D. C., Hemrick-Luecke S. K., Kallman M. J., Kendrick W. T., Leander J. D., Nelson D. L., Overshiner C. D., Wainscott D. B., Wolff M. C., Wong D. T., Branchek T. A., Zgombick J. M., Xu Y.-c., J. Pharmacol. Exp. Ther. 2000, 294, 688; [PubMed] [Google Scholar]
  • 1c. Macchia B., Macchia M., Martinelli A., Martinotti E., Orlandini E., Romagnoli F., Scatizzi R., Eur. J. Med. Chem. 1997, 32, 231–240; [Google Scholar]
  • 1d. Cheng A., Uyeno E., Polgar W., Toll L., Lawson J. A., DeGraw J. I., Loew G., Camerman A., Camerman N., J. Med. Chem. 1986, 29, 531–537. [DOI] [PubMed] [Google Scholar]
  • 2.For a recent review on enantioselective alkylations of enolates, see: T. B. Wright, P. A. Evans, Chem. Rev. 2021, 121, 9196–9242. [DOI] [PubMed]
  • 3.For reviews on the stereoselective synthesis of quaternary stereogenic centers, see:
  • 3a. Christoffers J., Baro A., Adv. Synth. Catal. 2005, 347, 1473–1482; [Google Scholar]
  • 3b. Corey E. J., Guzman-Perez A., Angew. Chem. Int. Ed. 1998, 37, 388–401; [DOI] [PubMed] [Google Scholar]; Angew. Chem. 1998, 110, 402–415; [Google Scholar]
  • 3c. Douglas C. J., Overman L. E., Proc. Natl. Acad. Sci. USA 2004, 101, 5363; [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 3d. Fuji K., Chem. Rev. 1993, 93, 2037–2066; [Google Scholar]
  • 3e. Hawner C., Alexakis A., Chem. Commun. 2010, 46, 7295–7306; [DOI] [PubMed] [Google Scholar]
  • 3f. Marek I., Minko Y., Pasco M., Mejuch T., Gilboa N., Chechik H., Das J. P., J. Am. Chem. Soc. 2014, 136, 2682–2694. [DOI] [PubMed] [Google Scholar]
  • 4.For examples of the stereoselective synthesis of tertiary ethers, see:
  • 4a. Uozumi Y., Kato K., Hayashi T., J. Am. Chem. Soc. 1997, 119, 5063–5064; [Google Scholar]
  • 4b. Trost B. M., Shen H. C., Dong L., Surivet J.-P., J. Am. Chem. Soc. 2003, 125, 9276–9277; [DOI] [PubMed] [Google Scholar]
  • 4c. Zhang K., Peng Q., Hou X.-L., Wu Y.-D., Angew. Chem. Int. Ed. 2008, 47, 1741–1744; [DOI] [PubMed] [Google Scholar]; Angew. Chem. 2008, 120, 1765–1768; [Google Scholar]
  • 4d. Tanaka S., Seki T., Kitamura M., Angew. Chem. Int. Ed. 2009, 48, 8948–8951; [DOI] [PubMed] [Google Scholar]; Angew. Chem. 2009, 121, 9110–9113; [Google Scholar]
  • 4e. Xie J., Guo W., Cai A., Escudero-Adán E. C., Kleij A. W., Org. Lett. 2017, 19, 6388–6391; [DOI] [PubMed] [Google Scholar]
  • 4f. Shibatomi K., Kotozaki M., Sasaki N., Fujisawa I., Iwasa S., Chem. Eur. J. 2015, 21, 14095–14098; [DOI] [PubMed] [Google Scholar]
  • 4g. Cefalo D. R., Kiely A. F., Wuchrer M., Jamieson J. Y., Schrock R. R., Hoveyda A. H., J. Am. Chem. Soc. 2001, 123, 3139–3140; [Google Scholar]
  • 4h. Wong O. A., Shi Y., Chem. Rev. 2008, 108, 3958–3987; [DOI] [PubMed] [Google Scholar]
  • 4i. Denmark S. E., Kornfilt D. J. P., Vogler T., J. Am. Chem. Soc. 2011, 133, 15308–15311; [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 4j. Corbett M. T., Johnson J. S., Chem. Sci. 2013, 4, 2828–2832; [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 4k. Zhang S.-Y., Zhang F.-M., Tu Y.-Q., Chem. Soc. Rev. 2011, 40, 1937–1949. [DOI] [PubMed] [Google Scholar]
  • 5. 
  • 5a. Brown A. R., Kuo W.-H., Jacobsen E. N., J. Am. Chem. Soc. 2010, 132, 9286–9288; [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 5b. Lalonde M. P., Chen Y., Jacobsen E. N., Angew. Chem. Int. Ed. 2006, 45, 6366–6370; [DOI] [PubMed] [Google Scholar]; Angew. Chem. 2006, 118, 6514–6518; [Google Scholar]
  • 5c. Mase N., Thayumanavan R., Tanaka F., Barbas C. F., Org. Lett. 2004, 6, 2527–2530; [DOI] [PubMed] [Google Scholar]
