Diels–Alder reaction of β-fluoro-β-nitrostyrenes with cyclic dienes

  1. 1 ORCID Logo ,
  2. 1 ,
  3. 1 ORCID Logo ,
  4. 2 ORCID Logo ,
  5. 2 ,
  6. 3 ,
  7. 3 ORCID Logo and
  8. 1,§ ORCID Logo
1Department of Chemistry, Lomonosov Moscow State University, Leninskie gory 1, Moscow, 119991, Russian Federation
2N. D. Zelinsky Institute of Organic Chemistry, Russian Academy of Sciences Leninsky prosp. 47, Moscow 119991, Russian Federation
3Department of Chemistry Johannes Gutenberg-University, Duesbergweg 10–14, D-55128 Mainz, Germany
  1. Corresponding author email
§ Fax: (+)7-495-9328846
Guest Editor: D. O'Hagan
Beilstein J. Org. Chem. 2021, 17, 283–292. https://doi.org/10.3762/bjoc.17.27
Received 11 Nov 2020, Accepted 13 Jan 2021, Published 27 Jan 2021
Full Research Paper
cc by logo

Abstract

The Diels–Alder reaction of β-fluoro-β-nitrostyrenes with cyclic 1,3-dienes was investigated. A series of novel monofluorinated norbornenes was prepared in up to 97% yield. The reaction with 1,3-cyclohexadiene permits the preparation of monofluorinated bicyclo[2.2.2]oct-2-enes. The kinetic data of the reactions with 1,3-cyclopentadiene and 1,3-cyclohexadiene were used to calculate activation parameters. Furthermore, the synthetic utility of the cycloadducts obtained was demonstrated.

Introduction

Organofluorine compounds play an exceptionally important role in various fields of science and technology. The incorporation of fluorine into molecules can significantly influence their pharmacokinetic and physicochemical properties and enhance their metabolic and chemical stability [1-5]. For instance, nearly a quarter of the currently manufactured agrochemical and pharmaceutical products contains at least one fluorine atom [6-8]. Fluorinated functional materials have also found wide application as durable ion exchange membranes, e.g., in fuel cells [9-11], as thermoplastic polymers [12-14], in electronic and optoelectronic technologies [15], and in liquid crystal display applications [16-21], etc [22]. The use of fluorinated building blocks is a very convenient approach and in many cases represents an indispensable alternative to late-stage fluorinations in the preparation of such unique materials [23].

The Diels–Alder reaction is considered a versatile and powerful tool for assembling a variety of fluorinated carbo- and heterocycles using either the diene [24-30] or the dienophile component [31-39] as fluorine-containing building blocks. The application of [4 + 2] cycloadditions for the preparation of fluorinated bicyclic compounds has attracted much attention [40-47]. In this regard, the development of new protocols to relevant monofluorinated bicyclic molecules involving novel versatile fluorine-containing building blocks is of key importance. Fluoroalkenes are recognized to be one of the most widely used fluorine-containing building blocks [48,49]. Recently, we have developed an efficient stereoselective synthesis of β-fluoro-β-nitrostyrenes 1 based on the radical nitration of 2-bromo-2-fluorostyrenes [50]. This process takes place with simultaneous elimination of bromine, and gives the target structures solely in the Z-isomeric form in high yields (up to 92%). These fluorine-containing olefins activated by a nitro group proved to be the appealing building blocks for the construction of numerous monofluorinated compounds [51-56]. This paper is devoted to a new synthetic approach to novel monofluorinated bicyclic compounds, namely norbornenes and bicyclo[2.2.2]oct-2-enes and their subsequent functionalization. The present study is our follow-up work on the Diels–Alder reaction involving β-fluoro-β-nitrostyrenes [57].

A recent review reported that by 2018, the total number of publications and patents related to the production and use of norbornene and norbornadiene derivatives had exceeded 30,000 [58]. Indeed, norbornene and its derivatives have found application in medicine, agriculture, microelectronics, and rocket technology as well as in production of polymeric materials, efficient gas separation membranes and solar energy converters [58]. Considering the high interest in such structures and the unique role of fluorine, we believe that novel norbornene derivatives obtained in the framework of this study can become relevant compounds in practical use.

Results and Discussion

Initially, we studied the Diels–Alder reaction of β-fluoro-β-nitrostyrenes 1 with 1,3-cyclopentadiene (CPD) to prepare a series of novel monofluorinated norbornenes. The starting nitrostyrenes were prepared and used in the Z-isomer form. The transformations were conducted in screw-top vials in o-xylene at 110 °C using a fivefold excess of the diene (Scheme 1). The reaction proceeded smoothly under these conditions to give the target cycloadducts 2 as a mixture of exo and endo-isomers in high isolated yield (up to 97%). It should be noted that in the present work we indicate an isomer as exo or endo according to the stereo-position of the fluorine atom. Thus exo-2 and endo-2 means 5-exo-fluoro-5-endo-nitro-6-exo-arylnorbornene and 5-endo-fluoro-5-exo-nitro-6-endo-arylnorbornene, respectively (Figure 1). Both diastereomers are formed in a nearly 1:1 ratio for the majority of the substituents on the aryl group of the nitrostyrenes 1. However, a higher diastereoselectivity towards the endo-isomer was observed when strong electron-withdrawing groups (EWGs) were present in the dienophile. For example, in the case of the 4-cyano and the 3-nitro-substituted derivative, the ratio of endo/exo was 2:1.

