Abstract
C-Glycosides are important carbohydrate mimetics found in natural products, bioactive compounds, and marketed drugs. However, stereoselective preparation of this class of glycomimetics remains a significant challenge in organic synthesis. Herein, we report an excited-state palladium-catalyzed α-selective C-ketonylation strategy using readily available 1-bromosugars to access a range of C-glycosides. The reaction features excellent α-selectivity and mild conditions that tolerate a wide range of functional groups and complex molecular architectures. The resulting α-ketonylsugars can serve as versatile precursors for their β-isomers and other C-glycosides. Preliminary experimental and computational studies of the mechanism suggest a radical pathway involving the formation of palladoradical and glycosyl radical that undergoes polarity-mismatched coupling with silyl enol ether, affording the desired α-ketonylsugars. Insight into the reactivity and mechanism will inspire new reaction development and provide straightforward access to both α- and β-C-glycosides, greatly expanding the chemical and patent spaces of glycomimetics.
Keywords: Excited-state catalysis, Palladium, glycosyl radical, ketonylation, silyl enol ethers
Graphical Abstract
Introduction
Glycoconjugates are important constituents of living organisms where they regulate indispensable functions in a wide range of biological processes, including cell-cell recognition, cell-matrix interactions, and detoxification processes.1 Over the past several decades, intensive efforts have been made to develop methods of constructing and utilizing C─O and C─C glycosidic bonds.2-8 C─C glycosidic bonds are not subject to hydrolysis and are inert toward the hydrolytic enzymes in vivo, enabling C-glycosides to be used as artificial surrogates for potential therapeutic agents by mimicking their native O-glycoside structure.9-12 For example, the C-glycoside analog of KRN7000 is metabolically stable and exhibits excellent antitumor activity (Figure 1A).13 Other bioactive C-glycosides include Pro-Xylane,14 (−)-neodysiherbaine A,15 (+)-varitriol,16 and (+)-ambruticin S.17 Although chemical methods to construct the anomeric C─C bond have been reported, the development of a synthetic strategy that has a broad substrate scope and enables stereoselective access to both α- and β-anomers remains a challenge in the field.2, 18-21
Figure 1.
Development and exploration of ketonylation of carbohydrates.
C1-ketonylation of carbohydrates (i.e., replacing the C1 leaving group with a ketone) is a fundamental strategy with which to synthesize C-glycosides.2 The resulting C1-ketonylsugars could serve as versatile precursors for the preparation of a range of C-glycosides. Catalytic C1-ketonylation reactions that have a high level of α-selectivity are highly desired because the corresponding α-ketonylsugars can be epimerized to thermodynamically stable β-ketonylsugars,22-23 granting selective access to both α- and β-glycosides from the same starting materials. Most of the existing strategies, including Knoevenagel condensation,24-28 Horner-Wadsworth-Emmons olefination/Michael addition cascade reactions,29 Claisen rearrangements,30 and nucleophilic substitutions31-34 are non-catalytic transformations and afforded the thermodynamic β-ketonylsugars. Catalytic C1-ketonylation is rare, and the current state-of-the-art approach involves Au-catalyzed nucleophilic substitution of an in situ generated oxocarbenium ion by silyl enol ethers through an ionic, 2-electron reaction pathway, producing the desired C1-ketonylsugars with 2:1 to 10:1 α/β-selectivity (Figure 1B).35 This elegant strategy took place at room temperature and finished within 30 min, but substrates bearing C2-O-acyl protected moiety (disarmed sugars) failed.36-39 Presumably, this is due to the formation of the oxocarbenium ion intermediate, which is disfavored by the electron withdrawing property of the acyl protecting group and complicated by the neighbouring group participation of the C2-OAc/OBz moiety.40-41 Given that radical reactions often proceed under mild reaction conditions and the corresponding glycosyl radicals stereoelectronically favor the formation of the α-epimer,8, 42-45 we posited that an open-shell, radical C1-ketonylation of carbohydrate derivatives could provide a complementary strategy to achieve C-glycosylation with high levels of α-selectivity and functional group compatibility.
While developing the excited-state Pd-catalyzed C2-ketonylation of 1-bromosugars,46 we observed the formation of C1 α-ketonylated adduct as a side-product. We questioned whether we could establish a general and highly α-selective C1-ketonylation through excited-state palladium catalysis.47 This catalytic approach allows access to both close- and open-shell species under mild conditions and has recently been utilized in a number of radical transformations of aryl or alkyl halides.48-63 Herein, we report our success in exploiting this excited-state catalytic platform and readily accessible 1-bromosugars to generate the palladoradical intermediate ([PdI]Br) and glycosyl radicals, which can undergo polarity-mismatched coupling with nucleophilic silyl enol ethers, selectively forming the desired α-coupling product (Figure 1C). Preliminary mechanistic studies suggest that the subsequent reaction proceeds through bromine atom transfer followed by H-Br elimination, furnishing the silyl enol ether that is hydrolyzed to deliver the desired C1-ketonylated carbohydrates. In addition to these mechanistic insights, the reaction (i) tolerates disarmed sugars,64-65,66-67 (ii) exhibits broad substrate scope and wide functional group compatibility, (iii) features excellent α-selectivity, (iv) is amenable to late-stage functionalization of complex molecules, and (v) allows rapid access to a range of α- and β-C-glycosides via various post-functionalization reactions.
