Next Article in Journal
Microwave-Assisted Stereoselective Heterocyclization to Novel Ring d-fused Arylpyrazolines in the Estrone Series
Previous Article in Journal
Comparison of Bioactive Compounds and Antioxidant Activities of Maclura tricuspidata Fruit Extracts at Different Maturity Stages
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

An Example of Polynomial Expansion: The Reaction of 3(5)-Methyl-1H-Pyrazole with Chloroform and Characterization of the Four Isomers

1
Chemistry Department and QOPNA and LAQV-REQUIMTE, University of Aveiro, 3810-193 Aveiro, Portugal
2
Departamento de Química Orgánica y Bio-Orgánica, Facultad de Ciencias, UNED, Paseo Senda del Rey, 9, E-28040 Madrid, Spain
3
Departamento de Cristalografía y Biología Estructural, Instituto de Química-Física Rocasolano, CSIC, Serrano, 119, E-28006 Madrid, Spain
4
Instituto de Química Médica, CSIC, Juan de la Cierva, 3, E-28006 Madrid, Spain
*
Authors to whom correspondence should be addressed.
Molecules 2019, 24(3), 568; https://doi.org/10.3390/molecules24030568
Submission received: 5 January 2019 / Revised: 25 January 2019 / Accepted: 27 January 2019 / Published: 4 February 2019

Abstract

:
The reaction in phase-transfer catalyzed conditions of 3(5)-methyl-1H-pyrazole with chloroform affords four isomers 333, 335, 355 and 555 in proportions corresponding to the polynomial expansion (a + b)3, with a = 0.6 and b = 0.4, a and b being 3-methyl and 5-methyl proportions. The up (u) and down (d) conformation of the pyrazolyl rings with regard to the Csp3–H atom was established by X-ray crystallography and by 1H-, 13C- and 15N-NMR in solution combined with gauge-including atomic orbitals (GIAO)/B3LYP/6-311++G(d,p) calculations. A comparison with other X-ray structures of tris-pyrazolylmethanes was carried out.

1. Introduction

N-unsubstituted pyrazoles, and in general N-unsubstituted azoles, react with chloroform in phase-transfer catalysis (PTC) conditions to afford trispyrazolylmethanes [1], the neutral equivalents of anionic scorpionates [2]. The reaction of 3(5)-methyl-1H-pyrazole (1) with chloroform was reported in 1984 and the only compound isolated in a pure state was the tris(3-methylpyrazol-1-yl)methane 333 derivative (1H, CDCl3, Csp3-H: 8.11 ppm) (Figure 1) [3]. We have used a double nomenclature, 3 or 5 to define the position of the methyl group (3 precedes 5) and down (d) and up (u) to define the position of N2 (N4, N6) with regard to the H of the Csp3–H group (down on opposite sides, up on the same side). For instance, structure types 335 duu and 355 duu should be read 3Me-d, 3Me-u, 5Me-u and 3Me-d, 5Me-u, 5Me-u. We have used this nomenclature in previous papers [4,5,6].
The reaction was reported again in 1999 and, surprisingly, the only isolated isomer (17% yield) was the 335 isomer (1H, CDCl3, Csp3-H: 8.21 ppm) [7]. In 2012 the more hindered 555 derivative was prepared from tri(pyrazol-1-yl)methane (tpzm) by alkylation of the lithium derivative (1H, CDCl3, Csp3-H: 8.31 ppm) [8].
Finally, the reaction was repeated again in 2012 and, although the yields of different isomers were not discussed, using the information concerning the 1H-NMR of the Csp3-H proton from this and others papers the relative yields of the four isomers can be determined (Figure 2) [8,9]. The authors also demonstrated that the mixture of the four isomers could be isomerized under the action of p-toluenesulfonic acid to a mixture of 333 and 335 in a 2:1 ratio (crude yield 82%), proving that they are the most stable isomers. The utility of the 333 ligand prompted Anwander et al. to prepare it from the mixture and isolate it by recrystallization [10].
The authors point out that the 333 isomer, δ = 8.13 ppm, represents approximately 22% of the crude. From the data of Figure 2 reported and from the integration (total 2.324) it is easy to determine the proportions: 333, δ = 8.13 ppm, 22%; 335, δ = 8.21 ppm, 43%; 355, δ = 8.26 ppm, 28.5%; and 555, δ = 8.32 ppm, 6.5%.
We have shown that in the reaction of polyhaloalkanes, such as dichloromethane, and chloroform with non-symmetrical 1H-pyrazoles (different substituents at position 3 and 5), the proportion of isomers in the crude always follows a binomial expansion, in this case (a + b)3, where a + b = 1 [11,12].
Figure 1 shows the results obtained with a = 0.6 and b = 0.4, i.e., more 3-methyl than 5-methyl isomers, a ratio consistent with the proportions obtained by alkylation of 3(5)-methyl-1H-pyrazole [13,14,15,16,17,18].
The calculated values (21.6%, 43.2%, 28.8% and 6.4%) are remarkably consistent, indicating that at every step the ratio 0.6/0.4 is constant. The accuracy is so remarkable that it is possible to conclude that a = 0.603, b = 0.397 yields better results: 21.9%, 28.5%, 43.3% and 6.3%.
The 333 compound prepared as in [3] was used to prepare complexes with Fe(II) to study their spin-transition temperatures [19]. The few relevant data concerning these compounds are 13C-NMR data (unassigned) of 555 [7], and the crystal structure of 333 (obtained by the general procedure followed by the isomerization of the mixture using p-toluenesulfonic acid) [20]. The structure of 333 (TUYZEU, Figure 3) [21] corresponds to a uud conformation. These isomerization experiments, together with those reported previously [9], prove that the stability decreases in the order 333 > 335 > 355 or 555.
The melting points of three isomers were known: 333, 107–109 °C [3], 335, 114–115 °C [7], and 555, 145–148 °C [8].

