Substrate Scope. The scope of the reaction under optimized
conditions was investigated. Various aryl ethylenes were tested as
nucleophiles, and the corresponding ionic insertion products delivered
moderate to excellent reaction yields (Figure 2A). For substrates with
substitutions on the para position, electron-deficient substrates
usually gave higher reaction yields, which were primarily ascribed to
their relatively low reactivity toward the newly generated benzyl
chloride. Further, excellent functional-group tolerance was observed
(3ac–3ae ). For example, the reaction with substrates bearing
–OCF3 and –SCF3 substitutions, often
embedded in pharmaceutical and agrochemical products, afforded the
corresponding products in excellent yields (3af and3ag ). The –Br (3ae ), –I (3ap ), alkyne
(3aj ), and boronate (3an and 3ao )
substitutions, which may have been incompatible in previous
transition-metal catalyzed processes, were all well tolerated, thereby
providing handles for further derivation. Some reactive functional
groups in conventional synthetic methods, including aryl esters,
aldehydes, carboxyl, and benzyl chloride groups, were well accommodated
(3ak–3am and 3aq ). Furthermore, functional groups atmeta or ortho positions of the phenyl ring were acceptable
(3ar–3av ) and showed a similar electronic preference compared
with the para substitutions.
Subsequently, the feasibilities of benzyl chlorides 2 were
explored, beginning with the examination of the electronic effects of
the substituents (Figure 2B). After comparing 3ba and3be , we found that the electron-donating substituents at thepara position increased the reaction yield. The electron-donating
conjugation effect on the aromatic ring decreased the activation energy
of the heterolytic cleavage of the C–Cl bond. The substrates featuring
halogen atom, such as –F, –Cl, –Br, and –I, at the phenyl ring
(3ba–3bd ), reacted with styrene to afford products in good
yields.
Further, steric effect was analyzed by changing the position of the
methyl group. This showed that the steric effect of an aryl ring has a
slight influence on the proposed reaction. Theortho -methyl-substituted substrate provided the product in 92%
yield (3bg ). In addition, long alkyl chains provided products
(3bi and 3bj ) in moderate yields.
Figure 2. Insertion of alkenes into C–Cl bonds.aReaction conditions: 1 (0.6 mmol) and2 (1.8 mmol) at 80 °C–120 °C for 1–5 h under air.
R1 = 1-phenylethyl, R2 = benzhydryl,
Ar1 = Ph, Ar2 =
4-CF3-Ph, and Ar3 = 4-MeO-Ph. SI
provides the detailed conditions for each substrate.bA gram scale reaction was performed under air using1 (10.0 mmol).
Diphenylmethane has highly important applications in the synthesis of
bioactive compounds;72 thus, a series of diarylmethyl
chlorides as substrates were investigated. Owing to the stability of a
carbocation, excellent compatibility of these substrates was observed
for the proposed reaction. Both electron-withdrawing (3bl ) and
electron-donating (3br ) substituents underwent insertion
reactions with good yields. A heteroaromatic thienyl group
(3bq ) was also well tolerated in the reaction. In addition to
benzyl chlorides, primary and secondary allyl chlorides were verified as
suitable coupling partners (3bs and 3bt ). Furthermore,
the scalability of the proposed reaction was evaluated by performing a
reaction with 10.0 mmol of 1b . An improved yield of 95% was
obtained compared with small-scale trials.
Figure 3. Insertion into C–Br and C–I bonds.aAr2 = 4-CF3-Ph. SI
provides the detailed conditions for each substrate.
We extended the reaction scope of benzyl bromides and iodides (Figure
3). The insertion reaction was explored using styrene as the
nucleophile. Alkyl bromide 5aa was successfully obtained with a
satisfactory yield under standard reaction conditions. Introduction of
an electron-withdrawing substituent decreased the reaction yield
(5ab ), which can be attributed to an increased difficulty for
the heterolytic cleavage of the C–Br bond. The replacement of methyl
with ethyl provided 5ac but with a slight reduction in yield.
