Alkylarylation of alkenes with arylsulfonylacetate as bifunctional reagent via photoredox radical addition/Smiles rearrangement cascade

Chonglong He Yulong Wang Quan-Xin Li Zichen Yan Keyuan Zhang Shao-Fei Ni Xin-Hua Duan Le Liu

Citation:  Chonglong He, Yulong Wang, Quan-Xin Li, Zichen Yan, Keyuan Zhang, Shao-Fei Ni, Xin-Hua Duan, Le Liu. Alkylarylation of alkenes with arylsulfonylacetate as bifunctional reagent via photoredox radical addition/Smiles rearrangement cascade[J]. Chinese Chemical Letters, 2025, 36(5): 110253. doi: 10.1016/j.cclet.2024.110253 shu

Alkylarylation of alkenes with arylsulfonylacetate as bifunctional reagent via photoredox radical addition/Smiles rearrangement cascade

English

  • Difunctionalization of alkenes provides a robust tool for converting simple alkenes into complex molecules, which is of high interest and challenging from both academic and industrial perspectives [1-7]. The simultaneous formation of multiple bonds using just one bifunctional reagent has become state-of-the-art in achieving this target [8-17]. Numerous examples of bifunctional reagents-mediated difunctionalization of alkenes have been disclosed in recent decades due to the rapid development of electrocatalysis [18-20] and photocatalysis [18,21-30]. Among them, the sulfonyl bifunctional reagent stands out for enabling alkene difunctionalization via a radical-induced functional group migration process, proving to be one of the most efficient approaches [31-34]. In general, those reactions could be categorized mechanistically into the following two types: (1) Single-electron oxidation of the electron-rich alkenes to generate the key active radical cation species. Then, bifunctional reagents nucleophilic attach the radical cation and engage in the following intramolecular Smiles rearrangement to achieve the difunctionalization (Fig. 1A) [35-37]. (2) Instead of alkene oxidation, single-electron reduction of sulfonyl alkyl bromide bifunctional reagents yields the electrophilic radical species, which subsequently undergo radical addition to styrenes. Then, the benzyl radical goes through ipso attack, leading to 1,4-functional group migration to deliver the difunctionalized products (Fig. 1B) [32,33,38-46]. Pioneered by Stephenson and Zhu, various efficient desulfonylative difunctionalizations of alkenes with these activation modes have been reported for incorporating amino or alkyl units and (hetero)aryl units across alkenes. In addition, an asymmetric radical difunctionalization was also realized by the Nevado group, employing chiral sulfinamide as the auxiliary [47,48]. Nevertheless, both the mechanistic activation mode and the substrate scopes in alkene radical difunctionalizations are still limited. More easily accessible and functional group-compatible bifunctional reagents, as well as complementary mechanistic activation modes, are still highly desirable.

    Figure 1

    Figure 1.  Previous work on sulfonyl bifunctional reagent mediated difunctionalization of alkenes and our working hypothesis.

    Recently, arylsulfonylacetate has been independently developed as a bifunctional reagent for the difunctionalization of unsaturated carbon-carbon bonds by both our research group [49,50] and others [51]. As an extension of our ongoing exploration into radical chemistry [52-57], we conceived a strategy employing arylsulfonylacetate as a bifunctional reagent for the difunctionalization of styrenes to access γ,γ-diaryl and γ-aryl ester compounds, which are frequently encountered in bioactive compounds and natural products (Fig. 1D) [58-60]. Subsequently, a crucial mechanistic question arose, specifically concerning the activation mode that would govern the anticipated transformation (Fig. 1C). This consideration holds significant implications for both the substrate scope and the overall efficiency of the proposed method. Theoretically, the acidic proton (pKa 13–16) in the active methylene site would be readily deprotonated under basic conditions to give a nucleophilic enolate, which would trap alkene radical cation via alkene activation mode. Conversely, the bond dissociation energy of the C‒H bond at the active methylene site falls within the range of approximately 96 kcal/mol, rendering selective hydrogen atom abstraction (HAA) thermodynamically feasible [50]. Then, the generation of a highly electrophilic radical species, facilitated through either a suitable hydrogen atom abstraction (HAA) or a single electron oxidation process appears operational to initiate the difunctionalization via bifunctional reagent activation mode.

