Abstract

•Access to a wide range of linear alcohols and arylated alkanes•Compatible with aliphatic and electronically deactivated epoxides•No pre-activation required for primary aliphatic alcohols•Hexafluoroisopropanol is key to the reactivity Substitution reactions are a common way to couple two molecular fragments at a sp3-hybridized carbon atom. The pre-activation of one or both coupling partners is typically required, which adds additional chemical steps and generates waste at each stage. Efforts to develop catalytic substitution reactions of arenes that bypass pre-activation, starting from common feedstocks, such as alcohols or epoxides, have been limited to specific structural subclasses. We report the discovery of catalytic conditions for the direct substitution of arenes with numerous classes of alcohols and epoxides that were not previously accessible, allowing for one-step access to branched alcohols and complex arylated products. Furthermore, since the products of epoxide substitution are alcohols, this discovery enables the direct bis-substitution of epoxides with two different arenes in one pot. Alcohols and epoxides are arguably ideal electrophiles for the Friedel-Crafts alkylation, since they are widely available, require no pre-activation, and produce no stoichiometric waste beyond water. However, neither primary aliphatic alcohols nor most classes of terminal epoxides are compatible with existing intermolecular Friedel-Crafts methodologies, and sequential Friedel-Crafts reactions starting from epoxides consequently remain underexplored. Here, we report that these limitations are easily overcome using Brønsted acid catalysis in hexafluoroisopropanol (HFIP) as a solvent. Electron-poor aromatic epoxides and aliphatic epoxides undergo stereospecific arylation to give an alcohol which, depending on the reaction conditions, can partake in a second nucleophilic substitution with a different arene in one pot. Phenyl ethanols react through a phenonium intermediate, whereas simple aliphatic alcohols participate in a rare intermolecular SN2 Friedel-Crafts process, delivering linear products exclusively. This work provides an alternative to metal-catalyzed cross-couplings for accessing important scaffolds, widening the range of applications of the Friedel-Crafts reaction. Alcohols and epoxides are arguably ideal electrophiles for the Friedel-Crafts alkylation, since they are widely available, require no pre-activation, and produce no stoichiometric waste beyond water. However, neither primary aliphatic alcohols nor most classes of terminal epoxides are compatible with existing intermolecular Friedel-Crafts methodologies, and sequential Friedel-Crafts reactions starting from epoxides consequently remain underexplored. Here, we report that these limitations are easily overcome using Brønsted acid catalysis in hexafluoroisopropanol (HFIP) as a solvent. Electron-poor aromatic epoxides and aliphatic epoxides undergo stereospecific arylation to give an alcohol which, depending on the reaction conditions, can partake in a second nucleophilic substitution with a different arene in one pot. Phenyl ethanols react through a phenonium intermediate, whereas simple aliphatic alcohols participate in a rare intermolecular SN2 Friedel-Crafts process, delivering linear products exclusively. This work provides an alternative to metal-catalyzed cross-couplings for accessing important scaffolds, widening the range of applications of the Friedel-Crafts reaction. Epoxides and primary aliphatic alcohols both represent important building blocks in synthetic chemistry,1Weissermel K. Arpe H.