When bromodiphenylmethane was used as the electrophile, the yield
(5ad ) was improved because C–Br bonds are easier to cleave
owing to the conjugation effect. Halogens on the alkyl chains were well
tolerated (5ae and 5af ), and their insertion into
unactivated carbon–halogen bonds was not observed. Further, benzyl
iodides were suitable substrates in the insertion reaction
(5ag ), for which the insertion process was
difficult.73
One-pot Reaction. To demonstrate the practical application of
the proposed reaction, a series of one-pot transformation sequences was
investigated (Figure 4). The C–C, C–O, C–S, C–N, and C–B
difunctionalizations of alkenes were realized through this reaction
(6aa–6af ). Reductive coupling (6ag ) and deprotonation
(6ah ) products were obtained after treating 3bk with a
reductant and base, respectively.
Figure 4. Carbohalogenative 1,2-difunctionalization of double
bonds.
Mechanistic Investigations. To shed light on the reaction
mechanism, some control experiments were conducted. The reactions of1b and 2k were performed in the presence of a radical
inhibitor, 2,6-bis(1,1-dimethylethyl)-4-methylphenol (BHT). The
insertion product was obtained in 86% yield, indicating that the
proposed reaction is unlikely to proceed via a radical pathway
(Figure 5A). The increased yield (compared with Figure 2B) may be
explained by the inhibition of the radical polymerization of alkene by
BHT. Subsequently, a cation-exchange experiment was designed (Figure
5B), where benzhydrol was introduced as a cation donor. As expected, the
normal insertion product 3ab was obtained in 21% yield, and3bk that was generated using benzhydrol as the benzhydryl
cation donor was obtained in 55% yield. The formation of these products
implied that the proposed reaction proceeded through an ionic
mode.74 In addition, the intermolecular competition
between styrene 1b and its dideuterated analogd8-1b showed a kinetic isotope effect
kH/kD = 1.00 (Figure 5C); however, a
secondary isotope effect was not observed, suggesting that C–C and
C–Cl bond formations were not involved in the rate-determining
step.75 In addition, solvent quenching experiments
were designed to elaborate the influence of the aggregated states on
such ionic insertion processes. As shown in Figure 5D, while a trace
amount (three equivalents) of commonly used solvent was introduced into
the reaction system, the desired transformation could be interrupted,
which may be ascribed to the disconnection of the aggregated state of
reactants by the solvent molecules. Notably, even nonpolar, weak
coordinating solvents such as mesitylene were competent blockers.
Possibly, there are three reasons why the solvent prevents the above
reaction. First, the cage effect of the solvent can reduce effective
collision between substrates and increase the activation energy of the
insertion process (restricted intermolecular collision, RIC). Next,
induction and coordination effects will reduce the transient effective
charge produced by the system, which is not conducive to the reaction
(charge dispersion, CD). Finally, the solvent effect may quench the
intermediates or transition states produced by the reaction, inhibiting
the reaction (intermediate quenching or transition state quenching, IQ
or TSQ). To confirm the above speculation, density functional theory
(DFT) calculations and symmetry-adapted perturbation theory (SAPT)
analysis were conducted.
Figure 5. Preliminary mechanistic experiments.
DFT calculations provide mechanistic insights into the abovementioned
insertion reactions of alkenes to C–X bonds. For C–Cl bond insertion
(Figure 6A), the benzyl chloride substrate 2k heterolytically
dissociated to carbocation and chloride anion before styrene attacked
the so-generated carbocation via a chloride anion bound
transition state (TS). Such an SN1-like reaction pathway
was exergonic (3.5 kcal/mol) with an overall free-energy barrier of 26.2
kcal/mol. Consistent with the mechanistic studies, the radical mechanism
was ruled out because the homolysis of the C–Cl bond in 2kseemed unfavorable as it required higher energy (by >20
kcal/mol) than that in the heterolysis of the C–Cl bond (Figure S4).
Moreover, the SN2 mechanism in which a styrene directly
attacks the C(sp3) atom of 2k to release a
chloride anion was ruled out because it encounters a 5 kcal/mol higher
energy barrier than that in the SN1-like mechanism.
