Abstract
A short, enantioselective synthesis of (–)-maximiscin, a structurally-intriguing metabolite of mixed biosynthetic origin, is reported. A retrosynthetic analysis predicated on maximizing ideality and efficiency led to several unusual disconnections and tactics. Formation of the central highly-oxidized pyridone ring through a convergent coupling at the end of the synthesis simplified the route considerably. The requisite building blocks could be prepared from feedstock materials (derived from shikimate and mesitylene). Strategies rooted in hidden symmetry recognition, C–H functionalization, and radical retrosynthesis played key roles developing this concise route.
Graphical Abstract
Natural products derived from mixed biosynthetic lineages have historically provided chemists with some of the most exotic molecular architectures imaginable (e.g. staurosporine, hyperforin, and reserpine).1 The unique structure of (–)-maximiscin 1 (Figure 1A) is no exception, resulting from the rare union of three separate metabolic pathways.2 Thus, a central 1,4-dihydroxy-2-pyridone (derived from tyrosine 3) is linked to both a shikimate derivative (derived from shikimic acid 2) and a trisubstituted cyclohexyl fragment of polyketide origin (derived from 4). As such, 1 exists as an equilibrating mixture of atropisomers about the C-3,7 bond. The challenge associated with synthesizing such a structure is compounded by its documented instability, as it tends to fragment between the shikimate and pyridone residues.3 4-hydroxy-2-pyridone alkaloids such as 1 have historically represented an exciting class of natural products for chemical synthesis, due to their intriguing structural properties, and biological activiites.4 Although no synthesis of 1 has been reported, simpler variants lacking the shikimate subunit have been prepared.5 In this Communication, an abiotic, convergent, enantioselective preparation of 1 is reported; this synthesis is enabled by exploiting hidden symmetry and leveraging the logic of C–H functionalization6 and radical retrosynthesis strategies7.
To maximize convergency, retrosynthetic scission of the central pyridone ring produced two equally sized fragments (R=shikimate derivative 5, and 6). Their union through a non-canonical Guareschi-Thorpe-type condensation (Tactic 1, Figure 1B) could potentially forge this core motif at a late-stage. While such condensations normally require an electron-withdrawing group on the enamine fragment,8 in this variant a β-silicon atom was employed, reminiscent of the Sakurai-type allylation.9 This provided the added benefit of building in the requisite N-oxide motif, which would otherwise require subsequent oxidation from the parent pyridone.5, 10 Fragment 5 could be traced back to a known shikimate-derived epoxide in a few simple steps. Fragment 6 was envisaged to arise from a decarboxylative radical homologation sequence (Tactic 2) to forge the hindered C-3,7-bond and set its relative trans-orientation. During this process, a remarkably efficient radical cascade was developed to achieve not only this bond formation but also a redox relay11 to set the proper C-13 oxidation state. Recognizing that if the C-13 position were simply a methyl group, an enantiocontrolled desymmetrizing C–H activation could be invoked (Tactic 3) to cement the absolute configuration of four centers in one step.12 Such a tactic is not without risk, as the required C-H activation would enlist a challenging 6-membered palladacycle intermediate.13 Finally, the all-cis stereochemistry needed for this step could originate from the hydrogenation of an inexpensive mesitylene-derived carboxylic acid 714 (Tactic 4).
