Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2022 Jun 23.
Published in final edited form as: J Am Chem Soc. 2021 Jun 2;143(24):8962–8969. doi: 10.1021/jacs.1c03178

A Transient Directing Group Strategy Enables Enantioselective Multicomponent Organofluorine Synthesis

Zhonglin Liu 1, Lucas J Oxtoby 1, Mingyu Liu 1, Zi-Qi Li 1, Van T Tran 1, Yang Gao 1, Keary M Engle 1,*
PMCID: PMC8222155  NIHMSID: NIHMS1709637  PMID: 34077182

Abstract

The vicinal fluorofunctionalization of alkenes represents an expedient strategy for converting feedstock olefins into valuable fluorinated molecules and as such has garnered significant attention from the synthetic community; however, current methods remain limited in terms of scope and selectivity. Here we report the site-selective palladium-catalyzed three-component coupling of alkenylbenzaldehydes, arylboronic acids, and N-fluoro-2,4,6-trimethylpyridinium hexafluorophosphate facilitated by a transient directing group. The synthetically enabling methodology constructs vicinal stereocenters with excellent regio-, diastereo-, and enantioselectivities, forging products that map onto bioactive compounds.

Graphical Abstract

graphic file with name nihms-1709637-f0003.jpg


The incorporation of carbon–fluorine (C–F) bonds into drug molecules can often improve their pharmacokinetic properties, including increasing oral bioavailability, protein binding affinities, and metabolic stability, especially in the case of replacement of benzylic C–H bonds prone to metabolic oxidation.1,2 As such, the development of strategies that enable the enantioselective formation of C–F bonds has become a major research area of both industrial and academic importance in recent years.35 In particular, intermolecular 1,2-carbofluorination of alkenes is an attractive transformation as it allows for the conversion of alkene feedstocks into fluorinated molecules with potential applications in the pharmaceutical, agrochemical, and material sectors6; however, this type of transformation remains challenging to execute due to issues with regio-, stereo-, and chemoselectivity. In early work, the groups of Ma,7 Gagné,8 Alexakis,9 and Gouverneur10 reported pioneering examples of asymmetric fluorocyclizations of prochiral alkenes, in which a functional group tethered to the alkene reacts in the cyclization process (Figure 1A).

Figure 1.

Figure 1.

(A) Pioneering examples of asymmetric fluorocyclizations of alkenes. (B) State of the art in palladium-catalyzed alkene arylfluorination. (C) General depiction of an arylfluorination facilitated by an in situ formed imine using a chiral TDG strategy.

More recently, Toste and coworkers have reported an elegant series of intermolecular (three-component) asymmetric arylfluorination reactions to construct chiral benzyl fluorides using palladium/N,N-ligand systems. This strategy has been used for both 1,1-arylfluorination (where regioselectivity is governed by substrate electronics)11,12 and 1,2-arylfluorination (where selectivity is governed by substrate directivity)13 as depicted in Figure 1B. While the aforementioned work represents a great deal of progress, significant limitations remain. Palladium-catalyzed arylfluorination reactions are sensitive to alkene substitution patterns; for instance, disubstituted alkenes require double activation to enhance reactivity,14 and no existing methods are able to construct fully substituted C(sp3)–F or C(sp3)–Ar stereocenters. Additionally, achieving high levels of pathway selectivity for a given substrate class (favoring 1,2-arylfluorination over 1,1-arylfluorination, β-hydride elimination, or other side reactions) often requires extensive ligand optimization and the use of potentially synthetically restrictive directing groups.15

With the previous efforts in mind, we wondered if we could address these issues by implementing a chiral transient directing group (TDG) strategy (Figure 1C). The viability of catalytic TDGs has previously been established in several mechanistically distinct transition-metal catalyzed reactions, including notably in the field of C–H activation;16 however, the scope of transiently directed asymmetric alkene functionalizations remains quite limited, and 1,2-difunctionalization reactions using a TDG approach remain unknown. Herein, we report a highly enantioselective 1,2-arylfluorination of alkenyl benzaldehydes that is able to form two vicinal chiral centers, including fully substituted C(sp3)–F and C(sp3)–Ar stereocenters, in synthetically useful yields with broad functional group tolerance.