Figure 6. DFT calculations and SAPT analysis.
SAPT analysis evaluates the intermolecular interaction energy between
the fragments of the obtained TS in the C–Cl bond insertion reaction
(Figure 6B). The results from this energy decomposition calculation
clearly show a strong electrostatic stabilization interaction of
~77.2 kcal/mol between the chloride anion and its
bounded benzylic counterpart. Interestingly, almost half of the
electrostatic stabilization (i.e., ~35.6 kcal/mol) was
found between styrene and the remaining fragments in the TS.
Furthermore, second-order perturbation analysis based on natural bond
orbital calculations showed that a strong orbital interaction existed
between the pz orbital of the carbocation and p bonding orbital of the
olefin fragment of styrene. The calculated second-order perturbation
energy was 108.2 kcal/mol. Therefore, the temporary anchorage of the
chloride anion provided a channel for the nucleophilic attack of an
olefin and facilitated its subsequent combination with a newly generated
carbocation, providing an almost synergetic reaction mode while
suppressing the possible side reactions associated with the carbocation.
Therefore, the insertion reaction of alkenes into C–X bonds was
controlled by intrinsic factors such as strong orbital interactions and
electrostatic stabilizations between the fragments involved in the
reactions. However, if the reaction is performed under a non-aggregated
state, the abovementioned TSs or intermediates would be excluded owing
to their inevitable and much stronger interactions with solvent.
Discussion
In summary, a new strategy relying on
aggregated state has enabled the
catalyst-free insertion of alkenes into C–X bonds. The
difunctionalization of alkene can be realized without using any
transition metal catalysts in the aggregated state. Practically, alkenes
were inserted into C–X (X = Cl, Br, and I) bonds via ionic mode.
The method exhibits excellent atom and step economy and environmental
sustainability. Moreover, its practicality is highlighted by a broad
substrate scope, excellent functional-group tolerance, and extremely
simple operation. The method tolerates active functional groups such as
CHO, B(OH)2, CO2H,
Me2SiH, alkynes, and CO2Me, which are
often incompatible in transition-metal catalyzed or Lewis-acid catalyzed
reactions. This work is the first attempt to apply aggregate science to
the field of synthetic chemistry, which further expands the application
reaction of aggregation strategy as well as provides a new idea for
designing new reactions in organic chemistry.
Materials and Methods
General procedure for probing the scope of insertion of alkenes into
carbon-halogen bonds. To an oven-dried, 10 mL Teflon-lined screw-capped
Pyrex test tube was added aryl ethylenes (0.6 mmol) and benzyl halides
(1.8 mmol) without argon protection. A magnetic stir bar were added to
the tube carefully and the mixture was stirred slowly for 5 min at room
temperature. Then increase the temperature to 100 °C and continue to
stir for extra 5 h. After being cooled down to room temperature, it was
purified by flash chromatography on silica gel to afford pure product.
DFT calculations were performed using the ORCA program and SAPT analysis
was performed using the PSI4 code at the SAPT2+/aug-cc-pVDZ level.
Acknowledgements
This work was supported by the National Natural Science Foundation of
China (91940305, 21933009, 81874181, 22271195, 21871284), Natural
Science Foundation of Fujian Province (2021J01525), Major Scientific and
Technological Special Project for ”Significant New Drugs Creation”
(2019ZX09301158), Emerging Frontier Program of Hospital Development
Centre (SHDC12018107), Shanghai Science and Technology Development Fund
from Central Leading Local Government (YDZX20223100001004), and Shanghai
Municipal Health Commission/Shanghai Municipal Administration of
Traditional Chinese Medicine (ZY(2021-2023)-0501). The authors would
like to thank Shiyanjia Lab (www.shiyanjia.com) for the language editing
service.
Conflict of Interests
The authors declare no conflict of interests.
Supporting Information
Supporting Information is available from the Wiley Online Library or
from the author.