Results and Discussion
At the outset of our investigation, we selected the challenging combination of a readily available acetylated α-glucosyl bromide (1a) and acetophenone trimethylsilyl enol ether (2a) as model substrates for initial optimization studies, expecting that a broad reaction scope, including disarmed sugars would be observed (Table 1). When 1a and 2a were treated with 5.00 mol% Pd(PPh3)4, 6.00 mol% Xantphos [(9,9-dimethyl-9H-xanthene-4,5-diyl)bis(diphenylphosphane)], 1.50 equiv KOAc, and 10.0 equiv H2O in dioxane (0.10 M) at room temperature (rt) under irradiation from 24 W blue LEDs for 24 h (entry 1), the desired product (3a) was obtained in 95% yield with >20:1 α-selectivity. The Pd(PPh3)4 catalyst was critical for the desired reactivity: no reaction occurred in its absence, and only 14% of the desired product was formed when it was replaced with Pd(OAc)2 (entries 2 & 3). Removing Xantphos or replacing it with BINAP decreased the reaction yield (entries 4 & 5). While other common photocatalysts, such as Ru(bpy)3(PF6)2 and Eosin Y free acid failed, Ir(ppy)3 gave only 24% yield (entries 6-8). Using MeCN as the solvent instead of dioxane significantly lowered the yield (entry 9). Water, which is presumed to hydrolyze the silyl enol ether intermediate to form the ketone product, was crucial for a high yield of the product (entry 10). Finally, oxygen-free conditions and light were essential for high reaction efficiency (entries 11 and 12).
Table 1.
Selected optimization experimentsa
![]() | |||
---|---|---|---|
Entry | Deviation from standard conditions | Yield (%) | α/β |
1 | None | 95 | >20:1 |
2 | Without Pd(PPh3)4 | N.R. | - |
3 | Pd(OAc)2 instead of Pd(PPh3)4 | 14 | - |
4 | Without Xantphos | 69 | >20:1 |
5 | BINAP instead of Xantphos | 70 | 18:1 |
6 | Ir(ppy)3 as photocatalyst | 24 | - |
7 | Ru(bpy)3(PF6)2 as photocatalyst | N.R. | - |
8 | Eosin Y free acid as photocatalyst | N.R. | - |
9 | MeCN as solvent | 7 | - |
10 | Without H2O | 45 | - |
11 | Air | N.R. | - |
12 | Keep in dark | N.R. | - |
See Supplementary Information (SI) for experimental details. Reaction yields and α/β ratios were determined by 1H-NMR using CH2Br2 as an internal standard. Ac, acetyl; BINAP, 2,2′-bis(diphenylphosphino)-1,1′-binaphthyl; LED, light-emitting diode; N.R., no reaction.
With the optimized reaction conditions in hand, we examined the generality of the reaction and found that a range of silyl enol ethers reacted with α-glucosyl bromide, forming the desired products in moderate to excellent yields (Table 2A). Different silyl enol ethers with electron-neutral (2a), electron-withdrawing (2b-2e), or electron-donating (2f) substituents on the aryl ring delivered the corresponding α-ketonyl glucosides (3a-3f) in 55-87% yields and with up to 20:1 α-selectivity. Aryl silyl enol ethers with an extended conjugation (2g) or multiple substituents (2h-2i) were compatible. Notably, medicinally relevant heteroaryl derivatives, including pyridyl (2j), furanyl (2k), and thienyl substituted silyl enol ethers (2l), were viable substrates, furnishing the desired products (3j-3l) in good to excellent yields and with high levels of α-selectivity.68 The reaction is amenable to gram scale synthesis with similar efficiency (3l, Fig. S1). The absolute stereochemistry of the product 3a was confirmed by a single-crystal analysis,69 as shown in Table 3, and 2D-NMR spectroscopy. The stereochemistry of the other products was tentatively assigned by analogy.
Table 2.
Scope of excited-state palladium-catalyzed α-selective ketonylation of carbohydratesa
![]() |
See SI for experimental details. Isolated yield and α/β ratio are indicated below each entry
Table 3.
Post-functionalization of α-ketonylsugarsa
![]() |
See SI for experimental details. Isolated yield and α/β ratio are indicated below each entry.
We evaluated the scope of α-bromosugars under the standard reaction conditions. A broad array of α-bromosugars derived from D-galactose, D-xylose, D-glucose, D-mannose, L-fucose, and L-rhamnose (3m-3q, 3s, 3t) reacted with silyl enol ether 2a, affording the desired products in 64-87% yields and with up to >20:1 α-selectivity (Table 2B). Protecting groups such as tert-butyldiphenylsilyl, acetal, and ketal were well tolerated (3o-3q). A D-mannofuranose derivative reacted smoothly to generate the desired product 3r in 95% yield and with complete α-selectivity. An analog of D-glucuronic acid also proved to be compatible with the standard conditions (3u). Acetyl-protected disaccharides, such as cellobiose, maltose, and melibiose, underwent C1-ketonylation, furnishing the corresponding products 3v-3x in 74-87% yields and with 10:1 to >20:1 stereoselectivity.
Late-stage modification of complex molecules is often a key to identify medicinal agents.70 To demonstrate the applicability of the excited-state Pd-catalyzed C1-ketonylation to late-stage syntheses, natural product- and drug-conjugated sugars were subjected to the standard reaction conditions (Table 2C). For example, α-bromosugar derivatives of L-menthol, Febuxostat, Adapalene, Probenecid, Ibuprofen, Indomethacin, oleanolic acid, and Zaltoprofen reacted, affording the desired products 5a-5h in 58-90% yields and with up to >20:1 α-selectivity.