-J. Industrial Organic Chemistry.Fourth Edition. Wiley-VCH Verlag GmbH, 2008Google Scholar the former especially serving as a gateway to densely functionalized molecules in medicinal chemistry, crop science, and material science.2Yudin A.K. Aziridines and Epoxides in Organic Synthesis. Wiley-VCH Verlag GmbH, 2006Google Scholar In principle, the intermolecular Friedel-Crafts reaction would represent an ideal way to form C–C bonds between arenes and epoxides or primary aliphatic alcohols,3Rueping M. Nachtsheim B.J. A review of new developments in the Friedel-Crafts alkylation - from green chemistry to asymmetric catalysis.Beilstein J. Org. Chem. 2010; 6: 6https://doi.org/10.3762/bjoc.6.6Google Scholar, 4Kumar R. Van der Eycken E.V. Recent approaches for C−C bond formation via direct dehydrative coupling strategies.Chem. Soc. Rev. 2013; 42: 1121-1146https://doi.org/10.1039/C2CS35397KGoogle Scholar, 5Dryzhakov M. Richmond E. Moran J. Recent advances in direct catalytic dehydrative substitution of alcohols.Synthesis. 2016; 48: 935-959https://doi.org/10.1055/s-0035-1560396Google Scholar, 6Estopiña-Durán S. Taylor J.E. Brønsted acid-catalysed dehydrative substitution reactions of alcohols.Chemistry. 2021; 27: 106-120https://doi.org/10.1002/chem.202002106Google Scholar since it would prevent pre-activation steps of the substrates and produce no stoichiometric waste beyond water, but this type of reactivity remains challenging in both cases:(1)Ring-opening Friedel-Crafts reactions of terminal epoxides, which give branched products, are mainly limited to electron-rich styrene oxides and arenes when Lewis or Brønsted acid promoters are employed (Scheme 1).7Li G.-X. Qu J. Friedel-Crafts alkylation of arenes with epoxides promoted by fluorinated alcohols or water.Chem. Commun. (Camb). 2010; 46: 2653-2655https://doi.org/10.1039/B926684DGoogle Scholar, 8Talukdar R. Synthetically important ring opening reactions by alkoxybenzenes and alkoxynaphthalenes.RSC Adv. 2020; 10: 31363-31376https://doi.org/10.1039/D0RA05111JGoogle Scholar, 9Dover T.L. McDonald F.E. Fluorinated alcohols: powerful promoters for ring-opening reactions of epoxides with carbon nucleophiles.Arkivoc iii. 2021; : 85-114https://doi.org/10.24820/ark.5550190.p011.449Google Scholar Consequently, epoxide arylation strategies based on transition metal catalysis or photocatalysis have been developed but suffer from poor access to branched products,10Nielsen D.K. Doyle A.G. Nickel-catalyzed cross-coupling of styrenyl epoxides with boronic acids.Angew. Chem. Int. Ed. Engl. 2011; 50: 6056-6059https://doi.org/10.1002/anie.201101191Google Scholar, 11Zhao Y. Weix D.J. Nickel-catalyzed regiodivergent opening of epoxides with aryl halides: co-catalysis controls regioselectivity.J. Am. Chem. Soc. 2014; 136: 48-51https://doi.org/10.1021/ja410704dGoogle Scholar, 12Wang Z. Kuninobu Y. Kanai M. Palladium-catalyzed oxirane-opening reaction with arenes via C−H bond activation.J. Am. Chem. Soc. 2015; 137: 6140-6143https://doi.org/10.1021/jacs.5b02435Google Scholar, 13Lu X.-Y. Yang C.-T. Liu J.-H. Zhang Z.-Q. Lu X. Lou X. Xiao B. Fu Y. Cu-catalyzed cross-coupling reactions of epoxides with organoboron compounds.Chem. Commun. (Camb). 2015; 51: 2388-2391https://doi.org/10.1039/C4CC09321FGoogle Scholar, 14Lu X.-Y. Yan L.-Y. Li J.-S. Li J.-M. Zhou H.-p. Jiang R.-C. Liu C.-C. Lu R. Hu R. Base-free Ni-catalyzed Suzuki-type cross-coupling reactions of epoxides with boronic acids.Chem. Commun. 2020; 56: 109-112https://doi.org/10.1039/C9CC08079AGoogle Scholar, 15Parasram M. Shields B.J. Ahmad O. Knauber T. Doyle A.G. Regioselective cross-electrophile coupling of epoxides and (hetero)aryl iodides via Ni/Ti/photoredox catalysis.ACS Catal. 