2. Results and Discussion

We decided to repeat the reaction shown in Figure 1 to determine if it was possible to isolate other isomers and establish their structures, and to discuss their conformations.

2.1. Chemistry

The synthesis of N,N’,N”-3(5)-trimethylpyrazolylmethanes was performed following the protocol described by Juliá et al. for the synthesis of N,N’,N”-triazolylmethanes (see Section 3.2) [3]. The percentages were determined by 1H-NMR using the signals of the Csp3-H atom in CDCl3 at 400 MHz (Table 1): 21.8% 333, 47.8% 335, 25.2% 355 and 5.2% 555. These percentages correspond to (a + b)3 for a = 0.60 and b = 0.40 (21.6%; 43.2%; 28.8%; 6.4%) with a little worse agreement than the literature results [8].

2.2. NMR Studies

Table 1 (CDCl3) and Table 2 (C6D6) contain all the NMR information concerning the four isomers. The calculated values are in the Electronic Supplementary Information (ESI), Supplementary Materials Tables S1 to S4. In CDCl3 equilibration between isomers occurs that can be due to the presence of DCl, although we keep the solvent over Ag wire. For this reason, only the “fast” 1H-NMR experiments are reported in this solvent (see later). To avoid the problems encountered with deuterochloroform we move to another solvent. We select hexadeuterobenzene because in this solvent (in earlier work C6H6 was used) the methyl groups of pyrazoles have rather different 1H-NMR chemical shifts [22,24].
The methyl groups appear in C6D6 at ~2.10 (3-methyl) and ~1.85 ppm (5-methyl), very close to the values reported in the literature, 2.25 and 1.80 ppm, respectively [22,23,24]. These values together with the relative intensities allow an immediate identification of the four isomers by 1H-NMR.
Another useful criterion is that 3JHH has a value of 1.8 Hz between H3 and H4 protons and 2.6 Hz between H3 and H4 protons. When a methyl group was involved, then 4JH4Me5 > 4JH4Me3 [22,23,24].
The GIAO calculated chemical shifts reported in the ESI agree well with the experimental values. To determine the major conformers, the problem is that the variation of chemical shift inter-conformers are small compared with the variation within each isomer that results in correlation coefficients R = 1.000 or 0.999 in most cases (correlation coefficient matrix). However, considering not only R2, but also an intercept as small as possible and a slope as close to 1.00 as possible, the results of Table 3 were obtained.
They are not all consistent but clear preferences are observed. We have tried a mixture of the conformations of lower energy; for the 335 isomer we have preferred the udd (12.6 kJ·mol–1) to the uuu (12.1 kJ·mol–1).
333 = −(0.2 ± 0.3) + (0.95 ± 0.16) uud + (0.06 ± 0.16) uuu, n = 20, R2 = 1.000, RMS = 0.8 ppm
335 = −(0.4 ± 0.2) + (0.30 ± 0.07) udd + (0.72 ± 0.07) uud, n = 20, R2 = 1.000, RMS = 0.8 ppm
355 = −(0.7 ± 0.5) + (0.75 ± 0.04) ddu + (0.25 ± 0.04) duu, n = 20, R2 = 1.000, RMS = 2.0 ppm
In the case of the 555 isomers, the regression leads to a negative coefficient for the uud isomer that does not have a physical meaning, which indicates that the only isomer present is the udd isomer.
The mixtures (sum ≈ 1.00) correspond to about 3/4 of the lower energy isomers and 1/4 of the higher energy ones for 335 and 355; in the case of the 333 isomer there is 95% of the uud isomer.
We found that a solution of an almost pure sample of the 335 isomer in CDCl3 slowly isomerizes into the more stable 333 isomer (Figure 4). This could be due to the presence of DCl in the solvent produced by photodecomposition of CDCl3, and is related to the already reported isomerization in acid media [9]. The mechanism should proceed by protonation of one of the pyrazoles (formation of a pyrazolium salt), leaving this ring as neutral 3(5)-methyl-1H-pyrazole, and the resulting carbocation reacting with 3(5)-methyl-1H-pyrazole.

