Subscribe to RSS
DOI: 10.1055/a-1961-6102
Catalytic Enantioselective Dihalogenation of Alkenes
This research was supported by the China Postdoctoral Science Foundation (2020M682280). We also thank the National Natural Science Foundation of China (NSFC, grant nos. 22071149 and 21871178), and the Science and Technology Commission of Shanghai Municipality (STCSM, grant no. 19JC1430100) for financial support. This research was also supported by The Program for Professor of Special Appointment (Eastern Scholar) at Shanghai Institutions of Higher Learning.
Abstract
Vicinal dihalides not only emerge as reactive intermediates in synthetic organic chemistry, but also are extensively existing in bioactive marine natural products. The dihalogenation of alkenes is the most direct and effective method for the synthesis of vicinal dihalides. Because there is always an exchange process between the chiral haloniums and the unreacted olefins to cause racemization, the development of catalytic enantioselective dihalogenation of alkenes is of great difficulty. Recently, great progress has been made in catalytic asymmetric manner. However, there is a lack of related review of discussions of the mechanisms and reaction systems. This review is aimed at summarizing enantioselective dihalogenation of alkenes, including 1,2-dichlorination, 1,2-dibromination, and 1,2-difluorination, which is expected to encourage more researchers to participate in this field.
1 Introduction
2 Enantioselective 1,2-Dichlorination and 1,2-Dibromination of Alkenes
2.1 Chiral-Boron-Complex-Promoted Enantioselective 1,2-Dichlorination
2.2 Organocatalytic Asymmetric 1,2-Dichlorination and 1,2-Dibromination
2.3 Chiral-Titanium-Complex-Catalyzed 1,2-Dihalogenation
3 Chiral-Iodide-Catalyzed Enantioselective Oxidative 1,2-Difluorination
4 Summary and Outlook
#
Biographical Sketches


Jia-Wei Dong received his BSc from Nanjing University in 2013. He obtained his PhD in organic chemistry in 2020 from Shanghai Jiao Tong University under the supervision of Prof. Yong-Qiang Tu. He started his research at Henan University in 2020. His current research focuses mainly on development of targeting treatment and pharmaceuticals for glioblastoma and Alzheimer’s.


Ren-Fei Cao received his BSc from Zhengzhou University in 2019. Currently, he is a second-year PhD student under the supervision of Prof. Zhi-Min Chen at Shanghai Jiao Tong University.


Zhi-Min Chen received his BSc from Fuzhou University in 2009. He obtained his PhD in organic chemistry in 2014 from Lanzhou University under the supervision of Prof. Yong-Qiang Tu. After spending three years (2014–2017) as a postdoctoral fellow at Shanghai Jiao Tong University (Mentor: Prof. Yong-Qiang Tu) and University of Utah (2015–2017, Mentor: Prof. Matthew S. Sigman), he started his independent research at Shanghai Jiao Tong University in 2017. His current research focuses mainly on chiral organochalcogen chemistry and asymmetric catalysis.
Introduction
Vicinal dihalides are extensively existing in marine natural products with bioactivity and also emerged as reactive intermediates in synthetic organic chemistry.[1] [2] As a classic textbook reaction, the dihalogenation of alkenes have been utilized to afford vicinal dihalides for more than a century, which is shown to be the most direct and effective method. Nevertheless, the development of catalytic, enantioselective dihalogenation of alkenes is of great difficulty and successful examples are still limited.[3,4]
According to experimental and computational evidence, the dihalogenation of alkenes proceeds via haliranium ion intermediate in most cases except difluorination. As shown in Scheme [1], alkene dihalogenation is generally featured with a two-step mechanism. The dihalogenation reaction proceeds through initial electrophilic addition to Re or Si face of alkenes and forms enantiomeric haliranium ion intermediates I and ent-I. Subsequent nucleophilic attack of haliranium ion is confronted with the regioselectivity of two carbon termini, which results in enantiomeric vicinal dihalides II and ent-II. It implies that, even if electrophilic addition process is rendered facial-selective, the next step may lead to racemic products due to poor regioselectivity. In addition, it is potential for causing racemization of enantioenriched haliranium ions resulting from alkene-to-alkene haliranium ion transfer processes. Hence, it is quite difficult to achieve highly enantioselective dihalogenation of alkenes owing to the challenges mentioned above. As for difluorination of alkenes, unidentical linear cation intermediates are generated due to the small atomic radius and maximum electronegativity, which results in extremely high energy to form the cyclic onium intermediates.