    To test the feasibility of our hypothesis and inspire the future development of novel bifunctional reagents for alkene difunctionalization, we selected arylsulfonyl acetate 1aa and trans-anethole 2aa as model substrates for our initial investigations. To our delight, when 4CzIPN was employed as the photocatalyst and K3PO4 as the base, the desired γ,γ-diaryl ester 3aa could be formed in 76% yield with excellent diastereoselectivity (Fig. 2A). Then, cyclic voltammetry (CV) measurements were carried out to elucidate the oxidative process underlying this photocatalytic event. Compound 1aa displayed the first distinct oxidation at Ep/2 = + 0.68 V (vs. SCE in CH3CN) in the presence of base, while that of trans-anethole 2aa was determined to be Ep/2 = + 1.30 V (vs. SCE in CH3CN) which is consistent with the reported data (Fig. 2B) [61,62]. Given the redox potential of the photocatalyst 4CzIPN (*E1/2 (PC*/PC- = +1.35 V vs. SCE in MeCN) enables the thermodynamically favorable oxidation of both reactants, further experiments were performed. Stern-Volmer fluorescence quenching experiments revealed a notably faster quenching rate for 1aa in the presence of a base compared to 2aa, while pure 1aa is entirely unresponsive (Fig. 2C). Moreover, upon the introduction of 2.0 equiv. of the radical scavenger TEMPO into the reaction mixture, the reaction was entirely inhibited, concomitant with the detection of TEMPO adduct 5 (Fig. 2D). These results demonstrate that the crucial radical species derived from bifunctional reagent 1aa actively participated in this cascade transformation. As anticipated, Ir(ppy)3 possessing a relatively low oxidative potential (EIr(Ⅱ)/Ir(Ⅲ)* = +0.31 V) [63,64] exhibited completely inertness in the reaction. While Ru(bpy)3Cl2 with an oxidation potential positioned between those of 1aa and 2aa was employed as the photocatalyst, the desired product 3aa can be obtained with a 72% yield. Intriguingly, switching the photocatalyst to Mes-Acr+-MeClO4- or TPT which were commonly used for alkene radical cation species generation [65-70], resulted in no detection of 3aa (Fig. 2E). Taken together, these data provide support for single-electron oxidation of bifunctional reagent 1aa to initiate this radical difunctionalization reaction.

    Figure 2

    Figure 2.  Preliminary investigation and proof of the concept.

    Further optimization of reaction parameters, including the photocatalyst, base, and solvent, led us to identify the optimal conditions to achieve an 84% yield of product 3aa with >20:1 diastereoselectivity (For details, see Supporting information). Control experiments indicated that the photocatalyst, light irradiation, and K3PO4·3H2O were all ciritical to the cascade transformation, as the absence of any of these components resulted in no detection of the desired product (Fig. 2F). To the best of our knowledge, this reaction not only presents a novel alternative for the consecutive formation of multiple C‒C bonds, leading to the synthesis of γ,γ-diaryl esters but also introduces a complementary mechanistic activation mode for the sulfonyl bifunctional reagent-mediated difunctionalization of alkenes.

    With the optimal conditions in hand, our initial focus was on exploring the structural diversity concerning the migrating aromatic ring in arylsulfonylacetate bifunctional reagents (Scheme 1). Phenyl rings containing various substituents, including electron-donating groups such as methyl, tert–butyl, and methoxy, as well as electron-withdrawing groups (halogens), were all found to be compatible to provide the corresponding γ,γ-diarylesters (3aa-3ap) in 20% to 84% yields with excellent diastereoselectivity. It is worth noting that the structure of 3aa was ambiguously confirmed by X-ray analysis of the corresponding acid derivative (CCDC: 2323478). Generally, bifunctional reagents featuring electron-withdrawing groups on the phenyl rings exhibited more efficient reactions with alkenes, resulting in higher yields of the corresponding γ,γ-diarylesters compared to those with electron-donating groups. This phenomenon was attributed to the favorable π-π stacking interaction between the electron-deficient aryl ring of the bifunctional reagent and the electron-rich PMP rings [71]. This interaction is believed to contribute to lowering the energy barrier for the aryl migration process [36]. Significantly, functional groups like halides, amenable to subsequent derivatization, were well-tolerated, yielding the corresponding products in good yields. Furthermore, heteroaryl groups such as pyridine and thiophene were found to be compatible under the standard conditions, yielding γ,γ-diarylesters 3ao and 3ap in moderate yields.