2.3. Crystallography

The crystal structure of 335 presents one independent molecule in its asymmetric unit with two of their pyrazole N2 atoms pointing to the CH direction and the third one pointing into the opposite direction; therefore, these molecules present an uud conformation displaying HCN1N2 dihedral angles of 39.0(6)°, 17.6(7)° and −168.1(3)°. As the crystal space group contains inversion centers, both enantiomers coexist in a 1:1 ratio.
A summary of the crystal data and structure refinement is included in the ESI as Table S1. Table 4 contains geometrical parameters of 3(5)-methylpyrazolylmethane isomers, the one recorded in the Cambridge Structural Database (CSD) [21] and the new one reported in this manuscript. A view of their molecule structure with their atom labeling is depicted in Figure 3.
There are no significant differences in the observed geometry parameters for the pyrazole rings for the 3-methyl rings in both 333 and 335 isomers; however, the 5-methyl ring in 335 presents unusual geometries, especially for the intra-ring bond angles, N2-C3-C4, C3-C4-C5 and C4-C5-N1, and also for the N1-C5-C6 and C4-C5-C6 angles. There is not any apparent reason to justify this, but X-ray data collected for other crystals showed a possible occupancy disorder of the two isomers, 335 and 333. Therefore, pz_C can, overall, be of 5-methylpyrazolyl with high occupancy and 3-methylpyrazolyl at low occupancy. It also justifies the difference peaks observed around the pz_C ring (Figure 5). A disorder model has been able to be refined with a crystal collected at low temperature but, due to the problems in reaching refinement convergence with the data, and the many geometrical restraints and constraints that were necessary, we do not think that it adds any knowledge to the results presented in this manuscript.
Both compounds present a twist of the pz-planes describing a propeller structure independently of the N2 position (up or down).
Compound 335 forms dimers through weak CH···N hydrogen bonds (C1-H···N2(B), C6(B)-H···N2(A)) that expand into chains along (1–10) axis by C5(B)-H···N2(C) contacts. Saturating the three N-acceptors in the molecules, these chains join to form (001) layers by C-H···π-pz non-bonded interactions (C3(C)-H···π-pz(A), C4(C)-H···π-pz(B)). Methyl groups of pyrazoles A and B point out of these layers forming lines along the b-direction with an a-axis separation between methyl lines where a methyl line, from a consecutive layer, fits as a zipper to pack the layers and to build the crystal. C6(A)-H···π-pz(C) and van der Waal interactions glue the layers (Figure 6).
In the Cambridge Structural Database (CSD, Version 5.39, updates to Feb 2018) [21], 19 structures of 17 trispyrazolylmethane compounds are recorded (Figure 7). Five of these structures contain 3(5)-pyrazoles, and the observed isomers are 333 in four cases (AKUSAA, DUDFUF, TUYZEU, XIVVAA) and 335 in the remaining one (XIVVEE). The one we report here is the second example of a 335 isomer.
All trispyrazolylmethane molecules display a propeller structure with a wide range of twisting in their pyrazole planes, from 3° to 68°. The most frequent conformations are udd (observed in 10 structures) and uud (observed in seven structures).

2.4. Theoretical Calculations

The geometries of the two most relevant isomers are depicted in Figure 8 while the energies are reported in Table 5.
The isomers’ stability decreases in the order 333 (0.0) > 335 (1.7) > 355 (17.9) > 555 (23.8 kJ·mol−1), in agreement with the experimental results; this is the thermodynamic order that is unrelated to the kinetic order of the percentages measured on the crude. The acid-catalyzed isomerization from the crude leads to a mixture of 333 and 335 which have very close energies (Figure 7) [8]. Concerning the up/down isomerism, the most stable are the uud ones (333, 335), the ddu one (355) and the udd one (555). The 355 uud is 9.9 kJ·mol−1 above the 355 ddu and the 555 uud is 5.9 kJ·mol−1 above the 555 udd one.
Concerning the calculated geometries, the most interesting parameters are the torsion angles (Table 6).
The chemical shifts we have used for the interpolations (Equations (1–3)) have been obtained transforming the absolute shielding calculated with the B3LYP/6-311++G(d,p)/GIAO methods for the gas phase and then transformed through empirical equations to chemical shifts (see Section 3.5). These δ values do not correspond to the gas phase but to solution, because the empirical equations were established using gas phase σ and solution δ.
When comparing the experimental chemical shifts in solution to the GIAO calculated ones, remember that the four 3/5 isomers correspond to different molecules that are stable in the NMR time scale. On the other hand, the up/down rotational isomers are separated by low rotational barriers and in solution only averaged signals will be observed.

3. Materials and Methods

3.1. Experimental

High-resolution mass spectra were recorded on a Quadrupole Time-of-Flight (QTOF) mass spectrometer under Electrospray Ionization (ESI) conditions.

3.2. Chemistry

A mixture of 3(5)-methyl-1H-pyrazole (1.93 mL, 24 mmol), anhydrous K2CO3 (16.58 g, 120 mmol) and (Bu)4NHSO4 (0.41 g, 1.2 mmol) was vigorously stirred and refluxed in dry CHCl3 (25 mL) for 24 h. Then the mixture was filtered and the residue washed with hot CHCl3 (3 × 25 mL). The organic solution was evaporated and the crude product was purified by column chromatography (using silica 1:130) and then by crystallization using a mixture of diethyl ether/hexane. Four isomers of N,N’,N”-3(5)-trimethylpyrazolylmethanes were formed: the 3,3,3-trimethylpyrazolylmethane (333), the 3,3,5-trimethylpyrazolylmethane (335), which was the major isomer, the 3,5,5-trimethylpyrazolylmethane (355) and the 5,5,5-trimethylpyrazolylmethane (555), which was the minor isomer. The overall reaction yield calculated as the sum of all the isomers isolated by column chromatography was 46% (0.94 g), corresponding to 0.20 g of 333, 0.45 g of 335, 0.24 g of 355 and 0.05 g of 555. Melting points were determined under microscope.
Only the 335 isomer was isolated pure in enough quantity; the 355 was isolated in a very small amount only enough to measure its melting point and record its exact mass; the two other isomers, 333 and 555 only as mixtures of two isomers enriched in one of them.
Tris(3-methyl-1H-pyrazol-1-yl)methane (333): Not isolated pure. According to literature it melts at 107–109 °C [3].
Bis(3-methyl-1H-pyrazol-1-yl)(5-methyl-1H-pyrazol-1-yl)methane (335): M.p. = 109–111 °C, literature: 114–115 °C [7]. HRMS (ESI) [M + H]+ Calcd for [C13H17N6+] 257.1509, found 257.1506.
Bis(5-methyl-1H-pyrazol-1-yl)(3-methyl-1H-pyrazol-1-yl)methane (355): M.p. = 55–57 °C. HRMS (ESI) [M + H]+ Calcd for [C13H17N6+] 257.1509, found 257.1512.
Tris(5-methyl-1H-pyrazol-1-yl)methane (555): Not isolated pure. According to literature it melts at 145–148 °C [8].