To reduce this idea to practice, we based our initial reaction design on our recently reported enantioselective reductive Heck hydroarylation of alkenyl benzaldehydes using an amino acid TDG.17 In our previous work, a stabilized alkylpalladium(II) intermediate is intercepted with formate, which decarboxylates to generate an alkylpalladium(II)–hydride species that subsequently undergoes reductive elimination. In the case of the envisioned transiently directed arylfluorination, the stabilized alkylpalladium(II) intermediate would react with a fluorinating oxidant (an [F+] reagent) to generate a palladium(IV) species (which can more readily undergo stereoretentive C–F reductive elimination than a palladium(II) fluoride).18 This seemingly simple extension is fraught with challenges, including undesired oxidation of the native aldehyde functional handle by [F+], competitive homocoupling of arylboronic acids in the presence of palladium(II) and oxidant, and deleterious interactions between the TDG and [F+].

In a series of pilot experiments, we found that L-tert-leucine (the optimal TDG in our previous system) did not lead to product formation, suggesting that a different TDG design was needed to support palladium(IV) formation. In their study on enantioselective C–H fluorination of electron-deficient benzaldehydes, Yu and coworkers found that switching from an LX-type amino acid to an L2-type α-amino amide TDG promoted C–F reductive elimination through formation of a pentacoordinate cationic Pd(IV) complex.19 We reasoned our system would benefit from the same effect and carried out a new screen using a library of TDGs that offered the potential of L2-type binding after aldehyde condensation (see Supplementary Figure S1). An initial hit was observed reacting alkene starting material 1 with N-fluoro-2,4,6-trimethylpyridinium salt ([F+]) and phenylboronic acid in the presence of a palladium(II) catalyst, previously unreported TDG-A, silver fluoride additive, and water in a 2:1 mixture of DCM/MeCN. A variety of unproductive side reactions were observed, including formation of palladium black and decomposition of the [F+] reagent and benzaldehyde SM. Other [F+] oxidants including Selectfluor and NFSI resulted in oxidation to the carboxylic acid. Because it was expected that most of the components would have strong interactions with each other that would impact the final yield of the reaction, our system appeared better suited for optimization via Design of Experiments (DoE) as opposed to typical “One Variable At a Time” (OVAT) screening.20,21

We elected to use a modified Definitive Screening Design (DSD) (Table 1), which allowed us to develop a linear regression model that describes the sensitivity of a response (in this case reaction yield) to a variety of input parameters with continuous levels (reagent loadings)22. The high- and low-end values for each reagent loading in the subsequent experiments were set based on what we expected to be the most extreme values the reaction would tolerate. We then ran just 18 experiments using values within those ranges to train a model, which subsequently predicted conditions that more than doubled our initial yield. While the arylboronic acid loading of 4.67 equivalents is relatively high, we deemed this acceptable in our system, as the arylboronic acid was not among the most valuable components of the reaction.23

Table 1.

Optimization of reaction conditions.

graphic file with name nihms-1709637-t0004.jpg

Having optimized the conditions, we began investigating the scope of alkene substitution (Table 2). The reaction with 1,2,2-trisubstituted alkenes to form quaternary carbon–aryl bonds proceeded in moderate to good yields with excellent enantioselectivity and broad functional group tolerance. Successful reaction of 1,1,2-trisubstituted alkene substrates with our conditions would form tertiary C(sp3)–F stereocenters, which would be a valuable addition to previously reported methods;24 however, this requires an unprecedented asymmetric formation of a fully substituted C(sp3)–F stereocenter through C–F reductive elimination. Gratifyingly, the reaction with 1,1,2-trisubstituted alkenes was able to form tertiary C(sp3)–F stereocenters, with somewhat lower, but still high enantioselectivity ranging from 90–96% ee. Next, we explored the effects of other alkene and benzaldehyde substitutions on the reaction. Both electron-rich and electron-poor benzaldehydes gave good yields with excellent enantioselectivity, with performance being similar between E- and Z-isomers of the alkene starting material (2y to 3y and 2z to 3z, respectively). Unfortunately, the method is ineffective for alkenes that are tetrasubstituted or whose functionalization results in dearomatization (see Supplementary Figure S5).