    Scheme 1

    Scheme 1.  Variation of the arylsulfonylacetates: reactions were carried out on 0.2 mmol scale and isolated yields are given. dr was determined by crude NMR spectroscopy, for details see Supporting information.

    We extended our investigation to explore the diversity of substituents in the electron-rich internal alkenes (Scheme 2). Both mono- and multi-substituted alkoxy phenyl rings in olefins resulted in the production of the target products with high yields (3ba-3bd). Notably, various protecting groups for the phenol moiety, including benzoyl, Boc, and triflate, were all compatible under the standard conditions (3bd-3bg). Particularly, the free hydroxyl group remained intact, yielding the corresponding difunctionalized product 3bh in a moderate yield of 65%. Furthermore, alkyl chains containing halide and cyano functionalities were also well-tolerated under the standard conditions, affording the products 3bj-3bl. It is noteworthy that an electron-rich styrene containing a terminal alkene chain exhibits pronounced selectivity to give the expected product 3bm in moderate 65% yield. This selectivity can be attributed to the favorable π-π stacking interactions within the electron-deficient aromatic phenyl ring of 1a and the electron-rich alkene 2m. Increasing the length of the alkyl chain in the internal alkene showed negligible impact on the transformation (3bn). Trisubstituted internal alkene also proved to be effective substrate under the standard conditions, yielding the difunctionalized product 3bo bearing a quaternary center with moderate yield and diastereoselectivity. Internal alkene was also examined, but poor reactivity and diastereoselective was observed (3bp). This was attributed to the steric hindrance of the cyclic internal alkene for both the radical addition and aryl migration process. Furthermore, methylthiol and dialkylamino groups were proven to be suitable electron-donating groups under this photoredox conditions, furnishing the corresponding γ,γ-diarylesters (3bq-3bs) in moderate yields with good diastereoselectivities.

    Scheme 2

    Scheme 2.  Variation of the electron-rich alkenes: reactions were carried out on 0.2 mmol scale and isolated yields are given. dr was determined by crude NMR spectroscopy, for details see Supporting information.

    Considering the initiation of the reaction through the bifunctiomal reagent activation mode, it can be inferred that radical acceptors are not restricted solely to electron-rich anethole derivatives (Scheme 3). To validate our hypothesis, (E)–prop-1-en-1-ylbenzene, with an oxidation potential Ep/2 = +1.74 V vs. SCE in CH3CN surpassing that of the excited photocatalyst 4CzIPN (E1/2 (PC*/PC- = +1.35 V vs. SCE in MeCN), was subjected to the standard conditions. As expected, the desired difunctionalized product 3ca was formed in a moderate yield of 65%. Substituents bearing different electronic propertied were also effective precursor for the preparation of the γ,γ-diarylesters 3cb-3d In addition, heterocycle substituted internal alkene can also be tolerated under the standard conditions, affording the thiophene substituted product 3ce with a yield of 34%. The disubstituted terminal alkene exhibited relatively low reactivity to produce a quaternary-carbon-bearing γ,γ-diarylester 3cf in 20% yield. Furthermore, styrenes turned out to be completely inert under the standard conditions. Moreover, this photocatalyzed radical addition/Smiles rearrangement cascade reaction is applicable to a variety of unactivated olefins, such as terminal and internal alkenes, yielding the corresponding alkylarylated products (3ci-3cl) with moderate yields. This underscores the synthetic versatility of our approach. Notably, cyclopentene and 1,2-dihydrofuran selectively generated cis-products, which likely attributed to the formation of the most stable cis-fused intermediates from these substrates [51].