3.3. NMR Spectroscopy

Solution NMR spectra were recorded on a 9.4 Tesla Bruker spectrometer (Bruker Española S.A., Madrid, Spain), 400.13 MHz for 1H, 100.62 MHz for 13C and 40.54 MHz for 15N) at 300 K with a 5-mm inverse detection H-X probe equipped with a z-gradient coil. Chemical shifts (δ in ppm) are given from internal solvents: CDCl3 7.26 for 1H; C6D6 7.16 for 1H and 128.39 for 13C. Nitromethane was used as external reference for 15N. Coupling constants (J in Hz) are accurate to ±0.2 Hz for 1H and ±0.6 Hz for 13C. CDCl3 contains 0.5 wt% silver wire as stabilizer.
Typical parameters for 1H-NMR spectra were spectral width 4000 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 10.6 μs at an attenuation level of −6 dB and relaxation delay 2 s. WALTZ 16 was used for broadband proton decoupling; the FIDs were multiplied by an exponential weighting (lb = 2 Hz) before Fourier transformation. In some cases, for resolution enhancement processing a Gaussian multiplication of the FID prior to Fourier transformation was applied.
2D (1H-13C) gs-HMQC, (1H-13C) gs-HMBC and (1H-15N) gs-HMBC, were acquired and processed using Bruker NMR software suite (Bruker, Karlsruhe, Germany) 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 gs-HMBC spectra were: spectral width 4000 Hz for 1H and 20 kHz for 13C, 1024 × 256 data set, number of scans 2 (HMQC) or 4 (HMBC) and relaxation delay 1s. In the gs-HMQC experiments GARP modulation of 13C was used for decoupling. 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.
Selected parameters for gs-HMBC spectra were: spectral width 4000 Hz for 1H and 15 kHz for 15N, 2048 × 1024 data set, number of scans 4, relaxation delay 1s. In the gs-HMBC delays of 60 and 100 ms for the evolution of the 15N-1H long-range coupling were used. 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.

3.4. Crystallography

For compound 335 [bis(3-methylpyrazolyl)(5-methylpyrazolyl)methane], X-ray single crystal diffraction data were collected on a Bruker APEX-II CCD diffractometer (Bruker Española S.A., Madrid, Spain) [25].
From initial data collected at room temperature, a possible occupancy disorder of the two isomers, 335 and 333 was identified. Data collected at 150 K of a different crystal allowed us to build a disorder model that showed a mixture of these two isomers. Due to the many problems to refine this disordered model and to reach convergence, we kept the best data for the work presented in this manuscript. It corresponds to data collected at room temperature and it shows the lowest proportion of the 333 isomer (so it could be omitted).
Using Olex2 (v1.2, Durham University, Durham, UK) [26], the structure was solved with the ShelXS (v4-2016, Universität Göttingen, Göttingen, Germany) [27] structure solution program using Direct Methods and refined with the ShelXL (v4-2016, Universität Göttingen, Göttingen, Germany) [28] refinement package using Least Squares minimization. A summary of the crystal data and structure refinement is included in Table S5. For the visualization and analysis of crystal structures the Mercury program was used [29].

3.5. Theoretical Calculations

Density Functional Theory (DFT) calculations were carried out using the Becke, three-Parameter, Lee, Yang and Parr (B3LYP) Gaussian 09 (Version D.01, Wallingford, CT, USA, 2009) [30,31,32], together with the 6-311++G/(d,p) basis set [33,34]. Absolute shieldings were calculated within the GIAO approximation [35,36]. All the calculations were carried out using the Gaussian 09 package (Version D.01, Wallingford, CT, USA, 2009) [37]. Empirical equations were used to transform the 1H, 13C and 15N absolute shieldings into chemical shifts [38,39].

4. Conclusions

The accuracy of the polynomial expansion (a + b)3 is extraordinary because it was unexpected. It implies that the ratios of the reactivity of the chlorine atoms with 3(5)-methyl-1H-pyrazole are the same for CHCl3, CH(Mepz)Cl2 and CH(Mepz)2Cl. This work reported the first systematic study of the structure of the four tris[3(5)-methyl]pyrazol-1-ylmethanes, a series of ligands used in coordination chemistry.
The solid-solution [40] structure of the 335 isomer, actually a mixture of 335 (major) and 333 (minor) isomers, results from the isomerization of the 335 isomer into the 333 isomer during the crystallization process. In the crystallization batch there are different crystals but a clear predominance of the 335 isomer is always found by crystallography.

Supplementary Materials

Electronic supplementary information (ESI) is available online at https://www.mdpi.com/1420-3049/24/3/568/s1: calculated NMR chemical shifts (GIAO), theoretical calculations (Tables S1 to S4), Crystallographic details (Table S5), CCDC 1875403 for compound 335.

Author Contributions

J.E. conceptualized the work and co-wrote the manuscript; V.L.M.S. performed the synthetic experimental work; L.I. co-conceptualized the work, participated in the methodology and wrote the original draft preparation; A.M.S.S. co-conceptualized the work and co-wrote the manuscript; R.M.C. co-wrote the manuscript; I.A. responsible for the software studies; A.M.L., D.S. and F.R. participated in the methodology.

Funding

This research was funded by the Spanish Ministerio de Economía, Industria y Competitividad (CTQ2014-56833-R, CTQ2015-63997-C2-2-P) and Comunidad Autónoma de Madrid (Project Fotocarbon, S2013/MIT-2841) and by “Fundação para a Ciência Tecnologia/Ministério da Ciência e Ensino Superior” (QOPNA research unit—FCTU ID/QUI/00062/2019 and Portuguese NMR network) through national funds and, where applicable, co-financed by the FEDER, within the PT2020 Partnership.