The successful execution of these tactics to access synthetic 1 for the first time is outlined in Scheme 1. Preparation of fragment 6 (Scheme 1A) began with hydrogenation of 7 using Adam’s catalyst, which furnished the all-cis carboxylic acid 8 with complete selectivity in 97% isolated yield (gram-scale).15 This set the stage for the development of a desymmetrizing C–H activation reaction that was to define four chiral centers from this meso-acid.16 An exhaustive set of directing groups and C–H functionalization conditions were explored (see SI), culminating in the identification of a chiral PIP-type directing group (9, X = H, inset table 1).17 Amide 10d, derived from acid 8 and amine 9 (84% isolated yield, gram-scale), conferred reasonable yield and high diastereoselectivity under Pd-catalyzed methoxylation conditions to deliver 11.18 The influence of various substituents on the pyridyl ring of the directing group was then explored. Although diastereoselectivity remained high, ring electronics exerted a profound effect on reaction yield (inset table 1). A 4-Cl substituted analog was identified as the optimal directing group, and upon further refinement of the reaction conditions, a scalable desymmetrizing methoxylation to access 11 was realized (58% isolated yield + 23% recovered 10d, gram-scale). This, to the best of our knowledge, represents the most complex desymmetrizing C–H activation reported to-date.19 It defines 4 stereocenters in a single step, including the distal δ-methyl substituent, which is remote from the directing group. Removal of the directing group from 11 was accomplished using HBr at elevated temperature (99% isolated yield, gram-scale). This reaction cleanly delivered lactone 12, along with a substantial amount of recovered directing group 9 (80%). The latter material could be recycled, and used to prepare another batch of amide 10d for the C–H activation sequence, without any erosion in enantiopurity (see SI). With lactone 12 in hand, the next tactic involved effecting a decarboxylative homologation for a key C–C bond forming step, drawing further utility from the C–H functionalization handle. Phenyl vinyl sulfone was selected as the homologation reagent of choice. It represents a simple, inexpensive 2-carbon extension unit, and has precedent for such applications.20 Initial success was achieved using a Ni-catalyzed decarboxylative Giese addition protocol21 (performed on a derivative of 12, see SI), but the yield was hampered by competing 1,5-hydrogen atom transfer (1,5-HAT) from C–2 to C–13 which resulted in double addition of the radical acceptor onto C–13 (see SI). It was hypothesized that transitioning from a reductive to oxidative decarboxylative manifold could enable this transposed radical to be oxidized, thus producing aldehyde 14.22 A number of exotic redox strategies were evaluated before a simple solution emerged (see SI) using Minisci-type conditions23 directly from lactone 12, which was hydrolyzed in-situ. Thus, following saponification of lactone 12 with NaOH (1.2 equiv), sequential addition of AgNO3 (0.3 equiv), Na2S2O8 (2.5 equiv), NaHSO4 (1.13 equiv), Fe2(SO4)3 (0.2 equiv), and the vinyl sulfone led to the formation of aldehyde 14. The reaction was exceptionally clean (see SI for crude NMR) and proved scalable, resulting in a 91% yield of 14 on gram scale. Some features of this radical translocation cascade are worth noting. The addition of NaHSO4 enabled the tandem hydrolysis/decarboxylation by buffering the resulting carboxylate. Without it, a heterogeneous mixture resulted, leading to aggregate formation and diminished product yield. The standard Ag+/persulfate combination proved ineffective (ca. 6% yield), prompting the exploration of additives (inset table 2). Fe2(SO4)3 uniquely served as a highly efficient co-catalyst, the first use we are aware of in concert with a Ag-catalyzed Minisci-reaction.24 The reaction requires both metals to be present, as no desired product is formed in the absence of Ag. Literature on the reactivity of Fe-salts in free radical chemistry suggest that the Fe3+ can assist in the selective oxidation of the intermediate α-oxy alkyl radical.25 With aldehyde 14 in hand, access to key building block 6 could be rapidly achieved. Classical Wittig conditions provided an olefin intermediate which was elaborated to 15 through a one-pot sulfone oxidation and methyl ester formation sequence (50% yield overall).26 Next, acylation of 15 proved challenging, with most acyl electrophile/base combinations leaving the hindered ester untouched. Use of LDA with Mander’s reagent in diethyl ether proved uniquely effective,27 and in-situ hydrolysis of the resulting diester provided 6 in 74% isolated yield. The crude diacid could be purified by simple trituration, and yielded crystals suitable for x-ray diffraction, which confirmed its absolute configuration.
table 1.
(X =) | yield | d.r. |
---|---|---|
10a (H) | 44% | >10:1 |
10b (6-OMe) | n.d. | - |
10c (4-Me) | 31% | >10:1 |
10d (4-CI)* | 58% | 20:1 |
10e (5-CI) | 53% | >10:1 |
10f (quin.) | 23% | >10:1 |
using (R)-DG
table 2.