[1860-5397-17-27-i1]

Scheme 1: Scope of the nitrostyrenes 1 in the Diels–Alder reaction with CPD (dr = exo:endo).

[1860-5397-17-27-1]

Figure 1: The structure assignment of norbornenes 2 by 1H (a) and NOE (b) NMR spectroscopy.

The stereochemistry of the products 2a–l can be unambiguously assigned using 1H NMR spectroscopy. According to the literature data [59] the dienophile-derived proton at C6 resonates at lower field in the exo-form than the corresponding proton of the endo-isomer. For example, in the case of 2l, the 1H NMR spectrum shows a doublet of doublet signal for H6 at 3.89 ppm for the minor isomer and at 4.27 ppm for the major isomer (Figure 1a). A significant chemical shift difference is observed for the aryl proton signals of the exo- and endo-isomeric norbornenes 2. Most probably such significant difference in the chemical shifts can be explained by the double bond anisotropy of the norbornene molecule [60]. The stereochemical assignments are in full accordance with the values of vicinal (3JH6-F and 3JH1-H6) and long-range coupling constants (4JH6-H7b). According to the literature data, the value of 3JH1-H6 is larger than that of 4JH6-H7b [32,61-63]. For example, the 1H NMR spectrum of the minor isomer of 2l showed the coupling constants 3JH6-F = 10.7 Hz and 4JH6-H7b = 2.9 Hz consistent with an exo-geometry. In contrast, the major isomer having constants 3JH6-F = 9.1 Hz and 3JH1-H6 = 3.1 Hz was ascribed to the endo-form (Figure 1). It should be noted that this observation applies for all cases investigated. The value of the coupling constant 3JH6-F between the exo-F and endo-H6 was always larger than the corresponding value between the endo-F and exo-H6. The stereochemical assignments were additionally confirmed by nuclear Overhauser effect spectroscopy (NOE). The peak of H6 was selected to be selectively excited for each isomer. As expected, in the case of endo-2l the NOE peaks resulted from the interaction of exo-H6 with H1 and H7a were observed. Whereas for exo-2l due to the opposite side position of endo-H6 there was no interaction observed. Thus, using these spectral data all the pairs of exo- and endo-isomers 2 obtained can be assigned unambiguously.

Moreover, the 13C NMR spectra of the exo and endo-isomers exhibit a significant difference (approximately 3 ppm) in the chemical shifts for some carbon atoms (Figure 2). A considerable difference in the chemical shifts was observed for C-7 of the methylene bridge (46.1 for the exo- vs 48.9 ppm for the endo-isomer), C-6 (51.3 for the exo vs 53.4 for the endo-isomer), C-4 (52.4 for the exo- vs 55.3 ppm for the endo-isomer), and C-2 (139.7 for the endo- vs 143.0 ppm for the exo-isomer). The same pattern in the chemical shifts and coupling constants was observed for all structures 2 synthesized.

[1860-5397-17-27-2]

Figure 2: 13C NMR spectrum of the mixture of exo- and endo-isomers of norbornene 2l.

For further insights into the mechanistic background of the endo-exo selectivity, the Diels–Alder reaction of CPD with the model nitrostyrene 1h was simulated in silico to predict the reaction pathway, the reaction rate constants, and the activation enthalpies. Density functional theory calculations were conducted for the reactants, products, and transition states using the B3LYP [64-66] and M062X [67] level of theory in combination with a Pople basis set and the IEFPCM [68] solvation model for o-xylene. Both functionals are already known in the literature for the investigation of cycloadditions [69-72]. For the computational details the reader is referred to Supporting Information File 1.

The predicted reaction pathways for the formation of the exo- and endo-isomeric norbornene 2h using M062X are displayed in Figure 3. For each isomer one transition state exo-TS and endo-TS was identified. The former transition state is higher in energy and leads to the less exergonic product exo-2h. The exo and endo-isomers were predicted to have free energies of activation ([Graphic 1]) of 120.62 and 119.64 kJ mol−1, respectively. The corresponding predicted reaction free energies (ΔG383.15) are −39.66 and −42.07 kJ mol−1. With the former values of ΔG, the reaction rate coefficient k can be calculated using the Eyring equation (Equation 1) [73,74]:

[1860-5397-17-27-3]

Figure 3: The predicted reaction pathway for the Diels–Alder reaction of nitrostyrene 1h with CPD is displayed. The path leading to the endo-2h isomer has a lower lying transition state and is more exergonic than the path resulting in the exo-2h isomer. Energies are given in kJ mol−1 and were calculated with M062X/6-311+G(d,p) at 383.15 K in o-xylene (the corresponding B3LYP reaction pathway is not shown).

