Sunteți pe pagina 1din 6

Research Article

Cite This: ACS Catal. 2018, 8, 6401−6406 pubs.acs.org/acscatalysis

Visible Light Mediated, Redox Neutral Remote 1,6-


Difunctionalizations of Alkenes
Wei Shu,† Estíbaliz Merino,† and Cristina Nevado*,†

Department of Chemistry, University of Zurich, Winterthurerstrasse 190, CH-8057 Zurich, Switzerland
*
S Supporting Information

ABSTRACT: A photoinduced cascade strategy is presented


here for the remote functionalization of alkenes under redox
neutral conditions. A broad portfolio of alkyl groups has been
added to double bonds to produce, upon 1,5-HAT, remote C-
Downloaded from pubs.acs.org by UNIV OF ZURICH on 07/18/18. For personal use only.

centered radicals which can be harvested in the presence of O-


or C-nucleophiles to efficiently form Csp3−O and Csp3−Csp2
bonds at room temperature.
KEYWORDS: photocatalysis, redox neutral, remote functionalization, alkene, radical, oxy-alkylation, carbo-alkylation

1. INTRODUCTION this context resulted in successful 1,6-oxy- and 1,6-azidotri-


Alkene difunctionalizations play a prominent role in the fluoromethylations of alkenes.11 These protocols utilize
ACS Catal. 2018.8:6401-6406.

efficient assembly of molecular complexity.1 A large degree of stoichiometric amounts of hypervalent iodine reagents12 at
diversity in terms of the functional groups added across double elevated temperature to trigger the addition of the trifluor-
bonds has been achieved by means of oxidative conditions, omethyl group onto the alkene. The remote functionalization
including, but not limited to oxy-arylation,2 carbo-alkylation,3 of the Csp3−H bonds requires a vicinal heteroatom-containing
amino-alkylation,4 and oxy-alkylation5,6 reactions (Scheme 1A). functionality (such as O, N, CO) in combination with
MeOH and TMSN3 as nucleophiles (Scheme 1B, left).11a,c,d
Scheme 1. Strategies for the Difunctionalization of Alkenes Photoredox catalysis has emerged as a powerful tool to
enable single electron transfer (SET) reactions under mild
conditions by taking advantage of photoexcited catalysts.6,7e−j,13
Thus, as part of our ongoing interest in the remote
functionalization of Csp3−H bonds,14 we set out to develop a
visible-light mediated, redox neutral 1,6-difunctionalization of
unactivated alkenes. The mild reaction conditions enabled us to
expand the portfolio of C-based functionalities that can be
added across the π-system. Further, the functionalization of the
remote Csp3−H bonds could be performed with both O- as
well as, for the first time, C-based nucleophiles, thus
demonstrating the complementarity and extended functional
group compatibility of this light-mediated protocol.

2. RESULTS AND DISCUSSION


The reaction of diethyl 2-allyl-2-(4-methoxyphenethyl)-
1,n-Hydrogen atom transfer (HAT) processes provide an malonate in the presence of ethyl 2-bromo-2,2-difluoroacetate
exceptional opportunity for the remote functionalization of was selected to find the optimal conditions. After initial
inert Csp3−H bonds.7 However, examples of 1,n-difunctional- screening of different photocatalysts to reduce the C−Br
izations of alkenes as a result of the combination of olefin bond,15 we found that the desired 1,6-acetoxy alkylation
functionalization and 1,n-HAT are still scarce (Scheme 1B).8 product 1 could be obtained in 59% yield using 2 mol % of
Multiple challenges hamper the successful development of Ir(ppy)3 upon irradiation with blue LED (34 W) in the
these transformations including the promiscuous reactivity of presence of AgOAc in acetonitrile at room temperature (Table
different intermediates (M1/M1′, M2/M2′) as well as their 1, entries 1−3). Different solvents were tested next: while
propensity to undergo competitive side reactions. Thus, toluene or DMF decreased the reaction efficiency (Table 1,
intermediates M1 and M2 might undergo radical atom-transfer
processes to deliver hydroalkylation or haloalkylation prod- Received: February 20, 2018
ucts,9 whereas carbocationic intermediates M1′ and M2′ tend Revised: May 31, 2018
to undergo elimination reactions.10 Elegant work by Liu et al. in Published: June 14, 2018

