DOI:
10.1039/D5QO00292C
(Research Article)
Org. Chem. Front., 2025,
12, 2926-2934
Unraveling the α-effect in α-fluorinated carbanionic nucleophiles: origins and synthetic implications†
Received
11th February 2025
, Accepted 10th March 2025
First published on 11th March 2025
Abstract
The α-effect refers to the dramatically enhanced reactivity of α-nucleophiles bearing a heteroatom with an adjacent lone-pair (R–Y–X:−) compared to normal nucleophiles (R–X:−), as predicted by Brønsted-type correlation. Despite extensive research, the underlying mechanisms of this phenomenon remain debated, and previous studies have predominantly focused on O-, N-, and S-based nucleophiles, leaving carbanions—key intermediates in organic synthesis—comparatively underexplored. Here, we present an in-depth computational investigation into the intriguing influence of α-fluorine substitution on carbanion nucleophilicity. Despite fluorine's strong electron-withdrawing capability, our results reveal that α-fluorocarbanions could exhibit the α-effect when they satisfy Hamlin's two criteria originally proposed for heteroatom-based anionic α-nucleophiles. This study extends the scope of the α-effect from heteroatom-based nucleophiles to carbanionic nucleophiles, offering new insights into this fundamental chemical phenomenon.
Introduction
The α-effect in organic chemistry is a fundamental phenomenon that describes the significantly enhanced reactivity of nucleophiles possessing adjacent heteroatoms with lone pairs.1 Coined by Pearson and Edwards2 in 1962, the term refers to the downward deviation from the expected Brønsted-type correlation for normal nucleophiles (Fig. 1a, left column). The α-effect has been observed in numerous reactions, with its magnitude highly dependent on both the class of reaction and the type of α-nucleophiles involved.3 The origin of the α-effect has been a long-standing debate and remains a subject of controversy. Several explanations have been proposed regarding the origin of the α-effect, including ground state destabilization, transition state stabilization, thermodynamic product stability, and external solvent effects.3g,4 Recently, Hamlin et al.5 identified two key criteria that α-nucleophiles need to fulfill to exhibit the α-effect (Fig. 1a, right column): (1) a small HOMO lobe on the nucleophilic center to minimize Pauli repulsion with the substrate and (2) a sufficiently high-energy HOMO to engage in strong orbital interactions with the substrate. These criteria have been proposed for heteroatom-based anionic nucleophiles, including O- and N-based α-nucleophiles. However, to the best of our knowledge, their implications for carbanions—fundamental intermediates in organic synthesis—remain notably underexplored.
 |
| Fig. 1 (a) Definition and criteria (proposed by Hamlin et al.) for the α-effect. (b) Variable effects of α-fluorine on carbanions. | |
Fluorocarbon anions play a crucial role in nucleophilic fluoroalkylations.6 As Schlosser's seminal review highlights,7 “fluorine as a substituent always good for a surprise, but often completely unpredictable”. Variable effects of α-fluorine substitution on the thermal stability,8 Brønsted basicity,9 and nucleophilicity10 of carbanions have been observed (Fig. 1b). For instance, while the fluoromalonate anion was traditionally perceived as a weaker nucleophile compared to malonate in reactions with alkyl bromides,11 Shibata and colleagues12 highlighted that fluorodi(benzenesulfonyl)methane (FBSM: (PhSO2)2CFH)13 participates smoothly in palladium-catalyzed allylic methylation reactions, achieving remarkable enantioselectivity and excellent yields, contrasting sharply with the trace yields obtained with di(benzenesulfonyl)methane (BSM: (PhSO2)2CH2). Furthermore, recent studies by Hu and colleagues14 demonstrated that FBSM reacts with enones significantly faster than BSM. They also found that the nucleophilic addition of (PhSO2)2CFLi to benzaldehyde produces the corresponding alcohol in high yields, whereas BSM or (PhSO2)2CClLi yields no adducts under similar conditions. Moreover, the conventional view that α-fluorine generally diminishes carbanion nucleophilicity due to its strong electron-withdrawing effects has been challenged by studies from Prakash and Mayr,10a which suggest that α-fluorine substitution can, in fact, enhance the nucleophilicity of certain carbanions.
In this study, we present an in-depth computational investigation into the intriguing effects of α-fluorine substitution on the nucleophilicity of carbanions,10a aiming to elucidate why it can enhance the nucleophilicity of specific carbanions. The results not only extend the scope of the α-effect from heteroatom-containing nucleophiles to carbanionic nucleophiles but also provide valuable insights into the intriguing role of α-fluorine substitution in nucleophilic fluoroalkylation.