2020; 10: 5821-5827https://doi.org/10.1021/acscatal.0c01199Google Scholar, 16Potrząsaj A. Musiejuk M. Chaładaj W. Giedyk M. Gryko D. Cobalt catalyst determines regioselectivity in ring opening of epoxides with aryl halides.J. Am. Chem. Soc. 2021; 143: 9368-9376https://doi.org/10.1021/jacs.1c00659Google Scholar notably for alkyl epoxides.11Zhao Y. Weix D.J. Nickel-catalyzed regiodivergent opening of epoxides with aryl halides: co-catalysis controls regioselectivity.J. Am. Chem. Soc. 2014; 136: 48-51https://doi.org/10.1021/ja410704dGoogle Scholar For styrene oxides, more branch-selective arylation examples are known,14Lu X.-Y. Yan L.-Y. Li J.-S. Li J.-M. Zhou H.-p. Jiang R.-C. Liu C.-C. Lu R. Hu R. Base-free Ni-catalyzed Suzuki-type cross-coupling reactions of epoxides with boronic acids.Chem. Commun. 2020; 56: 109-112https://doi.org/10.1039/C9CC08079AGoogle Scholar,15Parasram M. Shields B.J. Ahmad O. Knauber T. Doyle A.G. Regioselective cross-electrophile coupling of epoxides and (hetero)aryl iodides via Ni/Ti/photoredox catalysis.ACS Catal. 2020; 10: 5821-5827https://doi.org/10.1021/acscatal.0c01199Google Scholar but those bearing strong electron-withdrawing groups remain inaccessible.(2)Dehydrative Friedel-Crafts reactions of primary aliphatic alcohols are also undeveloped. Owing to their stability, only two low-yielding (<10%) examples of Friedel-Crafts reactions of primary aliphatic alcohols longer than two carbons are known, both of which give complex mixtures of linear and branched products due to rearrangements.17Sieskind O. Albrecht P. Synthesis of alkylbenzenes by Friedel-Crafts reactions catalysed by K10-montmorillonite.Tetrahedron Lett. 1993; 34: 1197-1200https://doi.org/10.1016/S0040-4039(00)77527-7Google Scholar,18Deshmukh A.R.A.S. Gumaste V.K. Bhawal B.M. Alkylation of benzene with long chain (C8-C18) linear primary alcohols over zeolite-Y.Catal. Lett. 2000; 64: 247-250https://doi.org/10.1023/A:1019055507996Google Scholar A few transition-metal-catalyzed methods have also been described, but they require the specific use of phenol derivatives as reaction partners.19Lee D.-H. Kwon K.-H. Yi C.S. Dehydrative C-H alkylation and alkenylation of phenols with alcohols: expedient synthesis for substituted phenols and benzofurans.J. Am. Chem. Soc. 2012; 134: 7325-7328https://doi.org/10.1021/ja302710vGoogle Scholar, 20Frost J.R. Cheong C.B. Donohoe T.J. Iridium-catalyzed C4-alkylation of 2,6-Di-tert-butylphenol by using hydrogen-borrowing catalysis.Synthesis. 2016; 49: 910-916https://doi.org/10.1055/s-0035-1561439Google Scholar, 21Yu J. Li C.-J. Zeng H. Dearomatization-rearomatization strategy for ortho-selective alkylation of phenols with primary alcohols.Angew. Chem. Int. Ed. Engl. 2021; 60: 4043-4048https://doi.org/10.1002/anie.202010845Google Scholar In general, to assemble such products, several strategies based on cross-coupling,22Martin R. Fürstner A. Cross-coupling of alkyl halides with aryl Grignard reagents catalyzed by a low-valent iron complex.Angew. Chem. Int. Ed. Engl. 2004; 43: 3955-3957https://doi.org/10.1002/anie.200460504Google Scholar, 23Nakamura M. Matsuo K. Ito S. Nakamura E. Iron-catalyzed cross-coupling of primary and secondary alkyl halides with aryl Grignard reagents.J. Am. Chem. Soc. 2004; 126: 3686-3687https://doi.org/10.1021/ja049744tGoogle Scholar, 24Hatakeyama T. Fujiwara Y.-i. Okada Y. Itoh T. Hashimoto T. Kawamura S. Ogata K. Takaya H. Nakamura M. Kumada-Tamao-Corriu coupling of alkyl halides catalyzed by an iron-bisphosphine complex.Chem. Lett. 2011; 40: 1030-1032https://doi.org/10.1246/cl.2011.1030Google Scholar, 25Yang C.-T. Zhang Z.-Q. Liu Y.-C. Liu L. Copper-catalyzed cross-coupling reaction of organoboron compounds with primary alkyl halides and pseudohalides.Angew. Chem. Int. Ed. Engl. 2011; 50: 3904-3907https://doi.org/10.1002/anie.201008007Google Scholar, 26Yang C.-T. Zhang Z.-Q. Tajuddin H. Wu C.-C. Liang J. Liu J.