    Scheme 3

    Scheme 3.  Variation of the electron-rich alkenes: reactions were carried out on 0.2 mmol scale and isolated yields are given. dr was determined by crude NMR spectroscopy, for details see supporting information.

    To gain additional insights in the reaction mechanism, several control experiments were performed (Scheme 4). First, standard conditions were applied to two independent experiments using trans-anethole and a 1:9 mixture of trans- and cis-anethole. In both cases, identical products were formed in comparable yields, with almost identical diastereoselectivity. This indicated that a stepwise process is involved in the reaction. In addition, a deuterium labelling experiments was performed. Introducing 5 equiv. of D2O to the reaction mixture under standard conditions resulted in the isolation of 31% yield of deuterium-labeled 3aa-D, along with the formation of 52% yield of 3aa. However, treating 3aa with D2O under the standard conditions did not lead to the detection of 3aa-D. Furthermore, when the reaction was performed in CD3CN with K3PO4 as the base, no deuterium was observed on the product. Considering the rather hydridic nature of the proton on substrate 1aa and radical polarity-match effect, we concluded that a radical polar crossover-enabled protonation might be operational in this cascade process. Moreover, we further measured the oxidation potentials of the substrates. Generally, all the arylsulfonyl acetates exhibited good compatibility with the redox potential of the photocatalyst, ranging from 0.4 V to 0.73 V, which is considerably lower than that of alkenes (1.07–1.89 V). It is worth mentioning that all the oxidation potentials for the arylsulfonyl acetates were determined in the presence of a base. This aligns with their effective fluorescence quenching observed in experiments involving a base, which indicated a deprotonation and oxidation sequence is involved. Interestingly, the dialkylamino-substituted phenyl alkene exhibited an oxidation potential of 0.56 V, even lower than that of compound 1aa. Further fluorescence quenching experiments indicated that these two alkenes are much more efficient quenchers of the photocatalyst (see Supporting information). These results suggest that different mechanisms might be operating in the cascade transformation depending on the olefinic partner (Fig. S12 in Supporting information).

    Scheme 4

    Scheme 4.  Mechanism investigation.

    Drawing upon the control experiments and prior literatures [35-37], a plausible mechanism for the majority of the difunctionalization was proposed, as depicted in Scheme 5. And it was further confirmed by density functional theory (DFT) calculations. Initially, the photocatalyst (PC) is irradiated by blue light to its excited state, which undergoes a single electron transfer event with 1a in the presence of base to afford alkyl radical . Electrophilic radical addition to the double bond of alkene forms benzyl radical Int1. Subsequently, radical Int1 undergoes ipso-radical addition to the aryl ring via a π-π stacked transition state TS1 with a Gibbs free energy barrier of 14.6 kcal/mol, delivering the spiroradical intermediate Int2. Fragmentation, accompanied by the extrusion of SO2 via a stepwise sequence involving TS2 and TS3, produces alkyl radical . The exergonicity of this transition was computed to be −20.4 kcal/mol. The transient radical Int4 is then reduced via another SET event from PC-1 followed by protonation to give the desired alkene 3, concomitantly regenerating PC to complete the photo-redox cycle. For the dialkylamino-substituted alkenes, which represent some of the electron-rich alkenes with relatively low oxidation potentials, the alkene activation mode might be involved, and a detailed discussion of the reaction mechanism is proposed in Fig. S12 (Supporting information).

    Scheme 5

    Scheme 5.  Proposed mechanism.