Results and discussion
According to the nucleophilicity measurement experiment adopted by Mayr and Prakash et al.,10a the nucleophilic addition of a series of K[(PhSO2)2CY] to quinone was employed as a model reaction to investigate whether there is an α-effect of fluorocarbanions (Fig. 2a). The α-group Y can be divided into two series: one series bears lone pairs such as –F, –OMe, and –NHMe; the corresponding carbanions belong to “α-nucleophiles”. The other series has no lone-pair in the α-atom, and the corresponding carbanions are “normal nucleophiles”. Fig. 2a presents a Brønsted-type correlation analysis for the nucleophilic additions, showing a strong correlation between the free energy barriers and the basicity of normal nucleophiles (R2 = 0.94). α-Nucleophiles with substituents Y = OMe and NHMe still roughly follow the Brønsted-type correlation. However, the α-F-substituted carbanion a-F− exhibits a higher reactivity than that predicted based on its basicity despite the strong electron-withdrawing capability of fluorine. This results in a downward deviation from the expected Brønsted correlation, thus demonstrating the α-effect. Similar observations were made using nucleophilic SN2 reaction models (Fig. 2b),5 although varying degrees of the α-effect were observed when changing the carbanions (see Fig. S1 in the ESI,† for b-Y− [(PhSO2)(EtOOC)CY−] and c-Y− [(EtOOC)2CY−]).
 |
| Fig. 2 (a) Reaction between a nucleophilic reagent and quinone methide and its computational reaction model, and Brønsted-type correlation between free energy barrier (ΔG‡calc.) and basicity (ΔG0, the larger the value of ΔG0, the stronger the basicity). (b) Computational reaction model of SN2 reactions of carbanions and Brønsted-type correlation between free energy barrier and basicity. Energies in kcal mol−1. | |
The α-effect observed for fluorocarbanion a-F− is consistent with Hamlin's two criteria.5 As shown in Fig. 3a, the HOMO of a-F− exhibits higher energy and a smaller lobe on the nucleophilic center compared to that of the corresponding normal nucleophile a-H−. The higher HOMO energy and smaller lobe on the nucleophilic center of a-F− make it more reactive than normal nucleophiles due to two main factors: (1) the higher HOMO energy can lead to a small HOMONu:–LUMOsubstrate orbital energy gap, thus enhancing the orbital interaction and (2) the smaller lobe on the nucleophilic center reduces the repulsive occupied–occupied orbital overlap between the nucleophile and substrate. Indeed, energy decomposition analysis15 performed using sobEDA16 along the intrinsic reaction coordinate (IRC) of SN2 reaction17 reveals that the Pauli repulsion between a-F− and CH3Cl (Fig. 3b, the reaction coordinate defined in this case as the IRC projection onto the C1⋯C2 distance) is always smaller than that for a-H−. In contrast, (steric) Pauli repulsion values for a-OMe− and a-NHMe− are significantly larger than those for a-H−. Thus, Hamlin's criteria for the α-effect can apply to carbanions.
 |
| Fig. 3 (a) Isosurfaces (isovalues = 0.05), energies, and volumes (isovalues = 0.05) of the HOMO in a-H− and α-nucleophiles a-Y−. (b) Pauli repulsion energy (ΔErep) between the carbanion and CH3Cl along the IRC projected on the C1⋯C2 distance (dC1–C2). The carbanion moiety of TS-SN2-a-Y is defined as Fragment 1 (F1) and the remaining moiety is defined as Fragment 2 (F2). | |
The applicability of Hamlin's α-effect criteria for carbanions has significant implications for predicting their nucleophilicity. While the HOMO energy is a key parameter influencing the nucleophilicity, it alone cannot fully reflect the nucleophilicity of BSM anions, as indicated by the poor correlation between the barrier of nucleophilic addition and HOMO energy (Fig. 4a). Given that the HOMO of BSM anions is contributed not only from the carbanionic center but also from the α-group Y (for examples in Fig. 3a), we hypothesized that the degree of localization of the HOMO, which is influenced by the composition of each atom and can be quantitatively estimated by the orbital delocalization index (ODI: Fig. 4b),18 could also affect the nucleophilicity of the BSM anion. Indeed, a strong correlation was observed between DFT-predicted barriers and those predicted using a binary linear equation that combines both EHOMO and ODIHOMO (R2 = 0.97, Fig. 4c, and also see Fig. S2† for the cases of b-Y− and c-Y−). Notably, ODIHOMO alone could not adequately reflect the nucleophilicity either (Fig. S3 in the ESI†). This further corroborates that both the energy and shape of the HOMO influence the reactivity of nucleophiles. In addition, other α-fluorocarbanions with higher EHOMO or ODIHOMO also have stronger nucleophilicity (Table S4 in the ESI†), while for methyl anions, neither EHOMO nor ODIHOMO of CH2F− is higher than that of CH3−, causing weaker nucleophilicity.
 |
| Fig. 4 (a) Relationship between the free energy barrier (ΔG‡) of the SN2 reaction and EHOMO of carbanions a-Y−. (b) Illustration for the significance of the orbital delocalization index (ODI), and ODIHOMO of carbanions a-Y− (isovalues = 0.05 for isosurfaces, the composition of atoms below 10% is omitted for clarity). (c) The plot of computed vs. predicted free energy barrier of the SN2 reaction between a-Y− and CH3Cl using the multivariate linear regression model. Energies in kcal mol−1. | |
Natural bond orbital (NBO)19 analyses offer additional insights into the effects of α-fluorine on the carbanion nucleophilicity (Fig. 5). For α-fluorocarbanions with stronger nucleophilicity like a-F−, there is a significant lone-pair repulsion between the carbanionic center and α-fluorine, causing a higher EHOMO and/or a more pyramidal structure of α-fluorocarbanions (meaning higher ODIHOMO). In contrast, for α-fluorocarbanions with weaker nucleophilicity like CH2F−, the lone-pair repulsion between α-fluorine and the carbanionic center could largely be avoided due to the already highly pyramidal structure of the carbanionic center. As a result, CH2F− has both lower energy and localization degree of the HOMO than CH3−.