© 2018 American Chemical Society 6401 DOI: 10.1021/acscatal.8b00707


ACS Catal. 2018, 8, 6401−6406
ACS Catalysis Research Article

Table 1. Optimization of the Reaction Conditionsa

entry catalyst additive, solvent 1 (yield, %)


1 Ir(ppy)3 AgOAc, CH3CN 36
2 [Ir{dF(CF3)ppy}2(dtbpy)]PF6 AgOAc, CH3CN 0
3 Ru(bpy)3Cl·6H2O AgOAc, CH3CN 0
4 Ir(ppy)3 AgOAc, toluene 31
5 Ir(ppy)3 AgOAc, DMF 33
6 Ir(ppy)3 AgOAc, 1,2-DCE 66
7 Ir(ppy)3 AgOAc, CH2Cl2 71
8 Ir(ppy)3 K/LiOAc, CH2Cl2 0
9 Ir(ppy)3 TBAOAc, CH2Cl2 31
10b Ir(ppy)3 AgOAc, CH2Cl2 71 (67)
11 No [Ir] or no light AgOAc, CH2Cl2 0
a
Alkene (0.1 mmol), alkyl bromide (2 equiv), MOAc (2 equiv) in solvent (0.1 M), 34 W blue LED. The yield was determined by crude 1H NMR
using mesitylene as internal standard. In brackets: isolated yield after column chromatography. bAgOAc (1.5 equiv).

entries 4 and 5), chlorinated solvents had a positive effect on O-donor. 2-Bromo-2,2-difluoro-1-(piperidin-1-yl)ethan-1-one
the reaction outcome (Table 1, entries 6 and 7). Different delivered 5 in 84% yield. Less activated bromides also proved
additives were also investigated (Table 1, entries 8−9). The to be compatible with the reaction protocol. Thus, 2-bromo-1-
best yield in 1 was obtained when 1.5 equiv of AgOAc were (piperidin-1-yl)ethan-1-one delivered 6 in 56% yield, whereas
used (Table 1, entry 10). In contrast, in the absence of either Ir- ethyl 2-bromo-2-fluoroacetate furnished 7 in 72% as a 1:1
catalyst or light, 1 was not observed (Table 1, entry 11) thus mixture of diastereoisomers. Simple alkyl and aryl 2-
confirming the important role of both parameters (metal and bromoacetates efficiently furnished the desired remote
light) for a successful outcome in these transformations. difunctionalization products 8-11 in good yields, albeit in
Using diethyl 2-allyl-2-(4-methoxyphenethyl)malonate and some cases, longer reaction times were required to obtain
ethyl 2-bromo-2,2-difluoroacetate, different oxygen-based nu- complete conversion of the starting material. Interestingly, the
cleophiles were explored (Scheme 2). AgOBz delivered the reaction could be scaled up to 1 mmol without erosion of the
corresponding 1,6-oxocarbofunctionalization product 2 in 65% isolated yield. The structure of the remote 1,6-difunctionaliza-
yield. Interestingly, the reaction could be carried out in the tion products was confirmed by X-ray diffraction analysis of 9’
presence of Ag2CO3 and 2 equiv of propanoic or p-toluic acid, stemming from the basic hydrolysis of compound 9.16 α-Bromo
to deliver the corresponding esters in comparable yields butyrolactone furnished the corresponding lactone derivative
(compounds 3 and 4). The nature of the alkyl halide was 12 in 58% as a 1:1 mixture of diastereoisomers. α-Bromo-
explored next under the optimized conditions with AgOAc as aromatic as well as -alkyl ketones could be efficiently employed
in these transformations, delivering the remote functionalized
Scheme 2. Reaction Scope on Alkyl Halide and O-Based ketones 13−16 in excellent yields with complete regioselectiv-
Reagent for 1,6-Oxy-alkylation ity. Finally, we also demonstrated that perfluoroalkyl-containing
alkyl bromides and iodides could be efficiently engaged in the
reaction as shown by the efficient formation of compounds 17−
19 under the standard reaction conditions. These examples
showcase the high functional group compatibility of this redox
neutral remote difunctionalization as esters, amides, lactones,
ketones, C−Cl bonds were well tolerated under the optimized
protocols. For a summary of substrates that did not work, see
section S3.3 in the Supporting Information, SI.
Next, we decided to interrogate the flexibility of the reaction
in terms of the olefinic partner (Scheme 3). Alkenes bearing
electronically diverse groups on the remote aromatic ring
delivered the desired oxocarbofunctionalization products 20−
28 in excellent yields. In the presence of methyl 2-
bromoacetate, 1,6-oxyalkylation products 29 and 30 could be
obtained in good yields. Interestingly, halogen substituents in
the aromatic ring both in ortho, meta, and para position proved
to be well tolerated as demonstrated by the efficient reaction to
produce the corresponding remote oxocarbofunctionalization
products in synthetically useful yields (26−28, 31−33).
a
See Table 1, entry 10 for detailed conditions. bAg2CO3 (0.75 equiv) Substrates with different substituents on the aliphatic chain
was used. c36 h. dOne mmol scale. e48 h. fPerfluorohexyl iodide was were also accommodated in the reaction as demonstrated by
used. the isolation of 34 and 35 in 58% and 54% yield, respectively.
6402 DOI: 10.1021/acscatal.8b00707
ACS Catal. 2018, 8, 6401−6406
ACS Catalysis Research Article