Received: ((will be filled in by the editorial staff))
Revised: ((will be filled in by the editorial staff))
Published online: ((will be filled in by the editorial staff))
References
1. Luo, J.; Xie, Z.; Lam, J. W. Y.; Cheng, L.; Chen, H.; Qiu, C.; Kwok,
H. S.; Zhan, X.; Liu, Y.; Zhu, D.; Tang, B. Z. Aggregation-induced
emission of 1-methyl-1,2,3,4,5-pentaphenylsilole. Chem. Commun.1740−1741 (2001).
2. Yan, D. Y.; Wu, Q.; Wang, D.; Tang, B. Z., Innovative Synthetic
Procedures for Luminogens Showing Aggregation-Induced Emission.Angew. Chem. Int. Ed. 60, 15724−15742 (2021).
3. Liu, H.; Xiong, L. H.; Kwok, R. T. K.; He, X.; Lam, J. W. Y.; Tang,
B. Z., AIE Bioconjugates for Biomedical Applications. Adv. Opt.
Mater 8, 2000162 (2020).
4. Jiang, G.; Yu, J.; Wang, J.; Tang, B. Z., Ion−π interactions for
constructing organic luminescent materials. Aggregate 3,e285 (2022).
5. Kang, M.; Zhang, Z.; Song, N.; Li, M.; Sun, P.; Chen, X.; Wang, D.;
Tang, B. Z., Aggregation-enhanced theranostics: AIE sparkles in
biomedical field. Aggregate 1, 80−106 (2020).
6. Liu, H.; Kumar, S. K.; Douglas, J. F. Self-Assembly-Induced Protein
Crystallization. Phys. Rev. Lett. 103, 018101 (2009).
7. Gutberlet, A.; Schwaab, G.; Birer, O.; Masia, M.; Kaczmarek, A.;
Forbert, H.; Havenith, M.; Marx, D. Aggregation-Induced Dissociation of
HCl(H2O)4 Below 1 K: The Smallest
Droplet of Acid. Science 324, 1545−1548 (2009).
8. Forbert, H.; Masia, M.; Kaczmarek-Kedziera, A.; Nair, N. N.; Marx, D.
Aggregation-Induced Chemical Reactions: Acid Dissociation in Growing
Water Clusters. J. Am. Chem.
Soc. 133, 4062−4072 (2011).
9. Cintas, P.; Tabasso, S.; Veselov, V. V.; Cravotto, G. Alternative
reaction conditions: Enabling technologies in solvent-free protocols.Curr. Opin. Green. Sust. Chem. 21, 44−49 (2020).
10. Avila-Ortiz, C. G., Juaristi, E. Novel Methodologies for Chemical
Activation in Organic Synthesis under Solvent-Free Reaction Conditions.Molecules 25, 3579 (2020).
11. Beale, T. M.; Chudzinski, M. G.; Sarwar, M. G.; Taylor, M. S.
Halogen bonding in solution: thermodynamics and applications.Chem. Soc. Rev. 42 , 1667−1680 (2013).
12. Cavallo, G.; Metrangolo, P.; Milani, R.; Pilati, T.; Priimagi, A.;
Resnati, G.; Terraneo, G. The Halogen Bond. Chem. Rev.116 , 2478−2601 (2016).
13. Saikia, I.; Borah, A. J.; Phukan, P. Use of Bromine and
Bromo-Organic Compounds in Organic Synthesis. Chem. Rev.116 , 6837−7042 (2016).
14. Biffis, A.; Centomo, P.; Del Zotto, A.; Zeccal, M. Pd Metal
Catalysts for Cross-Couplings and Related Reactions in the 21st Century:
A Critical Review. Chem. Rev. 118 , 2249−2295 (2018).
15. Juhasz, K.; Magyar, A.; Hell, Z. Transition-Metal-Catalyzed
Cross-Coupling Reactions of Grignard Reagents. Synthesis53 , 983−1002 (2021).
16. Hernandes, M. Z.; Cavalcanti, S. M. T.; Moreira, D. R. M.; de
Azevedo, W. F.; Leite, A. C. L. Halogen Atoms in the Modern Medicinal
Chemistry: Hints for the Drug Design. Curr. Drug Targets11 , 303−314 (2010).