 |
| Fig. 5 Comparison of EHOMO and ODIHOMO among fluorocarbanions and hydrocarbanions, and NBO (isovalues = 0.08) analyses for selected carbanions. The isosurfaces of NBOs are the lone-pair (LP) orbitals for the carbanionic center and F atom, φ(C−) represents the pyramidalization angle around the carbanion center. | |
The α-effect in fluorocarbanion a-F− can effectively rationalize the previously observed intriguing outcomes in monofluoromethylation reactions shown in Fig. 1b. For the palladium-catalyzed enantioselective allylic fluoromethylation reaction, the key step is nucleophilic attack by carbanions to the allylpalladium complex via the transition state TSB-Y (Fig. 6a, and see Fig. S4† for overall mechanistic information). Due to the stronger nucleophilicity of fluorocarbanion a-F−, the barrier corresponding to TSB-F is 2.4 kcal mol−1 lower than that of TSB-H. For LiHMDS intermediated monofluoromethylation of FBSM on benzaldehyde, as shown in Fig. 6b, the barrier of nucleophilic addition by (PhSO2)2CCl− and (PhSO2)2CH− is much higher than that of (PhSO2)2CF−, and the formation of products C2-Cl and C2-H is significantly endergonic, explaining why these products were not obtained. The stronger nucleophilicity and the corresponding bonding ability of fluorocarbanion a-F− are further supported by distortion/interaction activation strain (DIAS)20 analysis for nucleophilic attack along the IRC path (Fig. 6c and d, the reaction coordinate defined in this case as the IRC projection onto the C1⋯C2 distance). For both reactions, it is the more favorable interaction between two fragments, rather than distortion, that governs the reactivity. The absolute value of Eint follows the order of Y = F > Y = H (>Y = Cl), which is consistent with the order of nucleophilicity of carbanions.
 |
| Fig. 6 Calculated free energy profiles for nucleophilic addition of [(PhSO2)2CY]− in the palladium-catalyzed allylic methylation reaction (a) and in LiHMDS intermediated methylation of benzaldehyde (b). DIAS analysis for nucleophilic attack in palladium-catalyzed allylic methylation (c) and LiHMDS intermediated methylation of benzaldehyde (d) along the IRC projected on the C1⋯C2 distance (dC1–C2). The carbanion moiety of TSB-Y or TSC-Y is defined as Fragment 1 (F1) and the remaining moiety is defined as Fragment 2 (F2). Etot = [Edist(F1) + Edist(F2)] + Eint. | |
Conclusions
In conclusion, this study sheds light on the intriguing influence of α-fluorine substitution on the nucleophilicity of carbanions, a class of intermediates fundamentally important in organic synthesis. Through computational analysis, we demonstrated that the α-effect, previously observed in O- and N-based nucleophiles, also applies to carbanions when Hamlin's criteria are met. The enhancement in nucleophilicity upon α-fluorination arises from the elevation of HOMO energy and reduction in the orbital lobe size at the nucleophilic center, thereby facilitating efficient orbital interactions and reducing repulsive overlaps with substrates. These insights may inspire further exploration of the effects of other substituents on nucleophilicity and facilitate the design of more efficient and selective synthetic strategies in organic chemistry.
Computational details
All the quantum chemical calculations were carried out using the Gaussian 1621 package. Optimizations of the geometries of minima and transition states (TSs) were carried out at the PBE0-D3(BJ)22,23/6-31G(d,p)24-SDD(Pd, Cs)25 level of theory. The PCM26 implicit solvation model was used to account for the solvation effects of DMSO or dichloromethane (DCM), which were used in the corresponding experiments. Frequency calculations were performed at the same level to ensure their nature as minima (no imaginary frequency) or transition states (only one imaginary frequency). Intrinsic reaction coordinate (IRC) calculations at the same level verified the connectivity of located intermediates and transition states. The single point energy calculations were performed at the PBE0-D3(BJ)/6-311++G(d,p)-SDD(Pd, Cs) level of theory with the SMD27 solvation model to obtain more accurate electronic energies. Grimme's quasi-RRHO correction28 for the frequencies that are below 100 cm−1 and concentration correction from 1 atm to 1 mol L−1 were implemented using GoodVibes 3.2.29 The wavefunction analyses were performed using Multiwfn 3.8(dev),18 including the orbital delocalization index (ODI) and the volume of isosurfaces. Energy decomposition analysis was performed using sobEDA.16 Distortion/interaction activation strain (DIAS) analysis was realized using a Python tool named autoDIAS.30 The images of the orbital isosurfaces were created using VMD31 combined with Multiwfn 3.8(dev).
Data availability
The authors confirm that the data supporting the findings of this study are available within the article and its ESI,† including computed energy and thermal corrections, and optimized Cartesian coordinates for all the stationary points.
Conflicts of interest
There are no conflicts to declare.
Acknowledgements
This work was supported by the National Natural Science Foundation of China (grant numbers 22122104, 22193012 and 21933004), the CAS Project for Young Scientists in Basic Research (grant numbers YSBR-095 and YSBR-052), and the Strategic Priority Research Program of the Chinese Academy of Sciences (grant number XDB0590000).
References
- W. P. Jencks and J. Carriuolo, Reactivity of Nucleophilic Reagents toward Esters, J. Am. Chem. Soc., 1960, 82, 1778–1786 CrossRef CAS.
