Next Article in Journal
Preparative Separation and Purification of Four Glycosides from Gentianae radix by High-Speed Counter-Current Chromatography and Comparison of Their Anti-NO Production Effects
Next Article in Special Issue
Synthesis of Non-Racemic Pyrazolines and Pyrazolidines by [3+2] Cycloadditions of Azomethine Imines
Previous Article in Journal
Different Inhibitory Potencies of Oseltamivir Carboxylate, Zanamivir, and Several Tannins on Bacterial and Viral Neuraminidases as Assessed in a Cell-Free Fluorescence-Based Enzyme Inhibition Assay
Previous Article in Special Issue
Synthesis and Structure–Activity Relationships of 4-Morpholino-7,8-Dihydro-5H-Thiopyrano[4,3-d]pyrimidine Derivatives Bearing Pyrazoline Scaffold
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

19F‐NMR Diastereotopic Signals in Two N-CHF2 Derivatives of (4S,7R)-7,8,8-Trimethyl-4,5,6,7-tetrahydro-4,7-methano-2H-indazole †

1
Departamento de Química Orgánica y Bio-Orgánica, Facultad de Ciencias, UNED, Senda del Rey 9, E-28040 Madrid, Spain
2
Instituto de Química Médica, CSIC, Juan de la Cierva 3, E-28006 Madrid, Spain
*
Authors to whom correspondence should be addressed.
Dedicated to our friend V. A. Ostrovskii of the St. Petersburg State Institute of Technology, Russia, on the occasion of his 70th birthday.
Molecules 2017, 22(11), 2003; https://doi.org/10.3390/molecules22112003
Submission received: 2 November 2017 / Revised: 14 November 2017 / Accepted: 16 November 2017 / Published: 17 November 2017
(This article belongs to the Special Issue Pyrazole Derivatives)

Abstract

:
In this paper, we report the anisochrony of the fluorine atoms of a CHF2 group when linked to a pyrazole ring. The pyrazole is part of (4S,7R)-7,8,8-trimethyl-4,5,6,7-tetrahydro-4,7-methano-2H-indazole also known as (4S,7R)-campho[2,3-c]pyrazole, which has two stereogenic centers. Gauge-Independent Atomic Orbital (GIAO)/Becke, 3-parameter, Lee-Yang-Parr (B3LYP)/6-311++G(d,f) calculated 19F chemical shifts of the minimum energy conformations satisfactorily agree with the experimental data. The energy differences between minima need to consider solvent effects (continuum model) to be satisfactorily reproduced.

Graphical Abstract

1. Introduction

Anisochrony in NMR is observed when a prochiral group is linked to a molecule possessing a stereogenic center. In these conditions, the studied nuclei became diastereotopic [1,2,3,4]. In the majority of cases, the literature reports concern 1H-NMR and often the protons of CH2X groups (e.g., benzyl groups) [5,6]. The phenomenon can be observed on the methyl groups of Me2X substituents (e.g., isopropyl groups), with both 1H- and 13C-NMR [7]. Much less common is the observation of the anisochrony of phenyl substituents in CPh2X groups, also with 1H- and 13C-NMR [8,9].
The observation of diastereotopic signals for other nuclei have been reported less often, but, for instance 31P [10,11,12,13,14,15,16,17,18] is much more common than for 15N, where only one example has been described [19]. Other seldom-explored nuclei are 2H [20], 3H [21], 7Li [22], and 17O [23].
In the present paper, we present our results concerning the observation of 19F diastereotopic signals. In 1957, anisochronous signals were already observed for F2BrC–C*HBrPh, before the phenomenon was clearly understood [24]. Since then, the phenomenon has been repeatedly described, mainly for CHF2 groups [25,26,27], but also for CRF2 groups [28,29] as well as CRAr2 (Ar = meta and para substituted with F atoms) and CR(CH2F)2 [30].
None of the examples reported in the preceding paragraph concern a chiral molecule containing an N-CHF2 substituent. There are many examples of azoles bearing a C-CHF2 substituent, mainly in agrochemistry [31,32,33], the field of N-CHF2 and N-CRF2 azoles is less studied although there are several articles dealing with the structures presented in Figure 1.
Imidazoles 1 and benzimidazoles 2 [34,35,36], pyrazoles 3 [37,38], indazoles 4 and 5 [35,39], benzotriazole 6 [34,35,36] were reported. Related compounds 912 with CXF2 substituents are described in reference [40].
The compounds we have prepared (Scheme 1) and studied, 13 and 14, are derivatives of (4S,7R)-7,8,8-trimethyl-4,5,6,7-tetrahydro-4,7-methano-2H-indazole also known as (4S,7R)-campho[2,3-c]pyrazole, a compound we have previously investigated [41,42,43,44].

2. Results and Discussion

2.1. Chemistry

As indicated in Scheme 1, compounds 13 and 14 were prepared for the first time by direct difluoromethylation of camphopyrazole 15 with sodium chlorodifluoroacetate (SCDA) [45], according to the Mehta and Greaney conditions [46] or by adding a phase transfer catalyst [47], in both cases using N,N′-dimethylformamide as solvent and K2CO3 as base. Both isomers were obtained in an 85:15 ratio (see Experimental Section). The only other paper where the N-substitution of 15 was reported (with 1,2-dichloroethane) yielded a 50:50 mixture of both isomers [48]. The structure elucidation of compounds 13 and 14 was based on the close correlation of the 13C chemical shifts of the pyrazole ring with those of a reference compound [48].

2.2. NMR Spectroscopy

In both configurational isomers, the fluorine atoms are diastereotopic, and two distinct signals were observed for each one. From the spectra (Figure 2 and Figure 3 and data given in Supplementary Materials), 2J(1H-19F) and 2J(19F-19F) coupling constants can be measured.
The 2JFF SSCC (spin-spin coupling constant) in F-C-F compounds is very sensitive to structural aspects, especially the C atom hybridization; for sp3 carbons range between 3.5 and 340 Hz [49]. There are no 2JFF values published for N-azolyl derivatives, and thus the values we have measured (about 225 Hz) are the only representatives of this kind of compound.
In 1H-NMR (see experimental part and Supplementary Material), the most interesting information concerning the CHF2 group where when the anisochrony is larger (compound 13) the two 2JHF couplings are different and when the anisochrony is smaller (compound 14) they are identical. Moreover, the signal of the 9-CH3 group in compound 13 shows a long-distance 6JHF coupling of 1.4 Hz (also measured in the 19F-NMR spectrum, see Figure 2); in compound 14, this coupling is not observed due to the additional bond (it would be a 7JHF).

