(Poly)Azines - American Chemical Society

ICREA, Passeig Lluís Companys, 23, 08010, Barcelona, Spain. Supporting Information Placeholder. ABSTRACT: A base-mediated protocol that allows for...
1 downloads 0 Views 1MB Size
Communication Cite This: J. Am. Chem. Soc. 2019, 141, 127−132

pubs.acs.org/JACS

A Mild and Direct Site-Selective sp2 C−H Silylation of (Poly)Azines Yiting Gu,†,¶,‡ Yangyang Shen,†,¶,‡ Cayetana Zarate,†,¶ and Ruben Martin*,†,§ †

Institute of Chemical Research of Catalonia (ICIQ), The Barcelona Institute of Science and Technology, Av. Països Catalans 16, 43007 Tarragona, Spain ¶ Departament de Química Analítica i Química Orgànica, Universitat Rovira i Virgili, c/Marcel·lí Domingo, 1, 43007 Tarragona, Spain § ICREA, Passeig Lluís Companys, 23, 08010 Barcelona, Spain Downloaded via IOWA STATE UNIV on January 9, 2019 at 12:31:07 (UTC). See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

S Supporting Information *

steric and electronic bias.11 If successful, we recognized that such a pathway would give access to elusive silylated advanced pharmaceutical ingredients (APIs) with predictable reactivity and site-selectivity. As part of our interest in catalytic C−Si bond-formations,12 we report herein the successful realization of this goal (Scheme 1, bottom). Our protocol is characterized by its simplicity, mild conditions and a predictable site-selectivity pattern, even in the context of late-stage silylation of azine drugs. Indeed, site-selectivity can be modulated by a judicious choice of the solvent employed while obviating the need for transition metal complexes or prefunctionalized substrates. Preliminary mechanistic studies suggest the intermediacy of silyl anions, offering orthogonal reactivity with existing silylation techniques. Our study began by evaluating the direct C4−silylation of pyridine (1a) with Et 3 SiBPin (Table 1). After some experimentation,13 we found that a transition metal-free protocol consisting of KHMDS in DME at rt afforded 2a in 74% isolated yield with excellent C4:C2 (10:1) ratio (entry 1). Interestingly, a strong solvent effect (entries 1−4) was noticed.

ABSTRACT: A base-mediated protocol that allows for the site-selective sp2 C−H silylation of azines is described. This method is distinguished by its mild conditions, simplicity and excellent site-selective modulation for a diverse set of azines, even in the context of late-stage functionalization, while exhibiting orthogonal reactivity with classical silylation reactions.

T

he prevalence of (poly)azines in pharmaceuticals and biologically active molecules has challenged chemists to design site-selective C−H functionalizations of these heterocycles.1,2 At present, these techniques typically rely on two-step procedures via preactivated (poly)azine intermediates (Scheme 1, path a).3 Despite the advances realized,2,3 a switchable and Scheme 1. Forging C−Heteroatom Bonds in (Poly)Azines

Table 1. Optimization of the Reaction Conditionsa

direct site-selective C−H functionalization capable of forging C−heteroatom bonds in (poly)azines still remains largely unexplored.4 For example, C4-selective functionalization is not as commonly practiced as one might initially anticipate, thus constituting a worthwhile endeavor.4−6 Driven by the versatility of (hetero)aryl silanes as synthons in organic synthesis and their interest in medicinal and materials science,7 the recent years have witnessed the design of metalcatalyzed sp2 C−H silylation reactions.8 While the coupling of electron-rich or electron-neutral (hetero)arenes has become routine,8,9 sp2 C−H silylation strategies of electron-deficient azines remain currently confined to C3-selective processes with noble Ir or Ru catalysts (Scheme 1, path b).10 Therefore, at the outset of our investigations it was unclear whether a switchable C2/C4−H silylation of pyridines could be developed without © 2018 American Chemical Society

entry

deviation standard conditions

yield (%)b

2a:3a

1 2 3 4 5 6 7 8 9 10 11

None diglyme as solvent 1a as solvent HMPA as solvent KOtBu instead of KHMDS Mg(HMDS)2 instead of KHMDS LiHMDS instead of KHMDS NaHMDS instead of KHMDS Et3SiH instead of Et3SiBPin no inert atmospheres (under air) from 1a N-oxide

80 (74)c 49 52 64 76 0 23 65 0 66 61

10:1 12:1 1.2:1 7:1 5:1 − 6:1 9:1 − 10:1 1:99

a 1a (0.40 mmol), Et3SiBPin (0.40 mmol), KHMDS (1 equiv), DME (2 mL) at rt. bGC yields using decane as internal standard. cIsolated yield, average of two independent runs.