- J. O. Edwards and R. G. Pearson, The Factors Determining Nucleophilic Reactivities, J. Am. Chem. Soc., 1962, 84, 16–24 CrossRef CAS.
-
(a) C. H. DePuy, E. W. Della, J. Filley, J. J. Grabowski and V. M. Bierbaum, Absence of an .alpha.-effect in the gas-phase nucleophilic reactions of hydroperoxide ion, J. Am. Chem. Soc., 1983, 105, 2481–2482 CrossRef CAS;
(b) K. R. Fountain, L. K. Hutchinson, D. C. Mulhearn and Y. B. Xu, α-Effect in Menschutkin alkylations, J. Org. Chem., 1993, 58, 7883–7890 CrossRef CAS;
(c) I.-H. Um, J.-S. Lee and S.-M. Yuk, An Unexpectedly Small α-Effect in Nucleophilic Attack at sp-Hybridized Carbon: Michael-Type Additions of Primary Amines to 3-Butyn-2-one, J. Org. Chem., 1998, 63, 9152–9153 CrossRef CAS;
(d) E. V. Patterson and K. R. Fountain, On Gas Phase α-Effects. 1. The Gas-Phase Manifestation and Potential SET Character, J. Org. Chem., 2006, 71, 8121–8125 CrossRef CAS PubMed;
(e) Y. Ren and H. Yamataka, The α-Effect in Gas-Phase SN2 Reactions Revisited, Org. Lett., 2006, 8, 119–121 CrossRef CAS PubMed;
(f) Y. Ren and H. Yamataka, G2(+) Investigation on the α-Effect in the SN2 Reactions at Saturated Carbon, Chem. – Eur. J., 2007, 13, 677–682 CrossRef CAS PubMed;
(g) Y. Ren and H. Yamataka, The α-Effect in Gas-Phase SN2 Reactions: Existence and the Origin of the Effect, J. Org. Chem., 2007, 72, 5660–5667 CrossRef CAS PubMed;
(h) A. M. McAnoy, M. R. L. Paine and S. J. Blanksby, Reactions of the hydroperoxide anion with dimethyl methylphosphonate in an ion trap mass spectrometer: evidence for a gas phase α-effect, Org. Biomol. Chem., 2008, 6, 2316–2326 RSC;
(i) S. M. Villano, N. Eyet, W. C. Lineberger and V. M. Bierbaum, Reactions of α-Nucleophiles with Alkyl Chlorides: Competition between SN2 and E2 Mechanisms and the Gas-Phase α-Effect, J. Am. Chem. Soc., 2009, 131, 8227–8233 CrossRef CAS PubMed;
(j) J. M. Garver, S. Gronert and V. M. Bierbaum, Experimental Validation of the α-Effect in the Gas Phase, J. Am. Chem. Soc., 2011, 133, 13894–13897 CrossRef CAS PubMed;
(k) Y. Ren, X.-G. Wei, S.-J. Ren, K.-C. Lau, N.-B. Wong and W.-K. Li, The α-effect exhibited in gas-phase SN2@N and SN2@C reactions, J. Comput. Chem., 2013, 34, 1997–2005 CrossRef CAS PubMed;
(l) D. L. Thomsen, J. N. Reece, C. M. Nichols, S. Hammerum and V. M. Bierbaum, Investigating the α-Effect in Gas-Phase SN2 Reactions of Microsolvated Anions, J. Am. Chem. Soc., 2013, 135, 15508–15514 CrossRef CAS PubMed;
(m) D. L. Thomsen, J. N. Reece, C. M. Nichols, S. Hammerum and V. M. Bierbaum, The α-Effect in Gas-Phase SN2 Reactions of Microsolvated Anions: Methanol as a Solvent, J. Phys. Chem. A, 2014, 118, 8060–8066 CrossRef CAS PubMed;
(n) W.-Y. Zhao, J. Yu, S.-J. Ren, X.-G. Wei, F.-Z. Qiu, P.-H. Li, H. Li, Y.-P. Zhou, C.-Z. Yin, A.-P. Chen, H. Li, L. Zhang, J. Zhu, Y. Ren and K.-C. Lau, Probing the reactivity of microhydrated α-nucleophile in the anionic gas-phase SN2 reaction, J. Comput. Chem., 2015, 36, 844–852 CrossRef CAS PubMed;
(o) E. Juaristi, G. dos Passos Gomes, A. O. Terent'ev, R. Notario and I. V. Alabugin, Stereoelectronic Interactions as a Probe for the Existence of the Intramolecular α-Effect, J. Am. Chem. Soc., 2017, 139, 10799–10813 CrossRef CAS PubMed;
(p) X. Wu, F. M. Bickelhaupt and J. Xie, Solvent-induced dual nucleophiles and the α-effect in the SN2 versus E2 competition, Phys. Chem. Chem. Phys., 2024, 26, 11320–11330 RSC.