2.3. Computational Results

We have calculated the energy of compounds 13 and 14 as a function of the torsion angle θ about the N-(CHF2) bond (defined as H-C-N1-N2, 30-29-7-3 or 30-29-3-7). There are two minima (0 imaginary frequencies)—one near 0° and the other near 180° (Figure 4).
According to the calculations, the 2-substituted isomer 14 is more stable than the 1-substituted isomer 13 by 10.8 kJ·mol–1 (both in their minima; i.e., having 0 imaginary frequencies). Note that in camphopyrazole, tautomer 2H is more stable than tautomer 1H [41,43,44] due to the Mills–Nixon effect [50,51]; once again, tautomerism and isomerism behave similarly.
When the energy was calculated as a function of the torsion angle θ about the N-(CHF2) bond, in both cases, the minimum energy conformation corresponds to θ = 0°; i.e., the H atom of the CHF2 group eclipsing the “pyridine-like” N atom of pyrazole, the so-called syn-periplanar conformation (Figure 5). The difference between the 0° and the 180° minima are for 13 15.7 kJ·mol–1 and for 14 11.8 kJ·mol–1, and the transition states are for 13 23.6 (θ = 104.4°) and 26.6 kJ·mol–1 (θ = 255.8°) and for 14 23.5 (θ = 114.9°) and 23.1 kJ·mol–1 (θ = 242.9°).
This conformational preference can most probably be explained by the dominance of vicinal hyperconjugation, with electron donation from the electron-rich sigma N-N bonding orbital into both of the very electron deficient vicinal C-F anti-bonding orbitals [52,53,54,55].
A natural bond orbital (NBO) analysis shows that the energetic difference between the conformations minima at 0° and 180° can be explained based on the stabilization due to the sum of the charge transfer between the lone pair of the pyridine-like nitrogen and the σ* C-H bond and between the σ N-N and the σ* C-F bonds. This stabilization amount is 6.6 kJ·mol–1 in the minima at 0° of 13 and 14, while in the minima at 180° it is between 1.1 and 1.0 kJ·mol–1, respectively.
Gauge-Independent Atomic Orbital (GIAO) calculated parameters (absolute shieldings) accounted for the experimental results obtained by multinuclear NMR (1H, 13C, 15N and 19F) (see Supplementary Materials). We will focus on the 19F chemical shifts (Table 1).
The four experimental values (−91.6, −89.2, −92.0, −90.8, ppm) are close to the calculated ones for 13 (0°) (−93.2, −89.7 ppm) and for 14 (0°) (−92.8, −88.7 ppm) than for the 180° assignment (−98.5, −90.1, −97.4, −85.7 ppm). Assuming the simplification that only the two minima contribute to the experimental values, a simple interpolation of the type Exp = a × (Calc. abs minima) + (1–a) × (Calc. second minima) lead to 13 = 91.8% of conformer θ ≈ 0° and 8.2% of conformer θ ≈ 180°, and 14 = 81.6% of conformer θ ≈ 0° and 18.4% of conformer θ ≈ 180°. This corresponds at 298.15 K to −6.0 and −3.7 kJ·mol−1, respectively—lower than the calculated differences between both rotamers, but of the same sign. To see if the inclusion of solvent effects improves the agreement, we calculated the differences of energy between minima in CHCl3 (Polarizable continuum model, PCM) obtaining for 13 and 14, −7.6 and −5.1 kJ·mol−1, respectively—much closer to the experimental results (the TS have very close values: 19.8 and 19.2 kJ·mol−1); the solvent slightly modifies the geometries, see θ values in Table 2.
We have calculated the chemical shifts in CHCl3, obtaining the values reported in Table 2. With these values, we have calculated that the difference of energies for 13 and 14 are −4.9 and −4.3 kJ·mol−1, respectively, comparable to those obtained for the gas phase (−6.0 and −3.7 kJ·mol−1) to be compared with −7.6 and −5.1 kJ·mol−1.
We have also calculated the 13C chemical shifts of the three carbon atoms of the pyrazole ring (C3, C3a, C7a named C4, C9, and C11 in Figure 3). The results are reported in Table 3 and correlates well with the experimental carbon signal shifts, and aided the assignment of the pyrazole ring carbons.