Table 2.

Alkenylbenzaldehyde scope of the arylfluorination reaction.

graphic file with name nihms-1709637-t0005.jpg

Reactions were carried out on 0.1 mmol scale. All reported yields are isolated yields. The use of TDG-A or TDG-B is designated by a superscript A or B, respectively.

We explored the scope of arylboronic acids with 1a as the model alkene substrate (Table 3). Both electron-rich and electron-poor para-substituted boronic acids afforded the desired products in good yields with excellent enantioselectivity. Arylboronic acids with electron-withdrawing meta-substituents (4f) performed markedly worse than those with electron-donating substituents (4g), requiring either prolonged reaction time or higher temperatures. The reaction did proceed with benzofuran and benzodioxole boronic acids (giving 4h and 4i, respectively), but did not tolerate other heterocycles, including substituted pyridines and pyrazoles (presumably due to competitive binding to the metal center). Additional limitations include both alkenyl and alkyl boronic acids. In order to demonstrate the potential synthetic utility of the reaction, multiple diversifications were performed. Both oxidation (5a) and reduction (5c) of the aldehyde proceeded in high yields while maintaining the high ee of 4a. Olefination (5b) and decarbonylation (5d) gave the desired products in moderate yields with little to no erosion in ee.

Table 3.

Arylboronic acid scope of the arylfluorination reaction and synthetic derivations.

graphic file with name nihms-1709637-t0006.jpg

Reactions for the arylboronic acid scope and synthetic diversifications were run on 0.1 mmol scale unless otherwise noted. See SI for details regarding scale and specific conditions for the synthesis of MCJ001F.

To further demonstrate the synthetically enabling nature of this method, we targeted the synthesis of the bioactive compound MCJ001F.25 The reported patent route to the (R,R)-stereoisomer involves a chiral resolution of a racemic intermediate and late-stage separation of diastereomers in a 10-step sequence. In contrast, our procedure is both enantio- and diastereoselective, affording MCJ001F-RR in fewer steps from commercially available materials and in overall higher yield. By starting from the opposite alkene isomer (Z-configured), we were also able to prepare the (R,S)-stereoisomer in higher overall yield than the reported patent route (see Supplementary Figure S7).

In an attempt to gain insight into the reaction mechanism, we performed several control reactions and kinetic experiments (Figure 2). First, we confirmed that the TDG was essential by excluding it from a standard reaction with 7f as well as subjecting ester 7e, which cannot condense with the TDG, to standard conditions. In both cases, only recovered starting material was observed. Next, we confirmed the reaction does not occur through C–H activation by subjecting substrate 7g to our standard conditions and observing that it was unreactive. The role of AgF was also briefly considered. While the reaction of model substrate 1a does proceed without silver fluoride, the reaction only reaches roughly 40% yield by 1H NMR, with the remaining mass balance of 1a being decomposed material. We observe the formation of palladium black when stirring a mixture of 1a, various loadings of TDG-A, and Pd(MeCN)4(BF4)2 (10 mol%) in the standard solvent conditions, raising the possibility of catalyst deactivation in the reaction system. Palladium black was qualitatively observed forming over the course of roughly 6 hours at 35 °C for loadings of TDG higher than 75 mol%. Competitive homocoupling of the boronic acid (confirmed by GC-MS) helps explain the high amount of boronic acid required. It may also indicate a potential beneficial effect of Ag(I) for reoxidation of inactive Pd(0).

Figure 2.

Figure 2.

Summary of control experiments and kinetics investigation (see SI for experimental details and all data) and proposed catalytic cycle.