cocatalyst | yield |
---|---|
none | 6% |
Cu(OAc)2 | 22% |
Co(CIO4)2 | 28% |
Fe(OAc)2 | 48% |
Fe2(SO4)3 | 71% |
Fe2(SO4)3 | 91%* |
gram scale
Synthesis of fragment 5 was accomplished starting from known epoxide 16, which is derived from shikimic acid (Scheme 1B).28 Opening of the epoxide intermediate to furnish 17 was achieved using N-Boc-hydroxylamine assisted by DBU in methylene chloride (70% isolated yield, gram-scale); use of Lewis acids led to mixtures of SN2/SN2’ products, while elimination/aromatization pathways dominated in different solvents. X-ray analysis confirmed the regioselectivity of the epoxide opening process. Intermediate 17 was converted to a TBS-protected hydroxylamine intermediate (72% isolated yield, gram-scale), which was condensed with acetaldehyde to give des-TMS-5 (not shown, 82% isolated yield). The final bond-forming step involved union of this fragment with 6 using a bold late-stage pyridone synthesis to forge the central ring of 1. Initial conditions surveyed for this step were based on a report for the synthesis of alkyl-fused hydroxypyridones, derived from diacid chlorides and ketoxime ethers.29 Combination of oxime ether des-TMS-5 with the diacid chloride derivative of 6 in toluene at 90°C generated intractable mixtures of decomposition products, with significant recovery of unreacted des-TMS-5. It was reasoned that the poor nucleophilicity of the oxime ether, combined with elevated reaction temperatures, promoted decomposition of the diacid chloride before it could engage with the oxime. To remedy this, two modifications were implemented: (1) The TMS derivative 5 was prepared, inspired by the venerable Hosomi-Sakurai reaction30 as it was envisaged that a β-silicon effect could enhance the nucleophilicity at nitrogen via σ(Si–C) to π donation.31 (2) The reaction was conducted at lower temperature, using AgOTf to promote activation of the diacid chloride electrophile. Ultimately, this push-pull system of Si/Ag+ activation delivered the condensation product 18 in moderate yield with surprising speed (reaction quenched after 7 minutes, see inset table 3 for selected optimization conditions). The reaction benefited from elevated temperature, although by-products became predominant above 50 °C. Acetonitrile proved to be the optimal solvent, and a methanol quench was employed to liberate product which had been acylated on the pyridone 4-OH by starting material. This enabled the facile recovery of the diacid fragment as a mixture of the mono- and di-methyl esters (25% + 22% respectively) which could be recycled to access additional 6. A reasonable mechanistic proposal involves silver mediated activation of the diacid chloride derived from 6, to produce a transient diacyl triflate, which is intercepted by oxime 5, assisted by the appended TMS moiety. This would produce an enamine intermediate which could rapidly engage the second acyl electrophile in an intramolecular fashion to generate the pyridone ring. Additionally, TMSOTf generated in situ may serve to further enhance the reactivity of the electrophile (TMS derivatization is observed by LCMS). The importance of the silyl substituent on 5 is highlighted by the observation that des-TMS-5 provides 18 in much lower yield (14%) under the same reaction conditions (see SI). This represents a unique pyridone synthesis and enables the construction of a remarkably hindered ring system, which exists as a mixture of interconverting atropisomers. Deprotection using TFA/MeOH afforded (–)-maximiscin 1 (= –147), completing a 10-step (LLS) total synthesis of this natural product (60% ideality).32 All spectral data were wholly consistent with the original isolation report.2
table 3.
solvent | temp. (°C) | yield |
---|---|---|
PhMe | RT | 9% |
PhMe | 50 | 12% |
PhMe | 50 | 21%* |
PhMe/MeNO2 | 50 | 26%* |
CH3CN | 50 | 28%* |
CH3CN; MeOH | 50 | 41%* |
AgOTf added last
Several unusual steps in this synthesis are worthy of note: (1) a scalable enantiocontrolled C–H activation-desymmetrization defined 4-stereocenters in a single step and enabled access to a highly-functionalized carbocycle from simple aromatic feedstock; (2) a radical disconnection, leveraging a Ag/Fe co-catalyzed stereoinvertive decarboxylative Giese addition, constructed a hindered C–C bond with concomitant radical translocation to a remote site; (3) a non-intuitive disconnection through the central hydroxypyridone ring enabled a highly convergent synthesis of 1 and led to the development of a new tactic to assemble such systems.