3. Experimental Section

3.1. Chemistry

General
All chemicals cited in the synthetic procedure are commercial compounds. Melting points were determined by differential scanning calorimetry (DSC) with a SEIKO DSC 220 C connected to a model SSC5200H disk station. Thermograms (sample size 0.003–0.005 g) were recorded with a scan rate of 5.0 °C. Column chromatography was performed on silica gel 60 (Merck KGaA, Darmstadt, Germany), 70–230 mesh), and elemental analyses using a Perkin-Elmer 240 apparatus (Madrid, Spain).
Preparation of (4S,7R)-1-(Difluoromethyl)-7,8,8-trimethyl-4,5,6,7-tetrahydro-4,7-methano-1H-indazole (13) and (4S,7R)-2-(Difluoromethyl)-7,8,8-trimethyl-4,5,6,7-tetrahydro-4,7-methano-1H -indazole (14).
Procedure A from Ref. [46]. Into a 100-mL round-bottom three-necked flask equipped with reflux condenser and magnetic stirring, 2 equivalents of sodium chlorodifluoroacetate (SCDA) and 1.5 equivalents of the base (K2CO3) were introduced. The vacuum was established for 15 min and then purged with argon for another 15 min (this process was repeated three times). Six milliliters of N,N-dimethylformamide (DMF) was added slowly with stirring and under an argon stream, and then 1 equivalent of (4S,7R)-7,8,8-trimethyl-4,5,6,7-tetrahydro-4,7-methano-2H-indazole (15) dissolved in 2 mL of DMF was added from an addition funnel over 15 min. The flask was immersed in a silicone bath previously heated to 100 °C and left stirring for 8 h. To control the temperature, a thermometer was used which was connected to the heating plate and immersed in the silicone oil bath. After the reaction time was completed, it was cooled to room temperature and EtOAc (15 mL) and water (15 mL) were added to the mixture. The organic fraction was washed with brine, and the aqueous fraction was extracted with EtOAc. The organic fractions were combined, dried over anhydrous MgSO4, and the solvent evaporated off. The yield of the reaction crude—in which both isomers are present in a ratio (85% of 13: 15% of 14)—is quantitative. The purification was carried out by column chromatography using dichloromethane/hexane (1:1) as eluent. Compound 14 was eluted first.
Procedure B from Ref [47]. Into a 100-mL round-bottom flask equipped with reflux condenser and magnetic stirring, 2 equivalents of SCDA, 3 equivalents of the base (K2CO3), 1 equivalent of (4S,7R)-7,8,8-trimethyl-4,5,6,7-tetrahydro-4,7-methano-2H-indazole (15), and 0.3 equivalents of tetraethylammonium bromide (TEAB) were dissolved in 10 mL of DMF and the mixture was stirred at 100 °C for 3 h. The resulting mixture was poured into water and extracted with EtOAc, the organic extract containing again an 85:15 mixture of both isomers (overall yield 90%) was treated as previously described in procedure A.
(4S,7R)-1-(Difluoromethyl)-7,8,8-trimethyl-4,5,6,7-tetrahydro-4,7-methano-1H-indazole (13). m.p.: 45.4 °C; 1H-NMR: (400.13 MHz, CDCl3) δ = 7.27 (s, H3), 7.14 (dd, 2JF = 59.5, 2JF = 60.5, CHF2), 2.81 (d, 3J = 3.8), 2.05 (cm, H5ec), 1.03 (cm, H5ax), 1.81(cm, H6ec), 1.18 (cm, H6ax), 1.37 (dd, 6JF = 1.4, CH3-9), 0.92 (s, CH3-10), 0.77 (s, CH3-11); 13C-NMR: (100.61 MHz, CDCl3) δ = 153.6 (dd, 3JF = 1.6, C7a), 134.3 (dd, 4JF = 2.3, C3), 132.1 (C3a), 111.6 (dd, 1JF = 246.0, 1JF = 248.7, CHF2), 63.2 (C8), 53.7 (C7), 47,6 (C4), 33.0 (C6), 27.4 (C5), 20.1 (CH3-11), 19.5 (CH3-10), 11.6 (dd, 5JF = 5JF = 1.4, CH3-9); 19F NMR: (376.50 MHz, CDCl3) δ = −89.16 (ddd, 2JF = 226.6, 2JH = 60.6, 6JH = 1.4), −91.64 (dd, 2JF = 226.6, 2JH = 59.4); 15N-NMR: (40.54 MHz, CDCl3) δ = −177.4 (dd, 2JF = 2JF = 27.9, N1), −79.9 (N2). Anal. calcd. for C12H16F2N2: C 63.70, H 7.13, N 12.38. Found: C 63.45, H 7.45, N 12.13.
(4S,7R)-2-(Difluoromethyl)-7,8,8-trimethyl-4,5,6,7-tetrahydro-4,7-methano-1H-indazole (14). m.p.: 40.7 °C; 1H-NMR: (400.13 MHz, CDCl3) δ = 7.28 (s, H3), 7.11 (dd, 2JF = 2JF = 60.9, CHF2), 2.79 (d, 3J = 4.1), 2.10 (cm, H5ec), 1.22 (cm, H5ax), 1.88 (cm, H6ec), 1.35 (cm, H6ax), 1.29 (s, CH3-9), 0.97 (s, CH3-10)), 0.65 (s, CH3-11); 13C-NMR: (100.61 MHz, CDCl3) δ = 169.1 (dd, 4JF = 4JF = 2.2, C7a), 130.2 (C3a), 117.9 (C3), 111.2 (dd, 1JF = 246.4, 1JF = 246.5, CHF2), 60.4 (C8), 50.1 (C7), 46.9 (C4), 33.3 (C6), 27.2 (C5), 20.4 (CH3-11), 18.9 (CH3-10), 10.4 (CH3-9); 19F-NMR: (376.50 MHz, CDCl3) δ = −90.80 (dd, 2JF = 225.5, 2JH = 61.3), −92.05 (dd, 2JF = 225.4, 2JH = 60.7); 15N-NMR: (40.54 MHz, CDCl3) δ = −177.2 (dd, 2JF = 2JF = 24.9, N2), N1 not detected. Anal. calcd. for C12H16F2N2: C 63.70, H 7.13, N 12.38. Found: C 63.37, H 7.48, N 11.98.

3.2. NMR

NMR spectra were recorded on a Bruker (Bruker Biospin GmbH, Rheinstetten, Germany) DRX 400 (9.4 Tesla, 400.13 MHz for 1H, 100.61 MHz for 13C and 40.54 MHz for 15N using a 5-mm inverse-detection H-X probe equipped with a z-gradient coil, at 300 K. Chemical shifts (δ in ppm) are given from internal solvent, CDCl3 7.26 for 1H and 77.0 for 13C and for 15N, nitromethane (0.00) was used as external reference. Signals were characterized as s (singlet), d (doublet), and cm (complex multiplet) and the J coupling constants are given in Hz.
Typical parameters for 1H-NMR spectra were spectral width 4800 Hz and pulse width 9.5 μs at an attenuation level of 0 dB. Typical parameters for 13C-NMR spectra were spectral width 21 kHz, pulse width 12.5 μs, at an attenuation level of −6 dB and relaxation delay 2 s, WALTZ-16 was used for broadband proton decoupling; the Free Induction Decays (FIDs) were multiplied by an exponential weighting (lb = 1 Hz) before Fourier transformation.
Inverse proton detected heteronuclear shift correlation spectra, (1H-13C) gs-HMQC, and (1H-13C) gs-HMBC were acquired and processed using standard Bruker NMR software and in non-phase-sensitive mode. Gradient selection was achieved through a 5% sine truncated shaped pulse gradient of 1 ms.
Selected parameters for (1H-13C) gs-HMQC and (1H-13C) gs-HMBC spectra were spectral width 4800 Hz for 1H and 20.5 kHz for 13C, 1024 × 256 data set, number of scans two (gs-HMQC) or four (gs-HMBC) and relaxation delay 1 s. The FIDs were processed using zero filling in the F1 domain and a sine-bell window function in both dimensions was applied prior to Fourier transformation. In the gs-HMQC experiments, Globally Optimized Alternating Phase Rectangular Pulse (GARP) modulation of 13C was used for decoupling. Selected parameters for (1H-15N) gs-HMQC, and (1H-15N) gs-HMBC spectra were spectral width 3500 Hz for 1H and 12.5 kHz for 15N, 1024 × 256 data set, number of scans four, relaxation delay 1 s, 37–60 ms delay for evolution of the 15N-1H long-range coupling. The FIDs were processed using zero filling in the F1 domain and a sine-bell window function in both dimensions was applied prior to Fourier transformation.
19F-NMR spectra were recorded on the same spectrometer (376.50 for 19F) using a 5 mm Quattro Nucleus Probe (QNP) direct-detection probehead equipped with a z-gradient coil, at 300 K. Chemical shifts (δ in ppm) are given from CFCl3 as external reference (one drop of CFCl3 in CDCl3 (0.00)). Typical parameters for 19F NMR spectra were spectral width of 55 kHz, pulse width of 13.75 µs at attenuation level of −6 dB and relaxation delay of 1 s. WALTZ-16 was used for broadband proton decoupling 19F{1H}, the FIDS were multiplied by an exponential weighting (lb = 1 Hz) before Fourier transformation.