Scheme 3. Reaction Scope on Alkene ring of the 6-aryl-1-hexene partner could be tolerated (36−38).
Both 2-bromo-2,2-difluoroketones and cyclohexyl amides could
also be successfully employed as alkyl partners in the
dicarbofunctionalization reaction (39−40). This 1,6-carbo-
alkylation process is also compatible with less activated alkyl
radical precursors (41−42) as well as with different substitution
patterns on the alkyl chain of the olefinic partner (43).
Different indoles were also tested yielding the desired reaction
products (44−47) in synthetically useful yields under the
standard reaction conditions with EtCN as solvent.
Mechanistic studies were also performed (Scheme 5).
Reactions in the presence of well-established radical traps

Scheme 5. Deuterium Labelling Experiments, DFT


Calculations and Proposed Reaction Mechanisma

a
See Table 1, entry 10 for detailed conditions.

Fully linear or heteroatom bridged alkyl chains did not furnish


the desired products, thus signalizing the importance of
Thorpe-Ingold effects in the reaction (see section S3.3 in the
SI).15,17
As discussed earlier, one of the major drawbacks of existing
alkene remote difunctionalizations stems from the use of
oxidative conditions, incompatible with C-based nucleophiles.11
Given the mild, redox neutral conditions of our protocol, we set
out to explore the incorporation of C-based nucleophiles in the
reaction (Scheme 4).
In the presence of 2-bromo-2,2-difluoroacetate and trime-
thoxybenzene, different substitution patterns on the aromatic

Scheme 4. Extension to 1,6-Carbo-alkylation of Alkene

a
Intermediates and transition states computed at B3LYP/6-311+
+G(d,p) (iefpcm, solvent = dichloromethane) level. Energies are given
in kcal/mol (relative to the sum of the starting materials, GI + Galkene =
0 kcal/mol).

showed no desired 1,6-difunctionalization products whereas the


alkene could be completely recovered, which is indicative of
radical intermediates along the reaction pathway (see SI).15
Deuterium labeling studies involving 48 and d2-48 yielded KIE
values of 1.2 and 1.0 for parallel (two-flask) and competitive
(one-pot) intermolecular experiments, respectively (Scheme
5a.1). In contrast, a KIE = 3.3 could be determined for the
intramolecular competition experiment with substrate d1-48
(Scheme 5a.2).18 DFT calculations15 revealed a slightly higher
energy barrier for the addition of the C-centered radical I to the
a
Alkene (0.1 mmol), alkyl bromide (0.3 mmol), arene (0.3 mmol), in alkene (TSI−II, ΔG‡ = 16.1 kcal/mol) compared to that of the
the presence of 3 mol % of Ir(ppy)3 and Ag2CO3 (0.075 mmol) in 1,5-HAT taking place on intermediate II (TSII−III, ΔG‡ = 11.8
DCM (0.1 M) at 25 °C. bEtCN was used as solvent. kcal/mol). Interestingly, the latter process is exergonic (ΔG =
6403 DOI: 10.1021/acscatal.8b00707
ACS Catal. 2018, 8, 6401−6406
ACS Catalysis Research Article