Supplementary Material
ACKNOWLEDGMENTS
Financial support for this work was provided by NIH (GM-118176), Tsinghua Xuetang Talent Program (predoctoral fellowship to F. W.), Fulbright Scholar Program (predoctoral fellowship to A. M.). NIH (1S10OD025208) for OSR instrument. We thank Johnson Matthey for providing samples of solid supported rhodium catalysts. Crystallographic structures were rendered using PyMOL, a product of Schrödinger LLC. We are grateful to Dr. D.-H. Huang and Dr. L. Pasternack (Scripps Research) for NMR spectroscopic assistance, Prof. A. L. Rheingold, Dr. C. E. Moore, Dr. M. Gembicky and Dr. J. B. Bailey and Gary J. Balaich (University of California San Diego) for X-ray crystallographic analysis. Dr. Jason Chen, Ms. Brittany Sanchez and Ms. Emily Sturgell (Scripps Research) for analytical support. Prof. Robert Cichewicz and Dr. Lin Du (University of Oklahoma) for providing samples of authentic (–)-maximiscin and useful discussion. David Hill (Scripps Research) for assistance with chiral GCMS analysis. Prof. Keary Engle (Scripps Research) for glovebox access. We also thank Yuzuru Kanda, Solomon Reisberg, Stephen Harwood, Tyler Saint-Denis, David Peters and Dr. Julien Vantourout (Scripps Research) for useful discussions.
Footnotes
ASSOCIATED CONTENT
Supporting Information. Experimental procedures, analytical data (1H and 13C NMR, MS) for all new compounds as well as optimization tables. This material is available free of charge via the Internet at http://pubs.acs.org.
The authors declare no competing financial interest.
REFERENCES
- (1).Walsh CT; Tang Y Natural Product Biosynthesis: Chemical Logic and Enzymatic Machinery. Royal Society of Chemistry: London, 2017. [Google Scholar]
- (2).Du L; Robles AJ; King JB; Powell DR; Miller AN; Mooberry SL; Cichewicz RH Crowdsourcing Natural Products Discovery to Access Uncharted Dimensions of Fungal Metabolite Diversity. Angewandte Chemie International Edition 2014, 53, 804–809. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (3).(a) Du L; Robles AJ; King JB; Powell DR; Miller AN; Mooberry SL; Cichewicz RH Corrigendum: Crowdsourcing Natural Products Discovery to Access Uncharted Dimensions of Fungal Metabolite Diversity. Angewandte Chemie International Edition 2015, 54, 6671–6671; [DOI] [PubMed] [Google Scholar]; (b) Du L; You J; Nicholas KM; Cichewicz RH Chemoreactive Natural Products that Afford Resistance Against Disparate Antibiotics and Toxins. Angewandte Chemie International Edition 2016, 55, 4220–4225. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (4).(a) Jessen HJ; Gademann K 4-Hydroxy-2-pyridone alkaloids: Structures and synthetic approaches. Natural Product Reports 2010, 27, 1168–1185; for biological activity of maximiscin and related N-hydroxypyridone alkaloids see: [DOI] [PubMed] [Google Scholar]; (b) Robles AJ; Du L; Cichewicz RH; Mooberry SL Maximiscin Induces DNA Damage, Activates DNA Damage Response Pathways, and Has Selective Cytotoxic Activity against a Subtype of Triple-Negative Breast Cancer. Journal of Natural Products 2016, 79, 1822–1827; [DOI] [PMC free article] [PubMed] [Google Scholar]; (c) Fürstner A; Feyen F; Prinz H; Waldmann H Synthesis and evaluation of the antitumor agent TMC-69–6H and a focused library of analogs. Tetrahedron 2004, 60, 9543–9558. [Google Scholar]
- (5).(a) Snider BB; Lu Q Total Synthesis of (.+−.)-Pyridoxatin. The Journal of Organic Chemistry 1994, 59, 8065–8070; [Google Scholar]; (b) Jones IL; Moore FK; Chai CLL Total Synthesis of (±)-Cordypyridones A and B and Related Epimers. Organic Letters 2009, 11, 5526–5529. [DOI] [PubMed] [Google Scholar]
- (6).Brückl T; Baxter RD; Ishihara Y; Baran PS Innate and Guided C–H Functionalization Logic. Accounts of Chemical Research 2012, 45, 826–839. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (7).Smith JM; Harwood SJ; Baran PS Radical Retrosynthesis. Accounts of Chemical Research 2018, 51, 1807–1817. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (8).(a) Reaction Guareschi. In Comprehensive Organic Name Reactions and Reagents, pp 1294–1297; [Google Scholar]; (b) Joule JA; Mills K Heterocyclic Chemistry. 5th ed.; John Wiley & Sons: 2010. [Google Scholar]