-
(a) N. J. Fina and J. O. Edwards, The alpha effect. A review, Int. J. Chem. Kinet., 1973, 5, 1–26 CrossRef CAS;
(b) J. F. Liebman and R. M. Pollack, Aromatic transition states and the .alpha. effect, J. Org. Chem., 1973, 38, 3444–3445 CrossRef CAS;
(c) S. Hoz, The .alpha. effect: on the origin of transition-state stabilization, J. Org. Chem., 1982, 47, 3545–3547 CrossRef CAS;
(d) K. R. Fountain, C. J. Felkerson, J. D. Driskell and B. D. Lamp, The α-Effect in Methyl Transfers from S-Methyldibenzothiophenium Fluoroborate to Substituted N-Methylbenzohydroxamates, J. Org. Chem., 2003, 68, 1810–1814 CrossRef CAS PubMed;
(e) E. Buncel and I.-H. Um, The α-effect and its modulation by solvent, Tetrahedron, 2004, 60, 7801–7825 CrossRef CAS.
- T. Hansen, P. Vermeeren, F. M. Bickelhaupt and T. A. Hamlin, Origin of the α-Effect in SN2 Reactions, Angew. Chem., Int. Ed., 2021, 60, 20840–20848 CrossRef CAS PubMed.
-
(a) C. Ni and J. Hu, Selective Nucleophilic Fluoroalkylations Facilitated by Removable Activation Groups, Synlett, 2011, 770–782 CAS;
(b) C. Ni and J. Hu, The unique fluorine effects in organic reactions: recent facts and insights into fluoroalkylations, Chem. Soc. Rev., 2016, 45, 5441–5454 RSC;
(c) A. Knieb and G. K. S. Prakash, Recent Advances in Fluoroalkylation Strategies: Exploring Novel Reactivities and Synthetic Applications of Sulfone- and Sulfinate-Based Reagents for Mono-, Di-, and Trifluoromethylations, Synthesis, 2024, 57, 331–361 Search PubMed;
(d) N. Shibata, Development of Shelf-Stable Reagents for Fluoro-Functionalization Reactions, Bull. Chem. Soc. Jpn., 2016, 89, 1307–1320 CrossRef CAS;
(e) A. Knieb, V. Krishnamurti, Z. Zhu and G. K. Surya Prakash, Synthesis of Monofluoromethylarenes: Direct Monofluoromethylation of Diaryliodonium Bromides using Fluorobis(phenylsulfonyl)methane (FBSM), J. Fluorine Chem., 2023, 267, 110095 CrossRef CAS;
(f) F.-L. Qing, X.-Y. Liu, J.-A. Ma, Q. Shen, Q. Song and P. Tang, A Fruitful Decade of Organofluorine Chemistry: New Reagents and Reactions, CCS Chem., 2022, 4, 2518–2549 CrossRef CAS;
(g) A. A. Zemtsov, V. V. Levin and A. D. Dilman, Allylic substitution reactions with fluorinated nucleophiles, Coord. Chem. Rev., 2022, 459, 214455 CrossRef CAS;
(h) B.-J. Li, Y.-L. Ruan, L. Zhu, J. Zhou and J.-S. Yu, Recent advances in catalytic enantioselective construction of monofluoromethyl-substituted stereocenters, Chem. Commun., 2024, 60, 12302–12314 RSC;
(i) W. B. Farnham, Fluorinated Carbanions, Chem. Rev., 1996, 96, 1633–1640 CrossRef CAS PubMed;
(j) G. K. S. Prakash and A. K. Yudin, Perfluoroalkylation with Organosilicon Reagents, Chem. Rev., 1997, 97, 757–786 CrossRef CAS PubMed;
(k) G. K. S. Prakash and J. Hu, Selective Fluoroalkylations with Fluorinated Sulfones, Sulfoxides, and Sulfides, Acc. Chem. Res., 2007, 40, 921–930 CrossRef CAS PubMed;
(l) R. Britton, V. Gouverneur, J.-H. Lin, M. Meanwell, C. Ni, G. Pupo, J.-C. Xiao and J. Hu, Contemporary synthetic strategies in organofluorine chemistry, Nat. Rev. Methods Primers, 2021, 1, 47 CrossRef CAS;
(m)
G. K. Surya Prakash, F. Wang, D. O'Hagan, J. Hu, K. Ding and L.-X. Dai, in Organic Chemistry – Breakthroughs and Perspectives, 2012, pp. 413–476, DOI:10.1002/9783527664801.ch12;
(n) X. Liu, C. Xu, M. Wang and Q. Liu, Trifluoromethyltrimethylsilane: Nucleophilic Trifluoromethylation and Beyond, Chem. Rev., 2015, 115, 683–730 CrossRef CAS PubMed.
- M. Schlosser, Parametrization of Substituents: Effects of Fluorine and Other Heteroatoms on OH, NH, and CH Acidities, Angew. Chem., Int. Ed., 1998, 37, 1496–1513 CrossRef CAS PubMed.
-
(a) F. M. Bickelhaupt, H. L. Hermann and G. Boche, α-Stabilization of Carbanions: Fluorine Is More Effective
than the Heavier Halogens, Angew. Chem., Int. Ed., 2006, 45, 823–826 CrossRef CAS PubMed;
(b) D. A. Dixon, T. Fukunaga and B. E. Smart, Structures and stabilities of fluorinated carbanions. Evidence for anionic hyperconjugation, J. Am. Chem. Soc., 1986, 108, 4027–4031 CrossRef CAS;
(c) W. Zhang, C. Ni and J. Hu, Selective Fluoroalkylation of Organic Compounds by Tackling the “Negative Fluorine Effect”, Top. Curr. Chem., 2012, 25–44, DOI:10.1007/128_2011_246;
(d) C. Ni, Y. Li and J. Hu, Nucleophilic Fluoroalkylation of Epoxides with Fluorinated Sulfones, J. Org. Chem., 2006, 71, 6829–6833 CrossRef CAS PubMed.