Experiments aimed at establishing the robustness of the palladium catalyst and the concentration dependencies of the reaction components were carried out according to the “same-excess” and “different-excess” protocols of Reaction Progress Kinetic Analysis (RPKA), respectively.26 RPKA allows for visual analysis of the kinetic data over the entire time course of the reaction at synthetically relevant conditions. The lack of overlay between these profiles indicates that product inhibition or catalyst deactivation is occurring.27 These possibilities can be discerned by running a third experiment with the amount of product generated by the reaction until this reaction point added. This reaction is identical by composition, but the catalyst has completed fewer turnovers; therefore, the lack of overlay in this case is indicative of mild catalyst deactivation.28

Having confirmed the presence of catalyst deactivation, initial rates from reaction progress profiles were employed to probe concentration dependencies of the reaction components. A series of “different-excess” experiments showed the reaction to have a positive dependence on [Pd] and a negative rate dependence on [TDG]. Studies established the absence of a nonlinear effect of TDG ee, showing that product ee varies linearly with TDG ee. This, coupled with an observed first-order dependence on palladium, confirms the absence of Pd dimers or other higher order Pd species either on or off the cycle, but does not preclude the possibility of monomeric off-cycle species.29 The reaction was shown to be zero-order in water, silver fluoride, and alkene. The initial rates for reactions with lower [PhB(OH)2] or lower [F+] were reduced compared to the reaction run under standard conditions, suggesting a positive rate dependence; however, when comparing increased concentrations of these reagent during the “different excess” experiments, the resulting rates showed an apparent zero-order dependence in [F+] and boronic acid, pointing to saturation kinetics in both of these components.

Taken together, the kinetics data are consistent with the mechanism proposed in Figure 2 where transmetalation is turnover-limiting at low concentrations of boronic acid and oxidative addition is turnover-limiting at low concentrations of [F+]. A zero-order dependence in all components under standard conditions is consistent with the C–F reductive elimination step being turnover-limiting. Negative-order rate dependence on the concentration of TDG is consistent with the TDG mediating decomposition of [F+] or potentially forming an off-cycle TDG•PdII complex. Analogous to our previous studies, we hypothesize that the chiral TDG mediates enantiodetermining migratory insertion by attenuating geometric distortion in the favored metallatricyclic transition state.17 The kinetics data as a whole underscore the mechanistic complexity of this TDG-mediated arylfluorination reaction and highlights the value of DoE for optimization of complicated dual catalytic systems of this type in the absence of a detailed a priori mechanistic picture. Not only does the highly enantioselective 1,2-arylfluorination of alkenyl benzaldehydes presented here allow expedient access to organofluorine compounds that are otherwise difficult to prepare; its successful development sets the stage for expansion of chiral TDG strategies across increasingly diverse alkene difunctionalization reactions.

Supplementary Material

Supplementary Materials

ACKNOWLEDGMENT

We thank Dr. Dee-Hua Huang and Dr. Laura Pasternack for assistance with NMR spectroscopy. We also thank Dr. Gary J. Balaich (UCSD) for X-ray crystallographic analysis. Dr. Jason Chen, Brittany Sanchez, and Emily Sturgell (Scripps Research Automated Synthesis Facility) are acknowledged for HPLC, SFC, and HRMS analysis. We also thank Dr. Joseph Derosa and Dr. Donna Blackmond for helpful discussion.

Funding Sources

Financial support for this work was provided by the National Institutes of Health (R35GM125052) and the Cottrell Scholars program. Additional support was provided through an SIOC Fellowship (Z.L.) and a National Science Foundation Graduate Research Fellowship (NSF/DGE-184247, L.J.O.).

Footnotes

Supporting Information

The Supporting Information is available free of charge on the ACS Publications website at DOI: https://pubs.acs.org/doi/10.1021/jacs.1c03178.

Experimental details, analytical data, and 1H, 13C, and 19F NMR spectra (PDF)

X-ray data for compound 3u (CIF)