The α-ketonylsugars produced in this reaction are versatile synthetic intermediates that can be used to prepare other valuable C-glycosides. For instance, the α-isomers could be epimerized to their β-isomers under mild basic conditions with up to 93% yield and >20:1 β-selectivity (Table 3A). The epimerization proceeds through the base-mediated ring-opening followed by a Michael addition reaction, aligning with our computational results (Fig. S13). Moreover, the ketonylsugars could also undergo hydrogenation, olefination, Baeyer-Villiger oxidation, Beckmann rearrangement, and reduction to afford the corresponding alkylated C-glucosides in moderate to good yields (Table 3B-3E and Fig. S1).
To shed more light on the reaction mechanism and the origin of the stereoselectivity, we conducted a series of experimental and computational studies. UV-Vis measurements showed that a mixture of Pd(PPh3)4 and Xantphos had a strong absorption at 347 nm (Figure 2A). This observation suggested that the ligand exchange forms [Pd0(PPh3)2Xantphos], which might be the active catalyst.55 Irradiating the reaction mixture with blue light for 8 min shifted the λmax from 347 nm to 360 nm, which implicates the formation of [PdII]-species.63 Stern-Volmer quenching studies demonstrated that α-glucosyl bromide (1a) quenches the excited palladium species more efficiently than phenyl silyl enol ether (2a, Figure 2B). The addition of a radical scavenger such as butylated hydroxytoluene (BHT) or 2,2,6,6-tetramethylpiperidine 1-oxyl (TEMPO) inhibited the reaction (Figure 2C). The TEMPO-glucosyl adduct (12) was also detected by HRMS (Fig. S5). In addition, the radical clock experiment using cyclopropyl silyl enol ether 13 afforded the ring-opening product 14 in 22% yield (Figure 2D). These results suggest that the reaction likely proceeds through a radical pathway. Moreover, radical chain propagation is unlikely because the quantum yield of the reaction is 0.02 (Fig. S10). We did not observe the formation of TMS-protected silyl enol ether products, presumably, the labile TMS group was deprotected under the reaction. Indeed, performing the reaction using TBS-protected silyl enol ether 2a’ in the absence of water additive furnished silyl enol ether product 3a’ in 61% yield accompanied by 22% ketone product 3a, suggesting that 3a’ is an intermediate en route to the ketone product (Figure 2E). Intermolecular kinetic isotope effect studies using 1:1 of 2a’:2a’-d2 furnished the product 3a’:3a’-d1 in a 1.1:1 ratio, showing that cleavage of the C-H bond is not the rate-determining step (Figure 2F).
Figure 2.
Mechanistic studies of C1-ketonylation of carbohydrates. aThe dip at around 350 nm is the instrumental artifact as the instrument switches the light sources. bAn additional 22% of ketone product 3a was obtained as well. cDFT calculations were performed at the M06/SDD-6-311+G(d,p)/SMD//B3LYP-D3/SDD-6-31G(d) level of theory using a simplified model of the glucosyl radical(1), where OMe groups were used in place of the OAc groups at the C3, 4, and 6 positions of the pyranose ring.
DFT calculations showed that α-radical addition to the silyl enol ether via transition state TS1α to form radical intermediate IIIα, is 1.4 kcal/mol more favorable than the formation of the β anomer IIIβ via TS1β (Fig. S11). This data corroborates the experimental results and literature reports.8, 42-44 The stereoselectivity originates from the more favorable chair-like conformation of TS1α, in which the silyl enol ether approaches from the axial position of the glycosyl radical. On the other hand, the chair conformer of the β-radical addition transition state TS1β′, where the silyl enol ether approaches from the less favorable equatorial position (ΔG‡ = 13.4 kcal/mol, Fig. S12), is less stable. The most stable conformer of the β-radical addition transition state has a twist-boat geometry (TS1β), and is 1.4 kcal/mol higher in energy than TS1α. Although the polarity-mismatched addition of nucleophilic radicals to electron-rich alkenes is rare, DFT calculations showed that the addition of nucleophilic anomeric radicals to silyl enol ethers is energetically feasible (ΔG≠ = 10.8 kcal/mol) and the formation of stabilized α-OTMS alkyl radical IIIα drives the stereo-determining, irreversible C-C bond-forming event (Fig. S11).71 Subsequent silyl enol ether formation from radical intermediate IIIα could proceed through at least four distinct reaction pathways: (P1) single electron transfer (SET) from IIIα to [PdI]Br to form carbocation followed by deprotonation; (P2) recombination of radical IIIα with [PdI]Br to form PdII intermediate IV followed by β-hydride elimination; (P3) a palladoradical β-H-atom abstraction;72-81 and (P4) bromine atom transfer from [PdI]Br to IIIα followed by HBr elimination (Figure 2G). DFT calculations showed that P1 is highly disfavored with ΔG = 66.6 kcal/mol with respect to IIIα (Fig. S14). For the P2, the recombination of radical IIIα with [PdI]Br to form PdII intermediate IV is endergonic by 4.2 kcal/mol (Fig. S11). Here, the [PdII]-C bond is weakened by steric repulsions between the tertiary carbon center and the large-bite-angle of the Xantphos ligand (BDE = 11.6 kcal/mol, Table S2). From IV, subsequent β-hydride elimination (TS4) requires a relatively higher barrier (ΔG‡ = 16.8 kcal/mol with respect to IIIα) because one of the P-arms of the Xantphos ligand needs to be dissociated prior to the β-hydride elimination.82 Consequently, intermediate IV is prone to undergo a backward reaction to form radical IIIα and [PdI]Br, especially under the photoexcitation conditions. DFT calculations also showed that the palladoradical β-H-atom abstraction (P3) has a higher energy barrier with ΔG‡ = 22.3 kcal/mol with respect to IIIα. These computational results implied that the rate-determining step of both the pathways P283 and P384 involve the C-H bond cleavage. However, the intermolecular KIE studies gave a non-first order KIE of 1.1 (Figure 2E),85 suggesting that mechanistic pathways P2 and P3 are unlikely. On the other hand, the non-first order KIE is more in line with the bromine atom transfer followed by a rapid H-Br elimination (P4).86-88 Although we failed to locate the transition state of the bromine atom transfer, DFT calculations showed that the formation of the α-OTMS alkyl bromide from [PdI]Br and radical IIIα is endergonic by only 4.6 kcal/mol, which is energetically feasible.