-
(a) H. Zheng, Z. Li, J. Jing, X. S. Xue and J. P. Cheng, The Acidities of Nucleophilic Monofluoromethylation Reagents: An Anomalous alpha-Fluorine Effect, Angew. Chem., Int. Ed., 2021, 60, 9401–9406 CrossRef CAS PubMed;
(b)
H. Yamamoto, T. Hiyama, K. Kanie, T. Kusumoto, Y. Morizawa and M. Shimzu, Organofluorine Compounds: Chemistry and Applications, Springer Berlin Heidelberg, 2000 Search PubMed;
(c) F. G. Bordwell, M. Van der Puy and N. R. Vanier, Carbon acids. 8. The trimethylammonio group as a model for assessing the polar effects of electron-withdrawing groups, J. Org. Chem., 1976, 41, 1883–1885 CrossRef CAS;
(d) X.-M. Zhang and F. G. Bordwell, Equilibrium acidities and homolytic bond dissociation enthalpies of the acidic C-H bonds in dialkyl malonates and related compounds, J. Phys. Org. Chem., 1994, 7, 751–756 CrossRef CAS;
(e) W. N. Olmstead and F. G. Bordwell, Ion-pair association constants in dimethyl sulfoxide, J. Org. Chem., 1980, 45, 3299–3305 CrossRef CAS;
(f) M. Suenaga, K. Nakata, J.-L. M. Abboud and M. Mishima, Negative Hyperconjugation in Acidity of Polyfluorinated Alkanes. A Natural Bond Orbital Analysis, Bull. Chem. Soc. Jpn., 2017, 90, 289–297 CrossRef CAS;
(g) A. Streitwieser Jr. and F. Mares, Acidity of hydrocarbons. XXIX. Kinetic acidities of benzal fluoride and 9-fluorofluorene. A pyramidal benzyl anion, J. Am. Chem. Soc., 1968, 90, 2444–2445 CrossRef;
(h) J. Hine, L. G. Mahone and C. L. Liotta, .alpha.-Fluoro and .alpha.-alkoxy substituents as deactivators in carbanion formation, J. Am. Chem. Soc., 1967, 89, 5911–5920 CrossRef CAS;
(i) H. G. Adolph and M. J. Kamlet, Fluoronitroaliphatics. I. The Effect of α Fluorine on the Acidities of Substituted Nitromethanes, J. Am. Chem. Soc., 1966, 88, 4761–4763 CrossRef CAS;
(j) M. Zhang, M. M. R. Badal, M. Pasikowska, T. Sonoda, M. Mishima, H. Fukaya, T. Ono, H.-U. Siehl, J.-L. M. Abboud and I. A. Koppel, Gas-Phase Acidity of Polyfluorinated Hydrocarbons. Effects of Fluorine and the Perfluoroalkyl Group on Acidity, Bull. Chem. Soc. Jpn., 2014, 87, 825–834 CrossRef CAS;
(k) M. Mishima and J.-L. M. Abboud, Substituent effects on acidity of aryl-substituted fluoroalkanes: a computational study, J. Phys. Org. Chem., 2016, 29, 77–83 CrossRef CAS;
(l) M. Mishima, J.-L. M. Abboud, M. Fujio, M. Suenaga, H. F. Koch and J. G. Koch, Gas-phase Acidities of 2-Aryl-2-chloro-1,1,1-trifluoroethanes and Related Compounds. Experimental and Computational Studies, J. Phys. Org. Chem., 2016, 29, 523–531 CrossRef CAS.
-
(a) Z. Zhang, A. Puente, F. Wang, M. Rahm, Y. Mei, H. Mayr and G. K. S. Prakash, The Nucleophilicity of Persistent alpha-Monofluoromethide Anions, Angew. Chem., Int. Ed., 2016, 55, 12845–12849 CrossRef CAS PubMed;
(b) L. A. Kaplan and H. B. Pickard, The α-fluorine effect and carbanion nucleophilicity, J. Chem. Soc., Chem. Commun., 1969, 1500–1501, 10.1039/C29690001500.
- F. L. M. Pattison, R. L. Buchanan and F. H. Dean, The synthesis of α-monofluoroalkanoic acids, Can. J. Chem., 1965, 43, 1700–1713 CrossRef CAS.
- T. Fukuzumi, N. Shibata, M. Sugiura, H. Yasui, S. Nakamura and T. Toru, Fluorobis(phenylsulfonyl)methane: a fluoromethide equivalent and palladium-catalyzed enantioselective allylic monofluoromethylation, Angew. Chem., Int. Ed., 2006, 45, 4973–4977 CrossRef CAS PubMed.
-
(a)
G. K. Surya Prakash, N. Shao, F. Wang and C. Ni, in Organic Syntheses, 2014, pp. 130–144. DOI:10.1002/0471264229.os090.13;
(b)
X. Ispizua-Rodriguez, V. Krishnamurti, M. Coe and G. K. S. Prakash, in Organic Syntheses, 2020, pp. 1–20. DOI:10.1002/0471264229.os096.29.