REFERENCES

  • (1).Purser S; Moore PR; Swallow S; Gouverneur V, Fluorine in medicinal chemistry. Chem. Soc. Rev 2008, 37, 320. [DOI] [PubMed] [Google Scholar]
  • (2).Gillis EP; Eastman KJ; Hill MD; Donnelly DJ; Meanwell NA, Applications of Fluorine in Medicinal Chemistry. J. Med. Chem 2015, 58, 8315. [DOI] [PubMed] [Google Scholar]
  • (3).Ma JA; Cahard D, Asymmetric fluorination, trifluoromethylation, and perfluoroalkylation reactions. Chem. Rev 2004, 104, 6119. [DOI] [PubMed] [Google Scholar]
  • (4).Brunet VA; O’Hagan D, Catalytic asymmetric fluorination comes of age. Angew. Chem. In. Ed 2008, 47, 1179. [DOI] [PubMed] [Google Scholar]
  • (5).Yang X; Wu T; Phipps RJ; Toste FD, Advances in catalytic enantioselective fluorination, mono-, di-, and trifluoromethylation, and trifluoromethylthiolation reactions. Chem. Rev 2015, 115, 826. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (6).Wolstenhulme JR; Gouverneur V, Asymmetric fluorocyclizations of alkenes. Acc. Chem. Res 2014, 47, 3560. [DOI] [PubMed] [Google Scholar]
  • (7).Nie J; Zhu HW; Cui HF; Hua MQ; Ma JA, Catalytic stereoselective synthesis of highly substituted indanones via tandem Nazarov cyclization and electrophilic fluorination trapping. Org. Lett 2007, 9, 3053. [DOI] [PubMed] [Google Scholar]
  • (8).Cochrane NA; Nguyen H; Gagne MR, Catalytic enantioselective cyclization and C3-fluorination of polyenes. J. Am. Chem. Soc 2013, 135, 628. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (9).Romanov-Michailidis F; Guenee L; Alexakis A, Enantioselective organocatalytic fluorination-induced Wagner-Meerwein rearrangement. Angew. Chem. In. Ed 2013, 52, 9266. [DOI] [PubMed] [Google Scholar]
  • (10).Wolstenhulme JR; Rosenqvist J; Lozano O; Ilupeju J; Wurz N; Engle KM; Pidgeon GW; Moore PR; Sandford G; Gouverneur V, Asymmetric electrophilic fluorocyclization with carbon nucleophiles. Angew. Chem. In. Ed 2013, 52, 9796. [DOI] [PubMed] [Google Scholar]
  • (11).He Y; Yang Z; Thornbury RT; Toste FD, Palladium-catalyzed enantioselective 1,1-fluoroarylation of aminoalkenes. J. Am. Chem. Soc 2015, 137, 12207. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (12).Miro J; Del Pozo C; Toste FD; Fustero S, Enantioselective Palladium-Catalyzed Oxidative beta, beta-Fluoroarylation of alpha, beta-Unsaturated Carbonyl Derivatives. Angew. Chem. In. Ed 2016, 55, 9045. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (13).Talbot EP; Fernandes Tde A; McKenna JM; Toste FD, Asymmetric palladium-catalyzed directed intermolecular fluoroarylation of styrenes. J. Am. Chem. Soc 2014, 136, 4101. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (14).Xi Y; Wang C; Zhang Q; Qu J; Chen Y, Palladium-Catalyzed Regio-, Diastereo-, and Enantioselective 1,2-Arylfluorination of Internal Enamides. Angew. Chem. In. Ed 2020, 60, 2699. [DOI] [PubMed] [Google Scholar]
  • (15).Thornbury RT; Saini V; Fernandes TA; Santiago CB; Talbot EPA; Sigman MS; McKenna JM; Toste FD, The development and mechanistic investigation of a palladium-catalyzed 1,3-arylfluorination of chromenes. Chem. Sci 2017, 8, 2890. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (16).Higham JI; Bull JA, Transient imine directing groups for the C-H functionalisation of aldehydes, ketones and amines: an update 2018-2020. Org. Biomol. Chem 2020, 18, 7291. [DOI] [PubMed] [Google Scholar]