- (9).Allylation Hosomi-Sakurai. In Comprehensive Organic Name Reactions and Reagents, pp 1491–1495. [Google Scholar]
- (10).(a) Snider BB; Lu Q Total Synthesis of (±)-Leporin A. The Journal of Organic Chemistry 1996, 61, 2839–2844; [DOI] [PubMed] [Google Scholar]; (b) Fürstner A; Feyen F; Prinz H; Waldmann H Total Synthesis and Reassessment of the Phosphatase-Inhibitory Activity of the Antitumor Agent TMC-69–6H. Angewandte Chemie International Edition 2003, 42, 5361–5364. [DOI] [PubMed] [Google Scholar]
- (11).Renata H; Zhou Q; Dünstl G; Felding J; Merchant RR; Yeh C-H; Baran PS Development of a Concise Synthesis of Ouabagenin and Hydroxylated Corticosteroid Analogues. Journal of the American Chemical Society 2015, 137, 1330–1340. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (12).(a) for examples of desymmetrizing C–H activation in total synthesis see: Johnson JA; Li N; Sames D Total Synthesis of (−)-Rhazinilam: Asymmetric C−H Bond Activation via the Use of a Chiral Auxiliary. Journal of the American Chemical Society 2002, 124, 6900–6903; [DOI] [PubMed] [Google Scholar]; (b) Siler DA; Mighion JD; Sorensen EJ An Enantiospecific Synthesis of Jiadifenolide. Angewandte Chemie International Edition 2014, 53, 5332–5335; [DOI] [PMC free article] [PubMed] [Google Scholar]; (c) Sharpe RJ; Johnson JS A Global and Local Desymmetrization Approach to the Synthesis of Steroidal Alkaloids: Stereocontrolled Total Synthesis of Paspaline. Journal of the American Chemical Society 2015, 137, 4968–4971; for a review of C–H activation in total synthesis see: [DOI] [PMC free article] [PubMed] [Google Scholar]; (d) Abrams DJ; Provencher PA; Sorensen EJ Recent applications of C–H functionalization in complex natural product synthesis. Chemical Society Reviews 2018, 47, 8925–8967. [DOI] [PubMed] [Google Scholar]
- (13).(a) for selected examples of C–H activation proceeding via a 6-membered palladacycle see: Reddy BVS; Reddy LR; Corey EJ Novel Acetoxylation and C−C Coupling Reactions at Unactivated Positions in α-Amino Acid Derivatives. Organic Letters 2006, 8, 3391–3394; [DOI] [PubMed] [Google Scholar]; (b) Lafrance M; Gorelsky SI; Fagnou K High-Yielding Palladium-Catalyzed Intramolecular Alkane Arylation: Reaction Development and Mechanistic Studies. Journal of the American Chemical Society 2007, 129, 14570–14571; [DOI] [PubMed] [Google Scholar]; (c) He G; Zhao Y; Zhang S; Lu C; Chen G Highly Efficient Syntheses of Azetidines, Pyrrolidines, and Indolines via Palladium Catalyzed Intramolecular Amination of C(sp3)–H and C(sp2)–H Bonds at γ and δ Positions. Journal of the American Chemical Society 2012, 134, 3–6; [DOI] [PubMed] [Google Scholar]; (d) Nadres ET; Daugulis O Heterocycle Synthesis via Direct C–H/N–H Coupling. Journal of the American Chemical Society 2012, 134, 7–10; [DOI] [PMC free article] [PubMed] [Google Scholar]; (e) He G; Zhang S-Y; Nack WA; Li Q; Chen G Use of a Readily Removable Auxiliary Group for the Synthesis of Pyrrolidones by the Palladium-Catalyzed Intramolecular Amination of Unactivated γ C(sp3)–H Bonds. Angewandte Chemie International Edition 2013, 52, 11124–11128; [DOI] [PubMed] [Google Scholar]; (f) Li S; Chen G; Feng C-G; Gong W; Yu J-Q Ligand-Enabled γ-C–H Olefination and Carbonylation: Construction of β-Quaternary Carbon Centers. Journal of the American Chemical Society 2014, 136, 5267–5270; [DOI] [PMC free article] [PubMed] [Google Scholar]; (g) Xu J-W; Zhang Z-Z; Rao W-H; Shi B-F Site-Selective Alkenylation of δ-C(sp3)–H Bonds with Alkynes via a Six-Membered Palladacycle. Journal of the American Chemical Society 2016, 138, 10750–10753; [DOI] [PubMed] [Google Scholar]; (h) Deb A; Singh S; Seth K; Pimparkar S; Bhaskararao B; Guin S; Sunoj RB; Maiti D Experimental and Computational Studies on Remote γ-C(sp3)–H Silylation and Germanylation of Aliphatic Carboxamides. ACS Catalysis 2017, 7, 8171–8175; [Google Scholar]; (i) Chen Y-Q; Wang Z; Wu Y; Wisniewski SR; Qiao JX; Ewing WR; Eastgate MD; Yu J-Q Overcoming the Limitations of γ- and δ-C–H Arylation of Amines through Ligand Development. Journal of the American Chemical Society 2018, 140, 17884–17894. [DOI] [PubMed] [Google Scholar]
- (14).$90 USD for 100 g from Oakwood Chemical.