[1860-5397-17-27-i7]
(1)

For T = 110 °C, the predicted ratio of kendo/kexo = 1.36 (1.68 for B3LYP) is in good accordance with the experimentally observed diastereomeric ratio of 1.22. The larger discrepancy in case of the B3LYP functional may be due to the fact that dispersion effects are not included, whereas M062X includes nonlocal effects of electronic dispersion [70,75].

We also demonstrate the preparation of norbornene structures substituted at the methylene bridge. The reaction of model nitrostyrene 1h with spiro[2.4]hepta-4,6-diene was carried out (Scheme 2). As a result, the corresponding norbornene 2m having a cyclopropane ring was obtained in moderate yield (44%). The cycloaddition proceeds much more slowly as a result of the high steric demand of the cyclopropyl ring of the spirodiene compared to the CH2 group of cyclopentadiene. We believe that this is the reason of the lower yield in comparison to the reaction with CPD. The stereochemical assignment was performed using 1H NMR spectroscopy (Scheme 2) to show similar peculiarities of the spectra. In contrast to the reaction with CPD, a slight prevalence in the formation of the exo-isomer (exo:endo = 56:44) was observed for the product 2m.

[1860-5397-17-27-i2]

Scheme 2: The Diels–Alder reaction of nitrostyrene 1h with spiro[2.4]hepta-4,6-diene.

Next, the reaction with 1,3-cyclohexadiene (CHD) was investigated. It was found that the reaction is very sensitive to the structure of starting diene and in the case of CHD proceeds much more slowly. Both thermal and microwave (MW) activation (Scheme 3) was investigated to accelerate the reaction with CHD. However, in all cases, the yields of the target cycloadducts 3 were below 35% despite the full conversion of the nitrostyrenes 1 which is common for this type of dienophiles (Scheme 3). The stereochemical assignment was made similarly to the norbornene structures using 1H NMR spectroscopy (Scheme 3). Larger values of 3JH6-F were observed for the exo-F isomers. The presence of a strong EWG on the aryl substituent led to higher stereoselectivity. For example, approximately a 2:1 ratio was observed for the nitro- and carboxymethyl-substituted products 3b, 3c, whereas in the absence of a strong EWG, the ratio was about 1:1 (3a). However, in contrast to CPD derivatives, the major products formed in the reaction with CHD have exo-configuration.

[1860-5397-17-27-i3]

Scheme 3: Diels–Alder reaction of nitrostyrenes 1 with CHD (dr = exo:endo). (а) Reaction under microwave activation.

To gain deeper insights into the reaction, we carried out some kinetic studies to evaluate and compare the reactivities of CHD and CPD in the reactions with model nitrostyrene 1h (Scheme 4). All the kinetic runs were performed using a ≈43–49 molar excess of the diene in o-xylene (1:1) to provide pseudo-first order conditions. Conversions (F) of 1 were measured by 1H NMR spectroscopy. The reactions were found to proceed under the kinetic control since the isomer ratio remained constant throughout the reaction course regardless of the temperature (Table 1).

[1860-5397-17-27-i4]

Scheme 4: Kinetic study of reactions of 1h with CPD and CHD.

Table 1: Kinetic parameters for the reactions of 1h with CPD and CHD.

entry diene T, °C molar ratio
exo/endo
k*·104
s−1
ktotal·105
L/mol·s
kexo·105,
L/mol·s
kendo·105,
L/mol·s
Rcorr
1 CPD 50 46:54 1.00 1.67 0.090 0.077 0.997
2 CPD 80 46:54 5.78 9.72 5.26 4.46 0.998
3 CPD 110 46:54 35.62 59.91 32.39 27.53 0.990
4 CPD 130 46:54 46.20 77.71 41.99 35.72 0.999
5 CHD 110 61:39 0.12 0.224 0.137 0.087 0.999

The total effective pseudo-first order rate constants k* were obtained by plotting the experimental values of ln(c0/c) versus time with good correlations (Table 1). The overall second-order rate total constants ktotal were calculated from the effective k* and initial concentration of the diene (Table 1). The individual constants for the endo and exo-isomers (kendo and kexo) were evaluated by multiplication of ktotal with the molar fractions of the isomers (Table 1). The data obtained demonstrated that the overall reaction rate for CHD is 267 times lower than that for CPD at 110 °C (Figure 4, Table 1). Such a large difference in the reactivity of CHD and CPD was described in the literature. For example, in model reactions with tetracyanoethene, the difference was 2600-fold at 20 °C [76]. The activation parameters were estimated for the reaction of 1h with CPD by plotting ln(k/T) versus 1/T according to the Eyring equation (Equations 2–4) [77].

[1860-5397-17-27-4]

Figure 4: Kinetic curves for the reactions of nitrostyrene 1h with CPD (50–130 °C) and CHD at 110 °C.