3.3. Computational Details

Calculations were carried out at the B3LYP/6-311++G(d,p) level [56,57]. Subsequent frequency calculations verify that the structures obtained correspond to energetic minima (imaginary frequencies = 0) or to transition states (imaginary frequencies = 1). In the optimization process, the 0° and 180° angles get slightly modified (Table 1 and Table 2). These resulting geometries have been used for the calculation of the absolute chemical shieldings with the GIAO method [58,59]. Solvent effects were calculated within the PCM approximation (continuum model) [60,61,62]. All the calculations have been performed with the Gaussian-09 package [63].
Equations (1)–(4) [64,65,66] have been used to transform absolute shieldings into chemical shifts:
δ1H = 31.0 − 0.97 × σ1H, (reference TMS, 0.00 ppm)
δ13C = 175.7 − 0.963 × σ13C, (reference TMS, 0.00 ppm)
δ15N = −152.0 − 0.946 × σ15N, (reference TMS, 0.00 ppm)
δ19F = 162.1 − 0.959 × σ19F, (reference CFCl3, 0.00 ppm)
The natural bond orbital (NBO) method [67] has been used to obtain the stabilizing charge-transfer interactions in complexes using the NBO-6 program [68].

4. Conclusions

In summary, we have found a new and original example of diastereotopic fluorine atoms, measured two values of 2JFF in an original environment and successfully carried out GIAO/B3LYP/6-311++G(d,p) calculations of 19F chemical shifts that agree with the calculated energies of the two minima of the potential energy curve when solvent was taken into account.

Supplementary Materials

Supplementary materials are available online: Tables S1–S3 and Figures S1–S14.

Acknowledgments

This work has been supported by the Spanish Ministerio de Economía, Industria y Competitividad (CTQ2012-35513-C02-02, CTQ2014-56833-R) and Comunidad Autónoma de Madrid (S2013/MIT-2841, Fotocarbon). Computer, storage, and other resources from the CTI (CSIC) are gratefully acknowledged.