Acknowledgments

Computer, storage and other resources from the CTI (CSIC) are gratefully acknowledged.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Alkorta, I.; Claramunt, R.M.; Díez-Barra, E.; Elguero, J.; de la Hoz, A.; López, C. The organic chemistry of poly(1H-pyrazol-1-yl)methanes. Coord. Chem. Rev. 2017, 339, 153–182. [Google Scholar] [CrossRef]
  2. Trofimenko, S. Scorpionates: The Coordination Chemistry of Polypyrazolylborate Ligands; Imperial College Press: London, UK, 1999; ISBN 1860941729. [Google Scholar]
  3. Juliá, S.; del Mazo, J.M.; Ávila, L.; Elguero, J. Improved synthesis of polyazolylmethanes under solid-liquid phase-transfer catalysis. Org. Prep. Proc. Int. 1984, 16, 299–307. [Google Scholar] [CrossRef]
  4. Foces-Foces, C.; Cano, F.H.; Martínez Ripoll, M.; Faure, R.; Roussel, C.; Claramunt, R.M.; López, C.; Sanz, D.; Elguero, J. Complete energy profile of a chiral propeller compound: Tris-(2′-methylbenzimidazol-1′-yl) Methane (TMBM). Chromatographic resolution on triacetyl cellulose, x-ray structures of the racemic and one enantiomer, and dynamiC-NMR study. Tetrahedron Asymm. 1990, 1, 65–86. [Google Scholar] [CrossRef]
  5. Alkorta, I.; Elguero, J. A theoretical analysis of the conformational space of tris(2-methylbenzimidazol-1-yl)methane. Tetrahedron Asymm. 2010, 21, 437–442. [Google Scholar] [CrossRef] [Green Version]
  6. Alkorta, I.; Elguero, J.; García, M.A.; López, C.; Claramunt, R.M.; Andrade, G.A.; Yap, G.P.A. A tris(pyrazol-1-yl)methane bearing carboxylic acid groups at position 4: {1-[bis(4-carboxy-3,5-dimethyl-1H-pyrazol-1-yl)methyl]-3,5-dimethyl-1H-pyrazole-4-carboxylato}sodium dihydrate. Acta Crystallogr. Ser. C 2013, 69, 972–976. [Google Scholar] [CrossRef] [PubMed]
  7. Pettinari, C.; Pellei, M.; Cingolani, A.; Martini, D.; Drozdov, A.; Troyanov, S.; Panzeri, W.; Mele, A. Synthesis, spectroscopic, and X-ray diffraction structural studies of tin(IV) derivatives with tris(pyrazol-1-yl)methanes. Inorg. Chem. 1999, 38, 5777–5787. [Google Scholar] [CrossRef]
  8. Goodman, M.A.; Nazarenko, A.Y.; Casavant, B.J.; Li, Z.; Brennesel, W.W.; DeMarco, M.J.; Long, G.; Goodman, M.S. Tris(5-methylpyrazolyl)methane: Synthesis and properties of its iron(II) complex. Inorg. Chem. 2012, 51, 1084–1093. [Google Scholar] [CrossRef]
  9. Goodman, M.A.; DeMarco, M.J.; Tarasek, S.E.; Nazarenko, A.Y.; Brennessel, W.; Goodman, M.S. Iron complexes of tris(pyrazolyl)ethane ligands methylated in the 3-, 4-, and 5-positions. Inorg. Chim. Acta 2014, 423, 358–368. [Google Scholar] [CrossRef]
  10. Stuhl, C.; Maichle-Mössmer, C.; Anwander, R. Magnesium Stung by Nonclassical Scorpionate Ligands: Synthesis and Cone-Angle Calculations. Chem. Eur. J. 2018, 24, 14254–14268. [Google Scholar] [CrossRef]
  11. Avila, L.; Elguero, J.; Juliá, S.; del Mazo, J.M. N-Polyazolymethanes. IV. Reaction of Benzotriazole with Methylene Chloride and Chloroform under Phase Transfer Conditions. Heterocycles 1983, 20, 1787–1792. [Google Scholar] [CrossRef]
  12. Claramunt, R.M.; Santa María, M.D.; Elguero, J.; Trofimenko, S. Isomer distribution in thallium hydrotris(polymethylenepyrazol-1-yl)borates (thallium scorpionates): A multinuclear NMR study. Polyhedron 2004, 23, 2985–2991. [Google Scholar] [CrossRef]
  13. Schofield, K.; Grimmett, M.R.; Keene, B.R.T. Heteroaromatic Nitrogen Compounds: The Azoles; Cambridge University Press: Cambridge, UK, 1976; p. 75, ISBN-10 0521205190; ISBN-13 978-0521205191. [Google Scholar]
  14. Elguero, J. Pyrazoles and Their Benzo Derivatives. In Comprehensive Heterocyclic Chemistry; Potts, K.T., Katritzky, A.R., Rees, C.W., Eds.; Pergamon Press: Oxford, UK, 1984; Volume 5, p. 229. [Google Scholar]
  15. Grimmett, M.R.; Lim, K.H.R.; Weavers, R.T. The N-Alkylation and N-arylation of unsymmetrical pyrazoles. Austr. J. Chem. 1979, 32, 2203–2213. [Google Scholar] [CrossRef]