[1860-5397-17-27-i8]
(2)
[1860-5397-17-27-i9]
(3)
[1860-5397-17-27-i10]
(4)

The activation enthalpies (ΔH) for exo- and endo-1h were found to be identical for both reaction pathways (51.6 kJ mol−1). Whereas the entropies of activation (ΔS) were −181.8 and −183.1 J mol−1 K−1 for the formation of the endo and exo-isomers, respectively. The values obtained are typical for concerted [4 + 2]-cycloaddition reactions [60]. The free energies of activation ([Graphic 2]) were calculated for 121.26 kJ mol−1 for endo-1h and 121.75 kJ mol−1 for exo-1h and were consistent with the predicted ones.

Next, the reaction with some other cyclic dienes was investigated. The reaction with the unsymmetrical 1-methoxy-1,3-cyclohexadiene (Scheme 5) led to the formation of a mixture of four products (regioisomers and stereoisomers, respectively) 3d in 40% overall yield. Two pairs of regioisomers were partially separated by column chromatography with sufficiently slow elution and analyzed by 1H NMR spectroscopy. The structure assignment was made as depicted in Scheme 5. The structures of two pairs of regioisomers were assigned by chemical shifts of the singlet of the methoxy group. The products having the MeO and NO2 groups in the adjacent position have the signal of the methoxy protons shifted to a lower field. The assignment of the exo/endo-isomers was carried out by the position of the benzylic proton (H5 or H6) and its coupling constant to fluorine (3JH5-F or 3JH6-F).

[1860-5397-17-27-i5]

Scheme 5: The Diels–Alder reaction of the nitrostyrene 1h with 1-methoxy-1,3-cyclohexadiene.

The reaction with 7- and 8-membered cyclic dienes (1,3-cycloheptadiene and 1,3-cyclooctadiene) did not result in the formation of the corresponding cycloadducts confirming that the reaction is very sensitive to the structure of the dienes. Moreover, it was found that furan did not react with nitrostyrenes 1.

Furthermore, we performed some subsequent transformations of the fluorinated norbornenes prepared to investigate their chemical properties and to demonstrate their utility (Scheme 6). These reactions were carried out to involve either the double bond or the nitro group of the norbornene products. The treatment of cycloadducts 2 with m-chloroperbenzoic acid afforded a series of novel fluorinated epoxynorbornane derivatives 4 in high yields (up to 87%). In all cases, the formation of mixtures of only two products was observed in ratios similar to those of the starting mixture 2. We believe that this is a result of an exo-epoxidation which is preferred in norbornene systems [78,79]. Such a functionalization is very attractive to produce new reactive building blocks bearing the norbornane scaffold. This approach can pave a straightforward way to numerous fluorine-containing bicyclic compounds not previously available. The syn-dihydroxylation of compound 2f with the N-methylmorpholine-N-oxide (NMO)–OsO4 system resulted in a mixture of the corresponding diols 5 in a 36:64 ratio in 65% yield. Again, exo-dihydroxylation is to be expected [80-82]. The treatment of norbornene 2l with t-BuOK resulted in the selective elimination of nitrous acid to form the desired monofluorinated norbornadiene 6 in 77% yield. No competitive elimination of HF was observed. The Diels–Alder reaction–base-induced HNO2 elimination sequence opens a straightforward way to novel fluorinated norbornadienes from β-fluoro-β-nitrostyrenes and CPD.

[1860-5397-17-27-i6]

Scheme 6: Selected chemical transformations of norbornenes 2 (dr = exo:endo).

Conclusion

In summary, the Diels–Alder reaction of β-fluoro-β-nitrostyrenes with cyclic 1,3-dienes was investigated. A series of novel monofluorinated norbornenes was prepared in high yield up to 97%. A number of novel monofluorinated bicyclo[2.2.2]oct-2-enes was obtained in up to 40% yield. The reactivity of CPD and its homologues was evaluated and compared. The reaction rate for CHD proved to be 267 times lower than that for CPD in a model reaction, whereas 1,3-cycloheptadiene and 1,3-cyclooctadiene were found to be unable to react. The activation parameters of the reaction of nitrostyrene 1h with CPD were estimated. In addition, the synthetic utility of the norbornenes obtained was demonstrated. All the structures obtained in this work were elucidated by NMR spectroscopy and elemental analysis or HRMS.

Supporting Information

Supporting Information File 1: Copies of spectra, experimental section, and computational details of DFT calculations.
Format: PDF Size: 8.3 MB Download

Acknowledgements

The authors acknowledge the partial support in measuring of NMR spectra from the M. V. Lomonosov Moscow State University Program of Development. We thank Jan Brauer and Paul Eckhardt for their assistance with the computational calculations and evaluations. Parts of this research were conducted using the supercomputer MOGON and/or advisory services offered by the Johannes Gutenberg University Mainz (hpc.uni-mainz.de), which is a member of the AHRP (Alliance for High Performance Computing in Rhineland Palatinate, http://www.ahrp.info) and the Gauss Alliance e.V. The authors gratefully acknowledge the computing time granted on the supercomputer MOGON at Johannes Gutenberg University Mainz.

Funding

This work was supported by RFBR, project number 20-33-70132.