Received: November 9, 2018 Published: December 18, 2018 127

DOI: 10.1021/jacs.8b12063 J. Am. Chem. Soc. 2019, 141, 127−132

Communication

Journal of the American Chemical Society

regioselectivity. As for Table 1 (entry 11), site-selectivity could be tuned and controlled by subtle modulation of the electronic properties of bipyridine cores via N-oxide derivatives (5y, 5z).18,19 The flexibility and generality of our site-selective silylation suggested that our protocol should be applicable within the context of late-stage silylation of azine drugs without requiring additional processing. As shown in Scheme 2, this turned out to

Specifically, C4 silylation was predominantly observed with bidentate solvents (entries 1−2) whereas the use of HMPA or 1a as solvent eroded both yield and selectivity (entries 3−4). Equally striking was the influence of the base and escorting counterion, thus revealing the noninnocent behavior of these variables on both reactivity and site-selectivity (entries 5−8).14 Moreover, Et3SiH as silyl source failed to provide 3a (entry 9)15,16 whereas good yields and selectivities were obtained under air (entry 10), thus constituting a bonus in practical terms.17 Interestingly, the utilization of N-oxide derivatives resulted in 3a as single regioisomer in 61% yield (entry 11).18,19 As evident from the results compiled in Table 2, systematic structural editing revealed that the C−H silylation can be

Scheme 2. Late-Stage Silylation of Azine Drugs

Table 2. Site-Selective Silylation of Azinesa,b

a

As Table 1 (entry 1). bIsolated yields, average of two independent runs. cKHMDS (2 equiv). d0 °C. e5 mmol scale. fFrom N-oxide derivative. gC4:C5 = 2.7:1.

be the case. C2- and C3-substituted pyridine drugs such as Doxylamine, Loratadine, Estrone and Nicotine exclusively delivered the corresponding C4-silylated pharmaceuticals (6, 8, 9 and 12) in good yields and site-selectivities.17 As evident from NMR spectroscopy,13 C2−Si bond-formation was solely obtained from Metyrapone (7) at the most electron-poor pyridyl backbone. A similar C2-selectivity pattern was observed with antiretroviral Nevirapine (10), an observation that could be confirmed by X-ray crystallography.21 Exhaustive C4-silylation could be applied to TmPyPB (11) whereas late-stage silylation was within reach for pyridimine-containing drugs such as Piribedil or Buspirone, affording 13 and 14 in good yields, the latter on a gram scale, as the only observable products. The non-negligible role exerted by both the solvent and escorting counterion suggested that C4-selectivity might be dictated by a solvent-separated ion pair, in which initial

applied to a wide number of azines. C4-silylated pyridines were invariably obtained in good yields and site-selectivities with azines possessing aliphatic or aromatic substituents at either C2 (5a, 5e−5h, 5k) or C3 (5b−5d, 5g−5j, 5l−5o), even with particularly sterically hindered combinations (5c, 5d).17 As anticipated, substituents at C4 gave rise to C2−silylated compounds (5p−5s). Notably, pyrazines (5u), pyridimines (5t), imidazo[1,2-b]pyridazines (5v), pyridimines (5w) or quinolines (5x) posed no problems. Equally interesting was the observation that ketones (5f), free amines (5k, 5l) or aryl chlorides (5m) do not interfere with C−Si bond-formation. The latter is particularly noteworthy, leaving ample room for further functionalization via cross-coupling reactions.20 Importantly, 5p could be scaled up without compromising neither yield nor 128

DOI: 10.1021/jacs.8b12063 J. Am. Chem. Soc. 2019, 141, 127−132

Communication

Journal of the American Chemical Society complexation of the nitrogen atom to K+ coordinated to DME confers a steric bias at C2 (Scheme 3, lef t).22,23 Exposure of 4p