-H. Fu Y. Czyzewska M. Steel P.G. Marder T.B. Liu L. Alkylboronic esters from copper-catalyzed borylation of primary and secondary alkyl halides and pseudohalides.Angew. Chem. Int. Ed. Engl. 2012; 51: 528-532https://doi.org/10.1002/anie.201106299Google Scholar, 27Liu J.-H. Yang C.-T. Lu X.-Y. Zhang Z.-Q. Xu L. Cui M. Lu X. Xiao B. Fu Y. Liu L. Copper-catalyzed reductive cross-coupling of nonactivated alkyl tosylates and mesylates with alkyl and aryl bromides.Chemistry. 2014; 20: 15334-15338https://doi.org/10.1002/chem.201405223Google Scholar, 28Kim S. Goldfogel M.J. Gilbert M.M. Weix D.J. Nickel-catalyzed cross-electrophile coupling of aryl chlorides with primary alkyl chlorides.J. Am. Chem. Soc. 2020; 142: 9902-9907https://doi.org/10.1021/jacs.0c02673Google Scholar, 29Charboneau D.J. Barth E.L. Hazari N. Uehling M.R. Zultanski S.L. A widely applicable dual catalytic system for cross-electrophile coupling enabled by mechanistic studies.ACS Catal. 2020; 10: 12642-12656https://doi.org/10.1021/acscatal.0c03237Google Scholar, 30Li Z. Sun W. Wang X. Li L. Zhang Y. Li C. Electrochemically enabled, nickel-catalyzed dehydroxylative cross-coupling of alcohols with aryl halides.J. Am. Chem. Soc. 2021; 143: 3536-3543https://doi.org/10.1021/jacs.0c13093Google Scholar, 31Dong Z. MacMillan D.W.C. Metallaphotoredox-enabled deoxygenative arylation of alcohols.Nature. 2021; 598: 451-456https://doi.org/10.1038/s41586-021-03920-6Google Scholar hydrodefluorination32Zhu J. Pérez M. Caputo C.B. Stephan D.W. Use of trifluoromethyl groups for catalytic benzylation and alkylation with subsequent hydrodefluorination.Angew. Chem. Int. Ed. Engl. 2016; 55: 1417-1421https://doi.org/10.1002/anie.201510494Google Scholar or C–H functionalization33Shi Z. He C. Direct functionalization of arenes by primary alcohol sulfonate esters catalyzed by gold(III).J. Am. Chem. Soc. 2004; 126: 13596-13597https://doi.org/10.1021/ja046890qGoogle Scholar,34Wilson A.S.S. Hill M.S. Mahon M.F. Dinoi C. Maron L. Organocalcium-mediated nucleophilic alkylation of benzene.Science. 2017; 358: 1168-1171https://doi.org/10.1126/science.aao5923Google Scholar have been devised, but they all require additional synthetic steps to pre-activate the alcohol and, in most cases, the arene coupling partner as well. The development of a Friedel-Crafts reaction broadly applicable to both terminal epoxides and primary aliphatic alcohols would have an additional benefit: since the products of Friedel-Crafts ring-opening of epoxides are themselves primary aliphatic alcohols, sequential Friedel-Crafts reactions could then be envisaged where two distinct arenes can be installed in one pot to provide a straightforward access to 1,1,2-triarylethane frameworks, which have applications ranging from the life sciences to feedstock precursors. Existing strategies to access these compounds largely start from alkenes as precursors and ensure positional control of the C–C bond forming reaction. However, they require either multi-step preparation of different cross-coupling partners or substrates, the use of directing groups, or complex reaction conditions under inert atmosphere.35Tolstoy P. Engman M. Paptchikhine A. Bergquist J. Church T.L. Leung A.W.-M. Andersson P.G. Iridium-catalyzed asymmetric hydrogenation yielding chiral diarylmethines with weakly coordinating or noncoordinating substituents.J. Am. Chem. Soc. 2009; 131: 8855-8860https://doi.org/10.1021/ja9013375Google Scholar, 36Crudden C.M. Ziebenhaus C. Rygus J.P.G. Ghozati K. Unsworth P.J. Nambo M. Voth S. Hutchinson M. Laberge V.S. Maekawa Y. Imao D. Iterative protecting group-free cross-coupling leading to chiral multiply arylated structures.Nat. Commun. 2016; 7: 11065https://doi.org/10.1038/ncomms11065Google Scholar, 37Guo J. Huang G.