While a precise reaction mechanism awaits further study, a plausible catalytic cycle is shown in Figure 2. A photoexcited [Pd0]* species abstracts the bromine atom from 1-bromosugar 1 to afford a [PdI]Br complex and 1-glycosyl radical intermediate IIa, which is in equilibrium with IIb under visible-light irradiation conditions. Guided by a stereoelectronic effect,8, 42-45 II adds to silyl enol ether 2 with a high level of α-selectivity, furnishing intermediate IIIα. Subsequent bromine atom transfer followed by H-Br elimination liberates [Pd0] and enol ether 3’’, which is hydrolyzed to deliver the desired α-selective C1-ketonylated product.
Conclusion
In summary, we reported a visible-light-induced excited-state palladium-catalyzed α-selective radical C-glycosylation of carbohydrates using readily available 1-bromosugars and silyl enol ethers. Preliminary experimental and computational studies of the reaction mechanism suggested a non-chain radical mechanism. The reaction features excellent stereoselectivity, broad substrate scope, and mild conditions that tolerate various functional groups and complex molecular structures. The resulting α-ketonylated products could serve as precursors for a range of valuable C-glycosides. This general and α-selective strategy, which allows facile access to both α- and β-glycosides, should find applications in drug discovery and chemical biology research.
Supplementary Material
Acknowledgments (required, if applicable)
We thank Dr Vincent M. Lynch (University of Texas, Austin) for the X-ray data of 3a.
Funding Information
The research reported in this publication was supported by the National Institutes of Health (R35-GM119652 to M.-Y.N. and R35-GM128779 to P.L.). DFT calculations were performed at the Center for Research Computing at the University of Pittsburgh, the TACC Frontera supercomputer, and the Extreme Science and Engineering Discovery Environment (XSEDE) supported by the National Science Foundation grant number ACI1548562. The Shimadzu UPLC/MS used for portions of this work were purchased with funds from NIGMS equipment administrative supplement (R35-GM119652-04S1), Shimadzu Scientific Instruments grant, and Office of the Vice President for Research at Stony Brook University.
Footnotes
Supporting Information
Supporting Information is available and includes all the experimental details and compound characterizations.
Conflict of Interest
There is no conflict of interest to report.
References
- (1).Varki A; Cummings RD; Esko JD; Stanley P; Hart GW; Aebi M; Darvill AG; Kinoshita T; Packer NH; Prestegard JH; Schnaar RL; Seeberger PH, Essentials of Glycobiology. 3rd ed.; Cold Spring Harbor Press: Cold Spring Harbor, New York, 2017. [PubMed] [Google Scholar]
- (2).Yang Y; Yu B Recent advances in the chemical synthesis of C-glycosides. Chem. Rev 2017, 117, 12281–12356. [DOI] [PubMed] [Google Scholar]
- (3).Wang H-Y; Blaszczyk SA; Xiao G; Tang W Chiral reagents in glycosylation and modification of carbohydrates. Chem. Soc. Rev 2018, 47, 681–701. [DOI] [PubMed] [Google Scholar]
- (4).Nielsen MM; Pedersen CM Catalytic glycosylations in oligosaccharide synthesis. Chem. Rev 2018, 118, 8285–8358. [DOI] [PubMed] [Google Scholar]
- (5).Kulkarni SS; Wang C-C; Sabbavarapu NM; Podilapu AR; Liao P-H; Hung S-C “One-pot” protection, glycosylation, and protection–glycosylation strategies of carbohydrates. Chem. Rev 2018, 118, 8025–8104. [DOI] [PubMed] [Google Scholar]
- (6).Bennett CS; Galan MC Methods for 2-deoxyglycoside synthesis. Chem. Rev 2018, 118, 7931–7985. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (7).Hettikankanamalage AA; Lassfolk R; Ekholm FS; Leino R; Crich D Mechanisms of stereodirecting participation and ester migration from near and far in glycosylation and related reactions. Chem. Rev 2020, 120, 7104–7151. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (8).Xu L-Y; Fan N-L; Hu X-G Recent development in the synthesis of C-glycosides involving glycosyl radicals. Organic & Biomolecular Chemistry 2020, 18, 5095–5109. [DOI] [PubMed] [Google Scholar]
- (9).Zou W C-glycosides and aza-C-glycosides as potential glycosidase and glycosyltransferase inhibitors. Curr. Top. Med. Chem 2005, 5, 1363–1391. [DOI] [PubMed] [Google Scholar]
- (10).Compain P; Martin OR Carbohydrate mimetics-based glycosyltransferase inhibitors. Bioorg. Med. Chem 2001, 9, 3077–3092. [DOI] [PubMed] [Google Scholar]