−11.7 kcal/mol) as a result of the stabilized benzylic radical 9′. The authors would like to acknowledge the support by the
produced in III.19 COST Action CHAOS (CA15106).
In line with these results, our mechanistic proposal involves
the activation of the Ir(III)-photocatalyst upon irradiation
which in turn, reduces the C−Br bond delivering an alkyl
■ REFERENCES
(1) For reviews on alkene difunctionalizations, see: (a) Romero, R.
radical I under redox neutral conditions. Subsequent addition M.; Woeste, T. H.; Muñiz, K. Vicinal Difunctionalization of Alkenes
of I onto the double bond produces C-centered radical II. A with Iodine(III) Reagents and Catalysts. Chem.−Asian J. 2014, 9,
1,5-Hydrogen atom transfer, presumably aided by the 972−983. (b) Courant, T.; Masson, G. Recent Progress in Visible-
preorganization of the starting material,17 takes place to Light Photoredox-Catalyzed Intermolecular 1,2-Difunctionalization of
produce a more stable C-centered benzylic radical III in a Double Bonds via an ATRA-Type Mechanism. J. Org. Chem. 2016, 81,
6945−6952. (c) Koike, T.; Akita, M. A Versatile Strategy for
remote and site-selective manner. Oxidation of the radical to Difunctionalization of Carbon−Carbon Multiple Bonds by Photo-
the corresponding carbocation takes place in the presence of an redox Catalysis. Org. Chem. Front. 2016, 3, 1345−1349. (d) Yin, G.;
Ir(IV) complex intermediate to produce IV, which is thus Mu, X.; Liu, G. Palladium(II)-Catalyzed Oxidative Difunctionalization
subsequently quenched in the presence of either an O- or a C- of Alkenes: Bond Forming at a High-Valent Palladium Center. Acc.
based nucleophile to produce the remote dicarbo- or Chem. Res. 2016, 49, 2413−2423. (e) Yan, M.; Lo, J. C.; Edwards, J.
oxocarbofunctionalization products observed in these trans- T.; Baran, P. S. Radicals: Reactive Intermediates with Translational
formations. The results of the deuterium labeling experiments, Potential. J. Am. Chem. Soc. 2016, 138, 12692−12714.
consistent with this proposal, point toward the Csp3−H bond (2) For selected examples, see: (a) Hartmann, M.; Li, Y.; Studer, A.
cleavage via 1,5-HAT being irreversible (with a primary KIE Transition-Metal-Free Oxyarylation of Alkenes with Aryl Diazonium
observed in the intramolecular experiment, Scheme 5a.2), but Salts and TEMPONa. J. Am. Chem. Soc. 2012, 134, 16516−16519.
(b) Fumagalli, G.; Boyd, S.; Greaney, M. F. Oxyarylation and
taking place after the RDS, which involves the irreversible Aminoarylation of Styrenes Using Photoredox Catalysis. Org. Lett.
functionalization of the substrate bearing the C−H/D bond 2013, 15, 4398−4401. (c) Kindt, S.; Wicht, K.; Heinrich, M. R.
(i.e., no primary KIE observed in the two intermolecular Thermally Induced Carbohydroxylation of Styrenes with Aryldiazo-
deuterium labeling experiments summarized in Scheme 5a.1).20 nium Salts. Angew. Chem., Int. Ed. 2016, 55, 8744−8747.
(3) For selected examples, see: (a) Ouyang, X.-H.; Song, R.-J.; Hu,
3. CONCLUSIONS M.; Yang, Y.; Li, J.-H. Silver-Mediated Intermolecular 1,2-Alkylar-
ylation of Styrenes with α-Carbonyl Alkyl Bromides and Indoles.
In summary, a visible light mediated, redox neutral protocol is Angew. Chem., Int. Ed. 2016, 55, 3187−3191. (b) Yatham, V. R.; Shen,
presented here toward the efficient remote functionalization of Y.; Martin, R. Catalytic Intermolecular Dicarbofunctionalization of
alkenes. The mild reaction conditions enable the addition of a Styrenes with CO2 and Radical Precursors. Angew. Chem., Int. Ed.
wide variety of C-centered radicals to the double bond 2017, 56, 10915−10919.
producing a vicinal radical intermediate that is not quenched (4) For selected examples, see: (a) Panchaud, P.; Chabaud, L.;
under the reaction conditions. A subsequent 1,5-HAT results in Landais, Y.; Ollivier, C.; Renaud, P.; Zigmantas, S. Radical Amination
the site selective generation of a remote benzylic radical which with Sulfonyl Azides: A Powerful Method for the Formation of C-N
can be trapped with O- or C-nucleophiles to produce, in a Bonds. Chem.−Eur. J. 2004, 10, 3606−3614. (b) Weidner, K.; Giroult,
highly efficient manner, new Csp3−O and Csp3−Csp2 bonds at A.; Panchaud, P.; Renaud, P. Efficient Carboazidation of Alkenes
room temperature. Using a Radical Desulfonylative Azide Transfer Process. J. Am. Chem.