-B. Wu Q.-L. Xie Y. Weng J. Lu G. An efficient approach to access 1,1,2-triarylethanes enabled by the organo-photoredox-catalyzed decarboxylative addition reaction.Org. Chem. Front. 2019; 6: 1955-1960https://doi.org/10.1039/C9QO00380KGoogle Scholar, 38Urkalan K.B. Sigman M.S. Palladium-catalyzed oxidative intermolecular difunctionalization of terminal alkenes with organostannanes and molecular oxygen.Angew. Chem. Int. Ed. Engl. 2009; 48: 3146-3149https://doi.org/10.1002/anie.200900218Google Scholar, 39Kuang Z. Yang K. Song Q. Pd-catalyzed 1,2-diarylation of vinylarenes at ambient temperature.Org. Chem. Front. 2017; 4: 1224-1228https://doi.org/10.1039/C7QO00097AGoogle Scholar, 40Shrestha B. Basnet P. Dhungana R.K. Kc S. Thapa S. Sears J.M. Giri R. Ni-catalyzed regioselective 1,2-dicarbofunctionalization of olefins by intercepting Heck intermediates as imine-stabilized transient metallacycles.J. Am. Chem. Soc. 2017; 139: 10653-10656https://doi.org/10.1021/jacs.7b06340Google Scholar, 41Chen B. Cao P. Yin X. Liao Y. Jiang L. Ye J. Wang M. Liao J. Modular synthesis of enantioenriched 1,1,2-triarylethanes by an enantioselective arylboration and cross-coupling sequence.ACS Catal. 2017; 7: 2425-2429https://doi.org/10.1021/acscatal.7b00300Google Scholar, 42Gao P. Chen L.-A. Brown M.K. Nickel-catalyzed stereoselective diarylation of alkenylarenes.J. Am. Chem. Soc. 2018; 140: 10653-10657https://doi.org/10.1021/jacs.8b05680Google Scholar, 43Basnet P. Kc S. Dhungana R.K. Shrestha B. Boyle T.J. Giri R. Synergistic bimetallic Ni/Ag and Ni/Cu catalysis for regioselective γ,δ-diarylation of alkenyl ketimines: addressing β-H Elimination by in situ Generation of Cationic Ni(II) Catalysts.J. Am. Chem. Soc. 2018; 140: 15586-15590https://doi.org/10.1021/jacs.8b09401Google Scholar, 44Ouyang X.-H. Cheng J. Li J.-H. 1,2-Diarylation of alkenes with aryldiazonium salts and arenes enabled by visible light photoredox catalysis.Chem. Commun. (Camb). 2018; 54: 8745-8748https://doi.org/10.1039/C8CC04526GGoogle Scholar, 45Derosa J. Kleinmans R. Tran V.T. Karunananda M.K. Wisniewski S.R. Eastgate M.D. Engle K.M. Nickel-catalyzed 1,2-diarylation of simple alkenyl amides.J. Am. Chem. Soc. 2018; 140: 17878-17883https://doi.org/10.1021/jacs.8b11942Google Scholar, 46Anthony D. Lin Q. Baudet J. Diao T. Nickel-catalyzed asymmetric reductive diarylation of vinylarenes.Angew. Chem. Int. Ed. Engl. 2019; 58: 3198-3202https://doi.org/10.1002/anie.201900228Google Scholar, 47Derosa J. Kang T. Tran V.T. Wisniewski S.R. Karunananda M.K. Jankins T.C. Xu K.L. Engle K.M. Nickel-catalyzed 1,2-diarylation of alkenyl carboxylates: A gateway to 1,2,3-trifunctionalized building blocks.Angew. Chem. Int. Ed. Engl. 2020; 59: 1201-1205https://doi.org/10.1002/anie.201913062Google Scholar, 48Apolinar O. Tran V.T. Kim N. Schmidt M.A. Derosa J. Engle K.M. Sulfonamide directivity enables Ni-catalyzed 1,2-diarylation of diverse alkenyl amines.ACS Catal. 2020; 10: 14234-14239https://doi.org/10.1021/acscatal.0c03857Google Scholar, 49Qin J.-H. Luo M.-J. An D.-L. Li J.-H. Electrochemical 1,2-diarylation of alkenes enabled by direct dual C-H functionalizations of electron-rich aromatic hydrocarbons.Angew. Chem. Int. Ed. Engl. 2021; 60: 1861-1868https://doi.org/10.1002/anie.202011657Google Scholar Therefore, the direct and straightforward Friedel-Crafts strategy presented here offers a complementary alternative to current methods. Herein, we describe the expansion of the Friedel-Crafts reaction to include most classes of terminal epoxides (ring-opening arylation), primary aliphatic alcohols (dehydroarylation), and a sequential dehydrodiarylation process stemming from their combination. The key to the reactivity is the use of the solvent