In practice, there are inevitable limitations to the screening of reaction conditions. The readily leaving tendency of halogen atom, especially in the case of the benzylic position, lead to the halogenation incompatible with somewhat nucleophilicity and alkalinity.[5] Moreover, the strategy to create a chiral environment for the halonium is still rare, and certain known canals might result in the racemization of the chiral onium intermediate, all of which caused great challenges to the development of the enantioselective dihalogenation and there are only a handful of relevant reports so far.
Since pioneering work of Nicolaou and co-workers in 2011,[6] some prominent studies were subsequently reported. Herein, we sort out and generalize from the remarkable reports according to our superficial understanding rather than simple overview. We look forward to attracting chemist’s in-depth attention and interest in this area and discovering different reaction types and mechanisms.
# 2
Enantioselective 1,2-Dichlorination and 1,2-Dibromination of Alkenes
2.1Chiral-Boron-Complex-Promoted Enantioselective 1,2-Dichlorination
During the synthetic work of napyradiomycin A1 by Snyder and co-workers in 2009, a substrate-involving chiral boron complex intermediate strategy was adopted to introduce chirality.[7] The (S)-[1,1′-biphenanthrene]-2,2′-diol ((S)-3) and oxygen atoms in the substrate coordinate with boron simultaneously and thus creates asymmetry between the two faces of the olefin. Subsequently introduced chlorine under low temperature of –78 °C promotes an electrophilic addition to the double bond and generates the cyclic chloronium on the side with less steric hindrance. Notably, the biphenanthrenol also protected the aryl ring of substrate from electrophilic chlorination while generating chiral steric hindrance. The regioselectivity of the final chloride-attack-producing ring opening is mainly determined by the steric hindrance of the geminal dimethyl group on the substrate. Undoubtedly, the phenolic hydroxyl group and quinone carbonyl group on the substrate are crucial for the chirality as the grasper for the binding of boron, as well as the π stacking between the large conjugated planar structure of the substrate and the fused aromatic ring of biphenanthrenol are favorable for interaction of substrate and chiral boron and yielding stable enantioselectivity. Since the most expensive (S)-[1,1′-biphenanthrene]-2,2′-diol in this catalytic system can be synthesized from the upstream cheap phenanthrene, and easily recovered and recycled, this method is of general significance and can be implemented in most laboratories and industries (Scheme [2]).[8]


# 2.2
Organocatalytic Asymmetric 1,2-Dichlorination and 1,2-Dibromination
In 2010, Snyder and co-workers reported a stable adduct from chloronium and chiral thiolane, which proved capable to confer weak enantioselectivity in the 1,2-dichlorination of the alkenes but with only up to 14% ee. Background reaction caused by reagent decomposition is most likely to be suspected. However, no product obtained in the absence of stabilization by SbCl5 as well as the ee is positively correlated with yield suggested it may be more than that.[9] From the subsequent relevant experimental results of control variables can be inferred that the influence of the steric hindrance of the reagent as well as the lack of stabilizing interaction between chiral sulfur compounds and substrates is most important to the result.[10] Denmark and co-workers conceived the diselenide catalysts for enantioselective 1,2-dichlorination, which include the selenium atom of the same family with sulfur as the center of catalytic activity.[11] Notably, the terminal reactive species is the adduct from multistep oxidative addition of chlorine to the selenide in the presence of N-fluoropyridinium tetrafluoroborate. Therefore, we discuss the reaction with the sulfur reagent together for their high similarities in mechanism. The structure of the catalyst has been intensively optimized apart with mechanistic validations. Despite there is omission of reaction-rate details and the experimental results without catalyst, we believe that there is remote probability for direct oxidative addition to double bond with chlorine source in the presence of oxidant. Under the catalysis of isopropyl silicon substitute the highest ee of 52% was obtained. Subsequent optimization requires substantial investment considering that the diverse synthesis of sulfur- and selenium-based compounds is still extremely challenging (Scheme [3]).