- (11).Franck RW; Tsuji M α-C-galactosylceramides: synthesis and immunology. Acc. Chem. Res 2006, 39, 692–701. [DOI] [PubMed] [Google Scholar]
- (12).Koester DC; Holkenbrink A; Werz DB Recent advances in the synthesis of carbohydrate mimetics. Synthesis 2010, 2010, 3217–3242. [Google Scholar]
- (13).Yang G; Schmieg J; Tsuji M; Franck RW The C-glycoside analogue of the immunostimulant α-galactosylceramide (KRN7000): synthesis and striking enhancement of activity. Angew. Chem. Int. Ed 2004, 43, 3818–3822. [DOI] [PubMed] [Google Scholar]
- (14).Sok J; Pineau N; Dalko-Csiba M; Breton L; Bernerd F Improvement of the dermal epidermal junction in human reconstructed skin by a new c-xylopyranoside derivative. Eur. J. Dermatol 2008, 18, 297–302. [DOI] [PubMed] [Google Scholar]
- (15).Donohoe TJ; Winship PC; Tatton MR; Szeto P A short and efficient synthesis of neodysiherbaine A by using catalytic oxidative cyclization. Angew. Chem 2011, 123, 7746–7748. [DOI] [PubMed] [Google Scholar]
- (16).Zeng J; Vedachalam S; Xiang S; Liu X-W Direct C-glycosylation of organotrifluoroborates with glycosyl fluorides and its application to the total synthesis of (+)-Varitriol. Org. Lett 2011, 13, 42–45. [DOI] [PubMed] [Google Scholar]
- (17).Liu P; Jacobsen EN Total synthesis of (+)-ambruticin. J. Am. Chem. Soc 2001, 123, 10772–10773. [DOI] [PubMed] [Google Scholar]
- (18).Nishikawa T; Adachi M; Isobe M C-Glycosylation. Glycoscience 2008, 755. [Google Scholar]
- (19).Zhu F; Rourke MJ; Yang T; Rodriguez J; Walczak MA Highly stereospecific cross-coupling reactions of anomeric stannanes for the synthesis of C-aryl glycosides. J. Am. Chem. Soc 2016, 138, 12049–12052. [DOI] [PubMed] [Google Scholar]
- (20).Liao H; Ma J; Yao H; Liu X-W Recent progress of C-glycosylation methods in the total synthesis of natural products and pharmaceuticals. Organic & biomolecular chemistry 2018, 16, 1791–1806. [DOI] [PubMed] [Google Scholar]
- (21).Sakamoto K; Nagai M; Ebe Y; Yorimitsu H; Nishimura T Iridium-Catalyzed Direct Hydroarylation of Glycals via C–H Activation: Ligand-Controlled Stereoselective Synthesis of α-and β-C-Glycosyl Arenes. ACS Catal. 2019, 9, 1347–1352. [Google Scholar]
- (22).Allevi P; Anastasia M; Ciuffreda P; Fiecchi A; Scala A Epimerization of α-to β-C-glucopyranosides under mild basic conditions. J. Chem. Soc., Perkin Trans 1 1989, 1275–1280. [Google Scholar]
- (23).Shao H; Wang Z; Lacroix E; Wu S-H; Jennings HJ; Zou W Novel Zinc (II)-Mediated Epimerization of 2 ‘-Carbonylalkyl-α-C-glycopyranosides to Their β-Anomers. J. Am. Chem. Soc 2002, 124, 2130–2131. [DOI] [PubMed] [Google Scholar]
- (24).Riemann I; Fessner W-D; Papadopoulos M; Knorst M C-Glycosides by aqueous condensation of β-dicarbonyl compounds with unprotected sugars. Aust. J. Chem 2002, 55, 147–154. [Google Scholar]
- (25).Hersant Y; Abou-Jneid R; Canac Y; Lubineau A; Philippe M; Semeria D; Radisson X; Scherrmann M-C One-step synthesis of β-C-glycolipid derivatives from unprotected sugars. Carbohydr. Res 2004, 339, 741–745. [DOI] [PubMed] [Google Scholar]
- (26).Wang J; Li Q; Ge Z; Li R A versatile and convenient route to ketone C-pyranosides and ketone C-furanosides from unprotected sugars. Tetrahedron 2012, 68, 1315–1320. [Google Scholar]
- (27).Khatri V; Kumar A; Singh B; Malhotra S; Prasad AK Synthesis of β-C-Glycopyranosyl Aldehydes and 2, 6-Anhydro-heptitols. J. Org. Chem 2015, 80, 11169–11174. [DOI] [PubMed] [Google Scholar]
- (28).Johnson S; Bagdi AK; Tanaka F C-Glycosidation of Unprotected Di-and Trisaccharide Aldopyranoses with Ketones Using Pyrrolidine-Boric Acid Catalysis. J. Org. Chem 2018, 83, 4581–4597. [DOI] [PubMed] [Google Scholar]
- (29).Ranoux A; Lemiegre L; Benvegnu T Synthesis and evaluation of C-glycosides as hydrotropes and solubilizing agents. Sci. China Chem 2010, 53, 1957–1962. [Google Scholar]
- (30).Godage HY; Chambers DJ; Evans GR; Fairbanks AJ Stereoselective synthesis of C-glycosides from carboxylic acids: the tandem Tebbe–Claisen approach. Organic & biomolecular chemistry 2003, 1, 3772–3786. [DOI] [PubMed] [Google Scholar]
- (31).Lichtenthaler FW; Lergenmüller M; Schwidetzky S C-Glycosidations of a 2-Ketohexosyl Bromide with Electrophilic, Radical, and Nucleophilic Anomeric Carbons. Eur. J. Org. Chem 2003, 2003, 3094–3103. [Google Scholar]
- (32).Ling J; Bennett CS Versatile Glycosyl Sulfonates in β-Selective C-Glycosylation. Angew. Chem. Int. Ed 2020, 59, 4304–4308. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (33).Liu M; Li B-H; Li T; Wu X; Liu M; Xiong D-C; Ye X-S C-Glycosylation enabled by N-(glycosyloxy) acetamides. Organic & biomolecular chemistry 2020, 18, 3043–3046. [DOI] [PubMed] [Google Scholar]