References

  1. Purser, S.; Moore, P. R.; Swallow, S.; Gouverneur, V. Chem. Soc. Rev. 2008, 37, 320–330. doi:10.1039/b610213c
    Return to citation in text: [1]
  2. Shah, P.; Westwell, A. D. J. Enzyme Inhib. Med. Chem. 2007, 22, 527–540. doi:10.1080/14756360701425014
    Return to citation in text: [1]
  3. Wang, B.-C.; Wang, L.-J.; Jiang, B.; Wang, S.-Y.; Wu, N.; Li, X.-Q.; Shi, D.-Y. Mini-Rev. Med. Chem. 2017, 17, 683–692. doi:10.2174/1389557515666151016124957
    Return to citation in text: [1]
  4. Gillis, E. P.; Eastman, K. J.; Hill, M. D.; Donnelly, D. J.; Meanwell, N. A. J. Med. Chem. 2015, 58, 8315–8359. doi:10.1021/acs.jmedchem.5b00258
    Return to citation in text: [1]
  5. Meanwell, N. A. J. Med. Chem. 2018, 61, 5822–5880. doi:10.1021/acs.jmedchem.7b01788
    Return to citation in text: [1]
  6. Politanskaya, L. V.; Selivanova, G. A.; Panteleeva, E. V.; Tretyakov, E. V.; Platonov, V. E.; Nikul’shin, P. V.; Vinogradov, A. S.; Zonov, Y. V.; Karpov, V. M.; Mezhenkova, T. V.; Vasilyev, A. V.; Koldobskii, A. B.; Shilova, O. S.; Morozova, S. M.; Burgart, Y. V.; Shchegolkov, E. V.; Saloutin, V. I.; Sokolov, V. B.; Aksinenko, A. Y.; Nenajdenko, V. G.; Moskalik, M. Y.; Astakhova, V. V.; Shainyan, B. A.; Tabolin, A. A.; Ioffe, S. L.; Muzalevskiy, V. M.; Balenkova, E. S.; Shastin, A. V.; Tyutyunov, A. A.; Boiko, V. E.; Igumnov, S. M.; Dilman, A. D.; Adonin, N. Y.; Bardin, V. V.; Masoud, S. M.; Vorobyeva, D. V.; Osipov, S. N.; Nosova, E. V.; Lipunova, G. N.; Charushin, V. N.; Prima, D. O.; Makarov, A. G.; Zibarev, A. V.; Trofimov, B. A.; Sobenina, L. N.; Belyaeva, K. V.; Sosnovskikh, V. Y.; Obydennov, D. L.; Usachev, S. A. Russ. Chem. Rev. 2019, 88, 425–569. doi:10.1070/rcr4871
    Return to citation in text: [1]
  7. Inoue, M.; Sumii, Y.; Shibata, N. ACS Omega 2020, 5, 10633–10640. doi:10.1021/acsomega.0c00830
    Return to citation in text: [1]
  8. Han, J.; Remete, A. M.; Dobson, L. S.; Kiss, L.; Izawa, K.; Moriwaki, H.; Soloshonok, V. A.; O’Hagan, D. J. Fluorine Chem. 2020, 239, 109639. doi:10.1016/j.jfluchem.2020.109639
    Return to citation in text: [1]
  9. Mauritz, K. A.; Moore, R. B. Chem. Rev. 2004, 104, 4535–4586. doi:10.1021/cr0207123
    Return to citation in text: [1]
  10. Schmidt-Rohr, K.; Chen, Q. Nat. Mater. 2008, 7, 75–83. doi:10.1038/nmat2074
    Return to citation in text: [1]
  11. Kraytsberg, A.; Ein-Eli, Y. Energy Fuels 2014, 28, 7303–7330. doi:10.1021/ef501977k
    Return to citation in text: [1]
  12. Ferrari, M. C.; Galizia, M.; De Angelis, M. G.; Sarti, G. C. Ind. Eng. Chem. Res. 2010, 49, 11920–11935. doi:10.1021/ie100242q
    Return to citation in text: [1]
  13. Mastronardi, F.; Gutmann, B.; Kappe, C. O. Org. Lett. 2013, 15, 5590–5593. doi:10.1021/ol4027914
    Return to citation in text: [1]
  14. Kubota, T. J. Synth. Org. Chem., Jpn. 2013, 71, 196–206. doi:10.5059/yukigoseikyokaishi.71.196
    Return to citation in text: [1]
  15. Babudri, F.; Farinola, G. M.; Naso, F.; Ragni, R. Chem. Commun. 2007, 1003–1022. doi:10.1039/b611336b
    Return to citation in text: [1]
  16. Hird, M. Chem. Soc. Rev. 2007, 36, 2070–2095. doi:10.1039/b610738a
    Return to citation in text: [1]
  17. Yokokoji, O.; Miyajima, T.; Irisawa, J.; Shimizu, T.; Inoue, S. Liq. Cryst. 2009, 36, 799–807. doi:10.1080/02678290903062986
    Return to citation in text: [1]
  18. Goodby, J. W.; Hindmarsh, P.; Hird, M.; Lewis, R. A.; Toyne, K. J. Mol. Cryst. Liq. Cryst. Sci. Technol., Sect. A 2001, 364, 889–898. doi:10.1080/10587250108025062