  • 5d. Witten M. R., Jacobsen E. N., Org. Lett. 2015, 17, 2772–2775; [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 5e. Xue F., Liu L., Zhang S., Duan W., Wang W., Chem. Eur. J. 2010, 16, 7979–7982; [DOI] [PubMed] [Google Scholar]
  • 5f. Yu F., Jin Z., Huang H., Ye T., Liang X., Ye J., Org. Biomol. Chem. 2010, 8, 4767–4774; [DOI] [PubMed] [Google Scholar]
  • 5g. Quintard A., Belot S., Marchal E., Alexakis A., Eur. J. Org. Chem. 2010, 2010, 927–936; [Google Scholar]
  • 5h. List B., Čorić I., Grygorenko O. O., Kaib P. S. J., Komarov I., Lee A., Leutzsch M., Chandra Pan S., Tymtsunik A. V., van Gemmeren M., Angew. Chem. Int. Ed. 2014, 53, 282–285; [DOI] [PubMed] [Google Scholar]; Angew. Chem. 2014, 126, 286–289. [Google Scholar]
  • 6. Mukherjee S., Yang J. W., Hoffmann S., List B., Chem. Rev. 2007, 107, 5471–5569. [DOI] [PubMed] [Google Scholar]
  • 7. 
  • 7a. Shao Z., Zhang H., Chem. Soc. Rev. 2009, 38, 2745–2755; [DOI] [PubMed] [Google Scholar]
  • 7b. Du Z., Shao Z., Chem. Soc. Rev. 2013, 42, 1337–1378; [DOI] [PubMed] [Google Scholar]
  • 7c. Deng Y., Kumar S., Wang H., Chem. Commun. 2014, 121, 9196–9242; [Google Scholar]
  • 7d. Afewerki S., Córdova A., Chem. Rev. 2016, 116, 13512–13570; [DOI] [PubMed] [Google Scholar]
  • 7e. Afewerki S., Córdova A., Top. Curr. Chem. 2019, 377, 38; [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8.For example, see:
  • 8a. Krautwald S., Sarlah D., Schafroth M. A., Carreira E. M., Science 2013, 340, 1065; [DOI] [PubMed] [Google Scholar]
  • 8b. Mo X., Hall D. G., J. Am. Chem. Soc. 2016, 138, 10762–10765; [DOI] [PubMed] [Google Scholar]
  • 8c. Zhou H., Wang Y., Zhang L., Cai M., Luo S., J. Am. Chem. Soc. 2017, 139, 3631–3634; [DOI] [PubMed] [Google Scholar]
  • 8d. Cruz F. A., Dong V. M., J. Am. Chem. Soc. 2017, 139, 1029–1032; [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8e. Jiang G., List B., Angew. Chem. Int. Ed. 2011, 50, 9471–9474; [DOI] [PubMed] [Google Scholar]; Angew. Chem. 2011, 123, 9643–9646. [Google Scholar]
  • 9.For reviews on allenamides, see:
  • 9a. Manoni E., Bandini M., Eur. J. Org. Chem. 2016, 3135–3142; [Google Scholar]
  • 9b. Mascareñas J. L., Varela I., López F., Acc. Chem. Res. 2019, 52, 465–479. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10. 
  • 10a. Carbery D. R., Org. Biomol. Chem. 2008, 6, 3455–3460; [DOI] [PubMed] [Google Scholar]
  • 10b. Yang B., Zhai X., Feng S., Hu D., Deng Y., Shao Z., Org. Lett. 2019, 21, 330–334; [DOI] [PubMed] [Google Scholar]
  • 10c. Yang X., Toste F. D., Chem. Sci. 2016, 7, 2653–2656. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11. 
  • 11a. Fernández-Casado J., Nelson R., Mascareñas J. L., López F., Chem. Commun. 2016, 52, 2909–2912; [DOI] [PubMed] [Google Scholar]
  • 11b. Ballesteros A., Morán-Poladura P., González J. M., Chem. Commun. 2016, 52, 2905–2908. [DOI] [PubMed] [Google Scholar]
  • 12. Loh C. C. J., Enders D., Chem. Eur. J. 2012, 18, 10212–10225. [DOI] [PubMed] [Google Scholar]
  • 13. 
  • 13a. Wiesner M., Revell J. D., Tonazzi S., Wennemers H., J. Am. Chem. Soc. 2008, 130, 5610–5611; [DOI] [PubMed] [Google Scholar]
  • 13b. Wiesner M., Revell J. D., Wennemers H., Angew. Chem. Int. Ed. 2008, 47, 1871–1874; [DOI] [PubMed] [Google Scholar]; Angew. Chem. 2008, 120, 1897–1900; [Google Scholar]
  • 13c. Duschmalé J., Wennemers H., Chem. Eur. J. 2012, 18, 1111–1120; [DOI] [PubMed] [Google Scholar]
  • 13d. Kastl R., Wennemers H., Angew. Chem. Int. Ed. 2013, 52, 7228–7232; [DOI] [PubMed] [Google Scholar]; Angew. Chem. 2013, 125, 7369–7373; [Google Scholar]