Author Contributions

J.E. conceived and C.L. designed the experiments; D.G.-P. performed the experiments; C.L. and R.M.C. analyzed the data; I.A. contributed with the computational calculations; R.M.C. and J.E. wrote the paper.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Devriese, G.; Ottinger, R.; Zimmermann, D.; Reisse, J.; Mislow, K. On the sign of the anisochrony of diastereotopic groups. Bull. Soc. Chim. Belg. 1976, 85, 167–178. [Google Scholar] [CrossRef]
  2. Nasipuri, D. Stereochemistry of Organic Compounds. Principles and Applications, 2nd ed.; New Age International: New Dehli, India, 1994. [Google Scholar]
  3. Eliel, E.L.; Wilen, S.H. Stereochemistry of Organic Compounds; John Wiley & Sons: New York, NY, USA, 1994. [Google Scholar]
  4. Alkorta, I.; Elguero, J. Essential versus accidental isochrony of diastereotopic nuclei in NMR spectroscopy. Struct. Chem. 2016, 27, 671–679. [Google Scholar] [CrossRef]
  5. Elguero, J.; Marzin, C.; Tizané, D. Nuclear Magnetic Resonance of Asymmetrical Nitrogen in a Pentagonal Ring—1,2-Dimethylpyrazolidine. Tetrahedron Lett. 1969, 10, 513–514. [Google Scholar] [CrossRef]
  6. Elguero, J.; Marzin, C.; Tizané, D. Magnetic Non-equivalence and Nitrogen Inversion in a Series of Dinitrogen Pentagonal Herocycles, 2- and 3-Pyrazolines, Pyrazolidines and Pyrazolidones. Org. Magn. Reson. 1969, 1, 249–275. [Google Scholar] [CrossRef]
  7. Zvilichovsky, G. The Effect of Solvent on Chemical-shift Non-equivalence of Diastereotopic Geminal Nuclei in (pro)1-Chiral N,N-Disubstituted 5-Oxo-4-phenyl-2.5-dihydroisoxazol-2-ium-3-olates. J. Chem. Soc. Perkin Trans. 1988, 2, 2015–2019. [Google Scholar] [CrossRef]
  8. Molina, P.; Alajarín, M.; López-Leonardo, C.; Hernandez Cano, F.; Llamas-Saiz, A.L.; Concepcion Foces-Foces, C.; Rosa Maria Claramunt, R.M.; Elguero, J. 2,4-Bisimino-1,3-diazetidines: Iminophosphoranes, Carbodiimides and Related Betaines. J. Chem. Soc. Perkin Trans. 1992, 1, 199–210. [Google Scholar] [CrossRef]
  9. Fowler, K.G.; Littlefield, S.L.; Baird, M.C. Synthesis, Structures, and Properties of the Phosphonium-1-indenylide (PHIN) Ligands 1-C9H6PPh3, 1-C9H6PMePh2, and 1-C9H6PMe2Ph and of the Corresponding Ruthenium(II) Complexes [Ru(η5-C5H5)(η5-PHIN)]PF6. Organometallics 2011, 30, 6098–6107. [Google Scholar] [CrossRef]
  10. Powell, G.L.; Jacobus, J. The Nonequivalence of the Phosphorus Atoms in Cardiolipin. Biochemistry 1974, 13, 4024–4026. [Google Scholar] [CrossRef] [PubMed]
  11. Pastor, S.D.; Shum, S.P.; Rodebaugh, R.K.; Debellis, A.D.; Clarke, F.H. Sterically Congested Phosphite Ligands: Synthesis, Crystallographic Characterization, and Observation of Unprecedented Eight-Bond 31P, 31P Coupling in the 31P-NMR Spectra. Helv. Chim. Acta 1993, 76, 900–914. [Google Scholar] [CrossRef]
  12. King, S.A.; DiMichele, L. Multinuclear NMR Observation of a Ru(II)-BINAP Catalyst and Possible Intermediates in the Reduction of Ketoesters. In Catalysis of Organic Reactions; Scaros, M.G., Prunier, M.L., Eds.; Marcel Dekker: New York, NY, USA, 1995. [Google Scholar]
  13. Brunel, J.M.; Faure, B. A New 31P NMR Method for the Enantiomeric Excess Determination of Diols and Secondary Diamines with C2 Symmetry. Tetrahedron Asymmetry 1995, 6, 2353–2356. [Google Scholar] [CrossRef]
  14. Reich, H.J.; Kulicke, K.J. Dynamics of Solvent Exchange in Organolithium Reagents. Lithium as a Center of Chirality. J. Am. Chem. Soc. 1996, 118, 273–274. [Google Scholar] [CrossRef]
  15. Yamada, I.; Ohkouchi, M.; Yamaguchi, M.; Yamagishi, T. Asymmetric Hydrogenation of Acrylic Acid Derivatives by Novel Chiral Rhodium-Phosphinediamine Complex Catalysts by Selective Ligation Between Two Amino Units of the Ligand and Electrostatic Interaction. J. Chem. Soc. Perkin Trans. 1997, 1, 1869–1873. [Google Scholar] [CrossRef]
  16. Bilge, S.; Demiriz, S.; Okumus, A.; Kilic, Z.; Tercan, B.; Kökelek, T.; Buyukgungör, O. Phosphorus-Nitrogen Compounds. Part 13. Syntheses, Crystal Structures, Spectroscopic, Stereogenic, and Anisochronic Properties of Novel Spiro-Ansa-Spiro-, and Spiro-Crypta Phosphazene Derivatives. Inorg. Chem. 2006, 45, 8755–8767. [Google Scholar] [CrossRef] [PubMed]
  17. Pregosin, P.S. 31P- and 13C-NMR Studies on Metal Complexes of Phosphorus-donors: Recognizing Surprises. Coord. Chem. Rev. 2008, 252, 2156–2170. [Google Scholar] [CrossRef]
  18. Kruck, M.; Munoz, M.P.; Bishop, H.L.; Frost, C.G.; Chapman, C.J.; Kociok-Köhn, G.; Butts, C.P.; Lloyd-Jones, G.C. BINOL-3,3’-Triflone N,N-Dimethyl Phosphoramidites Through-Space 19F, 31P Spin-Spin Coupling with a Remarkable Dependency on Temperature and Solvent Internal Pressure. Chem. Eur. J. 2008, 14, 7808–7812. [Google Scholar] [CrossRef] [PubMed]
  19. Alkorta, I.; Dardonville, C.; Elguero, J. Observation of Diastereotopic Signals in 15N-NMR Spectroscopy. Angew. Chem. Int. Ed. 2015, 54, 3997–4000. [Google Scholar] [CrossRef] [PubMed]
  20. Richards, J.C.; Spenser, I.D. 2H-NMR Spectroscopy as a Probe of the Stereochemistry of Enzymic Reactions at Prochiral Centres. Tetrahedron 1983, 39, 3549–3568. [Google Scholar] [CrossRef]
  21. Allen, B.D.; Cintrat, J.C.; Faucher, N.; Berthault, P.; Rousseau, B.; O'Leary, D.J. An Isosparteine Derivative for Stereochemical Assignment of Stereogenic (Chiral) Methyl Groups Using Tritium NMR: Theory and Experiment. J. Am. Chem. Soc. 2005, 127, 412–420. [Google Scholar] [CrossRef] [PubMed]
  22. Salomone, A.; Perna, F.M.; Falcicchio, A.; Nilsson Lill, S.O.; Moliterni, A.; Michel, R.; Florio, S.; Stalke, D.; Capriati, V. Direct Observation of a Lithiated Oxirane: A Synergistic Study Using Spectroscopic, Crystallographic, and Theoretical Methods on the Structure and Stereodynamics of Lithiated ortho-Trifluoromethyl Styrene Oxide. Chem. Sci. 2014, 5, 528–538. [Google Scholar] [CrossRef]
  23. Powers, T.A.; Evans, S.A., Jr. Lanthanide Induced 17O NMR Shifts of Diastereotopic Oxygen Atoms in 1-Thiodecalin 1,1-Dioxide and Related Compounds. Tetrahedron Lett. 1990, 31, 5835–5838. [Google Scholar] [CrossRef]
  24. Drysdale, J.J.; Phillips, W.D. Restricted Rotation in Substituted Ethanes as Evidenced by Nuclear Magnetic Resonance. J. Am. Chem. Soc. 1957, 79, 319–322. [Google Scholar] [CrossRef]
  25. Xu, Y.; Tang, P.; Firestone, L.; Zhang, T.T. 19F Nuclear Magnetic Resonance Investigation of Stereoselective Binding of Isoflurane to Bovine Serum Albumin. Biophys. J. 1996, 70, 532–538. [Google Scholar] [CrossRef]