Soc. 2010, 132, 17511−17515. (c) Wang, F.; Qi, X.; Liang, Z.; Chen,
P.; Liu, G. Copper-Catalyzed Intermolecular Trifluoromethylazidation
ASSOCIATED CONTENT of Alkenes: Convenient Access to CF3-Containing Alkyl Azides.
*
S Supporting Information Angew. Chem., Int. Ed. 2014, 53, 1881−1886. (d) Dagousset, G.;
The Supporting Information is available free of charge on the Carboni, A.; Magnier, E.; Masson, G. Photoredox-Induced Three-
ACS Publications website at DOI: 10.1021/acscatal.8b00707. Component Azido- and Aminotrifluoromethylation of Alkenes. Org.
Lett. 2014, 16, 4340−4343. (e) Liu, Y.-Y.; Yang, X.-H.; Song, R.-J.;
General information, experimental procedures, DFT Luo, S.; Li, J.-H. Oxidative 1,2-Carboamination of Alkenes with Alkyl
calculations, detailed descriptions for products, and Nitriles and Amines toward γ-Amino Alkyl Nitriles. Nat. Commun.
copies of NMR spectra (PDF) 2017, 8, 14720−14726. (f) Bunescu, A.; Ha, T. M.; Wang, Q.; Zhu, J.
Crystallographic data (CIF) Copper-Catalyzed Three-Component Carboazidation of Alkenes with


Acetonitrile and Sodium Azide. Angew. Chem., Int. Ed. 2017, 56,
10555−10558.
AUTHOR INFORMATION (5) For selected examples, see: (a) Hara, T.; Iwahama, T.; Sakaguchi,
Corresponding Author S.; Ishii, Y. Catalytic Oxyalkylation of Alkenes with Alkanes and
Molecular Oxygen via a Radical Process Using N-Hydroxyphthalimide.
*E-mail: cristina.nevado@chem.uzh.ch (C.N.). J. Org. Chem. 2001, 66, 6425−6431. (b) Wetter, C.; Jantos, K.; Woithe,
ORCID K.; Studer, A. Intermolecular Radical Addition and Addition/
Estíbaliz Merino: 0000-0002-2960-5404 Cyclization Reactions of Alkoxyamines onto Nonactivated Alkenes.
Cristina Nevado: 0000-0002-3297-581X Org. Lett. 2003, 5, 2899−2902. (c) Feng, C.; Loh, T.-P. Copper-
Catalyzed Olefinic Trifluoromethylation of Enamides at Room
Notes Temperature. Chem. Sci. 2012, 3, 3458−3462. (d) Yi, H.; Zhang, X.;
The authors declare no competing financial interest. Qin, C.; Liao, Z.; Liu, J.; Lei, A. Visible Light-Induced γ-Alkoxynitrile


Synthesis via Three-Component Alkoxycyanomethylation of Alkenes.
ACKNOWLEDGMENTS Adv. Synth. Catal. 2014, 356, 2873−2877. (e) Liao, Z.; Yi, H.; Li, Z.;
Fan, C.; Zhang, X.; Liu, J.; Deng, Z.; Lei, A. Copper-Catalyzed Radical
We thank the European Research Council (ERC Starting grant Carbooxygenation: Alkylation and Alkoxylation of Styrenes. Chem. -
agreement no. 307948) and the Swiss National Science Asian J. 2015, 10, 96−99. (f) Chatalova-Sazepin, C.; Wang, Q.;
Foundation (SNF 200020_146853) for financial support. We Sammis, G. M.; Zhu, J. Copper-Catalyzed Intermolecular Carboether-
thank Prof. Anthony Linden for the X-ray diffraction analysis of ification of Unactivated Alkenes by Alkyl Nitriles and Alcohols. Angew.