- (34).Ye W; Stevens CM; Wen P; Simmons CJ; Tang W Mild Cu (OTf) 2-Mediated C-Glycosylation with Chelation-Assisted Picolinate as a Leaving Group. J. Org. Chem 2020, 85, 16218–16225. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (35).Chen X; Wang Q; Yu B Gold (I)-catalyzed C-glycosylation of glycosyl ortho-alkynylbenzoates: the role of the moisture sequestered by molecular sieves. Chem. Commun 2016, 52, 12183–12186. [DOI] [PubMed] [Google Scholar]
- (36).“Armed” sugars are sugars with the ether protecting group and “disarmed” sugars are sugars with the acyl protecting group or with cyclic protecting groups that disfavor the formation of oxocarbenium ion intermediate, see Mootoo DR; Konradsson P; Udodong U; Fraser-Reid B, Armed and Disarmed N-Pentenyl Glycosides in Saccharide Couplings Leading to Oligosaccharides. J. Am. Chem. Soc 1988, 110, 5583–5584. See also refs 37-39. [Google Scholar]
- (37).Fraser-Reid B; Wu Z; Andrews CW; Skowronski E; Bowen JP Torsional effects in glycoside reactivity: saccharide couplings mediated by acetal protecting groups. J. Am. Chem. Soc 1991, 113, 1434–1435. [Google Scholar]
- (38).Crich D; Li M Revisiting the armed− disarmed concept: the importance of anomeric configuration in the activation of s-benzoxazolyl glycosides. Org. Lett 2007, 9, 4115–4118. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (39).Fraser-Reid B; López JC Armed–disarmed effects in carbohydrate chemistry: history, synthetic and mechanistic studies. Reactivity Tuning in Oligosaccharide Assembly 2010, 1–29. [DOI] [PubMed] [Google Scholar]
- (40).Levy DE; Fügedi P The organic chemistry of sugars; CRC Press, 2005. [Google Scholar]
- (41).Zhu X; Schmidt RR New principles for glycoside-bond formation. Angew. Chem. Int. Ed 2009, 48, 1900–1934. [DOI] [PubMed] [Google Scholar]
- (42).Dupuis J; Giese B; Rüegge D; Fischer H; Korth HG; Sustmann R Conformation of Glycosyl Radicals: Radical Stabilization by β-CO Bonds. Angew. Chem. Int. Ed 1984, 23, 896–898. [Google Scholar]
- (43).Korth H-G; Sustmann R; Gröninger KS; Witzel T; Giese B Electron spin resonance spectroscopic investigation of carbohydrate radicals. Part 3. Conformation in deoxypyranosan-2-, −3-, and −4-yl radicals. J. Chem. Soc., Perkin Trans 2 1986, 1461–1464. [Google Scholar]
- (44).Giese B The stereoselectivity of intermolecular free radical reactions [new synthetic methods (78)]. Angew. Chem. Int. Ed 1989, 28, 969–980. [Google Scholar]
- (45).Andrews RS; Becker JJ; Gagné MR Intermolecular addition of glycosyl halides to alkenes mediated by visible light. Angew. Chem 2010, 122, 7432–7434. [DOI] [PubMed] [Google Scholar]
- (46).Zhao G; Mukherjee U; Zhou L; Wu Y; Yao W; Mauro JN; Liu P; Ngai M-Y C2-ketonylation of carbohydrates via excited-state palladium-catalyzed 1, 2-spin-center shift. Chem. Sci 2022, 13, 6276–6282. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (47).A beautiful work by Liang showed the feasibility of excited-state Pd-catalyzed C1-ketonylation in 30% yield with β-selectivity. Li M; Qiu Y-F; Wang C-T; Li X-S; Wei W-X; Wang Y-Z; Bao Q-F; Ding Y-N; Shi W-Y; Liang Y-M, Visible-Light-Induced Pd-Catalyzed Radical Strategy for Constructing C-Vinyl Glycosides. Org. Lett 2020, 22, 6288–6293. [DOI] [PubMed] [Google Scholar]
- (48).Parasram M; Chuentragool P; Sarkar D; Gevorgyan V Photoinduced Formation of Hybrid Aryl Pd-Radical Species Capable of 1,5-HAT: Selective Catalytic Oxidation of Silyl Ethers into Silyl Enol Ethers. J. Am. Chem. Soc 2016, 138, 6340–6343. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (49).Kurandina D; Parasram M; Gevorgyan V Visible Light-Induced Room-Temperature Heck Reaction of Functionalized Alkyl Halides with Vinyl Arenes/Heteroarenes. Angew. Chem. Int. Ed 2017, 56, 14212–14216. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (50).Wang G-Z; Shang R; Cheng W-M; Fu Y Irradiation-induced Heck reaction of unactivated alkyl halides at room temperature. J. Am. Chem. Soc 2017, 139, 18307–18312. [DOI] [PubMed] [Google Scholar]
- (51).Zhou WJ; Cao GM; Shen G; Zhu XY; Gui YY; Ye JH; Sun L; Liao LL; Li J; Yu DG Visible-Light-Driven Palladium-Catalyzed Radical Alkylation of C–H Bonds with Unactivated Alkyl Bromides. Angew. Chem. Int. Ed 2017, 56, 15683–15687. [DOI] [PubMed] [Google Scholar]