    In conclusion, we have unveiled a novel oxidative bifunctional reagent activation mode for the alkylarylation of alkenes under photoredox conditions using arylsulfonyl acetate as the bifunctional reagent. This metal-free radical process enables the simultaneous incorporation of carboxylate-bearing alkyl groups and (hetero)aryl rings into a wide range of olefins, thereby facilitating the synthesis of a diverse library of synthetically valuable γ,γ-diarylester derivatives. This method features mild reaction conditions, high atom- and step-economy, excellent functional group compatibility and great structural diversity. Given the current easy availability of arylsulfonylacetate bifunctional reagents, along with the ubiquity of alkenes as feedstock substrates, we anticipate this method would serve as a highly enabling platform for research endeavors aimed at synthesizing synthetic useful γ,γ-diarylester and γ-arylesters in a single operation. The success of this strategy utilizing bifunctional reagents for the difunctionalization of alkenes is expected to stimulate further investigations into this concept.

    The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

    Chonglong He: Investigation, Data curation. Yulong Wang: Methodology, Investigation. Quan-Xin Li: Software, Investigation. Zichen Yan: Methodology, Formal analysis, Data curation. Keyuan Zhang: Methodology, Data curation. Shao-Fei Ni: Writing – review & editing, Writing – original draft, Formal analysis. Xin-Hua Duan: Writing – original draft, Supervision, Project administration. Le Liu: Writing – review & editing, Writing – original draft, Supervision, Formal analysis, Conceptualization.

    L. Liu thanks the National Natural Science Foundation of China (No. 21901199), National Training Program of Innovation and Entrepreneurship for Undergraduates (No. S202310698011), and Xi’an Jiaotong University (No. 7121192002) for financial support. We thank Prof. Li-Na Guo and Miss Kehui Wang for vaulable discussions. We thank Dr. Lu Bai and Dr. Chao Feng at Instrument Analysis Center of Xi’an Jiaotong University for HRMS and NMR analysis.

    Supplementary material associated with this article can be found, in the online version, at doi:10.1016/j.cclet.2024.110253.


    1. [1]

      R.N. McDonald, G. Liu, S.S. Stahl, Chem. Rev. 111 (2011) 2981–3019. doi: 10.1021/cr100371y

    2. [2]

      C. Zhang, C. Tang, N. Jiao, Chem. Soc. Rev. 41 (2012) 3464. doi: 10.1039/c2cs15323h

    3. [3]

      X.H. Ouyang, R. Song, M. Hu, Y. Yang, J.H. Li, Angew. Chem. Int. Ed. 55 (2016) 3187–3191. doi: 10.1002/anie.201511624

    4. [4]

      X. Wang, Y.F. Han, X.H. Ouyang, R. Song, J.H. Li, Chem. Commun. 55 (2019) 14637–14640. doi: 10.1039/c9cc07494e

    5. [5]

      D.C. Blakemore, L. Castro, I. Churcher, et al., Nat. Chem. 10 (2018) 383–394. doi: 10.1038/s41557-018-0021-z

    6. [6]

      K.R. Campos, P.J. Coleman, J.C. Alvarez, et al., Science 363 (2019) eaat0805. doi: 10.1126/science.aat0805

    7. [7]

      B. Trost, Science 254 (1991) 1471–1477. doi: 10.1126/science.1962206

    8. [8]

      H.M. Huang, P. Bellotti, J. Ma, T. Dalton, F. Glorius, Nat. Rev. Chem. 5 (2021) 301–321. doi: 10.1038/s41570-021-00266-5

    9. [9]

      E. Piers, Pure Appl. Chem. 60 (1988) 107–114. doi: 10.1351/pac198860010107

    10. [10]

      X. Yu, Q.Y. Meng, C.G. Daniliuc, A. Studer, J. Am. Chem. Soc. 144 (2022) 7072–7079. doi: 10.1021/jacs.2c01735

    11. [11]

      Z. Li, L. Jiao, Y. Sun, et al., Angew. Chem. Int. Ed. 59 (2020) 7266–7270. doi: 10.1002/anie.202001262

    12. [12]

      A. Link, C. Fischer, C. Sparr, Angew. Chem. Int. Ed. 54 (2015) 12163–12166. doi: 10.1002/anie.201505414

    13. [13]

      J.E. Erchinger, R. Hoogesteger, R. Laskar, et al., J. Am. Chem. Soc. 145 (2023) 2364–2374. doi: 10.1021/jacs.2c11295

    14. [14]