17. Gribble, G. W. A recent survey of naturally occurring organohalogen
compounds. Environ. Chem. 12 , 396−405 (2015).
18. Matulja, D.; Wittine, K.; Malatesti, N.; Laclef, S.; Turks, M.;
Markovic, M. K.; Ambrozic, G.; Markovic, D. Marine Natural Products with
High Anticancer Activities. Curr. Med. Chem. 27 ,
1243−1307 (2020).
19. Hohlman, R. M.; Sherman, D. H., Recent advances in hapalindole-type
cyanobacterial alkaloids: biosynthesis, synthesis, and biological
activity. Nat. Prod. Rep. 38 , 1567−1588 (2021).
20. Vigalok, A. Metal-mediated formation of carbon-halogen bonds.Chem. Eur. J. 14 , 5102−5108 (2008).
21. Vigalok, A. C-X Bond Formation ; Springer: Heidelberg,
Germany, 2010 .
22. Jiang, X.; Liu, H.; Gu, Z., Carbon-Halogen Bond Formation by the
Reductive Elimination of Pd-II Species. Asian J. Org. Chem.1 , 16−24 (2012).
23. Petrone, D. A.; Ye, J.; Lautens, M. Modern
Transition-Metal-Catalyzed Carbon-Halogen Bond Formation. Chem.
Rev. 116 , 8003−8104 (2016).
24. Marchese, A. D.; Adrianov, T.; Lautens, M. Recent Strategies for
Carbon-Halogen Bond Formation Using Nickel. Angew. Chem. Int. Ed.60 , 16750−16762 (2021).
25. Petrone, D. A.; Le, C. M.; Newman, S. G.; Lautens, M.
Pd(0)-Catalyzed Carboiodination: Early Developments and Recent
Advancements. In New Trends in Cross-Coupling: Theory and
Applications ; Colacot, T. J., Ed.; Royal Society of Chemistry:
Cambridge, 2015.
26. Jones, D. J.; Lautens, M.; McGlacken, G. P. The emergence of
Pd-mediated reversible oxidative addition in cross coupling,
carbohalogenation and carbonylation reactions. Nat. Catal.2 , 843−851 (2019).
27. Xu, T.; Cheung, C. W.; Hu, X. L. Iron- Catalyzed 1,2-Addition of
Perfluoroalkyl Iodides to Alkynes and Alkenes. Angew. Chem. Int.
Ed. 53 , 4910−4914 (2014).
28. Xu, T.; Hu, X. L. Copper-Catalyzed 1,2-Addition of alpha-Carbonyl
Iodides to Alkynes. Angew. Chem. Int. Ed. 54 , 1307−1311
(2015).
29. Jiang, X.; Liu, H.; Gu, Z., Carbon-Halogen Bond Formation by the
Reductive Elimination of Pd-II Species. Asian J. Org. Chem.1 , 16−24 (2012).
30. Newman, S. G.; Howell, J. K.; Nicolaus, N.; Lautens, M.,
Palladium-Catalyzed Carbohalogenation: Bromide to Iodide Exchange and
Domino Processes. J. Am. Chem. Soc. 133 , 14916−14919
(2011).
31. Newman, S. G.; Lautens, M., Palladium-Catalyzed Carboiodination of
Alkenes: Carbon-Carbon Bond Formation with Retention of Reactive
Functionality. J. Am. Chem. Soc. 133 , 1778−1780 (2011).
32. Petrone, D. A.; Yoon, H.; Weinstabl, H.; Lautens, M., Additive
Effects in the Palladium-Catalyzed Carboiodination of Chiral N-Allyl
Carboxamides. Angew. Chem. Int. Ed. 53 , 7908−7912
(2014).
33. Le, C. M.; Menzies, P. J. C.; Petrone, D. A.; Lautens, M.,
Synergistic Steric Effects in the Development of a Palladium-Catalyzed
Alkyne Carbohalogenation: Stereodivergent Synthesis of Vinyl Halides.Angew. Chem. Int. Ed. 54 , 254−257 (2015).