hexafluoroisopropanol (HFIP)50Colomer I. Chamberlain A.E.R. Haughey M.B. Donohoe T.J. Hexafluoroisopropanol as a highly versatile solvent.Nat. Rev. Chem. 2017; 1: 0088https://doi.org/10.1038/s41570-017-0088Google Scholar, 51Sinha S.K. Bhattacharya T. Maiti D. Role of hexafluoroisopropanol in C−H activation.React. Chem. Eng. 2018; 4: 244-253https://doi.org/10.1039/C8RE00225HGoogle Scholar, 52Yu C. Sanjosé-Orduna J. Patureau F.W. Pérez-Temprano M.H. Emerging unconventional organic solvents for C−H bond and related functionalization reactions.Chem. Soc. Rev. 2020; 49: 1643-1652https://doi.org/10.1039/C8CS00883CGoogle Scholar, 53Pozhydaiev V. Power M. Gandon V. Moran J. Lebœuf D. Exploiting hexafluoroisopropanol (HFIP) in Lewis and Brønsted acid-catalyzed reactions.Chem. Commun. (Camb.). 2020; 56: 11548-11564https://doi.org/10.1039/D0CC05194BGoogle Scholar, 54Bhattacharya T. Ghosh A. Maiti D. Hexafluoroisopropanol: the magical solvent for Pd-catalyzed C-H activation.Chem. Sci. 2021; 12: 3857-3870https://doi.org/10.1039/D0SC06937JGoogle Scholar with triflic acid (TfOH) as a catalyst. Indeed, the association of Lewis or Brønsted acid catalysts with HFIP is known to trigger transformations with rather unreactive alcohols, alkenes, and cyclopropanes through the formation of highly reactive hydrogen-bond networks.55Vuković V.D. Richmond E. Wolf E. Moran J. Catalytic Friedel-Crafts reactions of highly electronically deactivated benzylic alcohols.Angew. Chem. Int. Ed. Engl. 2017; 56: 3085-3089https://doi.org/10.1002/anie.201612573Google Scholar, 56Qi C. Hasenmaile F. Gandon V. Lebœuf D. Calcium(II)-catalyzed intra- and intermolecular hydroamidation of unactivated alkenes in hexafluoroisopropanol.ACS Catal. 2018; 8: 1734-1739https://doi.org/10.1021/acscatal.7b04271Google Scholar, 57Richmond E. Yi J. Vuković V.D. Sajadi F. Rowley C.N. Moran J. Ring-opening hydroarylation of monosubstituted cyclopropanes enabled by hexafluoroisopropanol.Chem. Sci. 2018; 9: 6411-6416https://doi.org/10.1039/C8SC02126KGoogle Scholar, 58Qi C. Gandon V. Lebœuf D. Calcium(II)-catalyzed intermolecular hydroarylation of deactivated styrenes in hexafluoroisopropanol.Angew. Chem. Int. Ed. Engl. 2018; 57: 14245-14249https://doi.org/10.1002/anie.201809470Google Scholar, 59Wang S. Force G. Guillot R. Carpentier J.-F. Sarazin Y. Bour C. Gandon V. Lebœuf D. Lewis acid/hexafluoroisopropanol: A promoter system for selective Ortho-C-alkylation of anilines with deactivated styrene derivatives and unactivated alkenes.ACS Catal. 2020; 10: 10794-10802https://doi.org/10.1021/acscatal.0c02959Google Scholar, 60Qi C. Force G. Gandon V. Lebœuf D. Hexafluoroisopropanol-promoted haloamidation and halolactonization of unactivated alkenes.Angew. Chem. Int. Ed. Engl. 2021; 60: 946-953https://doi.org/10.1002/anie.202010846Google Scholar Under these reaction conditions, the ring-opening Friedel-Crafts reaction of most epoxides is found to be stereospecific, and the resulting alcohols can react in a second Friedel-Crafts reaction through the intermediacy of a well-established61Cram D.J. Studies in stereochemistry. I. The Stereospecific Wagner--Meerwein rearrangement of the isomers of 3-phenyl-2-butanol.J. Am. Chem. Soc. 1949; 71: 3863-3870https://doi.org/10.1021/ja01180a001Google Scholar, 62Cram D.J. Davis R. Studies in stereochemistry. II. The preparation and complete resolution of 3-phenyl-2-pentanol and 2-phenyl-3-pentanol.J. Am. Chem. Soc. 1949; 71: 3871-3875https://doi.org/10.1021/ja01180a002Google Scholar, 63Cram D.J. Studies in stereochemistry. III. The Wagner--Meerwein rearrangement in the 2-phenyl-3-pentanol and 3-phenyl-2-pentanol systems.J. Am. Chem. Soc. 1949; 71: 3875-3883https://doi.org/10.1021/ja01180a003Google Scholar, 64Cram D.J. Studies in