      L.L. Liao, G.M. Cao, Y.X. Jiang, et al., J. Am. Chem. Soc. 143 (2021) 2812–2821. doi: 10.1021/jacs.0c11896

    15. [15]

      H. Li, H. Neumann, M. Beller, X.F. Wu, Angew. Chem. Int. Ed. 53 (2014) 3183–3186. doi: 10.1002/anie.201311198

    16. [16]

      J.J. Farndon, T.A. Young, J.F. Bower, J. Am. Chem. Soc. 140 (2018) 17846–17850. doi: 10.1021/jacs.8b10485

    17. [17]

      D. Wu, N.A. Fohn, J.W. Bode, Angew. Chem. Int. Ed. 58 (2019) 11058–11062. doi: 10.1002/anie.201904576

    18. [18]

      X.Y. Yu, Q.Q. Zhao, J. Chen, W.J. Xiao, J.R. Chen, Acc. Chem. Res. 53 (2020) 1066–1083. doi: 10.1021/acs.accounts.0c00090

    19. [19]

      Y. Yuan, A. Lei, Acc. Chem. Res. 52 (2019) 3309–3324. doi: 10.1021/acs.accounts.9b00512

    20. [20]

      P. Xiong, H.C. Xu, Acc. Chem. Res. 52 (2019) 3339–3350. doi: 10.1021/acs.accounts.9b00472

    21. [21]

      K.L. Skubi, T.R. Blum, T.P. Yoon, Chem. Rev. 116 (2016) 10035–10074. doi: 10.1021/acs.chemrev.6b00018

    22. [22]

      J.C. Tellis, C.B. Kelly, D.N. Primer, et al., Acc. Chem. Res. 49 (2016) 1429–1439. doi: 10.1021/acs.accounts.6b00214

    23. [23]

      R.C. McAtee, E.J. McClain, C.R.J. Stephenson, Trends Chem. 1 (2019) 111–125. doi: 10.1016/j.trechm.2019.01.008

    24. [24]

      W. Lee, I. Park, S. Hong, Sci. China Chem 66 (2023) 1688–1700. doi: 10.1007/s11426-023-1589-7

    25. [25]

      M. Kim, Y. Koo, S. Hong, Acc. Chem. Res. 55 (2022) 3043–3056. doi: 10.1021/acs.accounts.2c00530

    26. [26]

      B. Lu, Z. Zhang, M. Jiang, et al., Angew. Chem. Int. Ed. 62 (2023) e202309460. doi: 10.1002/anie.202309460

    27. [27]

      S. Ohmura, K. Katagiri, H. Kato, et al., J. Am. Chem. Soc. 145 (2023) 15054–15060. doi: 10.1021/jacs.3c04010

    28. [28]

      T. Horibe, S. Ohmura, K. Ishihara, J. Am. Chem. Soc. 141 (2019) 1877–1881. doi: 10.1021/jacs.8b12827

    29. [29]

      B.G. Cai, C. Empel, W.Z. Yao, R.M. Koenigs, J. Xuan, Angew. Chem. Int. Ed. 62 (2023) e202312031. doi: 10.1002/anie.202312031

    30. [30]

      H. Fang, C. Empel, L. Atodiresei, R.M. Koenigs, ACS Catal. 13 (2023) 6445–6451. doi: 10.1021/acscatal.3c00938

    31. [31]

      Z.M. Chen, X.M. Zhang, Y.Q. Tu, Chem. Soc. Rev. 44 (2015) 5220–5245. doi: 10.1039/C4CS00467A

    32. [32]

      X. Wu, C. Zhu, Acc. Chem. Res. (2020) 1620–1636. doi: 10.1021/acs.accounts.0c00306

    33. [33]

      X. Wu, Z. Ma, T. Feng, C. Zhu, Chem. Soc. Rev. 50 (2021) 11577–11613. doi: 10.1039/d1cs00529d

    34. [34]

      A.R. Allen, E.A. Noten, C.R.J. Stephenson, Chem. Rev. 122 (2022) 2695–2751. doi: 10.1021/acs.chemrev.1c00388