- (15).Bennani YL; Chamberlin SA; Chemburkar SR; Chen J; Dart MJ; Gupta AK; Wang L Cycloalkylamides and their therapeutic applications. U.S. Patent application 20040209858A1, 2003.
- (16).Merad J; Candy M; Pons J-M; Bressy C Catalytic Enantioselective Desymmetrization of Meso Compounds in Total Synthesis of Natural Products: Towards an Economy of Chiral Reagents. Synthesis 2017, 49, 1938–1954. [Google Scholar]
- (17).Kim Y; Kim S-T; Kang D; Sohn T. -i.; Jang E; Baik M-H; Hong S Stereoselective construction of sterically hindered oxaspirocycles via chiral bidentate directing group-mediated C(sp3)–O bond formation. Chemical Science 2018, 9, 1473–1480. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (18).(a) Dick AR; Hull KL; Sanford MS A Highly Selective Catalytic Method for the Oxidative Functionalization of C−H Bonds. Journal of the American Chemical Society 2004, 126, 2300–2301; [DOI] [PubMed] [Google Scholar]; (b) Zhang S-Y; He G; Zhao Y; Wright K; Nack WA; Chen G Efficient Alkyl Ether Synthesis via Palladium-Catalyzed, Picolinamide-Directed Alkoxylation of Unactivated C(sp3)–H and C(sp2)–H Bonds at Remote Positions. Journal of the American Chemical Society 2012, 134, 7313–7316; [DOI] [PubMed] [Google Scholar]; (c) Chen F-J; Zhao S; Hu F; Chen K; Zhang Q; Zhang S-Q; Shi B-F Pd(ii)-catalyzed alkoxylation of unactivated C(sp3)–H and C(sp2)–H bonds using a removable directing group: efficient synthesis of alkyl ethers. Chemical Science 2013, 4, 4187–4192; [Google Scholar]; (d) Shan G; Yang X; Zong Y; Rao Y An Efficient Palladium-Catalyzed C-H Alkoxylation of Unactivated Methylene and Methyl Groups with Cyclic Hypervalent Iodine (I3+) Oxidants. Angewandte Chemie International Edition 2013, 52, 13606–13610. [DOI] [PubMed] [Google Scholar]
- (19).(a) for recent examples of desymmetrizing C–H activation see: Pedroni J; Cramer N Enantioselective C–H Functionalization–Addition Sequence Delivers Densely Substituted 3-Azabicyclo[3.1.0]hexanes. Journal of the American Chemical Society 2017, 139, 12398–12401; [DOI] [PubMed] [Google Scholar]; (b) Fu J; Ren Z; Bacsa J; Musaev DG; Davies HML Desymmetrization of cyclohexanes by site- and stereoselective C–H functionalization. Nature 2018, 564, 395–399; for reviews on desymmetrization using C–H activation/functionalization see: [DOI] [PubMed] [Google Scholar]; (c) Newton CG; Wang S-G; Oliveira CC; Cramer N Catalytic Enantioselective Transformations Involving C–H Bond Cleavage by Transition-Metal Complexes. Chemical Reviews 2017, 117, 8908–8976; [DOI] [PubMed] [Google Scholar]; (d) Saint-Denis TG; Zhu R-Y; Chen G; Wu Q-F; Yu J-Q Enantioselective C(sp3)‒H bond activation by chiral transition metal catalysts. Science 2018, 359, eaao4798. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (20).Barton DHR; Ching-Yuh C; Joseph JC Homologation of acids via carbon radicals generated from the acyl derivatives of N-hydroxy-2-thiopyridone. (The two-carbon problem). Tetrahedron Letters 1991, 32, 3309–3312. [Google Scholar]