stereochemistry. V. Phenonium sulfonate ion-pairs as intermediates in the intramolecular rearrangements and solvolysis reactions that occur in the 3-phenyl-2-butanol system.J. Am. Chem. Soc. 1952; 74: 2129-2137https://doi.org/10.1021/ja01129a001Google Scholar, 65Olah G.A. Porter R.D. Stable carbocations. CXXI. Carbon-13 magnetic resonance spectroscopy study of ethylenarenium ions (spiro[2.5]octadienyl cations).J. Am. Chem. Soc. 1971; 93: 6877-6887https://doi.org/10.1021/ja00754a031Google Scholar, 66Olah G.A. Head N.J. Rasul G. Prakash G.K.S. Protonation of benzocyclobutene with superacid: Cram’s phenonium ion (spiro[5.2]octa-5,7-dien-4-yl cation) revisited.J. Am. Chem. Soc. 1995; 117: 875-882https://doi.org/10.1021/ja00108a003Google Scholar, 67del Río E. Menéndez M.I. López R. Sordo T.L. On the structure of phenonium ions: the important role of back-bonding interaction in carbocation chemistry.J. Phys. Chem. A. 2000; 104: 5568-5571https://doi.org/10.1021/jp994068gGoogle Scholar, 68del Río E. Menéndez M.I. López R. Sordo T.L. Rearrangements involving the phenonium ion: A theoretical investigation.J. Am. Chem. Soc. 2001; 123: 5064-5068https://doi.org/10.1021/ja0039132Google Scholar, 69Tsuji Y. Richard J.P. Formation and mechanism for reactions of ring-substituted phenonium ions in aqueous solution.J. Phys. Org. Chem. 2016; 29: 557-564https://doi.org/10.1002/poc.3510Google Scholar but underexploited70Boye A.C. Meyer D. Ingison C.K. French A.N. Wirth T. Novel lactonization with phenonium ion participation induced by hypervalent iodine reagents.Org. Lett. 2003; 5: 2157-2159https://doi.org/10.1021/ol034616fGoogle Scholar, 71Ho T.-L. Chein R.-J. Intervention of phenonium ion in Ritter reactions.J. Org. Chem. 2004; 69: 591-592https://doi.org/10.1021/jo035561yGoogle Scholar, 72Lebœuf D. Huang J. Gandon V. Frontier A.J. Using Nazarov electrocyclization to stage chemoselective [1,2]-migrations: stereoselective synthesis of functionalized cyclopentenones.Angew. Chem. Int. Ed. Engl. 2011; 50: 10981-10985https://doi.org/10.1002/anie.201104870Google Scholar, 73Protti S. Dondi D. Mella M. Fagnoni M. Albini A. Looking for a paradigm for the reactivity of phenonium ions.Eur. J. Org. Chem. 2011; 2011: 3229-3237https://doi.org/10.1002/ejoc.201100305Google Scholar, 74Lebœuf D. Gandon V. Ciesielski J. Frontier A.J. Experimental and theoretical studies on the Nazarov cyclization/Wagner-Meerwein rearrangement sequence.J. Am. Chem. Soc. 2012; 134: 6296-6308https://doi.org/10.1021/ja211970pGoogle Scholar, 75Banik S.M. Medley J.W. Jacobsen E.N. Catalytic, asymmetric difluorination of alkenes to generate difluoromethylated stereocenters.Science. 2016; 353: 51-54https://doi.org/10.1126/science.aaf8078Google Scholar, 76Scheidt F. Neufeld J. Schäfer M. Thiehoff C. Gilmour R. Catalytic geminal difluorination of styrenes for the construction of fluorine-rich bioisosteres.Org. Lett. 2018; 20: 8073-8076https://doi.org/10.1021/acs.orglett.8b03794Google Scholar, 77Li J. Bauer A. Di Mauro G. Maulide N. α-arylation of carbonyl compounds through oxidative C−C bond activation.Angew. Chem. Int. Ed. Engl. 2019; 58: 9816-9819https://doi.org/10.1002/anie.201904899Google Scholar, 78Xu S. Holst H.M. McGuire S.B. Race N.J. Reagent control enables selective and regiodivergent opening of unsymmetrical phenonium ions.J. Am. Chem. Soc. 2020; 142: 8090-8096https://doi.org/10.1021/jacs.0c02095Google Scholar phenonium ion without pre-activation of the alcohol. Simple aliphatic alcohols can also undergo a Friedel-Crafts reaction through a rare intermolecular SN2-type mechanism, as supported by density functional theory (DFT) calculations. This broad-reaching (>160 examples) expansion of the Friedel-Crafts reaction, thus, allows the user to re-evaluate