- (52).Cheng WM; Shang R; Fu Y Irradiation-induced palladium-catalyzed decarboxylative desaturation enabled by a dual ligand system. Nat. Commun 2018, 9, 5215. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (53).Zhao B; Shang R; Wang G-Z; Wang S; Chen H; Fu Y Palladium-Catalyzed Dual Ligand-Enabled Alkylation of Silyl Enol Ether and Enamide under Irradiation: Scope, Mechanism, and Theoretical Elucidation of Hybrid Alkyl Pd (I)-Radical Species. ACS Catal. 2019, 10, 1334–1343. [Google Scholar]
- (54).Chuentragool P; Kurandina D; Gevorgyan V Catalysis with Palladium Complexes Photoexcited by Visible Light. Angew. Chem. Int. Ed 2019, 58, 11586–11598. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (55).Zhou W-J; Cao G-M; Zhang Z-P; Yu D-G Visible Light-induced Palladium-catalysis in Organic Synthesis. Chem. Lett 2019, 48, 181–191. [Google Scholar]
- (56).Chuentragool P; Yadagiri D; Morita T; Sarkar S; Parasram M; Wang Y; Gevorgyan V Aliphatic Radical Relay Heck Reaction at Unactivated C (sp3)–H Sites of Alcohols. Angew. Chem. Int. Ed 2019, 58, 1794–1798. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (57).Torres GM; Liu Y; Arndtsen BA A dual light-driven palladium catalyst: Breaking the barriers in carbonylation reactions. Science 2020, 368, 318–323. [DOI] [PubMed] [Google Scholar]
- (58).Ratushnyy M; Kvasovs N; Sarkar S; Gevorgyan V Visible-Light-Induced Palladium-Catalyzed Generation of Aryl Radicals from Aryl Triflates. Angew. Chem. Int. Ed 2020, 59, 10316–10320. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (59).Huang H-M; Bellotti P; Pflueger PM; Schwarz JL; Heidrich B; Glorius F A Three-Component, Interrupted Radical Heck/Allylic Substitution Cascade Involving Unactivated Alkyl Bromides. J. Am. Chem. Soc 2020, 142, 10173–10183. [DOI] [PubMed] [Google Scholar]
- (60).Shing Cheung KP; Kurandina D; Yata T; Gevorgyan V Photoinduced Palladium-Catalyzed Carbofunctionalization of Conjugated Dienes Proceeding via Radical-Polar Crossover Scenario: 1,2-Aminoalkylation and Beyond. J. Am. Chem. Soc 2020, 142, 9932–9937. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (61).Kvasovs N; Iziumchenko V; Palchykov V; Gevorgyan V Visible Light-Induced Pd-Catalyzed Alkyl-Heck Reaction of Oximes. ACS Catal. 2021, 11, 3749–3754. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (62).Cheung KPS; Sarkar S; Gevorgyan V Visible Light-Induced Transition Metal Catalysis. Chem. Rev 2022, 122, 1543–1625. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (63).Zhao G; Yao W; Mauro JN; Ngai MY Excited-State Palladium-Catalyzed 1,2-Spin-Center Shift Enables Selective C-2 Reduction, Deuteration, and Iodination of Carbohydrates. J. Am. Chem. Soc 2021, 143, 1728–1734. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (64).Glycosyl donors protected with ether protecting group (armed) can be activated in the presence of disarmed donors, which is critical in developing one-pot oligosaccharide protocols, see: Douglas NL; Ley SV; Lücking U; Warriner SL, Tuning Glycoside Reactivity: New Tool for Efficient Oligosaccharide Synthesis. J. Chem. Soc., Perkin Trans 1 1998, 51–66. See also ref 65. [Google Scholar]
- (65).Zhang Z; Ollmann IR; Ye X-S; Wischnat R; Baasov T; Wong C-H Programmable one-pot oligosaccharide synthesis. J. Am. Chem. Soc 1999, 121, 734–753. [Google Scholar]
- (66).The peracetylated monosaccharide derivatives are particularly useful because they have been shown to passively diffuse through mammalian cell membranes and undergo subsequent deacetylation by cytosolic or ER esterases, see: Sarkar AK; Fritz TA; Taylor WH; Esko JD, Disaccharide Uptake and Priming in Animal Cells: Inhibition of Sialyl Lewis X by Acetylated Gal Beta 1--> 4glcnac Beta-O-Naphthalenemethanol. Proc. Natl. Acad. Sci 1995, 92, 3323–3327. See also ref. 67. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (67).Jacobs CL; Yarema KJ; Mahal LK; Nauman DA; Charters NW; Bertozzi CR Metabolic labeling of glycoproteins with chemical tags through unnatural sialic acid biosynthesis. Methods Enzymol. 2000, 327, 260–275. [DOI] [PubMed] [Google Scholar]
- (68).Styrenes are viable substrates that afford C1-alkenylated sugars, see Table S1 in the ESI.
- (69).CCDC 2171960 contains the supplementary crystallographic data for this paper. These data can be obtained free of charge via www.ccdc.cam.ac.uk/data_request/cif, or by email to data_request@ccdc.cam.ac.uk, or by contacting The Cambridge Crystallographic Data Centre, 12 Union Road, Cambridge CB2 1EZ, UK; fax: +44 1223 336033.