Scheme 4. Intermediacy of Silyl Anion Species

Scheme 3. Regiodivergent Silylation of Pyridines

Scheme 5. Synthetic Application Profile

to KHMDS in THF-d8 corroborated this notion, showing a significant deshielding of the pyridine H atoms by 1H NMR spectroscopy.13 Although preactivated N-oxide azines offered a solution to enable a C2-silylation (Table 1, entry 11),19 we wondered whether a site-selectivity switch could be effected with unfunctionalized azines by a subtle modulation of the solvent denticity and aggregation.24 Interestingly, solvents lacking a proximal bidentate coordination mode such as dioxane predominantly resulted in a C2-silylation (3a−3e),25−27 suggesting the involvement of contacted ion pairs (Scheme 3, right).22,23 This interpretation gains credence by the erosion in C2-selectivity found with dioxane in the presence of DME or 18crown-6, with the latter providing exclusive access to 2a.13 The data summarized above suggest a scenario based on an initial complexation of the azine to K+ and the intermediacy of silyl anion species.28,29 This notion is supported by the reactivity found with 15 with Et3SiBPin (Scheme 4, right pathways). As shown, dearomatization was observed upon quenching with D2O or Me3SiCl (18−19), whereas their aromatized congeners (20−21) were cleanly obtained upon subsequent oxidation. To put these results into perspective, 16 and 17 were exclusively observed from 15 via KOtBu/Et3SiH-mediated silylation15a or Ni-catalyzed C−O cleavage (lef t pathways),12b thus showing the complementarity and orthogonal reactivity with other silylation events.30,31 The results shown in Scheme 5 further illustrates the prospective potential of our protocol, either triggering unusual defluorinative C−H silylations via nucleophilic attack followed by two consecutive [1,3]-H shifts (22)32 or as handles for subsequent functionalization via C−Si bond-cleavage (23− 24).33 In summary, we have developed a mild and site-selective C2/ C4-silylation of (poly)azines that obviates the need for transition metals, thus complementing existing sp2 C−H silylation events. The method can be applied for late-stage silylation of azine drugs, with a site-selectivity that can be

modulated by the nature of the solvent utilized. Further studies into the mechanism and related processes are currently ongoing.34



ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/jacs.8b12063. Data for C21H28N4OSi (CIF) Data for C27H45N5O2Si (CIF) Experimental procedures, crystallographic data and spectral data (PDF)



AUTHOR INFORMATION

Corresponding Author

*[email protected] ORCID

Yangyang Shen: 0000-0003-4607-9722 Cayetana Zarate: 0000-0002-4002-6147 Ruben Martin: 0000-0002-2543-0221 Author Contributions ‡

Y. Gu and Y. Shen contributed equally to this work.

Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS We thank ICIQ and MINECO (CTQ2015-65496-R) for support and the ICIQ X-ray Diffraction Unit for the crystallo129