  16. Elguero, J.; Ochoa, C.; Stud, M.; Esteban-Calderon, C.; Martínez-Ripoll, M.; Fayet, J.-P.; Vertut, M.-C. Synthesis and physicochemical properties of 1,2,6-thiadiazine 1,1-dioxides. A comparative study with pyrazoles. J. Org. Chem. 1982, 47, 536–544. [Google Scholar] [CrossRef]
  17. Ouk, S.; Thiébaud, S.; Borredon, E.; Chabaud, B. N-Methylation of nitrogen-containing heterocycles with dimethyl carbonate. Synth. Commun. 2005, 35, 3021–3026. [Google Scholar] [CrossRef]
  18. Almena, I.; Díez-Barra, E.; de la Hoz, A.; Ruíz, J.; Sánchez-Migallón, A. Alkylation and arylation of pyrazoles under solvent-free conditions: Conventional heating versus microwave irradiation. J. Heterocycl. Chem. 1998, 35, 1263–1268. [Google Scholar] [CrossRef]
  19. Paulsen, H.; Duelund, L.; Zimmermann, A.; Averseng, F.; Gerdan, M.; Winkler, H.; Toftlund, H.; Trautwein, A.X. Substituent Effects on the Spin-Transition Temperature in Complexes with Tris(pyrazolyl) Ligands. Monatsh. Chem. 2003, 134, 295–306. [Google Scholar] [CrossRef]
  20. Goodman, M.A.; Goodman, M.S.; Nazarenko, A.Y.; Patel, E.N. Crystal structure of tris-(3-methyl-1H-pyrazol-1-yl)methane. Acta Crystallogr. Sect. E 2015, 71, o816. [Google Scholar] [CrossRef]
  21. Groom, C.R.; Bruno, I.J.; Lightfoot, M.P.; Ward, S.C. The Cambridge Structural Database, Acta Crystallogr. Sect. B 2016, 72, 171–179. [Google Scholar] [CrossRef]
  22. Elguero, J.; Jacquier, R.; Tien Duc, H.C.N. Recherches sans la série des azoles. XIII. Spectres RMN de pyrazoles. Bull. Soc. Chim. Fr. 1966, 63, 3727–3743. [Google Scholar]
  23. Elguero, J.; Jacquier, R. Recherches dans la série des azoles. VII. Attribution des signaux RMN de pyrazoles N-substitués. J. Chim. Phys. 1966, 63, 1242–1246. [Google Scholar] [CrossRef]
  24. Bruix, M.; de Mendoza, J.; Elguero, J. The 1H and 13C n.m.r. rules for the assignment of 1, 3- and 1, 5-disubstituted pyrazoles: A revision. Tetrahedron 1987, 43, 4663–4668. [Google Scholar] [CrossRef]
  25. Bruker. APEX2, SAINT and SADABS; Bruker AXS Inc.: Madison, WI, USA, 2013. [Google Scholar]
  26. Dolomanov, O.V.; Bourhis, L.J.; Gildea, R.J.; Howard, J.A.K.; Puschmann, H. OLEX2: A complete structure solution, refinement and analysis program. J. Appl. Cryst. 2009, 42, 339–341. [Google Scholar] [CrossRef]
  27. Sheldrick, G.M. A short history of SHELX. Acta Crystallogr. Sect. A 2008, 64, 112–122. [Google Scholar] [CrossRef] [PubMed]
  28. Sheldrick, G.M. Crystal structure refinement with SHELXL. Acta Crystallogr. Sect. C 2015, 71, 3–8. [Google Scholar] [CrossRef] [PubMed]
  29. Macrae, C.F.; Bruno, I.J.; Chisholm, J.A.; Edgington, P.R.; McCabe, P.; Pidcock, E.; Rodriguez-Monge, L.; Taylor, R.; van de Streek, J.; Wood, P.A. Mercury CSD 2.0—New features for the visualization and investigation of crystal structures. J. Appl. Cryst. 2008, 41, 466–470. [Google Scholar] [CrossRef]
  30. Becke, A.D. Density-functional thermochemistry. III. The role of exact exchange. J. Chem. Phys. 1993, 98, 5648–5652. [Google Scholar] [CrossRef]
  31. Lee, C.; Yang, W.; Parr, R.G. Development of the Colle-Salvetti correlation-energy formula into a functional of the electron density. Phys. Rev. B 1988, 37, 785–789. [Google Scholar] [CrossRef]
  32. Miehlich, B.; Savin, A.; Stoll, H.; Preuss, H. Results obtained with the correlation energy density functionals of Becke and Lee, Yang and Parr. Chem. Phys. Lett. 1989, 157, 200–206. [Google Scholar] [CrossRef]
  33. Ditchfield, R.; Hehre, W.J.; Pople, J.A. Self-consistent molecular-orbital methods. IX. An extended Gaussian-type basis for molecular-orbital studies of organic molecules. J. Chem. Phys. 1971, 54, 724–728. [Google Scholar] [CrossRef]
  34. Frisch, M.J.; Pople, J.A.; Binkley, J.S. Self-consistent molecular orbital methods 25. Supplementary functions for Gaussian basis sets. J. Chem. Phys. 1984, 80, 3265–3269. [Google Scholar] [CrossRef]