34. ) Lee, Y. H.; Morandi, B. Palladium-Catalyzed Intermolecular
Aryliodination of Internal Alkynes. Angew. Chem. Int. Ed.58 , 6444−6448 (2019).
35. Zhang, Z. M.; Xu, B.; Wu, L. Z.; Zhou, L. J.; Ji, D. T.; Liu, Y.;
Li, Z. M.; Zhang, J. L. Palladium/XuPhos-Catalyzed Enantioselective
Carboiodination of Olefin-Tethered Aryl Iodides. J. Am. Chem.
Soc. 141 , 8110−8115 (2019).
36. Chen, X.; Zhao, J.; Dong, M.; Yang, N.; Wang, J.; Zhang, Y.; Liu,
K.; Tong, X., Pd(0)-Catalyzed Asymmetric Carbohalogenation:
H-Bonding-Driven C(sp(3))-Halogen Reductive Elimination under Mild
Conditions. J. Am. Chem. Soc. 143 , 1924−1931 (2021).
37. Yoon, H.; Marchese, A. D.; Lautens, M. Carboiodination Catalyzed by
Nickel. J. Am. Chem. Soc. 140 , 10950−10954 (2018).
38. Marchese, A. D.; Lind, F.; Mahon, A. E.; Yoon, H.; Lautens, M.
Forming Benzylic Iodides via a Nickel Catalyzed Diastereoselective
Dearomative Carboiodination Reaction of Indoles. Angew. Chem. Int.
Ed. 58 , 5095−5099 (2019).
39. Marchese, A. D.; Wollenburg, M.; Mirabi, B.; Abel-Snape, X.; Whyte,
A.; Glorius, F.; Lautens, M. Nickel-Catalyzed Enantioselective Carbamoyl
Iodination: A Surrogate for Carbamoyl Iodides. ACS Catal.10 , 4780−4785 (2020).
40. Takahashi, T.; Kurahashi, T.; Matsubara, S. Nickel-Catalyzed
Intermolecular Carbobromination of Alkynes. ACS Catal.10 , 3773−3777 (2020).
41. Marchese, A. D.; Adrianov, T.; Kollen, M. F.; Mirabi, B.; Lautens,
M. Synthesis of Carbocyclic Compounds via a Nickel-Catalyzed
Carboiodination Reaction. ACS Catalysis 11 , 925−931
(2021).
42. Pintauer, T.; Matyjaszewski, K. Atom transfer radical addition and
polymerization reactions catalyzed by ppm amounts of copper complexes.Chem. Soc. Rev. 37 , 1087−1097 (2008).
43. Pintauer, T. Catalyst Regeneration in Transition-Metal-Mediated
Atom-Transfer Radical Addition (ATRA) and Cyclization (ATRC) Reactions.Eur. J. Inorg. Chem. 2010 , 2449−2460 (2010).
44. Munoz-Molina, J. M.; Belderrain, T. R.; Perez, P. J. Atom Transfer
Radical Reactions as a Tool for Olefin Functionalization - On the Way to
Practical Applications. Eur. J. Inorg. Chem. 2011 ,
3155−3164 (2011).
45. Clark, A. J. Copper Catalyzed Atom Transfer Radical Cyclization
Reactions. Eur. J. Org. Chem. 2016, 2231−2243 (2016).
46. Kharasch, M. S.; Jensen, E. V.; Urry, W. H. Addition of carbon
tetrachloride and chloroform to olefins. Science 102 ,
128−128 (1945).
47. Curran, D. P.; Chen, M. H.; Spletzer, E.; Seong, C. M.; Chang, C. T.
Atom transfer addition and annulation reactions of iodomalonates.J. Am. Chem. Soc. 111 , 8872−8878 (1989).
48. Yorimitsu, H.; Nakamura, T.; Shinokubo, H.; Oshima, K.; Omoto, K.;
Fujimoto, H. Powerful solvent effect of water in radical reaction:
Triethylborane-induced atom-transfer radical cyclization in water.J. Am. Chem. Soc. 122 , 11041−11047 (2000).