    Return to citation in text: [1]
  19. Ishigure, T.; Koike, Y. Mol. Cryst. Liq. Cryst. Sci. Technol., Sect. A 2000, 353, 451–469. doi:10.1080/10587250008025681
    Return to citation in text: [1]
  20. Kirsch, P. J. Fluorine Chem. 2015, 177, 29–36. doi:10.1016/j.jfluchem.2015.01.007
    Return to citation in text: [1]
  21. Al-Maharik, N.; Kirsch, P.; Slawin, A. M. Z.; Cordes, D. B.; O'Hagan, D. Org. Biomol. Chem. 2016, 14, 9974–9980. doi:10.1039/c6ob01986b
    Return to citation in text: [1]
  22. Pagliaro, M.; Ciriminna, R. J. Mater. Chem. 2005, 15, 4981–4991. doi:10.1039/b507583c
    Return to citation in text: [1]
  23. Kondratov, I. S.; Tolmachova, N. A.; Haufe, G. Eur. J. Org. Chem. 2018, 3618–3647. doi:10.1002/ejoc.201800327
    Return to citation in text: [1]
  24. Hayashi, T.; Usuki, Y.; Wakamatsu, Y.; Iio, H. Synlett 2010, 2843–2846. doi:10.1055/s-0030-1259013
    Return to citation in text: [1]
  25. Shi, G.-q.; Cottens, S.; Shiba, S. A.; Manfred, S. Tetrahedron 1992, 48, 10569–10574. doi:10.1016/s0040-4020(01)88354-x
    Return to citation in text: [1]
  26. Patrick, T. B.; Rogers, J.; Gorrell, K. Org. Lett. 2002, 4, 3155–3156. doi:10.1021/ol026512v
    Return to citation in text: [1]
  27. Patrick, T. B.; Gorrell, K.; Rogers, J. J. Fluorine Chem. 2007, 128, 710–713. doi:10.1016/j.jfluchem.2007.03.010
    Return to citation in text: [1]
  28. Jin, F.-q.; Xu, Y.-y.; Huang, W.-y. J. Fluorine Chem. 1995, 71, 1–4. doi:10.1016/0022-1139(94)03186-4
    Return to citation in text: [1]
  29. Shi, G.-q.; Schlosser, M. Tetrahedron 1993, 49, 1445–1456. doi:10.1016/s0040-4020(01)90196-6
    Return to citation in text: [1]
  30. Amii, H.; Kobayashi, T.; Terasawa, H.; Uneyama, K. Org. Lett. 2001, 3, 3103–3105. doi:10.1021/ol0163631
    Return to citation in text: [1]
  31. Hanamoto, T.; Korekoda, K.; Nakata, K.; Handa, K.; Koga, Y.; Kondo, M. J. Fluorine Chem. 2002, 118, 99–101. doi:10.1016/s0022-1139(02)00198-7
    Return to citation in text: [1]
  32. Sridhar, M.; Leela Krishna, K.; Madhusudana Rao, J. Tetrahedron 2000, 56, 3539–3545. doi:10.1016/s0040-4020(00)00242-8
    Return to citation in text: [1] [2]
  33. de Meijere, A.; Teichmann, S.; Seyed-Mahdavi, F.; Kohlstruk, S. Liebigs Ann. 1996, 1989–2000. doi:10.1002/jlac.199619961208
    Return to citation in text: [1]
  34. Ito, H.; Saito, A.; Taguchi, T. Tetrahedron: Asymmetry 1998, 9, 1979–1987. doi:10.1016/s0957-4166(98)00195-5
    Return to citation in text: [1]
  35. Arany, A.; Crowley, P. J.; Fawcett, J.; Hursthouse, M. B.; Kariuki, B. M.; Light, M. E.; Moralee, A. C.; Percy, J. M.; Salafia, V. Org. Biomol. Chem. 2004, 2, 455–465. doi:10.1039/b314314g
    Return to citation in text: [1]
  36. Crowley, P. J.; Percy, J. M.; Stansfield, K. Tetrahedron Lett. 1996, 37, 8237–8240. doi:10.1016/0040-4039(96)01877-1
    Return to citation in text: [1]
  37. Yamada, S.; Noma, M.; Konno, T.; Ishihara, T.; Yamanaka, H. Org. Lett. 2006, 8, 843–845. doi:10.1021/ol052897g
    Return to citation in text: [1]
  38. Yamada, S.; Hondo, K.; Konno, T.; Ishihara, T. RSC Adv. 2016, 6, 28458–28469. doi:10.1039/c6ra00569a
    Return to citation in text: [1]
  39. Shastin, A. V.; Nenajdenko, V. G.; Muzalevskiy, V. M.; Balenkova, E. S.; Fröhlich, R.; Haufe, G. Tetrahedron 2008, 64, 9725–9732. doi:10.1016/j.tet.2008.07.097
    Return to citation in text: [1]
  40. Crowley, P. J.; Percy, J. M.; Stansfield, K. Tetrahedron Lett. 1996, 37, 8233–8236. doi:10.1016/0040-4039(96)01876-x
    Return to citation in text: [1]