  35. London, F. Quantum theory of interatomic currents in aromatic compounds. J. Phys. Radium 1937, 8, 397–409. [Google Scholar] [CrossRef]
  36. Ditchfield, R. Self-consistent perturbation theory of diamagnetism. Mol. Phys. 1974, 27, 789–807. [Google Scholar] [CrossRef]
  37. 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]
  38. 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]
  39. Blanco, F.; Alkorta, I.; Elguero, J. Spectral assignments and reference data—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]
  40. Compendium of Chemical Terminology. Available online: http://goldbook.iupac.org (accessed on 03 February 2019).
Figure 1. The four tris [3(5)-methyl]pyrazol-1-ylmethanes and the (0.6 + 0.4)3 proportions.
Figure 1. The four tris [3(5)-methyl]pyrazol-1-ylmethanes and the (0.6 + 0.4)3 proportions.
Molecules 24 00568 g001
Figure 2. The 1H-NMR spectrum in CDCl3 of the Csp3-H proton of the crude mixture (integration in the vertical scale) [8].
Figure 2. The 1H-NMR spectrum in CDCl3 of the Csp3-H proton of the crude mixture (integration in the vertical scale) [8].
Molecules 24 00568 g002
Figure 3. A view of the structure of (a) the 333 isomer (Cambridge Structural Database (CSD) [21] refcode: TUYZEU) and (b) the 335 isomer (bis(3-methylpyrazolyl, 5-methylpyrazolyl)methane). Displacement ellipsoids are drawn at 50% probability level. Hydrogen atoms are represented as spheres of 0.1Å radii.
Figure 3. A view of the structure of (a) the 333 isomer (Cambridge Structural Database (CSD) [21] refcode: TUYZEU) and (b) the 335 isomer (bis(3-methylpyrazolyl, 5-methylpyrazolyl)methane). Displacement ellipsoids are drawn at 50% probability level. Hydrogen atoms are represented as spheres of 0.1Å radii.
Molecules 24 00568 g003
Figure 4. Evolution of the 333/335 ratio with time. (a) Recorded after 1 h in CDCl3, about 95% of 335 isomer and 5% of 333 isomer; (b) recorded after one night in CDCl3, about 50/50 of 335 and 333 isomer.
Figure 4. Evolution of the 333/335 ratio with time. (a) Recorded after 1 h in CDCl3, about 95% of 335 isomer and 5% of 333 isomer; (b) recorded after one night in CDCl3, about 50/50 of 335 and 333 isomer.
Molecules 24 00568 g004
Figure 5. A view of the molecular structure of compound 335 showing the highest difference peaks (e Å−3). Dotted lines joining these peaks mimic the disorder model observed and refined for the other crystal collected at low temperature.
Figure 5. A view of the molecular structure of compound 335 showing the highest difference peaks (e Å−3). Dotted lines joining these peaks mimic the disorder model observed and refined for the other crystal collected at low temperature.
Molecules 24 00568 g005
Figure 6. Three views of the structure of compound 335.
Figure 6. Three views of the structure of compound 335.
Molecules 24 00568 g006
Figure 7. Molecular diagram of 17 trispyrazolylmethane derivatives recorded in the CSD version 5.39, updates to Feb 2018. (c cocrystallized with tmeda, s1 benzene and s2 n-hexane solvates).
Figure 7. Molecular diagram of 17 trispyrazolylmethane derivatives recorded in the CSD version 5.39, updates to Feb 2018. (c cocrystallized with tmeda, s1 benzene and s2 n-hexane solvates).
Molecules 24 00568 g007
Figure 8. The most relevant calculated structures (the geometries of all the structures are to be found in the ESI).
Figure 8. The most relevant calculated structures (the geometries of all the structures are to be found in the ESI).
Molecules 24 00568 g008aMolecules 24 00568 g008b
Table 1. 1H-NMR data of trispyrazolylmethanes at 400.13 MHz. Solvent CDCl3. Chemical shifts (δ, ppm) in ppm; 1H-1H spin-spin coupling constants J (Hz).
Table 1. 1H-NMR data of trispyrazolylmethanes at 400.13 MHz. Solvent CDCl3. Chemical shifts (δ, ppm) in ppm; 1H-1H spin-spin coupling constants J (Hz).
Isomer3-Methyl5-MethylCH
333H4: 6.11 (d, 3JH4H5 = 2.5)
H5: 7.38 (d, 3JH5H4 = 2.5)
CH3: 2.27 (s)
----8.12
8.11 [8]