6404 DOI: 10.1021/acscatal.8b00707


ACS Catal. 2018, 8, 6401−6406
ACS Catalysis Research Article

Chem., Int. Ed. 2015, 54, 5443−5446. (g) Wang, Y.; Zhang, L.; Yang, Visible-Light Photocatalytic Radical Alkenylation of α-Carbonyl Alkyl
Y.; Zhang, P.; Du, Z.; Wang, C. Alkene Oxyalkylation Enabled by Bromides and Benzyl Bromides. Chem. - Eur. J. 2013, 19, 5120−5126.
Merging Rhenium Catalysis with Hypervalent Iodine(III) Reagents via (c) Egami, H.; Usui, Y.; Kawamura, S.; Nagashima, S.; Sodeoka, M.
Decarboxylation. J. Am. Chem. Soc. 2013, 135, 18048−18051. (h) Ha, Product Control in Alkene Trifluoromethylation: Hydrotrifluorome-
T. M.; Chatalova-Sazepin, C.; Wang, Q.; Zhu, J. Copper-Catalyzed thylation, Vinylic Trifluoromethylation, and Iodotrifluoromethylation
Formal [2+2+1] Heteroannulation of Alkenes, Alkylnitriles, and Using Togni Reagent. Chem. - Asian J. 2015, 10, 2190−2199.
Water: Method Development and Application to a Total Synthesis (11) (a) Yu, P.; Lin, J.-S.; Li, L.; Zheng, S.-C.; Xiong, Y.-P.; Zhao, L.-
of (±)-Sacidumlignan D. Angew. Chem., Int. Ed. 2016, 55, 9249−9252. J.; Tan, B.; Liu, X.-Y. Enantioselective C-H Bond Functionalization
(i) Jian, W.; Ge, L.; Jiao, Y.; Qian, B.; Bao, H. Iron-Catalyzed Triggered by Radical Trifluoromethylation of Unactivated Alkene.
Decarboxylative Alkyl Etherification of Vinylarenes with Aliphatic Angew. Chem., Int. Ed. 2014, 53, 11890−11894. (b) Yu, P.; Zheng, S.-
Acids as the Alkyl Source. Angew. Chem., Int. Ed. 2017, 56, 3650−3654. C.; Yang, N.-Y.; Tan, B.; Liu, X.-Y. Phosphine-Catalyzed Remote β-
(6) For a redox-neutral protocol yielding broader scope at ambient C−H Functionalization of Amines Triggered by Trifluoromethylation
temperature, see: Tlahuext-Aca, A.; Garza-Sanchez, R. A.; Glorius, F. of Alkenes: One-Pot Synthesis of Bistrifluoromethylated Enamides and
Multicomponent Oxyalkylation of Styrenes Enabled by Hydrogen- Oxazoles. Angew. Chem., Int. Ed. 2015, 54, 4041−4045. (c) Huang, L.;
Bond-Assisted Photoinduced Electron Transfer. Angew. Chem., Int. Ed. Zheng, S.-C.; Tan, B.; Liu, X.-Y. Metal-Free Direct 1,6- and 1,2-
2017, 56, 3708−3711. Difunctionalization Triggered by Radical Trifluoromethylation of
(7) For reviews on 1,n-HAT, see: (a) Wolff, M. E. Cyclization of N- Alkenes. Org. Lett. 2015, 17, 1589−1592. (d) Huang, L.; Lin, J.-S.;
Halogenated Amines (The Hofmann-Löffler Reaction). Chem. Rev. Tan, B.; Liu, X.-Y. Alkene Trifluoromethylation-Initiated Remote α-
1963, 63, 55−64. (b) Breslow, R. Biomimetic Control of Chemical Azidation of Carbonyl Compounds toward Trifluoromethyl γ-Lactam
Selectivity. Acc. Chem. Res. 1980, 13, 170−177. (c) Chiba, S.; Chen, H. and Spirobenzofuranone-Lactam. ACS Catal. 2015, 5, 2826−2831.
sp3 C−H Oxidation by Remote H-Radical Shift with Oxygen- and (e) Li, L.; Ye, L.; Ni, S.-F.; Li, Z.-L.; Chen, S.; Du, Y.-M.; Li, X.-H.;
Nitrogen-Radicals: A Recent Update. Org. Biomol. Chem. 2014, 12, Dang, L.; Liu, X.-Y. Phosphine-Catalyzed Remote α-C−H Bond
4051−4060. (d) Hu, X.-Q.; Chen, J.-R.; Xiao, W.-J. Controllable Activation of Alcohols or Amines Triggered by the Radical
Remote C−H Bond Functionalization by Visible-Light Photocatalysis. Trifluoromethylation of Alkenes: Reaction Development and Mech-