  • (17).Oxtoby LJ; Li ZQ; Tran VT; Erbay TG; Deng R; Liu P; Engle KM, A Transient-Directing-Group Strategy Enables Enantioselective Reductive Heck Hydroarylation of Alkenes. Angew. Chem. In. Ed 2020, 59, 8885. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (18).Kalyani D; Satterfield AD; Sanford MS, Palladium-catalyzed oxidative arylhalogenation of alkenes: synthetic scope and mechanistic insights. J. Am. Chem. Soc 2010, 132, 8419. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (19).Park H; Verma P; Hong K; Yu JQ, Controlling Pd(IV) reductive elimination pathways enables Pd(II)-catalysed enantioselective C(sp(3))-H fluorination. Nat. Chem 2018, 10, 755. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (20).Leardi R, Experimental design in chemistry: A tutorial. Anal. Chim. Acta 2009, 652, 161. [DOI] [PubMed] [Google Scholar]
  • (21).Lendrem D; Owen M; Godbert S, DOE (Design of Experiments) in Development Chemistry: Potential Obstacles. Org. Process Res. Dev 2001, 5, 324. [Google Scholar]
  • (22).Jones B; Nachtsheim CJ, A Class of Three-Level Designs for Definitive Screening in the Presence of Second-Order Effects. J. Qual. Technol 2011, 43, 1. [Google Scholar]
  • (23).It should be noted that while this technique can be used for rapid optimization and identification of major interactions, it does not suggest an explanation for why a given response is dependent on a particular input parameter; moreover, the DSD model does not necessarily accurately predict reaction performance in totally different chemical space. Multiple powerful techniques have emerged for the statistical prediction of out-of-sample reactions, including multivariate regression of physical organic descriptors (see:; (A) Reid JP; Sigman MS, Holistic prediction of enantioselectivity in asymmetric catalysis. Nature 2019, 571, 343.) [DOI] [PMC free article] [PubMed] [Google Scholar]; and machine-learning based chemical featurization approaches (see:; (B) Ahneman DT; Estrada JG; Lin S; Dreher SD; Doyle AG, Predicting reaction performance in C-N cross-coupling using machine learning. Science 2018, 360, 186. [DOI] [PubMed] [Google Scholar]; (C) Zahrt AF; Henle JJ; Rose BT; Wang Y; Darrow WT; Denmark SE, Prediction of higher-selectivity catalysts by computer-driven workflow and machine learning. Science 2019, 363 (6424).), [DOI] [PMC free article] [PubMed] [Google Scholar]; but these methods rely on much larger data sets with more input parameters. Our model translated well to other disubstituted and trisubstituted alkenes, but we did find that it was necessary to change TDGs for terminal alkene substrates (see Supplementary Figure S3).
  • (24).A number of approaches for the enantioselective formation of fully substituted C(sp3)–F stereocenters have been developed. For an overview see:; (A) Zhu Y; Han J; Wang J; Shibata N; Sodeoka M; Soloshonok VA; Coelho JAS; Toste FD, Modern Approaches for Asymmetric Construction of Carbon-Fluorine Quaternary Stereogenic Centers: Synthetic Challenges and Pharmaceutical Needs. Chem. Rev 2018, 118, 3887. [DOI] [PMC free article] [PubMed] [Google Scholar]; For enolate-type reactions of carbonyl derivatives with electrophilic fluorinating reagents see:; (B) Zhu Y; Han J; Wang J; Shibata N; Sodeoka M; Soloshonok VA; Coelho JAS; Toste FD, Modern Approaches for Asymmetric Construction of Carbon-Fluorine Quaternary Stereogenic Centers: Synthetic Challenges and Pharmaceutical Needs. Chem. Rev 2018, 118, 3887. [DOI] [PMC free article] [PubMed] [Google Scholar]; (C) Shibata N; Suzuki E; Asahi T; Shiro M, Enantioselective fluorination mediated by cinchona alkaloid derivatives/Selectfluor combinations: reaction scope and structural information for N-fluorocinchona alkaloids. J. Am. Chem. Soc 2001, 123, 7001. [DOI] [PubMed] [Google Scholar]; (D) Mohar B; Baudoux