- (21).Qin T; Malins LR; Edwards JT; Merchant RR; Novak AJE; Zhong JZ; Mills RB; Yan M; Yuan C; Eastgate MD; Baran PS Nickel-Catalyzed Barton Decarboxylation and Giese Reactions: A Practical Take on Classic Transforms. Angewandte Chemie International Edition 2017, 56, 260–265. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (22).(a) Farney EP; Feng SS; Schäfers F; Reisman SE Total Synthesis of (+)-Pleuromutilin. Journal of the American Chemical Society 2018, 140, 1267–1270; [DOI] [PMC free article] [PubMed] [Google Scholar]; (b) Wang H; Zhang J; Shi J; Li F; Zhang S; Xu K Organic Photoredox-Catalyzed Synthesis of δ-Fluoromethylated Alcohols and Amines via 1,5-Hydrogen-Transfer Radical Relay. Organic Letters 2019, 21, 5116–5120. [DOI] [PubMed] [Google Scholar]
- (23).(a) Minisci F; Bernardi R; Bertini F; Galli R; Perchinummo M Nucleophilic character of alkyl radicals—VI: A new convenient selective alkylation of heteroaromatic bases. Tetrahedron 1971, 27, 3575–3579; [Google Scholar]; (b) Cui L; Chen H; Liu C; Li C Silver-Catalyzed Decarboxylative Allylation of Aliphatic Carboxylic Acids in Aqueous Solution. Organic Letters 2016, 18, 2188–2191. [DOI] [PubMed] [Google Scholar]
- (24).for an example of Ag/Fe co-catalysis see: Ouyang X-H; Song R-J; Hu M; Yang Y; Li J-H Silver-Mediated Intermolecular 1,2-Alkylarylation of Styrenes with α-Carbonyl Alkyl Bromides and Indoles. Angewandte Chemie International Edition 2016, 55, 3187–3191. [DOI] [PubMed] [Google Scholar]
- (25).Walling C Fenton’s reagent revisited. Accounts of Chemical Research 1975, 8, 125–131. [Google Scholar]
- (26).Bonaparte AC; Betush MP; Panseri BM; Mastarone DJ; Murphy RK; Murphree SS Novel Aerobic Oxidation of Primary Sulfones to Carboxylic Acids. Organic Letters 2011, 13, 1447–1449. [DOI] [PubMed] [Google Scholar]
- (27).Crabtree SR; Chu WLA; Mander LN C-Acylation of Enolates by Methyl Cyanoformate: An Examination of Site- and Stereoselectivity. Synlett 1990, 1990, 169–170. [Google Scholar]
- (28).Mizuki K; Iwahashi K; Murata N; Ikeda M; Nakai Y; Yoneyama H; Harusawa S; Usami Y Synthesis of Marine Natural Product (−)-Pericosine E. Organic Letters 2014, 16, 3760–3763. [DOI] [PubMed] [Google Scholar]
- (29).Ziegler E; Belegratis K Synthesen von Heterocyclen, 112. Mitt.: Zur Chemie der Ketoximäther. Monatshefte für Chemie / Chemical Monthly 1968, 99, 1454–1459. [Google Scholar]
- (30).(a) Hosomi A Characteristics in the reactions of allylsilanes and their applications to versatile synthetic equivalents. Accounts of Chemical Research 1988, 21, 200–206; [Google Scholar]; (b) Akira H; Katsukiyo M Development of New Reagents Containing Silicon and Related Metals and Application to Practical Organic Syntheses. Bulletin of the Chemical Society of Japan 2004, 77, 835–851. [Google Scholar]
- (31).Déléris G; Pillot JP; Rayez JC Influence of a silyl group on an allylic position. A theoretical approach. Tetrahedron 1980, 36, 2215–2218. [Google Scholar]
- (32).Gaich T; Baran PS Aiming for the Ideal Synthesis. The Journal of Organic Chemistry 2010, 75, 4657–4673. [DOI] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.