Angew. Chem., Int. Ed. 2017, 56, 1960−1962. For recent selected anistic Insights. Org. Chem. Front. 2017, 4, 2139−2146.
examples, see: (e) Choi, G. J.; Zhu, Q.; Miller, D. C.; Gu, C. J.; (12) (a) Eisenberger, P.; Gischig, S.; Togni, A. Novel 10-I-3
Knowles, R. R. Catalytic Alkylation of Remote C−H Bonds Enabled Hypervalent Iodine-Based Compounds for Electrophilic Trifluorome-
by Proton-Coupled Electron Transfer. Nature 2016, 539, 268−271. thylation. Chem. - Eur. J. 2006, 12, 2579−2586. (b) Koller, R.; Stanek,
(f) Chu, J. C.; Rovis, T. Amide-Directed Photoredox-Catalysed C−C K.; Stolz, D.; Aardoom, R.; Niedermann, K.; Togni, A. Zinc-Mediated
Bond Formation at Unactivated sp3 C−H Bonds. Nature 2016, 539, Formation of Trifluoromethyl Ethers from Alcohols and Hypervalent
272−275. (g) Mukherjee, S.; Maji, B.; Tlahuext-Aca, A.; Glorius, F. Iodine Trifluoromethylation Reagents. Angew. Chem., Int. Ed. 2009, 48,
Visible-Light-Promoted Activation of Unactivated C(sp3)−H Bonds 4332−4336. (c) Kieltsch, I.; Eisenberger, P.; Togni, A. Mild
and Their Selective Trifluoromethylthiolation. J. Am. Chem. Soc. 2016, Electrophilic Trifluoromethylation of Carbon- and Sulfur-Centered
138, 16200−16203. (h) Jiang, H.; Studer, A. α-Aminoxy-Acid- Nucleophiles by a Hypervalent Iodine(III)−CF3 Reagent. Angew.
Auxiliary-Enabled Intermolecular Radical γ-C(sp3)-H Functionaliza- Chem., Int. Ed. 2007, 46, 754−757.
tion of Ketones. Angew. Chem., Int. Ed. 2018, 57, 1692−1696. (13) For selected reviews, see: (a) Narayanam, J. M. R.; Stephenson,
(i) Dauncey, E. M.; Morcillo, S. P.; Douglas, J. J.; Sheikh, N. S.; C. R. J. Visible Light Photoredox Catalysis: Applications in Organic
Leonori, D. Photoinduced Remote Functionalisations by Iminyl Synthesis. Chem. Soc. Rev. 2011, 40, 102−113. (b) Shaw, M. H.;
Radical Promoted C−C and C−H Bond Cleavage Cascades. Angew. Twilton, J.; MacMillan, D. W. C. Photoredox Catalysis in Organic
Chem., Int. Ed. 2018, 57, 744−748. (j) Becker, P.; Duhamel, T.; Stein, Chemistry. J. Org. Chem. 2016, 81, 6898−6926. (c) Yoon, T. P.;
C. J.; Reiher, M.; Muñiz, K. Cooperative Light-Activated Iodine and Stephenson, C. R. J. Opportunities in Photocatalytic Synthesis. Adv.
Photoredox Catalysis for the Amination of C(sp3)-H Bonds. Angew. Synth. Catal. 2014, 356, 2739. (d) Chen, J.-R.; Hu, X.-Q.; Lu, L.-Q.;
Chem., Int. Ed. 2017, 56, 8004−8008. Xiao, W.-J. Visible Light Photoredox-Controlled Reactions of N-
(8) (a) Qiu, J.-K.; Jiang, B.; Zhu, Y.-L.; Hao, W.-J.; Wang, D.-C.; Sun, Radicals and Radical Ions. Chem. Soc. Rev. 2016, 45, 2044−2056.
J.; Wei, P.; Tu, S.-J.; Li, G. Catalytic Dual 1,1-H-Abstraction/Insertion (14) For previous efforts in our group towards remote functionaliza-
for Domino Spirocyclizations. J. Am. Chem. Soc. 2015, 137, 8928− tions, see: (a) Shu, W.; Lorente, A.; Gómez-Bengoa, E.; Nevado, C.
8931. (b) Hu, M.; Fan, J.-H.; Liu, Y.; Ouyang, X.-H.; Song, R.-J.; Li, J.- Expeditious Diastereoselective Synthesis of Elaborated Ketones via
H. Metal-Free Radical [2+2+1] Carbocyclization of Benzene-Linked Remote C(sp3)−H Functionalization. Nat. Commun. 2017, 8, 13832−
1,n-Enynes: Dual C(sp3)−H Functionalization Adjacent to a 13839. (b) Shu, W.; Nevado, C. Visible-Light-Mediated Remote