- (70).Cernak T; Dykstra KD; Tyagarajan S; Vachal P; Krska SW The medicinal chemist's toolbox for late stage functionalization of drug-like molecules. Chem. Soc. Rev 2016, 45, 546–576. [DOI] [PubMed] [Google Scholar]
- (71).We also explored the possibility of 1-glucosyl-Pd(II)Br species coordinates to a silyl enol ether followed by migratory insertion or σ-bond metathesis and reductive elimination steps. However, both of these steps are energetically less favorable than the addition of a 1-glucosyl radical to a silyl enol ether (see Fig. S13 in ESI).
- (72).Metalloradical β-H-atom abstraction is known in the case of transition metals, such as Co(II), Ni(I), and Cr(I). For examples of Co(II), see: Gridnev AA; Ittel SD; Fryd M; Wayland BB, Formation of Organocobalt Porphyrin Complexes from Reactions of Cobalt (II) Porphyrins and Dialkylcyanomethyl Radicals with Organic Substrates: Chemical Trapping of a Transient Cobalt Porphyrin Hydride. Organometallics 1993, 12, 4871–4880. See also refs. 73-76. [Google Scholar]
- (73).Gridnev AA; Ittel SD Catalytic chain transfer in free-radical polymerizations. Chem. Rev 2001, 101, 3611–3660. [DOI] [PubMed] [Google Scholar]
- (74).Tang L; Norton JR Factors affecting the apparent chain transfer rate constants of chromium metalloradicals: Mechanistic implications. Macromolecules 2006, 39, 8229–8235. [Google Scholar]
- (75).de Bruin B; Dzik WI; Li S; Wayland BB Hydrogen-Atom Transfer in Reactions of Organic Radicals with [CoII (por)].(por= Porphyrinato) and in Subsequent Addition of [Co (H)(por)] to Olefins. Chem. - Eur. J 2009, 15, 4312–4320. [DOI] [PubMed] [Google Scholar]
- (76).Crossley SW; Barabé F; Shenvi RA Simple, chemoselective, catalytic olefin isomerization. J. Am. Chem. Soc 2014, 136, 16788–16791. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (77).For an example of Ni(I), see: Kapat A; Sperger T; Guven S; Schoenebeck F E-Olefins through intramolecular radical relocation. Science 2019, 363, 391–396. [DOI] [PubMed] [Google Scholar]
- (78).For examples of Cr(I), see: Abramo GP; Norton JR Catalysis by C5Ph5Cr (CO) 3• of Chain Transfer during the Free Radical Polymerization of Methyl Methacrylate. Macromolecules 2000, 33, 2790–2792. [Google Scholar]
- (79).Tang L; Papish ET; Abramo GP; Norton JR; Baik M-H; Friesner RA; Rappé A Kinetics and thermodynamics of H• transfer from (η5-C5R5) Cr (CO) 3H (R= Ph, Me, H) to methyl methacrylate and styrene. J. Am. Chem. Soc 2003, 125, 10093–10102. [DOI] [PubMed] [Google Scholar]
- (80).Tang L; Norton JR Effect of steric congestion on the activity of chromium and molybdenum metalloradicals as chain transfer catalysts during MMA polymerization. Macromolecules 2004, 37, 241–243. [Google Scholar]
- (81).Choi J; Tang L; Norton JR Kinetics of Hydrogen Atom Transfer from (η5-C5H5) Cr (CO) 3H to Various Olefins: Influence of Olefin Structure. J. Am. Chem. Soc 2007, 129, 234–240. [DOI] [PubMed] [Google Scholar]
- (82).Xantphos-ligated Pd-catalyst could disfavor β-hydride elimination, see: Lopez-Perez A; Adrio J; Carretero JC Palladium-catalyzed cross-coupling reaction of secondary benzylic bromides with Grignard reagents. Org. Lett 2009, 11, 5514–5517. [DOI] [PubMed] [Google Scholar]
- (83).Previously reports showed that the two-electron Pd-catalyzed β-hydride elimination and has KIE = 3, see: Shmidt A; Smirnov V Kinetic study of the Heck reaction by the method of competing reactions. Kinet. Catal 2001, 42, 800–804. [Google Scholar]
- (84).Metalloradical β-H-atom abstractions often show first-order KIE, see ref. 80.
- (85).Simmons EM; Hartwig JF On the Interpretation of Deuterium Kinetic Isotope Effects in C-H Bond Functionalizations by Transition-Metal Complexes. Angew. Chem. Int. Ed 2012, 51, 3066–3072. [DOI] [PubMed] [Google Scholar]
- (86).In their detailed studies of excited-state Pd-catalyzed alkylation of arenes, Hong and coworkers also observed a non-first order KIE and proposed a bromine atom transfer from Pd(I)Br to cyclohexadienyl radical, furnishing an alkyl-bromide that undergoes fast H-Br elimination to form aromatic ring, see: Kim D; Lee GS; Kim D; Hong SH Direct C (sp2)–H alkylation of unactivated arenes enabled by photoinduced Pd catalysis. Nat. Commun 2020, 11, 1–13.31911652 [Google Scholar]
- (87).For example of chlorine-atom transfer from Pd(I)Cl, see: Lee GS; Kim D; Hong SH Pd-catalyzed formal Mizoroki–Heck coupling of unactivated alkyl chlorides. Nat. Commun 2021, 12, 991. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (88).For example of iodine-atom transfer from Pd(I)I, see: Ratushnyy M; Parasram M; Wang Y; Gevorgyan V Palladium-Catalyzed Atom-Transfer Radical Cyclization at Remote Unactivated C(sp3)−H Sites: Hydrogen-Atom Transfer of Hybrid Vinyl Palladium Radical Intermediates. Angew. Chem. Int. Ed 2018, 57, 2712–2715. [DOI] [PMC free article] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.