J; Plaquevent J-C; Cahard D, Electrophilic Fluorination Mediated by Cinchona Alkaloids: Highly Enantioselective Synthesis of α-Fluoro-α-phenylglycine Derivatives. Angew. Chem. In. Ed 2001, 40, 4214. [DOI] [PubMed] [Google Scholar]; (E) Marigo M; Fielenbach D; Braunton A; Kjaersgaard A; Jorgensen KA, Enantioselective formation of stereogenic carbon-fluorine centers by a simple catalytic method. Angew. Chem. In. Ed 2005, 44, 3703. [DOI] [PubMed] [Google Scholar]; (F) Steiner DD; Mase N; Barbas CF 3rd, Direct asymmetric alpha-fluorination of aldehydes. Angew. Chem. In. Ed 2005, 44, 3706. [DOI] [PubMed] [Google Scholar]; (G) Shibatomi K; Kitahara K; Okimi T; Abe Y; Iwasa S, Enantioselective fluorination of alpha-branched aldehydes and subsequent conversion to alpha-hydroxyacetals via stereospecific C-F bond cleavage. Chem. Sci 2016, 7, 1388. [DOI] [PMC free article] [PubMed] [Google Scholar]; (H) You Y; Zhang L; Luo S, Reagent-controlled enantioselectivity switch for the asymmetric fluorination of beta-ketocarbonyls by chiral primary amine catalysis. Chem. Sci 2017, 8, 621. [DOI] [PMC free article] [PubMed] [Google Scholar]; (I) Shibata N; Kohno J; Takai K; Ishimaru T; Nakamura S; Toru T; Kanemasa S, Highly enantioselective catalytic fluorination and chlorination reactions of carbonyl compounds capable of two-point binding. Angew. Chem. In. Ed 2005, 44, 4204. [DOI] [PubMed] [Google Scholar]; For enolate-type reactions of α-fluorinated carbonyl derivatives with arylating/alkylating reagents see:; (J) Reddy DS; Shibata N; Nagai J; Nakamura S; Toru T; Kanemasa S, Desymmetrization-like catalytic enantioselective fluorination of malonates and its application to pharmaceutically attractive molecules. Angew. Chem. In. Ed 2008, 47, 164. [DOI] [PubMed] [Google Scholar]; (K) Jiao Z; Beiger JJ; Jin Y; Ge S; Zhou JS; Hartwig JF, Palladium-Catalyzed Enantioselective alpha-Arylation of alpha-Fluoroketones. J. Am. Chem. Soc 2016, 138, 15980. [DOI] [PubMed] [Google Scholar]; (L) Belanger E; Cantin K; Messe O; Tremblay M; Paquin JF, Enantioselective Pd-catalyzed allylation reaction of fluorinated silyl enol ethers. J. Am. Chem. Soc 2007, 129, 1034. [DOI] [PubMed] [Google Scholar]; (M) Liang Y; Fu GC, Catalytic asymmetric synthesis of tertiary alkyl fluorides: Negishi cross-couplings of racemic alpha,alpha-dihaloketones. J. Am. Chem. Soc 2014, 136, 5520. [DOI] [PMC free article] [PubMed] [Google Scholar]; (N) Han X; Kwiatkowski J; Xue F; Huang KW; Lu Y, Asymmetric Mannich reaction of fluorinated ketoesters with a tryptophan-derived bifunctional thiourea catalyst. Angew. Chem. In. Ed 2009, 48, 7604. [DOI] [PubMed] [Google Scholar]; For the reaction of alkenes with electrophilic fluorinating reagents to form allylic fluoride products see:; (O) Xie C; Wu L; Han J; Soloshonok VA; Pan Y, Assembly of Fluorinated Quaternary Stereogenic Centers through Catalytic Enantioselective Detrifluoroacetylative Aldol Reactions. Angew. Chem. In. Ed 2015, 54, 6019. [DOI] [PubMed] [Google Scholar]; (P) Ishimaru T; Shibata N; Horikawa T; Yasuda N; Nakamura S; Toru T; Shiro M, Cinchona alkaloid catalyzed enantioselective fluorination of allyl silanes, silyl enol ethers, and oxindoles. Angew. Chem. In. Ed 2008, 47, 4157. [DOI] [PubMed] [Google Scholar]; (Q) Wu J; Wang YM; Drljevic A; Rauniyar V; Phipps RJ; Toste FD, A combination of directing groups and chiral anion phase-transfer catalysis for enantioselective fluorination of alkenes. Proc. Natl Scad. Sci 2013, 110, 13729. [DOI] [PMC free article] [PubMed] [Google Scholar]; (R) Lozano O; Blessley G; Martinez del