  26. Vaughan, M.D.; Cleve, P.; Robinson, V.; Duewel, H.S.; Honek, J.F. Difluoromethionine as a Novel 19F-NMR Structural Probe for Internal Amino Acid Packing in Proteins. J. Am. Chem. Soc. 1999, 121, 8475–8478. [Google Scholar]
  27. Qiu, X.L.; Qing, F.L. Synthesis of cis-4-Trifluoromethyl- and cis-4-Difluoromethyl-l-pyroglutamic Acids. J. Org. Chem. 2003, 68, 3614–3617. [Google Scholar] [CrossRef] [PubMed]
  28. Marchione, A.A.; Buck, R.C. Complete Multinuclear Magnetic Resonance Characterization of a Set of Polyfluorinated Acids and Alcohols. Magn. Reson. Chem. 2009, 47, 194–198. [Google Scholar] [CrossRef] [PubMed]
  29. El Dine, A.N.; Khalaf, A.; Grée, D.; Tasseau, O.; Fares, F.; Jaber, N.; Lesot, P.; Hachem, A.; Grée, R. Synthesis of Enones, Pyrazolines and Pyrrolines with gem-Difluoroalkyl Side Chains. Beilstein J. Org. Chem. 2013, 9, 1943–1948. [Google Scholar] [CrossRef] [PubMed]
  30. Pike, S.J.; De Poli, M.; Zawodny, W.; Raftery, J.; Webb, S.J.; Clayden, J. Diastereotopic Fluorine Substituents as 19F NMR Probes of Screw-sense Preference in Helical Foldamers. Org. Biomol. Chem. 2013, 11, 3168–3176. [Google Scholar] [CrossRef] [PubMed]
  31. Mykhailiuk, P.K. In Situ Generation of Difluoromethyl Diazomethane for [3 + 2] Cycloadditions with Alkynes. Angew. Chem. Int. Ed. 2015, 54, 6558–6561. [Google Scholar] [CrossRef] [PubMed]
  32. Du, S.; Tian, Z.; Yang, D.; Li, X.; Li, H.; Jia, C.; Che, C.; Wang, M.; Qin, Z. Synthesis, Antifungal Activity and Structure-Activity Relationships of Novel 3-(Difluoromethyl)-1-methyl-1H-pyrazole-4-carboxylic Acid Amides. Molecules 2015, 20, 8395–8408. [Google Scholar] [CrossRef] [PubMed]
  33. Han, W.Y.; Zhao, J.; Wang, J.S.; Xiang, G.Y.; Zhang, D.L.; Bai, M.; Cui, B.D.; Wan, N.W.; Chen, Y.Z. Diastereoselective [3 + 2] cycloaddition of 3-ylideneoxindoles with in situ generated CF2HCHN2: Syntheses of CF2H-containing spirooxindoles. Org. Biomol. Chem. 2017, 15, 5571–5578. [Google Scholar] [CrossRef] [PubMed]
  34. Wang, F.; Weizhou, H.; Jinbo, H. Difluoromethylation of O-, S-, N-, C-Nucleophiles Using Difluoromethyltri(n-butyl)ammonium Chloride as a New Difluorocarbene Source. Chin. J. Chem. 2011, 29, 2717–2721. [Google Scholar] [CrossRef]
  35. Thomoson, C.S.; Wang, L.; Dolbier, W.R. Use of Fluoroform as a Source of Difluorocarbene in the Synthesis of N-CF2H Heterocycles and Difluoromethoxypyridines. J. Fluor. Chem. 2014, 168, 34–39. [Google Scholar] [CrossRef]
  36. Prakash, G.K.; Krishnamoorthy, S.; Ganesh, S.K.; Kulkarni, A.; Haiges, R.; Olah, G.A. N-Difluoromethylation of Imidazoles and Benzimidazoles Using the Ruppert_Prakash Reagent under Neutral Conditions. Org. Lett. 2014, 16, 54–57. [Google Scholar] [CrossRef] [PubMed]
  37. Morimoto, K.; Makino, K.; Yamamoto, S.; Sakata, G. Synthesis of Fluoromethyl, Difluoromethyl and Trifluoromethyl Analogues of Pyrazosulfuron-ethyl as Herbicides. J. Heterocycl. Chem. 1990, 27, 807–810. [Google Scholar] [CrossRef]
  38. Petko, K.I.; Sokolenko, T.M.; Yagulpolskii, L.M. Chemical Properties of Derivatives of N-Difluoromethyl- and N-2H-Tetrafluoroethylpyrazoles. Chem. Heterocycl. Comp. 2006, 42, 1177–1184. [Google Scholar] [CrossRef]
  39. Pelc, M.; Huang, W.; Trujillo, J.; Baldus, J.; Turner, S.; Kleine, P.; Yang, S.; Thorarensen, A. An Efficient and Regioselective Difluoromethylation of 3-Iodoindazole with Chlorodifluoromethane. Synlett 2010, 219–222. [Google Scholar] [CrossRef]
  40. Yagupolskii, L.M.; Fedyuk, D.V.; Petko, K.I.; Troitskaya, V.I.; Rudyk, V.I.; Rudyuk, V.V. N-Trihalomethyl Derivatives of Benzimidazole, Benzotriazole and Indazole. J. Fluor. Chem. 2000, 106, 181–187. [Google Scholar] [CrossRef]
  41. Llamas-Saiz, A.; Foces-Foces, C.; Sobrados, I.; Elguero, J.; Meutermans, W. (4S,7R)-7,8,8-Trimethyl-4,5,6,7-Tetrahydro-4,7-methano-1H(2H)-indazole (Campho[2,3-c]Pyrazole): Comparison Between the X-Ray Structure and Carbon-13 NMR Data in the Solid State. Acta Crystallogr. Sect. C 1993, 49, 724–729. [Google Scholar] [CrossRef]
  42. Yap, G.P.A.; Claramunt, R.M.; López, C.; García, M.A.; Pérez-Nedina, C.; Alkorta, I.; Elguero, J. The Structures of Chiral and Racemate Campho[2,3-c]pyrazole: A Combined Crystallographic, Solid-State NMR and Computational Study. J. Mol. Struct. 2010, 965, 74–81. [Google Scholar] [CrossRef]
  43. Quesada-Moreno, M.M.; Avilés-Moreno, J.R.; López-González, J.J.; Claramunt, R.M.; López, C.; Alkorta, I.; Elguero, J. Chiral Self-Assembly of Enantiomerically Pure (4S,7R)-Campho[2,3-c]pyrazole in the Solid State: A Vibrational Circular Dichroism (VCD) and Computational Study. Tetrahedron Asymmetry 2014, 25, 507–515. [Google Scholar] [CrossRef]
  44. Webber, A.L.; Emsley, L.; Claramunt, R.M.; Brown, S.P. NMR Crystallography of Campho[2,3-c]pyrazole (Z’=6): Combining High-Resolution 1H-13C Solid-State MAS NMR Spectroscopy and GIPAW Chemical-Shift Calculations. J. Phys. Chem. A 2010, 114, 10435–10442. [Google Scholar] [CrossRef] [PubMed]
  45. Birchall, J.M.; Cross, G.W.; Haszeldine, R.N. Difluorocarbene. Proc. Chem. Soc. 1960, 81, 37–96. [Google Scholar]
  46. Mehta, V.P.; Greaney, M.F. S-, N-, and Se-Difluoromethylation Using Sodium Chlorodifluoroacetate. Org. Lett. 2013, 15, 5036–5039. [Google Scholar] [CrossRef] [PubMed]
  47. Wang, W.; Hua, M.; Huang, Y.; Zhang, Q.; Zhang, X.; Wu, J. Difluoromethylation of 2-Hydroxychalcones Using Sodium 2-Chloro-2,2-difluoroacetate as Difluoromethylating Agent. Chem. Res. Chin. Univ. 2015, 31, 362–366. [Google Scholar] [CrossRef]
  48. Cabildo, P.; Claramunt, R.M.; Cornago, P.; Lavandera, J.L.; Sanz, D.; Jagerovic, N.; Jimeno, M.L.; Elguero, J.; Gilles, I.; Aubagnac, J.L. Synthesis, structure (NMR and mass spectrometry) and conformational analysis of heterocyclic analogues of dibenzo[a,e]cycloocta-l,5-diene: 5,6,12,13-tetrahydrobispyrazolo[1,2-a:1’,2’-e] [1,2,5,6] tetraazocinediium dihalides. J. Chem. Soc. Perkin Trans. 1996, 2, 701–711. [Google Scholar] [CrossRef]
  49. Berger, S.; Braun, S.; Kalinowski, H.O. NMR Spectroscopy of the Non-Elements; Wiley: New York, NY, USA, 1997; p. 635. [Google Scholar]
  50. Martínez, A.; Jimeno, M.L.; Elguero, J.; Fruchier, A. An Experimental Approach to the Mills-Nixon Effect: Tautomerism of 3(5),4-Polymethylenepyrazoles. New J. Chem. 1994, 18, 269–277. [Google Scholar]
  51. Alkorta, I.; Elguero, J. Tautomerism and the Mills-Nixon Effect. Struct. Chem. 1997, 8, 189–195. [Google Scholar] [CrossRef]
  52. Brand, D.J.; Steenkamp, J.A.; Brandt, E.V.; Takeuchi, Y. Conformational studies of (−)-epicatechin-Mosher ester. Tetrahedron Lett. 2007, 48, 2769–2773. [Google Scholar] [CrossRef]