  • 13e. Grünenfelder C. E., Kisunzu J. K., Wennemers H., Angew. Chem. Int. Ed. 2016, 55, 8571–8574; [DOI] [PubMed] [Google Scholar]; Angew. Chem. 2016, 128, 8713–8716; [Google Scholar]
  • 13f. Schnitzer T., Wennemers H., J. Am. Chem. Soc. 2017, 139, 15356–15362; [DOI] [PubMed] [Google Scholar]
  • 13g. Mohler J. S., Schnitzer T., Wennemers H., Chem. Eur. J. 2020, 26, 15623–15628; [DOI] [PubMed] [Google Scholar]
  • 13h. Schnitzer T., Budinska A., Wennemers H., Nat. Catal. 2020, 3, 143–147. [Google Scholar]
  • 14.For a recent review on peptide catalysis, see: A. J. Metrano, C. R. Shugrue, B. Kim, A. J. Chinn, E. A. Stone, S. J. Miller, Chem. Rev. 2020, 120, 11479–11615. For tandem reactions of metal and peptide catalysis, see of [DOI] [PMC free article] [PubMed]
  • 14a. Akagawa K., Fujiwara T., Sakamoto S., Kudo K., Org. Lett. 2010, 12, 1804–1807; [DOI] [PubMed] [Google Scholar]
  • 14b. Akagawa K., Fujiwara T., Sakamoto S., Kudo K., Chem. Commun. 2010, 46, 8040–8042. [DOI] [PubMed] [Google Scholar]
  • 15. Rigling C., Kisunzu J. K., Duschmalé J., Häussinger D., Wiesner M., Ebert M.-O., Wennemers H., J. Am. Chem. Soc. 2018, 140, 10829–10838. [DOI] [PubMed] [Google Scholar]
  • 16. Suárez-Pantiga S., Hernández-Díaz C., Piedrafita M., Rubio E., González J. M., Adv. Synth. Catal. 2012, 354, 1651–1657. [Google Scholar]
  • 17.Additional details can be found in the Supporting Information.
  • 18. de Frémont P., Stevens E. D., Fructos M. R., Mar Díaz-Requejo M., Pérez P. J., Nolan S. P., Chem. Commun. 2006, 2045–2047. [DOI] [PubMed] [Google Scholar]
  • 19.For a review on gold catalysis, see: Dorel R., Echavarren A. M., Chem. Rev. 2015, 115, 9028–9072. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20. Schnitzer T., Ganzoni R. L., Wennemers H., Tetrahedron 2020, 76, 131184. [Google Scholar]
  • 21.The use of a prolinolsilyl ether instead of the peptide catalyst provided the product under the same conditions in yields of <20 %, Tables S20 and S21. Of note, previous reports that used prolinolsilyl ethers for the target the reaction used a higher catalyst loading and CH3CN or toluene as solvents, solvents that were suboptimal for the catalysis with the peptide 4d. The results are therefore not directly comparable.
  • 22.Deposition Number 2096749 (3b), 2096750 (13) contain the supplementary crystallographic data for this paper. These data are provided free of charge by the joint Cambridge Crystallographic Data Centre and Fachinformationszentrum Karlsruhe Access Structures service.
  • 23.Comparison of the analytical data of the 1,4-diol 7 with that of previously reported data strongly supports that the products formed by the alkyl-aryl aldehydes are (R)-configured (see Scheme S1). Note, although the absolute configuration of γ,δ-enamide aldehydes bearing alkyl versus alkoxy substituents is different following the CIP rules, the relative orientation of the substituents about the fully substituted stereogenic center in both products is the same.

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

As a service to our authors and readers, this journal provides supporting information supplied by the authors. Such materials are peer reviewed and may be re‐organized for online delivery, but are not copy‐edited or typeset. Technical support issues arising from supporting information (other than missing files) should be addressed to the authors.

Supporting Information


Articles from Chemistry (Weinheim an Der Bergstrasse, Germany) are provided here courtesy of Wiley

RESOURCES