Cinchona alkaloid derivative showed obvious advantages, such as high efficiency, enantioselectivity, flexibility, and versatility, in the catalysis of enantioselective 1,2-dihalogenation by forming appropriately reactive adducts with chloronium or bromonium. It is essential for stereoinduction that an additional interaction between the substrate and catalyst stabilizes the reactive transition state, for example, an alcohol hydroxyl employed by Nicolaou and co-workers, likewise analogical the aryl amide in the work of Borhan and co-workers.[6a] [12] Even for the more challenging 1,2-dihalogenation of unfunctionalized alkenes described by Hennecke and co-workers, a π stacking between the aryl of the substrate and naphthyl of the catalyst is speculated indispensable for stereoselectivity.[13] Actually, Nicolaou and co-workers have demonstrated that either blocking the hydrogen-bond donor on the substrate or replacing the hydrogen-bond acceptor on the catalyst all results in a greatly compromised enantioselectivity. Exchange between the haloniums and the unreacted olefins, which was discovered by Brown and Denmark et al. is recognized to cause racemization.[14] Low temperature and promotion of subsequent nucleophilic halogenation is conducive to mitigate the racemization as well as competitive nucleophilic solvation. Especially in TFE a 100 equivalent of LiCl was employed as nucleophilic chlorine source even if most undissolved. Most recently, Bayeh-Romero and co-workers achieved the chiral-alkaloid-catalyzed asymmetric chlorobromination of conjugated ketones with near stoichiometric amounts of NBS and thionyl chloride as halogen sources. By comparing several reaction systems, it can be found that the use of aprotic, non-nucleophilic solvent systems and organic homogeneous halogen sources can avoid the introduction of excess competing nucleophilic halide salts (Scheme [4]).[15]


Known as the anionic binder, the urea group can bind the halogen anion by bifurcated hydrogen-bond interaction. With this concept, Tan and co-workers conceived of the enantioselective 1,2-dihalogenation of alkenes with urea moiety, which enabled greatly acceleration of the rate of nucleophilic halogenation and thus circumvent the regioselectivity issue.[16] The amount of nucleophilic halogen source is reduced to 5 to 20 equivalents. Under the NBS as Br+ precursor, various dihalogenations with metal fluoride, chloride, and bromide were investigated obtaining generally excellent enantio- and regioselectivities. In the 1,2-dihalogenation of 2-tert-butylphenylacetylene derivatives, the chirality is converted from the exquisite face selectivity during forming of halonium into the axial chirality of the product and preserved well. This protocol owns general significance in both research and application for urea and is ready for synthesis, removal, and diverse transformation (Scheme [5]).


# 2.3
Chiral-Titanium-Complex-Catalyzed 1,2-Dihalogenation
Burns and co-workers developed the titanium-ligand catalysis system for enantioselective 1,2-dihalogenation of alkenes that featured differentiated regioselectivity (anti-Markovnikov) of the nucleophilic halogenation step.[17] This reaction system has undergone two generations of optimization. Initially, bidentate chiral diol ligand was resorted so as the bromine source of malonate derivative contribute two coordination numbers equally. For the five-coordinated Ti(IV) with sp3d hybrid orbitals that is too stable to sustain the catalytic cycle, it was hypothesized more responsible to form a six-coordinate octahedral titanium(IV) for the halogenation. In the second-generation system, chiral dianionic tridentate Schiff base ligand was introduced to replace the bidentate ligand, so that employment of monodentate NBS as a source of electrophilic bromine in place of halomalonates maintained a six-coordinate titanium(IV) intermediate. As a matter of fact, more reactive NBS makes obvious positive contribution to the conversion rates and yields. This system proved a broad substrate availability and good enantio- and regioselectivity under the optimized conditions. In addition, the trans effect of NBS coordination to the titanium–bromine bond cannot be ruled out as a condition for activation. Therefore, the possibility of using catalytic amount of titanium salt and economic metal bromides to replace the TiBr(O i Pr)3, which was prepared from the expensive TiBr4, should be considered for nucleophilic bromination reactions (Scheme [6]).