DOI: 10.1021/jacs.8b12063 J. Am. Chem. Soc. 2019, 141, 127−132

Communication

Journal of the American Chemical Society

Commun. 2014, 50, 9273. (g) Boursalian, G. B.; Ngai, M.-Y.; Hojczyk, K. N.; Ritter, T. Pd- Catalyzed Aryl C−H Imidation with Arene as the Limiting Reagent. J. Am. Chem. Soc. 2013, 135, 13278. (h) Fier, P. S.; Hartwig, J. F. Selective C-H Fluorination of Pyridines and Diazines Inspired by a Classic Amination Reaction. Science 2013, 342, 956. (6) For a selection of elegant two-step procedures via phosphonium salts intermediates, see: (a) Dolewski, R. D.; Hilton, M. C.; McNally, A. 4-Selective Pyridine Functionalization Reactions via Heterocyclic Phosphonium Salts. Synlett 2018, 29, 8. (b) Anderson, R. G.; Jett, B. M.; McNally, A. A Unified Approach to Couple Aromatic Heteronucleophiles to Azines and Pharmaceuticals. Angew. Chem., Int. Ed. 2018, 57, 12514. (c) Hilton, M. C.; Dolewski, R. D.; McNally, A. Selective Functionalization of Pyridines via Heterocyclic Phosphonium Salts. J. Am. Chem. Soc. 2016, 138, 13806. (7) (a) Parasram, M.; Gevorgyan, V. Silicon-Tethered Strategies for C-H Functionalization Reactions. Acc. Chem. Res. 2017, 50, 2038. (b) Hiyama, T. In Organometallics in Synthesis; Schlosser, M., Ed.; John Wiley & Sons, Inc.: Hoboken, NJ, 2013; p 373. (c) Franz, A. K.; Wilson, S. O. Organosilicon Molecules with Medicinal Applications. J. Med. Chem. 2013, 56, 388. (d) Liu, X. − M.; He, C.; Huang, J.; Xu, J. Highly Efficient Blue-Light-Emitting Glass-Forming Molecules Based on Tetraarylmethane/Silane and Fluorene: Synthesis and Thermal, Optical, and Electrochemical Properties. Chem. Mater. 2005, 17, 434. (e) Bains, W.; Tacke, R. Silicon Chemistry as a Novel Source of Chemical Diversity in Drug Design. Curr. Opin. Drug Discovery Dev. 2003, 6, 526. (f) Showell, G. A.; Mills, J. S. Chemistry Challenges in Lead Optimization: Silicon Isosteres in Drug Discovery. Drug Discovery Today 2003, 8, 551. (8) For recent reviews: (a) Cheng, C.; Hartwig, J. F. Catalytic Silylation of Unactivated C-H Bonds. Chem. Rev. 2015, 115, 8946. (b) Xu, Z.; Huang, W.-S.; Zhang, J.; Xu, L.-W. Recent Advances in Transition-Metal-Catalyzed Silylations of Arenes with Hydrosilanes: C−X Bond Cleavage or C−H Bond Activation Synchronized with Si− H Bond Activation. Synthesis 2015, 47, 3645. (9) For selected references: (a) Maji, A.; Guin, S.; Feng, S.; Dahiya, A.; Singh, V. K.; Liu, P.; Maiti, D. Experimental and Computational Exploration of para-Selective Silylation with a hydrogen template. Angew. Chem., Int. Ed. 2017, 56, 14903. (b) Modak, A.; Patra, T.; Chowdhury, R.; Raul, S.; Maiti, D. Organometallics 2017, 36, 2418. (c) Elsby, M. R.; Johnson, S. A. Nickel-Catalyzed C-H silylation of Arenes with Vinylsilanes: Rapid and Reversible β-Si Elimination. J. Am. Chem. Soc. 2017, 139, 9401. (d) Kitano, T.; Komuro, T.; Ono, R.; Tobita, H. Organometallics 2017, 36, 2710. (e) Omann, L.; Oestreich, M. Catalytic Access to Infole-Fused Benzosiloles by 2-Fold Electrophilic C-H Silylation with Dihydrosilanes. Organometallics 2017, 36, 767. (f) Cheng, C.; Hartwig, J. F. Rhodium-Catalyzed Intermolecular C−H Silylation of Arenes with High Steric Regiocontrol. Science 2014, 343, 853. and citations therein. (g) Shippey, M. A.; Dervan, P. B. Trimethylsilyl Anions. Direct Synthesis of Trimethylsilylbenzenes. J. Org. Chem. 1977, 42, 2654. (h) Barry, A. J.; Gilkey, J. W.; Hook, D. E. In Metal-Organic Compounds; Advances in Chemistry; American Chemical Society: Washington, DC, 1959; Vol. 23, pp 246−264. (10) (a) Rubio-Pérez, L.; Iglesias, M.; Munárriz, J.; Polo, V.; Passarelli, V.; Pérez-Torrente, J. J.; Oro, L. A. Chem. Sci. 2017, 8, 4811. (b) Cheng, C.; Hartwig, J. F. Iridium-Catalyzed Silylation of Aryl C−H Bonds. J. Am. Chem. Soc. 2015, 137, 592. (c) Wübbolt, S.; Oestreich, M. Catalytic Electrophilic C-H Silylation of Pyridines Enabled by Temporary Dearomatization. Angew. Chem., Int. Ed. 2015, 54, 15876. (11) For an elegant Pd-catalyzed silylation event, see: (a) Oshima, K.; Ohmura, T.; Suginome, M. Palladium-Catalyzed Regioselective Silaboration of Pyridines Leading to the Synthesis of Silylated Dihydropyridines. J. Am. Chem. Soc. 2011, 133, 7324. For a two-step strategy via phosphonium salts, see: (b) Dolewski, R. D.; Fricke, P. J.; McNally, A. Site-Selective Switching Strategies to Functionalize Polyazines. J. Am. Chem. Soc. 2018, 140, 8020. For radical silylation of a specific substrate, see: (c) Du, W.; Kaskar, B.; Blumbergs, P.; Subramanian, P.-K.; Curran, D. P. Semisynthesis of DB-67 and other Silatecans from Camptothecin by Thiol-Promoted Addition of Silyl Radicals. Bioorg. Med. Chem. 2003, 11, 451.

graphic data. Y.S. and Y.G. thank China Scholarship Council (CSC) and MINECO for predoctoral fellowships.