  53. Brand, D.J.; Steenkamp, J.A.; Omata, K.; Kabuto, K.; Fujiwara, T.; Takeuchi, Y. The Origin of an Unusually Large 19F Chemical Shift Difference Between the Diastereomeric α-Cyano-α-Fluoro-p-Tolylacetic Acid (CFTA) Esters of 3’,4’,5,7-Tetra-O-Methylepicatechin. Chirality 2008, 20, 351–356. [Google Scholar] [CrossRef] [PubMed]
  54. Chen, Z.; Corminboeuf, C.; Mo, Y. Direct Evaluation of the Hyperconjugative Interactions in 1,1,1-Trihaloethane (CH3CX3, X = F, Cl, and Br). J. Phys. Chem. A 2014, 118, 5743–5747. [Google Scholar] [CrossRef] [PubMed]
  55. Baranac-Stojanovic, M. Theoretical Analysis of the Rotational Barrier in Ethane: Cause and Consequences. Struct. Chem. 2015, 26, 989–996. [Google Scholar] [CrossRef]
  56. Becke, A.D. Density-Functional Exchange-Energy Approximation with Correct Asymptotic Behavior. Phys. Rev. A 1988, 38, 3098–3100. [Google Scholar] [CrossRef]
  57. Frisch, M.J.; Pople, J.A. Self-Consistent Molecular Orbital Methods. 25. Supplementary Functions for Gaussian Basis Sets. J. Chem. Phys. 1984, 80, 3265–3269. [Google Scholar]
  58. Ditchfield, R. Self-consistent Perturbation Theory of Diamagnetism. I. A Gauge-Invariant LCAO (Linear Combination of Atomic Orbitals) Method for NMR Chemical Shifts. Mol. Phys. 1974, 27, 789–807. [Google Scholar] [CrossRef]
  59. London, F. Quantum Theory of Interatomic Currents in Aromatic Compounds. J. Phys. Radium. 1937, 8, 397–409. [Google Scholar] [CrossRef]
  60. Miertus, S.; Scrocco, E.; Tomasi, J. Electrostatic interaction of a solute with a continuum. A direct utilization of AB initio molecular potentials for the prevision of solvent effects. Chem. Phys. 1981, 55, 117–129. [Google Scholar] [CrossRef]
  61. Tomasi, J.; Persico, M. Molecular Interactions in Solution: An Overview of Methods Based on Continuous Distributions of the Solvent. Chem. Rev. 1994, 84, 2027–2094. [Google Scholar] [CrossRef]
  62. Mennucci, B.; Cammi, R. (Eds.) Continuum Solvation Models in Chemical Physics. From Theory to Applications; John Wiley & Sons: Chichester, UK, 2007. [Google Scholar]
  63. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G.A.; et al. Gaussian 09, Revision D.01; Gaussian, Inc.: Wallingford, CT, USA, 2009. [Google Scholar]
  64. Silva, A.M.S.; Sousa, R.M.S.; Jimeno, M.L.; Blanco, F.; Alkorta, I.; Elguero, J. Experimental measurements and theoretical calculations of the chemical shifts and coupling constants of three azines (benzalazine, acetophenoneazine and cinnamaldazine). Magn. Reson. Chem. 2008, 46, 859–864. [Google Scholar] [CrossRef] [PubMed]
  65. Blanco, F.; Alkorta, I.; Elguero, J. Statistical analysis of 13C- and 15N-NMR chemical shifts from GIAO/B3LYP/6-311++G** calculated absolute shieldings. Magn. Reson. Chem. 2007, 45, 797–800. [Google Scholar] [CrossRef] [PubMed]
  66. Fresno, N.; Pérez-Fernández, R.; Jimeno, M.L.; Alkorta, I.; Sánchez-Sanz, G.; Elguero, J.; Del Bene, J.E. Multinuclear NMR Characterization of Cyanuric Fluoride (2,4,6-Trifluoro-1,3,5-triazine). J. Heterocycl. Chem. 2012, 49, 1257–1259. [Google Scholar] [CrossRef]
  67. Reed, A.E.; Curtiss, L.A.; Weinhold, F. Intermolecular Interactions from a Natural Bond Orbital, Donor–Acceptor Viewpoint. Chem. Rev. 1988, 88, 899–926. [Google Scholar] [CrossRef]
  68. Glendening, E.D.; Badenhoop, J.K.; Reed, A.E.; Carpenter, J.E.; Bohmann, J.A.; Morales, C.M.; Landis, C.R.; Weinhold, F. NBO 6.0; University of Wisconsin: Madison, WI, USA, 2013. [Google Scholar]
Sample Availability: Samples of the compounds 13, 14, 15, are available from the authors.
Figure 1. N-CHF2, N-CClF2, and N-CBrF2 azoles and benzazoles.
Figure 1. N-CHF2, N-CClF2, and N-CBrF2 azoles and benzazoles.
Molecules 22 02003 g001
Scheme 1. Synthesis of the N-difluoromethyl derivatives 13 and 14 of (4S,7R)-campho[2,3-c]pyrazole. SCDA: sodium chlorodifluoroacetate.
Scheme 1. Synthesis of the N-difluoromethyl derivatives 13 and 14 of (4S,7R)-campho[2,3-c]pyrazole. SCDA: sodium chlorodifluoroacetate.
Molecules 22 02003 sch001
Figure 2. 19F-NMR spectrum of 13 in CDCl3 at 300 K with signals at −89.16 ppm (ddd, 2JF = 226.6, 2JH = 60.6, 6JH = 1.4), and −91.64 (dd, 2JF = 226.6, 2JH = 59.4); the red arrows correspond to the amplification of the left and right side of the signal at −89.16 ppm.
Figure 2. 19F-NMR spectrum of 13 in CDCl3 at 300 K with signals at −89.16 ppm (ddd, 2JF = 226.6, 2JH = 60.6, 6JH = 1.4), and −91.64 (dd, 2JF = 226.6, 2JH = 59.4); the red arrows correspond to the amplification of the left and right side of the signal at −89.16 ppm.
Molecules 22 02003 g002
Figure 3. 19F-NMR spectrum of 14 in CDCl3 at 300 K with signals at −90.80 ppm (dd, 2JF = 225.5, 2JH = 61.3), and −92.05 (dd, 2JF = 225.4, 2JH = 60.7).
Figure 3. 19F-NMR spectrum of 14 in CDCl3 at 300 K with signals at −90.80 ppm (dd, 2JF = 225.5, 2JH = 61.3), and −92.05 (dd, 2JF = 225.4, 2JH = 60.7).
Molecules 22 02003 g003
Figure 4. The four minima.
Figure 4. The four minima.
Molecules 22 02003 g004
Figure 5. Energy profiles in kJ·mol–1 vs. the dihedral angle θ.
Figure 5. Energy profiles in kJ·mol–1 vs. the dihedral angle θ.
Molecules 22 02003 g005
Table 1. Calculated (gas phase) and experimental 19F-NMR chemical shifts (CDCl3).
Table 1. Calculated (gas phase) and experimental 19F-NMR chemical shifts (CDCl3).
Comp.θ (°)19F (31)19F (32)Δδ (31–32)19F (a)19F (b)Δδ (a–b) a
Calculated ValuesExperimental Values
13–3.6–93.18–89.73−3.45–91.64–89.16−2.48
13–179.3–90.07–98.52+8.45
1411.1–92.83–88.66–4.17–92.05–90.80–1.25
14179.3–85.73–97.40+11.67
a The sign is arbitrary because the assignment of a and b is also arbitrary.
Table 2. Calculated (CHCl3) and experimental 19F-NMR chemical shifts (CDCl3).
Table 2. Calculated (CHCl3) and experimental 19F-NMR chemical shifts (CDCl3).
Comp.θ (°)19F (31)19F (32)Δδ (31–32)19F (a)19F (b)Δδ (a–b) a
Calculated ValuesExperimental Values
13−4.0−94.07−90.23−3.84−91.64−89.16−2.48
13−179.5−91.97−99.34+7.37
1411.2−94.21−90.68−3.53−92.05−90.80−1.25
14179.4−86.35−97.99+11.64
a The sign is arbitrary because the assignment of a and b is also arbitrary.
Table 3. Comparison of experimental and calculated 13C chemical shifts.
Table 3. Comparison of experimental and calculated 13C chemical shifts.
Comp.13 exp. CDCl313 calc. Gas13 calc. CHCl314 exp. CDCl314 calc. Gas14 calc. CHCl3
C3 (C4)134.3133.2134.4117.9117.3118.2
C3a (C9)132.1133.0133.9130.2132.3133.6
C7a (C11)153.6153.7155.2169.1167.5169.3