- X. Shen, L. Zhang, Y. Zhao, L. Zhu, G. Li and J. Hu, Nucleophilic Fluoromethylation of Aldehydes with Fluorobis(phenylsulfonyl)methane: The Importance of Strong Li–O Coordination and Fluorine Substitution for C-C Bond Formation, Angew. Chem., Int. Ed., 2011, 50, 2588–2592 CrossRef CAS PubMed.
-
(a) M. v. Hopffgarten and G. Frenking, Energy decomposition analysis, Wiley Interdiscip. Rev.: Comput. Mol. Sci., 2012, 2, 43–62 Search PubMed;
(b)
T. A. Hamlin, P. Vermeeren, C. F. Guerra and F. M. Bickelhaupt, in Complementary Bonding Analysis, ed. G. Simon, De Gruyter, Berlin, Boston, 2021, pp. 199–212. DOI:10.1515/9783110660074-008.
- T. Lu and Q. Chen, Simple, Efficient, and Universal Energy Decomposition Analysis Method Based on Dispersion-Corrected Density Functional Theory, J. Phys. Chem. A, 2023, 127, 7023–7035 CrossRef CAS PubMed.
-
(a) T. Hansen, P. Vermeeren, A. Haim, M. J. H. van Dorp, J. D. C. Codée, F. M. Bickelhaupt and T. A. Hamlin, Regioselectivity of Epoxide Ring-Openings via SN2 Reactions Under Basic and Acidic Conditions, Eur. J. Org. Chem., 2020, 3822–3828 CrossRef CAS;
(b) P. Vermeeren, T. Hansen, M. Grasser, D. R. Silva, T. A. Hamlin and F. M. Bickelhaupt, SN2 versus E2 Competition of F− and PH2− Revisited, J. Org. Chem., 2020, 85, 14087–14093 CrossRef CAS PubMed;
(c) P. Vermeeren, T. Hansen, P. Jansen, M. Swart, T. A. Hamlin and F. M. Bickelhaupt, A Unified Framework for Understanding Nucleophilicity and Protophilicity in the SN2/E2 Competition, Chem. – Eur. J., 2020, 26, 15538–15548 CrossRef CAS PubMed;
(d) T. Hansen, P. Vermeeren, R. Yoshisada, D. V. Filippov, G. A. van der Marel, J. D. C. Codée and T. A. Hamlin, How Lewis Acids Catalyze Ring-Openings of Cyclohexene Oxide, J. Org. Chem., 2021, 86, 3565–3573 CrossRef CAS PubMed;
(e) T. Hansen, P. Vermeeren, K. W. J. Zijderveld, F. M. Bickelhaupt and T. A. Hamlin, SN2 versus E2 Competition of Cyclic Ethers, Chem. – Eur. J., 2023, 29, e202301308 CrossRef CAS PubMed;
(f) T. Hansen, A. Nin-Hill, J. D. C. Codée, T. A. Hamlin and C. Rovira, Rational Tuning of the Reactivity of Three-Membered Heterocycle Ring Openings via SN2 Reactions, Chem. – Eur. J., 2022, 28, e202201649 CrossRef CAS PubMed;
(g) W. A. Remmerswaal, T. de Jong, K. N. A. van de Vrande, R. Louwersheimer, T. Verwaal, D. V. Filippov, J. D. C. Codée and T. Hansen, Backside versus Frontside SN2 Reactions of Alkyl Triflates and Alcohols, Chem. – Eur. J., 2024, 30, e202400590 CrossRef CAS PubMed.
-
(a)
T. Lu, Multiwfn Software Manual Version 3.8 (dev), 2025, available at https://sobereva.com/multiwfn (Last update: 2025-Jan-5) Search PubMed;
(b) T. Lu and F. Chen, Multiwfn: A multifunctional wavefunction analyzer, J. Comput. Chem., 2012, 33, 580–592 CrossRef CAS PubMed;
(c) T. Lu, A comprehensive electron wavefunction analysis toolbox for chemists, Multiwfn, J. Chem. Phys., 2024, 161, 082503 CrossRef CAS PubMed.
-
(a) E. D. Glendening, C. R. Landis and F. Weinhold, Natural bond orbital methods, Wiley Interdiscip. Rev.: Comput. Mol. Sci., 2012, 2, 1–42 CAS;
(b)
E. R. Glendenning, A. Reed, J. Carpenter and F. Weinhold, NBO Version 3.1, Madison, WI, USA, 2001.
-
(a) P. Vermeeren, S. C. C. van der Lubbe, C. F. Guerra, F. M. Bickelhaupt and T. A. Hamlin, Understanding chemical reactivity using the activation strain model, Nat. Protoc., 2020, 15, 649–667 CrossRef CAS PubMed;
(b) F. M. Bickelhaupt, Understanding reactivity with Kohn–Sham molecular orbital theory: E2−SN2 mechanistic spectrum and other concepts, J. Comput. Chem., 1999, 20, 114–128 CrossRef CAS;
(c) W.-J. van Zeist, A. H. Koers, L. P. Wolters and F. M. Bickelhaupt, Reaction Coordinates and the Transition-Vector Approximation to the IRC, J. Chem. Theory Comput., 2008, 4, 920–928 CrossRef CAS PubMed;
(d) F. M. Bickelhaupt and K. N. Houk, Analyzing Reaction Rates with the Distortion/Interaction-Activation Strain Model, Angew. Chem., Int. Ed., 2017, 56, 10070–10086 CrossRef CAS PubMed.