    35. [35]

      T.M. Monos, R.C. McAtee, C.R.J. Stephenson, Science 361 (2018) 1369–1373. doi: 10.1126/science.aat2117

    36. [36]

      A.K. Allen, J. Poon, R.C. McAtee, et al., ACS Catal. 12 (2022) 8511–8526. doi: 10.1021/acscatal.2c02577

    37. [37]

      E.A. Noten, R.C. McAtee, C.R.J. Stephenson, Chem. Sci. 13 (2022) 6942–6949. doi: 10.1039/d2sc01228f

    38. [38]

      J. Yu, Z. Wu, C. Zhu, Angew. Chem. Int. Ed. 57 (2018) 17156–17160. doi: 10.1002/anie.201811346

    39. [39]

      M. Wang, H. Zhang, J. Liu, X. Wu, C. Zhu, Angew. Chem. Int. Ed. 58 (2019) 17646–17650. doi: 10.1002/anie.201910514

    40. [40]

      X. Zhang, X. Wu, Y. Chen, C. Zhu, Green Chem. 25 (2023) 4234–4238. doi: 10.1039/d3gc01040f

    41. [41]

      J. Yu, X. Zhang, X. Wu, et al., Chem 25 (2022) 472–482. doi: 10.1016/j.chempr.2022.10.020

    42. [42]

      H. Zhang, M. Wang, X. Wu, C. Zhu, Angew. Chem. Int. Ed. 60 (2021) 3714–3719. doi: 10.1002/anie.202013089

    43. [43]

      J. Wang, X. Wu, Z. Cao, et al., Adv. Sci. 11 (2024) 2309022. doi: 10.1002/advs.202309022

    44. [44]

      Y. Wei, H. Zhang, X. Wu, C. Zhu, Angew. Chem. Int. Ed. 60 (2021) 20215–20219. doi: 10.1002/anie.202106145

    45. [45]

      T. Niu, J. Liu, X. Wu, C. Zhu, Chin. J. Chem. 38 (2020) 803–806. doi: 10.1002/cjoc.202000088

    46. [46]

      Y. Chen, X. Wu, S. Yang, C. Zhu, Angew. Chem. Int. Ed. 61 (2022) e202201027. doi: 10.1002/anie.202201027

    47. [47]

      C. Hervieu, M.S. Kirillova, Y. Hu, et al., Nat. Chem. 16 (2024) 607–614. doi: 10.1038/s41557-023-01414-8

    48. [48]

      C. Hervieu, M.S. Kirillova, T. Suárez, et al., Nat. Chem. 13 (2021) 327–334. doi: 10.1038/s41557-021-00668-4

    49. [49]

      C. He, M. Wang, Y. Wang, et al., Sci. China Chem. 67 (2024) 2022–2028. doi: 10.1007/s11426-023-1930-6

    50. [50]

      Y. Wang, S. Shen, C. He, et al., Chem. Sci. 14 (2023) 6663–6668. doi: 10.1039/d3sc01752d

    51. [51]

      D. Babcock, A.J. Wolfram, J.L. Barney, et al., Chem. Sci. 15 (2024) 4031–4040. doi: 10.1039/d3sc06476j

    52. [52]

      Y. Zhou, Q. Jiang, Y. Cheng, et al., Org. Lett. 26 (2024) 2656–2661. doi: 10.1021/acs.orglett.4c00830

    53. [53]

      C. He, K. Zhang, D.N. Wang, et al., Org. Lett. 24 (2022) 2767–2771. doi: 10.1021/acs.orglett.2c00875

    54. [54]

      K. Zhang, Y. Wang, C. He, et al., Org. Chem. Front. 9 (2022) 5599–5605. doi: 10.1039/d2qo01035f

    55. [55]

      C. He, M. Wang, B. Dong, et al., Green Synth. Catal. 5 (2024) 117–212. doi: 10.1016/j.gresc.2022.12.005

    56. [56]