49. Hou, L. L.; Zhou, Z. Z.; Wang, D.; Zhang, Y. W.; Chen, X.; Zhou, L.
J.; Hong, Y.; Liu, W.; Hou, Y. D.; Tong, X. F. DPPF-Catalyzed
Atom-Transfer Radical Cyclization via Allylic Radical. Org. Lett.19 , 6328−6331 (2017).
50. Pu, W.; Sun, D.; Fan, W.; Pan, W.; Chai, Q.; Wang, X.; Lv, Y.
Cu-Catalyzed atom transfer radical addition reactions of alkenes with
alpha-bromoacetonitrile. Chem. Commun. 55 , 4821−4824
(2019).
51. Reiser, O. Shining Light on Copper: Unique Opportunities for
Visible-Light-Catalyzed Atom Transfer Radical Addition Reactions and
Related Processes. Acc. Chem. Res. 49 , 1990−1996 (2016).
52. Bag, D.; Kour, H.; Sawant, S. D. Photo-induced
1,2-carbohalofunctionalization of C−C multiple bonds via ATRA pathway.Org. Biomol. Chem. 18 , 8278−8293 (2020).
53. Rawner, T.; Lutsker, E.; Kaiser, C. A.; Reiser, O. The Different
Faces of Photoredox Catalysts: Visible-Light-Mediated Atom Transfer
Radical Addition (ATRA) Reactions of Perfluoroalkyl Iodides with
Styrenes and Phenylacetylenes. ACS Catal. 8 , 3950−3956
(2018).
54. Engl, S.; Reiser, O. Copper Makes the Difference: Visible
Light-Mediated Atom Transfer Radical Addition Reactions of Iodoform with
Olefins. ACS Catal. 10 , 9899−9906 (2020).
55. Kostromitin, V. S.; Zemtsov, A. A.; Kokorekin, V. A.; Levin, V. V.;
Dilman, A. D. Atom-transfer radical addition of fluoroalkyl bromides to
alkenes via a photoredox/copper catalytic system. Chem. Commun.57 , 5219−5222 (2021).
56. Miller; V. A. The Aluminum Chloride-Catalyzed Addition of t-Butyl
Chloride to Propylene. J. Am. Chem. Soc. 69 , 1764−1768
(1947).
57. Schmerling, L.; Meisinger, E. E. Condensation of Saturated Halides
with Unsaturated Compounds. X. Condensation of t-Butyl Halides with
Propene, 1-Butene and 2-Butene. J. Am. Chem. Soc. 75 ,
6217−6222 (1953).
58. Mayr, H.; Striepe, W. Scope and limitations of aliphatic
friedel-crafts alkylations-lewis acid-catalyzed addition-reactions of
alkyl chlorides to carbon carbon double-bonds. J. Org. Chem.48 , 1159−1165 (1983).
59. Mayr, H.; Schneider, R.; Irrgang, B.; Schade, C. Kinetics of the
reactions of the p-methoxy-substituted benzhydryl cation with various
alkenes and 1,3-dienes. J. Am. Chem. Soc. 112 , 4454−4459
(1990).
60. Rappopor, Z.; Gal, A. Vinylic cations from solvolysis .13. sn1 and
electrophilic addition-elimination routes in solvolysis of
alpha-bromo-4-methoxystyrenes and alpha-chloro-4-methoxystyrenes.J. Chem. Soc. Perkin Trans. 2 , 301−310 (1973).
61. Toteva, M. M.; Richard, J. P. Mechanism for nucleophilic
substitution and elimination reactions at tertiary carbon in largely
aqueous solutions: Lifetime of a simple tertiary carbocation. J.
Am. Chem. Soc. 118 , 11434−11445 (1996).
62. Fu, G. C. Transition-Metal Catalysis of Nucleophilic Substitution
Reactions: A Radical Alternative to S(N)1 and S(N)2 Processes. ACS
Cent. Sci. 3 , 692−700 (2017).
63. Beringer, F. M.; Gindler, E. M. Ion Pairs as Intermediates in
Irreversible Reactions of Electrolytes. J. Am. Chem. Soc.77 , 3200−3203 (1995).