REFERENCES

(1) (a) Blakemore, D. C.; Castro, L.; Churcher, I.; Rees, D. C.; Thomas, A. W.; Wilson, D. M.; Wood, A. Organic Synthesis Provides Opportunities to Transform Drug Discovery. Nat. Chem. 2018, 10, 383. (b) Vitaku, E.; Smith, D. T.; Njardarson, J. T. Analysis of the Structural Diversity, Substitution Patterns, and Frequency of Nitrogen Heterocycles among U.S. FDA Approved Pharmaceuticals. J. Med. Chem. 2014, 57, 10257. (c) Baumann, M.; Baxendale, I. R. An Overview of the Synthetic Routes to the Best-Selling Drugs Containing 6-Membered Heterocycles. Beilstein J. Org. Chem. 2013, 9, 2265. (d) Campeau, L.-C.; Fagnou, K. Applications of and Alternatives to π-Electron-Deficient Azine Organometallics in Metal-Catalyzed Cross-Coupling Reactions. Chem. Soc. Rev. 2007, 36, 1058. (2) (a) Das, R.; Kapur, M. Transition Metal-Catalyzed C−H Functionalization Reactions of π-Deficient Heterocycles. Asian J. Org. Chem. 2018, 7, 1217. (b) Wu, X.-F. Transition Metal-Catalyzed Heterocycle Synthesis via C−H Activation; Wiley: Hoboken, NJ, 2016. (c) Maes, J.; Maes, B. U. W. Chapter Five − A Journey Through Metalcatalyzed C−H Functionalization of Heterocycles: Insights and Trends. Adv. Heterocycl. Chem. 2016, 120, 137. (d) Yu, J. − T.; Pan, C. Radical C−H Functionalization to Construct Heterocyclic Compounds. Chem. Commun. 2016, 52, 2220. (e) Lapointe, D.; Markiewicz, T.; Whipp, C. J.; Toderian, A.; Fagnou, K. Predictable and Site-Selective Functionalization of Poly(hetero)arene Compounds by Palladium Catalysis. J. Org. Chem. 2011, 76, 749. (3) Selected references: (a) Ruiz-Castillo, P.; Buchwald, S. L. Applications of Palladium-Catalyzed C−N Cross-Coupling Reactions. Chem. Rev. 2016, 116, 12564. (b) Corcoran, E. B.; Pirnot, M. T.; Lin, S.; Dreher, S. D.; DiRocco, D. A.; Davies, I. W.; Buchwald, S. L.; MacMillan, D. W. Aryl Amination Using Ligand-Free Ni(II) Salts and Photoredox Catalysis. Science 2016, 353, 279. (c) Creutz, S. E.; Lotito, K. J.; Fu, G. C.; Peters, J. C. Photoinduced Ullmann C−N Coupling: Demonstrating the Viability of a Radical Pathway. Science 2012, 338, 647. (d) Bhunia, S.; Pawar, G. G.; Kumar, S. V.; Jiang, Y.; Ma, D. Selected Copper-Based Reactions for C−N, C−O, C−S and C−C Bond Formation. Angew. Chem., Int. Ed. 2017, 56, 16136. (e) Walsh, K.; Sneddon, H. F.; Moody, C. J. An Efficient Method for the Synthesis of Heteroaryl C−O Bonds in the Absence of Added Transition Metal Catalysts. RSC Adv. 2014, 4, 28072. (f) Martin, R.; Buchwald, S. L. Palladium-Catalyzed Suzuki-Miyaura Cross-Coupling Reactions Employing Dialkylbiaryl Phosphone Ligands. Acc. Chem. Res. 2008, 41, 1461. (g) Liu, J.; Robins, M. J. SNAr Displacements with 6-(Fluoro, Chloro, Bromo, Iodo and Alkylsulfonyl)Purine Nucleosides: Synthesis, Kinetics and Mechanism. J. Am. Chem. Soc. 2007, 129, 5962. (4) Selected references: (a) Murakami, K.; Yamada, S.; Kaneda, T.; Itami, K. C−H Functionalization of Azines. Chem. Rev. 2017, 117, 9302. (b) Nakao, Y. Transition Metal-Catalyzed C−H Functionalization for the Synthesis of Substituted Pyridines. Synthesis 2011, 2011, 3209. (5) For selected metal-catalyzed C−heteroatom bond-formations of azines, see: (a) Yang, L.; Semba, K.; Nakao, Y. Para-Selective C−H Borylation of (Hetero)Arenes by Cooperative Iridium/Aluminum Catalysis. Angew. Chem., Int. Ed. 2017, 56, 4853. (b) Ito, E.; Fukushima, T.; Kawakami, T.; Murakami, K.; Itami, K. Catalytic Dehydrogenative C-H Imidation of Arenes Enabled by Photo-generated Hole Donation to Sulfonimide. Chem. 2017, 2, 383. (c) Obligacion, J. V.; Semproni, S. P.; Chirik, P. J. Cobalt-Catalyzed C−H Borylation. J. Am. Chem. Soc. 2014, 136, 4133. (d) Foo, K.; Sella, E.; Thomé, I.; Eastgate, M. D.; Baran, P. S. A Mild, Ferrocene-Catalyzed C−H Imidication of (Hetero)Arenes. J. Am. Chem. Soc. 2014, 136, 5279. (e) Allen, L. J.; Cabrera, P. J.; Lee, M.; Sanford, M. S. N-Acyloxyphthalimides as Nitrogen Radical Precursors in the Visible Light Photocatalyzed Room Temperature C-H Amination of Arenes and Heteroarenes. J. Am. Chem. Soc. 2014, 136, 5607. (f) Kim, H.; Kim, T.; Lee, D. G.; Roh, S. W.; Lee, C. Nitrogen- Centered Radical-Mediated C−H Imidation of Arenes and Heteroarenes via Visible Light Induced Photocatalysis. Chem. 130