  41. Petrov, V.; Marchione, A. A.; Dooley, R. Chem. Commun. 2018, 54, 9298–9300. doi:10.1039/c8cc05075a
    Return to citation in text: [1]
  42. Chanteau, F.; Essers, M.; Plantier-Royon, R.; Haufe, G.; Portella, C. Tetrahedron Lett. 2002, 43, 1677–1680. doi:10.1016/s0040-4039(02)00107-7
    Return to citation in text: [1]
  43. Baum, K.; Archibald, T. G.; Tzeng, D.; Gilardi, R.; Flippen-Anderson, J. L.; George, C. J. Org. Chem. 1991, 56, 537–539. doi:10.1021/jo00002a011
    Return to citation in text: [1]
  44. Percy, J. M.; Stansfield, K.; Crowley, P. J.; Stansfield, K. Chem. Commun. 1997, 2033–2034. doi:10.1039/a704717g
    Return to citation in text: [1]
  45. Petrov, V.; Dooley, R. J.; Marchione, A. A.; Diaz, E. L.; Clem, B. S. J. Fluorine Chem. 2019, 225, 1–10. doi:10.1016/j.jfluchem.2019.06.003
    Return to citation in text: [1]
  46. Ernet, T.; Maulitz, A. H.; Würthwein, E.-U.; Haufe, G. J. Chem. Soc., Perkin Trans. 1 2001, 1929–1938. doi:10.1039/b102684b
    Return to citation in text: [1]
  47. Bogachev, A. A.; Kobrina, L. S.; Meyer, O. G. J.; Haufe, G. J. Fluorine Chem. 1999, 97, 135–143. doi:10.1016/s0022-1139(99)00040-8
    Return to citation in text: [1]
  48. Konev, A. S.; Khlebnikov, A. F. Collect. Czech. Chem. Commun. 2008, 73, 1553–1611. doi:10.1135/cccc20081553
    Return to citation in text: [1]
  49. Yanai, H.; Taguchi, T. Eur. J. Org. Chem. 2011, 5939–5954. doi:10.1002/ejoc.201100495
    Return to citation in text: [1]
  50. Motornov, V. A.; Muzalevskiy, V. M.; Tabolin, A. A.; Novikov, R. A.; Nelyubina, Y. V.; Nenajdenko, V. G.; Ioffe, S. L. J. Org. Chem. 2017, 82, 5274–5284. doi:10.1021/acs.joc.7b00578
    Return to citation in text: [1]
  51. Motornov, V. A.; Tabolin, A. A.; Novikov, R. A.; Nelyubina, Y. V.; Nenajdenko, V. G.; Ioffe, S. L. Org. Chem. Front. 2018, 5, 2588–2594. doi:10.1039/c8qo00623g
    Return to citation in text: [1]
  52. Motornov, V. A.; Tabolin, A. A.; Nelyubina, Y. V.; Nenajdenko, V. G.; Ioffe, S. L. Org. Biomol. Chem. 2019, 17, 1442–1454. doi:10.1039/c8ob03126f
    Return to citation in text: [1]
  53. Motornov, V. A.; Tabolin, A. A.; Nelyubina, Y. V.; Nenajdenko, V. G.; Ioffe, S. L. Org. Biomol. Chem. 2020, 18, 1436–1448. doi:10.1039/c9ob02668a
    Return to citation in text: [1]
  54. Aldoshin, A. S.; Tabolin, A. A.; Ioffe, S. L.; Nenajdenko, V. G. Eur. J. Org. Chem. 2018, 3816–3825. doi:10.1002/ejoc.201800385
    Return to citation in text: [1]
  55. Motornov, V. A.; Tabolin, A. A.; Novikov, R. A.; Nelyubina, Y. V.; Ioffe, S. L.; Smolyar, I. V.; Nenajdenko, V. G. Eur. J. Org. Chem. 2017, 6851–6860. doi:10.1002/ejoc.201701338
    Return to citation in text: [1]
  56. Aldoshin, A. S.; Tabolin, A. A.; Ioffe, S. L.; Nenajdenko, V. G. Eur. J. Org. Chem. 2019, 4384–4396. doi:10.1002/ejoc.201900573
    Return to citation in text: [1]
  57. Larkovich, R. V.; Ponomarev, S. A.; Aldoshin, A. S.; Tabolin, A. A.; Ioffe, S. L.; Nenajdenko, V. G. Eur. J. Org. Chem. 2020, 2479–2492. doi:10.1002/ejoc.202000054
    Return to citation in text: [1]
  58. Flid, V. R.; Gringolts, M. L.; Shamsiev, R. S.; Finkelshtein, E. S. Russ. Chem. Rev. 2018, 87, 1169–1205. doi:10.1070/rcr4834
    Return to citation in text: [1] [2]
  59. Pretsch, E.; Bühlmann, P.; Badertscher, M. Structure Determination of Organic Compounds, Tables of Spectral Data; Springer-Verlag: Berlin, Heidelberg, 2009; p 174. doi:10.1007/978-3-540-93810-1
    Return to citation in text: [1]
  60. Jasiński, R.; Kwiatkowska, M.; Barański, A. J. Phys. Org. Chem. 2011, 24, 843–853. doi:10.1002/poc.1853