64. Abraham, M. H. Solvent effects on the free energies of ion-pairs,
and of transition states in an SN1 and an
SN2 reaction. 11 , 5233−5236 (1970).
65. Chiappe, C.; Detert, H.; Lenoir, D.; Pomelli, C. S.; Ruasse, M. F.
Polarizability effects and dispersion interactions in Alkene-Br-2
pi-complexes. J. Am. Chem. Soc. 125 , 2864−2865 (2003).
66. Olah, G. A.; Melby, E. G., Onium ions. III. Alkylarylhalonium Ions.J. Am. Chem. Soc. 94 , 6220−6221 (1972).
67. Toteva, M. M.; Richard, J. P. Mechanism for nucleophilic
substitution and elimination reactions at tertiary carbon in largely
aqueous solutions: Lifetime of a simple tertiary carbocation. J.
Am. Chem. Soc. 118 , 11434−11445 (1996).
68. Richard, J. P.; Szymanski, P.; Williams, K. B. Solvent effects on
carbocation-nucleophile combination reactions: A comparison of
pi-nucleophilicity in aqueous and organic solvents. J. Am. Chem.
Soc. 120 , 10372−10378 (1998).
69. Gawande, M. B.; Bonifacio, V. D. B.; Luque, R.; Branco, P. S.;
Varma, R. S. Solvent-Free and Catalysts-Free Chemistry: A Benign Pathway
to Sustainability. ChemSusChem 7 , 24−44 (2014).
70. Sarkar, A.; Santra, S.; Kundu, S. K.; Hajra, A.; Zyryanov, G. V.;
Chupakhin, O. N.; Charushin, V. N.; Majee, A. A decade update on solvent
and catalyst-free neat organic reactions: a step forward towards
sustainability. Green Chem. 18 , 4475−4525 (2016).
71. Patron, F.; Adelman, S. A. Solvent cage effects and
chemical-dynamics in liquids. Chem. Phys. 152 , 121−131
(1991).
72. Silvestri, R.; Artico, M.; De Martino, G.; Ragno, R.; Massa, S.;
Loddo, R.; Murgioni, C.; Loi, A. G.; La Colla, P.; Pani, A. Synthesis,
biological evaluation, and binding mode of novel
1-2-(diarylmethoxy)ethyl -2-methyl-5-nitroimidazoles targeted at the
HIV-1 reverse transcriptase. J. Med. Chem. 45 , 1567−1576
(2002).
73. Barata-Vallejo, S.; Cooke, M. V.; Sebastian, A. Radical
Fluoroalkylation Reactions. ACS Catal . 8 , 7287−7307
(2018).
74. Yue, H.-L.; Wei, W.; Li, M.-M.; Yang, Y.-R.; Ji, J.-X., sp(3)-sp(2)
C-C Bond Formation via Bronsted Acid Trifluoromethanesulfonic
Acid-Catalyzed Direct Coupling Reaction of Alcohols and Alkenes.Adv. Syn. Catal. 353 , 3139−3145 (2011).
75. Halevi, E. A., Secondary Isotope Effects. Progress in Physical
Organic Chemistry 1 , 109−221 (1963).
Molecular aggregation affects the electronic interactions between
molecules and has emerged as a powerful tool in material science. In
this study, aggregation strategy was applied to synthetic chemistry,
ionic insertion of alkene into carbon–halogen bond can occur smoothly
in aggregated state without any catalysts. Results show that the
non-aggregated state may quench the transition state and terminate the
insertion process.
Keywords aggregation enabled alkene insertion, solvent-free and
catalyst-free, atom-and step-economy
Meng-Yao Li,1,2,4 Xiao-Mei Nong,1Han Xiao,3 Ao Gu,1 Shuyang
Zhai,1 Jiatong Li,1 Ge
Zhang,4 Ze-Jian Xue,4 Yingbin
Liu,1* Chunsen Li,3* Guo-Qiang
Lin,2,4 Chen-Guo Feng2,4*
Title Aggregation-enabled Alkene Insertion into Carbon–Halogen
Bonds
ToC figure