#
# 3
Chiral-Iodide-Catalyzed Enantioselective Oxidative 1,2-Difluorination
It was reported that appropriate fluorine source can form hypervalent iodine adduct with iodoarene under the oxidative conditions, which maintains a delicate stability and performs the catalytic cycle of difluorination of alkenes.[18] Due to certain previously mentioned properties it is difficult for fluorine to form the stabilized cyclic halonium intermediates similar to chlorine and bromine, therefore neighboring group assistance to extend lifetime of the cation intermediates is highly required. According to the reaction mechanisms studied by Jacobsen and co-workers, aryl migration via a phenonium intermediate occurred during the difluorination of cinnamic acid derivatives to afford geminal difluorinated product with chirality transferred to the vicinal carbon. With increase in steric hindrance of the migration destination or N-tert-butylation of cinnamamide substrate, products resulting from the neighboring group assistance of the phenyl can be almost completely inhibited. It is suggested that the amide group of cinnamamide capture carbocation to form the onium intermediate process is kinetically favorable.


Nevertheless, the aryl neighboring group assistance is unavoidable during the 1,2-difluorination of aryl vinyl type substrates reported by Gilmour and co-workers.[19] In this case, the kinetic choice of the nucleophilic fluorination in the second step is the key to determining the geminal or vicinal difluorination. Excessive hydrofluoric acid can form a hydrogen bridge with the fluorine in the onium intermediate and increase the electropositivity of the connected carbon leading to different selectivity (Scheme [7]).
#
Summary and Outlook
Although chlorine, bromine, and fluorine from the same element family possess similarities in reaction habits, each one is provided with its own individuality due to different atomic radii, electronegativity, number of electron shells, etc. This objective contradiction is also reflected in the relevant reaction mechanisms and conditions. Anyway, methodological development of enantioselective dihalogenation of alkenes enabled diverse synthesis of versatile chiral 1,2-dihalides. Currently achieved approaches owned great significance but this is still far more than the end of the story. Since there is never a catalyst that works magically on all kinds of substrates, the methodological developments always benefit from newly evolved catalytic system to expand the substrates, enhanced efficiency and enantio-/diastereoselectivity, and popularized synthetic applications. Therefrom it is worth looking forward to the fusion of some modern disciplines, such as computational chemistry and enzymatic catalysis to bring new impact to this research field.[20]
#
#
Conflict of Interest
The authors declare no conflict of interest.
-
References
- 1a Gribble GW. Naturally Occurring Organohalogen Compounds – A Comprehensive Survey . Progress in the Chemistry of Organic Natural Products, Vol. 68. Herz W, Kirby GW, Moore RE, Steglich W, Tamm C. Springer; Wien: 1996. DOI:
- 1b Balard A.-J. Archiv Gesammte Naturl. 1826; 4: 231
- 2a Chung W.-j, Vanderwal CD. Angew. Chem. Int. Ed. 2016; 55: 4396
- 2b Saikia I, Borah AJ, Phukan P. Chem. Rev. 2016; 116: 6837
- 2c Doobary S, Lennox AJ. J. Synlett 2020; 31: 1333
- 3a Landry ML, Burns NZ. Acc. Chem. Res. 2018; 51: 1260
- 3b Bock J, Guria S, Wedek V, Hennecke U. Chem. Eur. J. 2021; 27: 4517
- 4 He T, Zeng X. Chin. J. Org. Chem. 2017; 37: 798
- 5 Cresswell AJ, Eey ST.-C, Denmark SE. Angew. Chem. Int. Ed. 2015; 54: 15642
- 6a Nicolaou KC, Simmons NL, Ying Y, Heretsch PM, Chen JS. J. Am. Chem. Soc. 2011; 133: 8134
- 6b Sarie JC, Neufeld J, Daniliuc CG, Gilmour R. ACS Catal. 2019; 9: 7232
- 7 Snyder SA, Tang Z.-Y, Gupta R. J. Am. Chem. Soc. 2009; 131: 5744
- 8a Tanaka Y, Sakuraba H, Nakanishi H. J. Chem. Soc., Chem. Commun. 1983; 947
- 8b Tanaka K, Shiraishi R, Toda F. J. Chem. Soc., Perkin Trans. 1 1999; 3069
- 9 Snyder SA, Treitler DS, Brucks AP. J. Am. Chem. Soc. 2010; 132: 14303
- 10 Brucks AP, Treitler DS, Liu S.-A, Snyder SA. Synthesis 2013; 45: 1886
- 11a Gilbert BB, Eey ST.-C, Ryabchuk P, Garry O, Denmark SE. Tetrahedron 2019; 75: 4086
- 11b Cresswell AJ, Eey ST.-C, Denmark SE. Nat. Chem. 2015; 7: 146
- 12 Soltanzadeh B, Jaganathan A, Yi Y, Yi H, Staples RJ, Borhan B. J. Am. Chem. Soc. 2017; 139: 2132
- 13 Wedek V, Van Lommel R, Daniliuc CG, De Proft F, Hennecke U. Angew. Chem. Int. Ed. 2019; 58: 9239
- 14a Brown RS, Nagorski RW, Bennet AJ, McClung RE. D, Aarts GH. M, Klobukowski M, McDonald R, Santarsiero BD. J. Am. Chem. Soc. 1994; 116: 2448
- 14b Bennet AJ, Brown RS, McClung RE. D, Klobukowski M, Aarts GH. M, Santarsiero BD, Bellucci G, Bianchini R. J. Am. Chem. Soc. 1991; 113: 8532
- 14c Denmark SE, Burk MT, Hoover AJ. J. Am. Chem. Soc. 2010; 132: 1232
- 15 Lubaev AE, Rathnayake MD, Eze F, Bayeh-Romero L. J. Am. Chem. Soc. 2022; 144: 13294
- 16 Wu S, Xiang S.-H, Li S, Ding W.-Y, Zhang L, Jiang P.-Y, Zhou Z.-A, Tan B. Nat. Catal. 2021; 4: 692
- 17a Hu DX, Shibuya GM, Burns NZ. J. Am. Chem. Soc. 2013; 135: 12960
- 17b Hu DX, Seidl FJ, Bucher C, Burns NZ. J. Am. Chem. Soc. 2015; 137: 3795
- 17c Bucher C, Deans RM, Burns NZ. J. Am. Chem. Soc. 2015; 137: 12784
- 17d Landry ML, Hu DX, McKenna GM, Burns NZ. J. Am. Chem. Soc. 2016; 138: 5150
- 18a Banik SM, Medley JW, Jacobsen EN. Science 2016; 353: 51
- 18b Banik SM, Medley JW, Jacobsen EN. J. Am. Chem. Soc. 2016; 138: 5000
- 18c Zhou B, Haj MK, Jacobsen EN, Houk KN, Xue X.-S. J. Am. Chem. Soc. 2018; 140: 15206
- 18d Haj MK, Banik SM, Jacobsen EN. Org. Lett. 2019; 21: 4919
- 19 Scheidt F, Schäfer M, Sarie JC, Daniliuc CG, Molloy JJ, Gilmour R. Angew. Chem. Int. Ed. 2018; 57: 16431
For selected reviews, see:
A strategy to create asymmetric environment on substrate using cyclodextrins etc. was reported by Tanaka and co-workers from 1983 on. Here we won’t go into too much detail. See:
Corresponding Author
Publication History
Received: 02 July 2022
Accepted after revision: 14 October 2022
Accepted Manuscript online:
14 October 2022
Article published online:
24 October 2022
© 2022. Thieme. All rights reserved
Georg Thieme Verlag KG
Rüdigerstraße 14, 70469 Stuttgart, Germany