Heteroatom. Angew. Chem., Int. Ed. 2015, 54, 9577−9580. Aliphatic C−H Functionalizations through a 1,5-Hydrogen Transfer
(9) For selected examples on haloalkylation and hydroalkylation of Cascade. Angew. Chem., Int. Ed. 2017, 56, 1881−1884. (c) Shu, W.;
alkenes, see: (a) Choi, S.; Kim, Y. J.; Kim, S. M.; Yang, J. W.; Kim, S. Genoux, A.; Li, Z.; Nevado, C. γ-Functionalizations of Amines through
W.; Cho, E. J. Hydrotrifluoromethylation and Iodotrifluoromethyla- Visible-Light-Mediated, Redox-Neutral C−C Bond Cleavage. Angew.
tion of Alkenes and Alkynes Using an Inorganic Electride as a Radical Chem., Int. Ed. 2017, 56, 10521−10524.
Generator. Nat. Commun. 2014, 5, 4881−4887. (b) Wilger, D. J.; (15) For additional experiments, see SI.
Gesmundo, N. J.; Nicewicz, D. A. Catalytic Hydrotrifluoromethylation (16) CCDC-1581108 (9′) contain the supplementary crystallo-
of Styrenes and Unactivated Aliphatic Alkenes via an Organic graphic data for this paper. The data can be obtained free of charge
Photoredox System. Chem. Sci. 2013, 4, 3160−3165. (c) Wu, X.; from the Cambridge Crystallographic Data Centre via www.ccdc.cam.
Chu, L.; Qing, F.-L. Silver-Catalyzed Hydrotrifluoromethylation of ac.uk/structures.
Unactivated Alkenes with CF3SiMe3. Angew. Chem., Int. Ed. 2013, 52, (17) Jung, M. E.; Piizzi, G. gem-Disubstituent Effect: Theoretical
2198−2202. (d) Mizuta, S.; Verhoog, S.; Engle, K. M.; Khotavivattana, Basis and Synthetic Applications. Chem. Rev. 2005, 105, 1735−1766.
T.; O’Duill, M.; Wheelhouse, K.; Rassias, G.; Médebielle, M.; (18) Note that for d1-48, four diastereotopic TS can be obtained for
Gouverneur, V. Catalytic Hydrotrifluoromethylation of Unactivated the 1,5-HAT step. Only in two of them the C−D bond is properly
Alkenes. J. Am. Chem. Soc. 2013, 135, 2505−2508. oriented for abstraction by the remote C-centered radical. These two
(10) For selected examples, see: (a) Nishikata, T.; Noda, Y.; TS will be, to a different extent, responsible for the KIE value
Fujimoto, R.; Sakashita, T. An Efficient Generation of a Functionalized determined experimentally. See ref 15.
Tertiary-Alkyl Radical for Copper-catalyzed Tertiary-Alkylative Miz- (19) (a) Szabo, Z. G. Stabilization of Free Radicals: Its Importance in
oroki-Heck Type Reaction. J. Am. Chem. Soc. 2013, 135, 16372− Reaction Kinetics. Nature 1952, 170, 246−247. (b) Henry, D. J.;
16375. (b) Liu, Q.; Yi, H.; Liu, J.; Yang, Y.; Zhang, X.; Zeng, Z.; Lei, A. Parkinson, C. J.; Mayer, P. M.; Radom, L. Bond Dissociation Energies

6405 DOI: 10.1021/acscatal.8b00707


ACS Catal. 2018, 8, 6401−6406
ACS Catalysis Research Article

and Radical Stabilization Energies Associated with Substituted Methyl


Radicals. J. Phys. Chem. A 2001, 105, 6750−6756. (c) Zipse, H.; Hioe,
J. Radical stability and its role in synthesis and catalysis. Org. Biomol.
Chem. 2010, 8, 3609−3617.
(20) (a) Anslyn, E. V.; Dougherty, D. A. Modern Physical Organic
Chemistry; University Science Books: Sausalito, 2006. (b) Simmons, E.
M.; Hartwig, J. F. 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.

6406 DOI: 10.1021/acscatal.8b00707


ACS Catal. 2018, 8, 6401−6406

S-ar putea să vă placă și