-
M. J. Frisch, G. W. Trucks, H. B. Schlegel, G. E. Scuseria, M. A. Robb, J. R. Cheeseman, G. Scalmani, V. Barone, G. A. Petersson, H. Nakatsuji, X. Li, M. Caricato, A. V. Marenich, J. Bloino, B. G. Janesko, R. Gomperts, B. Mennucci, H. P. Hratchian, J. V. Ortiz, A. F. Izmaylov, J. L. Sonnenberg, D. Williams-Young, F. Ding, F. Lipparini, F. Egidi, J. Goings, B. Peng, A. Petrone, T. Henderson, D. Ranasinghe, V. G. Zakrzewski, J. Gao, N. Rega, G. Zheng, W. Liang, M. Hada, M. Ehara, K. Toyota, R. Fukuda, J. Hasegawa, M. Ishida, T. Nakajima, Y. Honda, O. Kitao, H. Nakai, T. Vreven, K. Throssell, J. A. Montgomery Jr., J. E. Peralta, F. Ogliaro, M. J. Bearpark, J. J. Heyd, E. N. Brothers, K. N. Kudin, V. N. Staroverov, T. A. Keith, R. Kobayashi, J. Normand, K. Raghavachari, A. P. Rendell, J. C. Burant, S. S. Iyengar, J. Tomasi, M. Cossi, J. M. Millam, M. Klene, C. Adamo, R. Cammi, J. W. Ochterski, R. L. Martin, K. Morokuma, O. Farkas, J. B. Foresman and D. J. Fox, Gaussian 16 Rev. C.01, 2016 Search PubMed.
- C. Adamo and V. Barone, Toward reliable density functional methods without adjustable parameters: The PBE0 model, J. Chem. Phys., 1999, 110, 6158–6170 CrossRef CAS.
- S. Grimme, J. Antony, S. Ehrlich and H. Krieg, A consistent and accurate ab initio parametrization of density functional dispersion correction (DFT-D) for the 94 elements H-Pu, J. Chem. Phys., 2010, 132, 154104 CrossRef PubMed.
- P. C. Hariharan and J. A. Pople, The influence of polarization functions on molecular orbital hydrogenation energies, Theor. Chim. Acta, 1973, 28, 213–222 CrossRef CAS.
-
(a) A. Henglein, Physicochemical properties of small metal particles in solution: “microelectrode” reactions, chemisorption, composite metal particles, and the atom-to-metal transition, J. Phys. Chem., 1993, 97, 5457–5471 CrossRef CAS;
(b) M. Kaupp, P. v. R. Schleyer, H. Stoll and H. Preuss, Pseudopotential approaches to Ca, Sr, and Ba hydrides. Why are some alkaline earth MX2 compounds bent?, J. Chem. Phys., 1991, 94, 1360–1366 CrossRef CAS;
(c) U. Steinbrenner, A. Bergner, M. Dolg and H. Stoll, On the transferability of energy adjusted pseudopotentiais: a calibration study for XH4 (X=C, Si, Ge, Sn, Pb), Mol. Phys., 1994, 82, 3–11 CrossRef CAS.
-
(a) J. Tomasi, B. Mennucci and R. Cammi, Quantum Mechanical Continuum Solvation Models, Chem. Rev., 2005, 105, 2999–3094 CrossRef CAS PubMed;
(b) B. Mennucci, E. Cancès and J. Tomasi, Evaluation of Solvent Effects in Isotropic and Anisotropic Dielectrics and in Ionic Solutions with a Unified Integral Equation Method: Theoretical Bases, Computational Implementation, and Numerical Applications, J. Phys. Chem. B, 1997, 101, 10506–10517 CrossRef CAS;
(c) S. Miertuš, E. Scrocco and J. Tomasi, Electrostatic interaction of a solute with a continuum. A direct utilizaion of AB initio molecular potentials for the prevision of solvent effects, Chem. Phys., 1981, 55, 117–129 CrossRef.
- A. V. Marenich, C. J. Cramer and D. G. Truhlar, Universal Solvation Model Based on Solute Electron Density and on a Continuum Model of the Solvent Defined by the Bulk Dielectric Constant and Atomic Surface Tensions, J. Phys. Chem. B, 2009, 113, 6378–6396 CrossRef CAS PubMed.
- S. Grimme, Supramolecular Binding Thermodynamics by Dispersion-Corrected Density Functional Theory, Chem. – Eur. J., 2012, 18, 9955–9964 CrossRef CAS PubMed.
- G. Luchini, J. V. Alegre-Requena, I. Funes-Ardoiz and R. S. Paton, GoodVibes: automated thermochemistry for heterogeneous computational chemistry data, F1000Research, 2020, 9, 291 Search PubMed.
- D. Svatunek and K. N. Houk, autoDIAS: a python tool for an automated distortion/interaction activation strain analysis, J. Comput. Chem., 2019, 40, 2509–2515 CrossRef CAS PubMed.
- W. Humphrey, A. Dalke and K. Schulten, VMD: Visual molecular dynamics, J. Mol. Graphics, 1996, 14, 33–38 CrossRef CAS PubMed.
Footnote |
† Electronic supplementary information (ESI) available: Supplementary computational results (PDF) and Cartesian coordinates for all the stationary points (XYZ). See DOI: https://doi.org/10.1039/d5qo00292c |
|
This journal is © the Partner Organisations 2025 |
Click here to see how this site uses Cookies. View our privacy policy here.