DOI: 10.1021/jacs.8b12063 J. Am. Chem. Soc. 2019, 141, 127−132

Communication

Journal of the American Chemical Society

Lanthanoid Ions. Z. Anorg. Allg. Chem. 2008, 634, 325. (b) Cole, M. L.; Jones, C.; Junk, P. C. Ether and crown ether adduct complexes of sodium and potassium cyclopentadienide and methylcyclopentadienide−molecular structures of [Na(dme)Cp]∞, [K(dme)0.5Cp]∞, [Na(15-crown-5)Cp], [Na(18-crown-6)CpMe] and the “naked Cp−” complex [K(15-crown-5)2][Cp]. J. Chem. Soc., Dalton Trans. 2002, 896. (c) Jones, C.; Thomas, R. C. The synthesis and structural characterisation of a diphosphastibolyl potassium complex, [{[K(DME)][1,4,2-P2SbC2But2]}∞], and a novel hetero-cage compound. J. Organomet. Chem. 2001, 622, 61. (24) (a) Maddock, L. C. H.; Nixon, T.; Kennedy, A. R.; Probert, M. R.; Clegg, W.; Hevia, E. Utilising Sodium-Mediated Ferration for Regioselective Functionalisation of Fluoroarenes via C−H and C−F Bond Activations. Angew. Chem., Int. Ed. 2018, 57, 187. (b) Algera, R. F.; Ma, Y.; Collum, D. B. Sodium Diisopropylamide: Aggregation, Solvation, and Stability. J. Am. Chem. Soc. 2017, 139, 7921. (c) Tai, O.; Hopson, R.; Williard, P. G. Ligand Binding Constants to Lithium Hexamethyldisilazide Determined by Diffusion-Ordered NMR Spectroscopy. J. Org. Chem. 2017, 82, 6223. (d) Baillie, S. E.; Clegg, W.; García-Á lvarez, P.; Hevia, E.; Kennedy, A. R.; Klett, J.; Russo, L. Synthesis, Structural Elucidation, and Diffusion-Ordered NMR Studies of Homoleptic Alkyllithium Magnesiates: Donor-Controlled Structural Variations in Mixed-Metal Chemistry. Organometallics 2012, 31, 5131. (e) Lucht, B. L.; Collum, D. B. Lithium Ion Solvation: Amine and Unsaturated Hydrocarbon Solvates of Lithium Hexamethyldisilazide (LiHMDS). J. Am. Chem. Soc. 1996, 118, 2217. (f) Lucht, B. L.; Collum, D. B. Ethereal Solvation of Lithium Hexamethyldisilazide: Unexpected Relationships of Solvation Number, Solvation Energy, and Aggregation State. J. Am. Chem. Soc. 1995, 117, 9863. (25) The higher solubility of KHMDS in DME vs dioxane does not allow us to completely rule out whether kinetics come into play for determining C2/C4-selectivity. (26) The erosion in regioselectivity pattern observed for electron-rich arenes at C3 (2e, 3e) do not allow us to rule out whether electronic effects come into play. (27) Edelmann, F. T.; Pauer, F.; Wedler, M.; Stalke, D. Preparation and Structural Characterization of Dioxane-Coordinated Alkali Metal Bis(trimethylsilyl)amides. Inorg. Chem. 1992, 31, 4143. (28) Clean formation of (Me3Si)2NBPin was detected by GC of the crude mixtures. This result suggests the activation of TESBPin by KHMDS, likely via the initial formation of discrete [(Me3Si)2N(BPin) SiEt3]K species. (29) The intermediacy of silyl anions was further confirmed by control experiments with cyclohexene, leading to ring-opening silylation in the absence of pyridine. For the test of silyl anions, see: Gilman, H.; Aoki, D.; Wittenberg, D. Some Reactions of Silyllithium Compounds with Epoxides. J. Am. Chem. Soc. 1959, 81, 1107. For details, see ref 13. (30) See for example: (a) Xu, Z.; Chai, L.; Liu, Z. − Q. Free-RadicalPromoted Site-Selective C−H Silylation of Arenes by Using Hydrosilanes. Org. Lett. 2017, 19, 5573. (b) Xu, P.; Würthwein, E. − V.; Daniliuc, C. G.; Studer, A. Transition-metal Free Ring-Opening Silylation of Indoles and Benzofurans with Diphenyl-tert-Butylsilyl Lithium. Angew. Chem., Int. Ed. 2017, 56, 13872. (31) Careful NMR monitoring under nitrogen atmospheres revealed the formation of dearomatized N-BPin silylated azines and the corresponding aromatized congeners. At present, the exact mechanism by which aromatization occurs with KHMDS is not fully understood. (32) For recent defluorinative events of trifluoromethyl groups, see: (a) Wang, P.; Pu, X.; Zhao, Y.; Wang, P.; Li, Z.; Zhu, C.; Shi, Z. Enantioselective Copper-Catalyzed Defluoroalkylation Using Arylboronate-Activated Alkyl Grignard Reagents. J. Am. Chem. Soc. 2018, 140, 9061. (b) Lang, S.; Wiles, R. J.; Kelly, C. B.; Molander, G. A. Photoredox Generation of Carbon- Centered Radicals Enables the Construction of 1,1-Difluoroalkene Carbonyl Mimic. Angew. Chem., Int. Ed. 2017, 56, 15073. (c) Xiao, T.; Li, L.; Zhou, L. Synthesis of Functionalized gem-Difluoroalkenes via a Photocatalytic Decarboxylative/Defluorinative Reaction. J. Org. Chem. 2016, 81, 7908. (d) Wang, H.; Jui, N. T. Catalytic Defluoroalkylation of Trifluoromethylaromatics with Unactivated Alkenes. J. Am. Chem. Soc. 2018,