Share and Cite

MDPI and ACS Style

García-Pérez, D.; López, C.; Claramunt, R.M.; Alkorta, I.; Elguero, J. 19F‐NMR Diastereotopic Signals in Two N-CHF2 Derivatives of (4S,7R)-7,8,8-Trimethyl-4,5,6,7-tetrahydro-4,7-methano-2H-indazole. Molecules 2017, 22, 2003. https://doi.org/10.3390/molecules22112003

AMA Style

García-Pérez D, López C, Claramunt RM, Alkorta I, Elguero J. 19F‐NMR Diastereotopic Signals in Two N-CHF2 Derivatives of (4S,7R)-7,8,8-Trimethyl-4,5,6,7-tetrahydro-4,7-methano-2H-indazole. Molecules. 2017; 22(11):2003. https://doi.org/10.3390/molecules22112003

Chicago/Turabian Style

García-Pérez, Diana, Concepción López, Rosa M. Claramunt, Ibon Alkorta, and José Elguero. 2017. "19F‐NMR Diastereotopic Signals in Two N-CHF2 Derivatives of (4S,7R)-7,8,8-Trimethyl-4,5,6,7-tetrahydro-4,7-methano-2H-indazole" Molecules 22, no. 11: 2003. https://doi.org/10.3390/molecules22112003

Article Metrics

Back to TopTop