Campo T; Thompson AL; Giuffredi GT; Bettati M; Walker M; Borman R; Gouverneur V, Organocatalyzed enantioselective fluorocyclizations. Angew. Chem. In. Ed 2011, 50, 8105. [DOI] [PubMed] [Google Scholar]; (S) Rauniyar V; Lackner AD; Hamilton GL; Toste FD, Asymmetric electrophilic fluorination using an anionic chiral phase-transfer catalyst. Science 2011, 334, 1681. [DOI] [PubMed] [Google Scholar]; (T) Shunatona HP; Fruh N; Wang YM; Rauniyar V; Toste FD, Enantioselective fluoroamination: 1,4-addition to conjugated dienes using anionic phase-transfer catalysis. Angew. Chem. In. Ed 2013, 52, 7724. [DOI] [PubMed] [Google Scholar]; For iridium-catalyzed allylic substitution reactions of alkenyl fluorides see:; (U) Egami H; Niwa T; Sato H; Hotta R; Rouno D; Kawato Y; Hamashima Y, Dianionic Phase-Transfer Catalyst for Asymmetric Fluorocyclization. J. Am. Chem. Soc 2018, 140, 2785. [DOI] [PubMed] [Google Scholar]; (V) Butcher TW; Hartwig JF, Enantioselective Synthesis of Tertiary Allylic Fluorides by Iridium-Catalyzed Allylic Fluoroalkylation. Angew. Chem. In. Ed 2018, 57, 13125. [DOI] [PubMed] [Google Scholar]; (W) He ZT; Jiang X; Hartwig JF, Stereodivergent Construction of Tertiary Fluorides in Vicinal Stereogenic Pairs by Allylic Substitution with Iridium and Copper Catalysts. J. Am. Chem. Soc 2019, 141, 13066. [DOI] [PubMed] [Google Scholar]; (X) Butcher TW; Yang JL; Amberg WM; Watkins NB; Wilkinson ND; Hartwig JF, Desymmetrization of difluoromethylene groups by C-F bond activation. Nature 2020, 583, 548. [DOI] [PMC free article] [PubMed] [Google Scholar]; For the redox-relay palladium-catalyzed enantioselective Heck reaction of acyclic alkenyl fluorides with arylboronic acids see:; (Y) Liu J; Yuan Q; Toste FD; Sigman MS, Enantioselective construction of remote tertiary carbon-fluorine bonds. Nat. Chem 2019, 11, 710. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (25).Valenta P; Carroll PJ; Walsh PJ, Stereoselective synthesis of beta-hydroxy enamines, aminocyclopropanes, and 1,3-amino alcohols via asymmetric catalysis. J. Am. Chem. Soc 2010, 132, 14179. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (26).Blackmond DG, Reaction progress kinetic analysis: a powerful methodology for mechanistic studies of complex catalytic reactions. Angew. Chem. In. Ed 2005, 44, 4302. [DOI] [PubMed] [Google Scholar]
  • (27).To identify if there is significant catalyst deactivation or product inhibition in our system, we compared temporal concentration profiles ([1a] vs. time) of reactions carried out under same-excess conditions. This comparison is facilitated by shifting the profile of a reaction started at lower concentration to the right on the timescale until the first point overlays with the second profile (the standard conditions). For any point on the time-adjusted profile, both reactions will have the same concentration of starting materials, but different concentrations of product and a different number of turnovers completed by the catalyst.
  • (28).Baxter RD; Sale D; Engle KM; Yu JQ; Blackmond DG, Mechanistic rationalization of unusual kinetics in Pd-catalyzed C-H olefination. J. Am. Chem. Soc 2012, 134, 4600. [DOI] [PubMed] [Google Scholar]
  • (29).Hill DE; Pei Q.-l.; Zhang E.-x.; Gage JR; Yu J-Q; Blackmond DG, A General Protocol for Addressing Speciation of the Active Catalyst Applied to Ligand-Accelerated Enantioselective C(sp3)–H Bond Arylation. ACS Catal. 2018, 8, 1528. [Google Scholar]

Associated Data

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

Supplementary Materials

Supplementary Materials

RESOURCES