the reactivity of primary aliphatic alcohols, epoxides, and related compounds toward new synthetic applications. We commenced our investigations by studying the monoarylation of highly electron-deficient styrene oxides, which are notoriously challenging to functionalize. Following a large survey of reaction conditions, notably Lewis and Brønsted acid catalysts (see Table S1), we observed that the reaction between (pentafluorophenyl)ethylene oxide and m-xylene (5 equiv) provided the target product 3 in 97% yield at room temperature in 6 h by conducting the reaction in HFIP (0.4 M) in the presence of TfOH (5 mol %) as a catalyst (Scheme 2). A similar reactivity was obtained by using the promoter system Bi(OTf)3/nBu4NPF679Qin H. Yamagiwa N. Matsunaga S. Shibasaki M. Bismuth-catalyzed intermolecular hydroamination of 1,3-dienes with carbamates, sulfonamides, and carboxamides.J. Am. Chem. Soc. 2006; 128: 1611-1614https://doi.org/10.1021/ja056112dGoogle Scholar instead of TfOH under otherwise identical reaction conditions (96% yield). The reaction could also be performed at higher concentration (1 M), albeit with lower selectivity. More common solvents (dichloromethane, 1,2-dichloroethane, toluene, and nitromethane) led to a significant decrease in yield (<55%) because of the oligomerization of the substrate, highlighting the critical role of HFIP in the transformation. Of note, reducing the number of equivalents of m-xylene caused the ring-opening of the epoxide by HFIP to become a competitive side reaction. a[1a] = 0.2 M. bRegioisomers were separated by flash column chromatography. cBi(OTf)3/nBu4NPF6 (5 mol %) as the catalyst. dUsing TfOH (0.1 mol %). eDiarylated product detected but not isolated (22% yield determined by 1H NMR using hexamethyldisiloxane as an external standard). fDiarylated product 107 isolated in 34% yield. gDiarylated product 120 isolated in 13% yield. h[1x] = 0.2 M.iDiarylated product 121 isolated in 8% yield. Mes = 1,3,5-trimethylphenyl. TMP = 1,3,5-trimethoxyphenyl. With optimized conditions in hand, we first explored the scope of (pentafluorophenyl)ethanol synthesis from styrene oxide 1a using an array of aryl and heteroaryl nucleophiles. The transformation was compatible with a wide range of mono- to tetrasubstituted arenes, incorporating either electron-donating or electron-withdrawing substituents to afford the corresponding products 3–29 in 42%–97% yields. The steric hindrance exhibited by the various functional groups on the nucleophile did not hamper the reactivity, as nearly quantitative yields was achieved in most cases (up to 96%). Although the reaction with 1,3,5-triethylbenzene as a nucleophile produced a mixture of mono- and diarylated products 8 and 84 in a 1:1.25 ratio, conducting the reaction at 0°C enabled the selective formation of 8 in 96% yield. The same applied to mesitylene (9, 95%). Moreover, the reaction could be extended to less electron-rich nucleophiles, such as benzene (18), fluorobenzene (19), and bromobenzene (20), providing arylated compounds in 53%–84% yields. On the other hand, 1,4-difluorobenzene (21) was not sufficiently reactive due to its reduced nucleophilicity. In this case, the epoxide underwent oligomerization and nucleophilic ring-opening by HFIP. Lower reaction concentrations (0.2 M) improved the yield up to 90% in the case of benzene adduct 18, in agreement with previous studies that identified the key role of H-bonded solvent clusters in Lewis and Brønsted acid-catalyzed reactions in HFIP.53Pozhydaiev V. Power M. Gandon V. Moran J. Lebœuf D. Exploiting hexafluoroisopropa

Full Text
Published version (Free)

Talk to us

Join us for a 30 min session where you can share your feedback and ask us any queries you have

Schedule a call