-
References
- 1a Gribble GW. Naturally Occurring Organohalogen Compounds – A Comprehensive Survey . Progress in the Chemistry of Organic Natural Products, Vol. 68. Herz W, Kirby GW, Moore RE, Steglich W, Tamm C. Springer; Wien: 1996. DOI:
- 1b Balard A.-J. Archiv Gesammte Naturl. 1826; 4: 231
- 2a Chung W.-j, Vanderwal CD. Angew. Chem. Int. Ed. 2016; 55: 4396
- 2b Saikia I, Borah AJ, Phukan P. Chem. Rev. 2016; 116: 6837
- 2c Doobary S, Lennox AJ. J. Synlett 2020; 31: 1333
- 3a Landry ML, Burns NZ. Acc. Chem. Res. 2018; 51: 1260
- 3b Bock J, Guria S, Wedek V, Hennecke U. Chem. Eur. J. 2021; 27: 4517
- 4 He T, Zeng X. Chin. J. Org. Chem. 2017; 37: 798
- 5 Cresswell AJ, Eey ST.-C, Denmark SE. Angew. Chem. Int. Ed. 2015; 54: 15642
- 6a Nicolaou KC, Simmons NL, Ying Y, Heretsch PM, Chen JS. J. Am. Chem. Soc. 2011; 133: 8134
- 6b Sarie JC, Neufeld J, Daniliuc CG, Gilmour R. ACS Catal. 2019; 9: 7232
- 7 Snyder SA, Tang Z.-Y, Gupta R. J. Am. Chem. Soc. 2009; 131: 5744
- 8a Tanaka Y, Sakuraba H, Nakanishi H. J. Chem. Soc., Chem. Commun. 1983; 947
- 8b Tanaka K, Shiraishi R, Toda F. J. Chem. Soc., Perkin Trans. 1 1999; 3069
- 9 Snyder SA, Treitler DS, Brucks AP. J. Am. Chem. Soc. 2010; 132: 14303
- 10 Brucks AP, Treitler DS, Liu S.-A, Snyder SA. Synthesis 2013; 45: 1886
- 11a Gilbert BB, Eey ST.-C, Ryabchuk P, Garry O, Denmark SE. Tetrahedron 2019; 75: 4086
- 11b Cresswell AJ, Eey ST.-C, Denmark SE. Nat. Chem. 2015; 7: 146
- 12 Soltanzadeh B, Jaganathan A, Yi Y, Yi H, Staples RJ, Borhan B. J. Am. Chem. Soc. 2017; 139: 2132
- 13 Wedek V, Van Lommel R, Daniliuc CG, De Proft F, Hennecke U. Angew. Chem. Int. Ed. 2019; 58: 9239
- 14a Brown RS, Nagorski RW, Bennet AJ, McClung RE. D, Aarts GH. M, Klobukowski M, McDonald R, Santarsiero BD. J. Am. Chem. Soc. 1994; 116: 2448
- 14b Bennet AJ, Brown RS, McClung RE. D, Klobukowski M, Aarts GH. M, Santarsiero BD, Bellucci G, Bianchini R. J. Am. Chem. Soc. 1991; 113: 8532
- 14c Denmark SE, Burk MT, Hoover AJ. J. Am. Chem. Soc. 2010; 132: 1232
- 15 Lubaev AE, Rathnayake MD, Eze F, Bayeh-Romero L. J. Am. Chem. Soc. 2022; 144: 13294
- 16 Wu S, Xiang S.-H, Li S, Ding W.-Y, Zhang L, Jiang P.-Y, Zhou Z.-A, Tan B. Nat. Catal. 2021; 4: 692
- 17a Hu DX, Shibuya GM, Burns NZ. J. Am. Chem. Soc. 2013; 135: 12960
- 17b Hu DX, Seidl FJ, Bucher C, Burns NZ. J. Am. Chem. Soc. 2015; 137: 3795
- 17c Bucher C, Deans RM, Burns NZ. J. Am. Chem. Soc. 2015; 137: 12784
- 17d Landry ML, Hu DX, McKenna GM, Burns NZ. J. Am. Chem. Soc. 2016; 138: 5150
- 18a Banik SM, Medley JW, Jacobsen EN. Science 2016; 353: 51
- 18b Banik SM, Medley JW, Jacobsen EN. J. Am. Chem. Soc. 2016; 138: 5000
- 18c Zhou B, Haj MK, Jacobsen EN, Houk KN, Xue X.-S. J. Am. Chem. Soc. 2018; 140: 15206
- 18d Haj MK, Banik SM, Jacobsen EN. Org. Lett. 2019; 21: 4919
- 19 Scheidt F, Schäfer M, Sarie JC, Daniliuc CG, Molloy JJ, Gilmour R. Angew. Chem. Int. Ed. 2018; 57: 16431
For selected reviews, see:
A strategy to create asymmetric environment on substrate using cyclodextrins etc. was reported by Tanaka and co-workers from 1983 on. Here we won’t go into too much detail. See:



