(12) (a) Somerville, R.; Hale, L.; Gomez-Bengoa, E.; Burés, J.; Martin, R. Intermediacy of Ni-Ni Species in sp2 C−O Bond Cleavage of Aryl Esters: Relevance in Catalytic C−Si Bond Formation. J. Am. Chem. Soc. 2018, 140, 8771. (b) Zarate, C.; Nakajima, M.; Martin, R. A Mild and Ligand-Free Ni-Catalyzed Silylation via C−OMe Cleavage. J. Am. Chem. Soc. 2017, 139, 1191. (c) Zarate, C.; Martin, R. A Mild Ni/CuCatalyzed Silylation via C−O Cleavage. J. Am. Chem. Soc. 2014, 136, 2236. (13) See Supporting Information for details. (14) For selected examples in which the escorting counterion plays a non-negligible role on reactivity, see: (a) Tobisu, M.; Takahira, T.; Morioka, T.; Chatani, N. Nickel-Catalyzed Alkylative Cross-Coupling of Anisoles with Grignard Reagents via C−O Bond Activation. J. Am. Chem. Soc. 2016, 138, 6711. (b) Cornella, J.; Martin, R. Ni-Catalyzed Stereoselective Arylation of Inert C-O bonds at Low Temperatures. Org. Lett. 2013, 15, 6298. (c) see ref 12b. (d) Casado, F.; Pisano, L.; Farriol, M.; Gallardo, I.; Marquet, J.; Melloni, G. J. Org. Chem. 2000, 65, 322. and citations therein. (15) The combination of Et3SiH/KOtBu has shown to promote the silylation of electron-rich heteroarenes via radical chain or pentacoordinated species: (a) Toutov, A. A.; Liu, W.-B.; Betz, K. N.; Fedorov, A.; Stoltz, B. M.; Grubbs, R. H. Silylation of C−H bonds in aromatic heterocycles by an Earth-abundant metal catalyst. Nature 2015, 518, 80. (b) Liu, W.-B.; Schuman, D. P.; Yang, Y.-F.; Toutov, A. A.; Liang, Y.; Klare, H. F. T.; Nesnas, N.; Oestreich, M.; Blackmond, D. G.; Virgil, S. C.; Banerjee, S.; Zare, R. N.; Grubbs, R. H.; Houk, K. N.; Stoltz, B. M. Potassium tert-Butoxide-Catalyzed Dehydrogenative C−H Silylation of Heteroaromatics: A Combined Experimental and Computational Mechanistic Study. J. Am. Chem. Soc. 2017, 139, 6867. (c) Banerjee, S.; Yang, Y.-F.; Jenkins, I. D.; Liang, Y.; Toutov, A. A.; Liu, W.-B.; Schuman, D. P.; Grubbs, R. H.; Stoltz, B. M.; Krenske, E. H.; Houk, K. N.; Zare, R. N. Ionic and Neutral Mechanisms for C−H Bond Silylation of Aromatic Heterocycles Catalyzed by Potassium tert-Butoxide. J. Am. Chem. Soc. 2017, 139, 6880. (16) It is worth noting that not even traces of 2a/3a were observed by exposing electron-poor azines to either KOtBu /Et3SiH (see ref 15) or KHMDS/Et3SiH, indicating that our silylation follows a different mechanistic rationale. (17) The addition of extra KHMDS/Et3SiBPin does not improve yields. The mass balance accounts for unreacted azine. (18) For a mechanistic rationale, see ref 13. (19) Albini, A.; Pietra, S. Heterocyclic N-Oxides; CRC Press: Boca Raton, FL, 1991. (20) Metal-Catalyzed Cross-Coupling Reactions; Diederich, F., de Meijere, A., Eds.; Wiley-VCH: Weinheim, 2004. (21) At present, we do not have a rationale behind the role exerted by the ketone backbone on site-selectivity en route to 7 and 10. (22) For the importance of coordination on the functionalization of azines, see: (a) Nagase, M.; Kuninobu, Y.; Kanai, M. 4-PositionSelective C−H Perfluoroalkylation and Perfluoroarylation of SixMembered Heteroaromatic Compounds. J. Am. Chem. Soc. 2016, 138, 6103. (b) Andou, T.; Saga, Y.; Komai, H.; Matsunaga, S.; Kanai, M. Cobalt-Catalyzed C4-Selective Direct Alylation of Pyridines. Angew. Chem., Int. Ed. 2013, 52, 3213. (c) Bull, J. A.; Mousseau, J. J.; Pelletier, G.; Charette, A. B. Synthesis of Pyridine and Dihydropyridine Derivatives by Regio- and Stereoselective Addition of N-Activated Pyridines. Chem. Rev. 2012, 112, 2642. (d) Tsai, C. − C.; Shih, W. − C.; Fang, C. − H.; Li, C. − Y.; Ong, T. − G.; Yap, G. P. A. Bimetallic Nickel Aluminum Mediated Para-Selective Alkenylation of Pyridine: Direct Observation of η2,η1-Pyridine Ni(0)−Al(III) Intermediates Prior to C−H Bond Activation. J. Am. Chem. Soc. 2010, 132, 11887. (e) Nakao, Y.; Yamada, Y.; Kashihara, N.; Hiyama, T. Selective C−4 Alkylation of Pyridine by Nickel/Lewis Acid Catalysis. J. Am. Chem. Soc. 2010, 132, 13666. (f) See ref 5a. (23) For selected references dealing with the complexation of K+ in DME, see: (a) Binda, P. I.; Delbridge, E. E.; Dugah, D. T.; Skelton, B. W.; White, A. H. Synthesis and Structural Characterization of Some Potassium Complexes of Some Bis(phenolate) Ligands and Some Novel Heterobimetallic Binuclear Arrays Formed with Trivalent 131

DOI: 10.1021/jacs.8b12063 J. Am. Chem. Soc. 2019, 141, 127−132

Communication

Journal of the American Chemical Society 140, 163. (e) Chen, K.; Berg, N.; Gschwind, R.; König, B. Selective Single C(sp3)−F Bond Cleavage in Trifluoromethylarenes: Merging Visible-Light Catalysis with Lewis Acid Activation. J. Am. Chem. Soc. 2017, 139, 18444. (33) (a) von Wolff, N.; Char, J.; Frogneux, X.; Cantat, T. Synthesis of Aromatic Sulfones from SO2 and Organosilanes Under Metal-free Conditions. Angew. Chem., Int. Ed. 2017, 56, 5616. (b) Komiyama, T.; Minami, Y.; Hiyama, T. Aryl(triethyl)silanes for Biaryl and Teraryl Synthesis by Copper(II)-Catalyzed Cross-Coupling Reaction. Angew. Chem., Int. Ed. 2016, 55, 15787. For an ipso-iodination event, see ref 13. (34) It is worth noting that even unactivated benzene can be silylated in moderate yield. See ref 13.

132

DOI: 10.1021/jacs.8b12063 J. Am. Chem. Soc. 2019, 141, 127−132