    Return to citation in text: [1] [2]
  61. De Tollenaere, C.; Ghosez, L. Tetrahedron 1997, 53, 17127–17138. doi:10.1016/s0040-4020(97)10126-0
    Return to citation in text: [1]
  62. Sridhar, M.; Krishna, K. L.; Srinivas, K.; Rao, J. M. Tetrahedron Lett. 1998, 39, 6529–6532. doi:10.1016/s0040-4039(98)01357-4
    Return to citation in text: [1]
  63. Smart, B. E. J. Org. Chem. 1973, 38, 2027–2035. doi:10.1021/jo00951a013
    Return to citation in text: [1]
  64. Lee, C.; Yang, W.; Parr, R. G. Phys. Rev. B 1988, 37, 785–789. doi:10.1103/physrevb.37.785
    Return to citation in text: [1]
  65. Becke, A. D. J. Chem. Phys. 1993, 98, 5648–5652. doi:10.1063/1.464913
    Return to citation in text: [1]
  66. Vosko, S. H.; Wilk, L.; Nusair, M. Can. J. Phys. 1980, 58, 1200–1211. doi:10.1139/p80-159
    Return to citation in text: [1]
  67. Zhao, Y.; Truhlar, D. G. Theor. Chem. Acc. 2008, 120, 215–241. doi:10.1007/s00214-007-0310-x
    Return to citation in text: [1]
  68. Tomasi, J.; Mennucci, B.; Cancès, E. J. Mol. Struct.: THEOCHEM 1999, 464, 211–226. doi:10.1016/s0166-1280(98)00553-3
    Return to citation in text: [1]
  69. Pieniazek, S. N.; Clemente, F. R.; Houk, K. N. Angew. Chem., Int. Ed. 2008, 47, 7746–7749. doi:10.1002/anie.200801843
    Return to citation in text: [1]
  70. Mardirossian, N.; Head-Gordon, M. Mol. Phys. 2017, 115, 2315–2372. doi:10.1080/00268976.2017.1333644
    Return to citation in text: [1] [2]
  71. Levandowski, B. J.; Houk, K. N. J. Org. Chem. 2015, 80, 3530–3537. doi:10.1021/acs.joc.5b00174
    Return to citation in text: [1]
  72. Opoku, E.; Tia, R.; Adei, E. J. Phys. Org. Chem. 2019, 32, e3992. doi:10.1002/poc.3992
    Return to citation in text: [1]
  73. McQuarrie, D. A.; Simon, J. D. Physical chemistry: a molecular approach; Sterling Publishing Company, Inc.: New York, NY, USA, 1997.
    Return to citation in text: [1]
  74. Eyring, H. J. Chem. Phys. 1935, 3, 107–115. doi:10.1063/1.1749604
    Return to citation in text: [1]
  75. Mardirossian, N.; Head-Gordon, M. J. Chem. Theory Comput. 2016, 12, 4303–4325. doi:10.1021/acs.jctc.6b00637
    Return to citation in text: [1]
  76. Fringuelli, F.; Taticchi, A. Dienes in the Diels–Alder reaction; John Wiley & Sons: Chichester, UK, 1990; p 179.
    Return to citation in text: [1]
  77. Chang, R. Physical Chemistry for the Biosciences; University Science Books: USA, 2005; pp 338–342.
    Return to citation in text: [1]
  78. Cadogan, J. I. G.; Cameron, D. K.; Gosney, I.; Millar, J. R. A.; Newlands, S. F.; Reed, D. J. Chem. Soc., Perkin Trans. 2 1996, 2309–2317. doi:10.1039/p29960002309
    Return to citation in text: [1]
  79. Afarinkia, K.; Mahmood, F. Tetrahedron Lett. 2000, 41, 1287–1290. doi:10.1016/s0040-4039(99)02270-4
    Return to citation in text: [1]
  80. Lorenzo, J.; Delgado, A.; Montaña, Á. M.; Mesas, J. M.; Alegre, M.-T.; Rodríguez, M. d. C.; Avilés, F.-X. Eur. J. Med. Chem. 2014, 83, 374–388. doi:10.1016/j.ejmech.2014.06.042
    Return to citation in text: [1]
  81. Kiss, L.; Nonn, M.; Sillanpää, R.; Haukka, M.; Fustero, S.; Fülöp, F. Chem. – Asian J. 2016, 11, 3376–3381. doi:10.1002/asia.201601046
    Return to citation in text: [1]
  82. Das, M.; Du, Y.; Ribeiro, O.; Hariharan, P.; Mortensen, J. S.; Patra, D.; Skiniotis, G.; Loland, C. J.; Guan, L.; Kobilka, B. K.; Byrne, B.; Chae, P. S. J. Am. Chem. Soc. 2017, 139, 3072–3081. doi:10.1021/jacs.6b11997
    Return to citation in text: [1]
Other Beilstein-Institut Open Science Activities