      Y. Zhou, L. Zhao, M. Hu, X.H. Duan, L. Liu, Org. Lett. 25 (2023) 5268–5272. doi: 10.1021/acs.orglett.3c01787

    57. [57]

      Y. Zheng, Y. You, Q. Shen, et al., Org. Chem. Front. 7 (2020) 2069–2074. doi: 10.1039/d0qo00497a

    58. [58]

      S. Kc, R.K. Dhungana, N. Khanal, R. Giri, Angew. Chem. Int. Ed. 59 (2020) 8047–8051. doi: 10.1002/anie.201913435

    59. [59]

      N. Katta, Q.Q. Zhao, T. Mandal, O. Reiser, ACS Catal. 12 (2022) 14398–14407. doi: 10.1021/acscatal.2c04736

    60. [60]

      T.V.T. Nguyen, A. Bossonnet, M.D. Wodrich, J. Waser, J. Am. Chem. Soc. 145 (2023) 25411–25421. doi: 10.1021/jacs.3c09789

    61. [61]

      M. Riener, D.A. Nicewicz, Chem. Sci. 4 (2013) 2625. doi: 10.1039/c3sc50643f

    62. [62]

      D. Nicewicz, H. Roth, N. Romero, Synlett 27 (2016) 714–723.

    63. [63]

      Y. Sakakibara, K. Murakami, ACS Catal. 12 (2022) 1857–1878. doi: 10.1021/acscatal.1c05318

    64. [64]

      K. Kwon, R.T. Simons, M. Nandakumar, J.L. Roizen, Chem. Rev. 122 (2021) 2353–2428.

    65. [65]

      T.M. Nguyen, D.A. Nicewicz, J. Am. Chem. Soc. 135 (2013) 9588–9591. doi: 10.1021/ja4031616

    66. [66]

      A.J. Perkowski, D.A. Nicewicz, J. Am. Chem. Soc. 135 (2013) 10334–10337. doi: 10.1021/ja4057294

    67. [67]

      T.M. Nguyen, N. Manohar, D.A. Nicewicz, Angew. Chem. Int. Ed. 53 (2014) 6198–6201. doi: 10.1002/anie.201402443

    68. [68]

      P.D. Morse, D.A. Nicewicz, Chem. Sci. 6 (2015) 270–274. doi: 10.1039/C4SC02331E

    69. [69]

      K.A. Margrey, D.A. Nicewicz, Acc. Chem. Res. 49 (2016) 1997–2006. doi: 10.1021/acs.accounts.6b00304

    70. [70]

      J.D. Griffin, C.L. Cavanaugh, D.A. Nicewicz, Angew. Chem. Int. Ed. 56 (2017) 2097–2100. doi: 10.1002/anie.201610722

    71. [71]

      T. Chen, M.X. Li, Liu J.Q, Cryst. Growth Des. 18 (2018) 2765–2783. doi: 10.1021/acs.cgd.7b01503

  • Figure 1  Previous work on sulfonyl bifunctional reagent mediated difunctionalization of alkenes and our working hypothesis.

    Figure 2  Preliminary investigation and proof of the concept.

    Scheme 1  Variation of the arylsulfonylacetates: reactions were carried out on 0.2 mmol scale and isolated yields are given. dr was determined by crude NMR spectroscopy, for details see Supporting information.

    Scheme 2  Variation of the electron-rich alkenes: reactions were carried out on 0.2 mmol scale and isolated yields are given. dr was determined by crude NMR spectroscopy, for details see Supporting information.

    Scheme 3  Variation of the electron-rich alkenes: reactions were carried out on 0.2 mmol scale and isolated yields are given. dr was determined by crude NMR spectroscopy, for details see supporting information.

    Scheme 4  Mechanism investigation.

    Scheme 5  Proposed mechanism.

  • 加载中
计量
  • PDF下载量:  2
  • 文章访问数:  170
  • HTML全文浏览量:  9
文章相关
  • 发布日期:  2025-05-15
  • 收稿日期:  2024-05-20
  • 接受日期:  2024-07-15
  • 修回日期:  2024-07-08
  • 网络出版日期:  2024-07-15
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

/

返回文章