335H4: 6.11 (d, 3JH4H5 = 2.5)
H5: 7.33 (d, 3JH5H4 = 2.5)
CH3: 2.27 (s)
H3: 7.55 (d, 3JH3H4 = 1.5)
H4: 6.11 (dq, 3JH4H3 = 1.5, 4JH4Me = 0.4)
CH3: 2.39 (d, 4JMeH4 = 0.4)
8.21
8.21 [8]
355H4: 6.12 (dq, 3JH4H5 = 2.5, 4JMeH4 = 0.4)
H5: 7.33 (d, 3JH5H4 = 2.5)
CH3: 2.29 (d)
H3: 7.37 (dq,3JH3H4 = 1.7, 5J H3Me = 0.4)
H4: 5.72 (dq, 3JH4H3 = 1.7, 4JH4Me = 0.8)
CH3: 2.23 (dd, 4JMeH4 = 0.8, 5JMeH3 = 0.4)
8.26 (s)
8.26 [8]
555----H3: 7.51 (ddq, 3JH3H4 = 1.6, 5JMeH3 = 0.4)
H4: 5.80 (dq, 3JH4H3 = 1.7, 4JH4Me = 0.8)
CH3: 2.09 (dd, 4JMeH4 = 0.8, 5JMeH3 = 0.4)
8.31 (s)
8.33 [8]
Table 2. 1H (400.13 MHz), 13C (100.62 MHz) and 15N (40.54 MHz) NMR data of trispyrazolylmethanes. Solvent C6D6. Chemical shifts (δ, ppm) in ppm; 1H-1H and 1H-13C spin-spin coupling constants J (Hz).
Table 2. 1H (400.13 MHz), 13C (100.62 MHz) and 15N (40.54 MHz) NMR data of trispyrazolylmethanes. Solvent C6D6. Chemical shifts (δ, ppm) in ppm; 1H-1H and 1H-13C spin-spin coupling constants J (Hz).
Isomer3-Methyl5-MethylCH
333N1: –173.3
N2: –79.5
--------
C3: 150.9
C4: 107.1, 1J = 175.3, 2J = 6.8, 3J = 3.4
C5: 130.6, 1J = 188.5, 2J = 9.5, 3J = 2.5
CH3: 14.0, 1J = 127.3
----83.8
1J = 166.4
H4: 5.76 (d, 3JH4H5 = 2.5)
H5: 7.27 (d, 3JH5H4 = 2.5)
CH3: 2.08 (s)
----8.22
335N1: –172.8
N2: –77.9
N1: –171.4
N2: –80.4
----
C3: 150.5
C4: 107.3, 1J = 175.0, 2J = 8.3, 3J = 3.3
C5: 130.3, 1J = 189.3, 2J = 9.5, 3J = 2.6
CH3: 13.9, 1J = 127.2
C3: 141.2, 1J = 185.2, 2J = 5.7
C4: 107.0, 1J = 175.1, 2J = 10.6, 3J = 3.8
C5: 140.3
CH3: 10.5, 1J = 128.5
81.4
1J = 164.2
H4: 5.82 (d, 3JH4H5 = 2.5)
H5: 7.56 (d, 3JH5H4 = 2.5)
CH3: 2.11 (s)
H3: 7.36 (d, 3JH3H4 = 1.7)
H4: 5.62 (dq, 3JH4H3 = 1.7, 4JH4Me = 0.8)
CH3: 1.82 (d, 4JMeH4 = 0.8)
8.40 (s)
355N1: –175.6
N2: –78.8
N1: –171.3
N2: –76.5
----
C3: 151.0
C4: 106.8, 1J = 175.3
C5: 131.3,1J = 190.3, 2J = 9.9
CH3: 14.0, 1J = 127.3
C3: 140.7, 1J = 185.1, 2J = 5.8
C4: 107.8, 1J = 174.8
C5: 140.5
CH3: 10.9, 1J = 129.1
81.6
1J = 164.6
H4: 5.82 (3JH4H5 = 2.5)
H5: 7.28 (3JH5H4 = 2.5)
CH3: 2.14(s)
H3: 7.37 (3JH3H4 = 1.7)
H4: 5.72 (dq, 3JH4H3 = 1.7, 4JH4Me = 0.8)
CH3: 1.87 (m)
8.46 (s)
555----N1: –173.4
N2: –75.4
----
----C3: 140.1, 1J = 185.0, J = 5.8
C4: 108.1, 1J = 174.7 (m)
C5: 140.5
CH3: 10.9, 1J = 128.6
82.1
1J = 163.1
----H3: 7.37 (3JH3H4 = 1.7, H3)
H4: 5.80 (dq, 3JH4H3 = 1.7, 4JH4Me = 0.8)
CH3: 1.84 (m)
8.43 (s)
Table 3. Best conformers according to the method. 1H in CDCl3; 13C and 15N in C6D6.
Table 3. Best conformers according to the method. 1H in CDCl3; 13C and 15N in C6D6.
CompConformationErel1H13C15NX-ray
333ddd37.1ddd
udd11.2 udd
uud0.0uuduuduuduud: TUYZEU
uuu2.4 uuu
335ddd37.4
ddu15.1ddu
duu17.3duu
udd12.6 uddudd
uud0.0 uuduuduud: this work
uuu12.1
355ddd23.2
ddu0.0ddu udd
duu2.5duu
udd4.1 udd
uud9.9
uuu10.7 uuu
555ddd19.3
udd0.0udduddudd
uud5.9uuduuduud
uuu18.5
Table 4. Selected geometrical parameters for 333 and 335.
Table 4. Selected geometrical parameters for 333 and 335.
333uud (TUYZEU)335uud
pz_Apz_Bpz_Cpz_Apz_Bpz_C
N1-N21.3591.3621.3551.359(5)1.364(5)1.346(6)
N2-C31.3301.3281.3351.328(6)1.332(5)1.351(7)
C3-C41.4011.4021.3951.375(7)1.383(6)1.338(8)
C4-C51.3611.3621.3571.360(7)1.343(6)1.339(7)
N1-C51.3491.3511.3471.326(5)1.350(5)1.402(6)
C1-N11.4451.4441.4401.453(5)1.455(5)1.483(6)
C5-N1-N2111.8112.1112.5111.2(4)111.8(3)112.0(4)
N1-N2-C3104.8104.6104.3105.1(4)104.1(3)105.0(4)
N2-C3-C4110.9111.0110.7110.5(4)110.9(4)108.8(5)
C3-C4-C5105.5105.8106.2106.1(4)106.9(4)112.1(5)
C4-C5-N1107.0106.5106.2107.2(4)106.4(4)102.1(5)
C1-N1-N2117.5117.4121.5117.0(3)118.3(3)119.6(3)
N2/1-C3/5-C6120.2120.4120.6120.8(5)120.5(4)120.5(5)
C4-C3/5-C6128.9128.5128.7128.6(5)128.7(4)137.5(5)
H1-C1-N1-N224.528.0−170.939.0(6)17.6(7)−168.1(3)
H1-C1-N1-C5−164.8−160.010.8−150.3(5)−163.4(5)15.3(5)
C1-N1-N2-C3173.4173.6−178.4172.6(4)179.9(4)−177.7(4)
Table 5. Energies (kJ·mol–1). The minimum energy conformation in black; the crystallographic structure, TUYZEU, in italics.
Table 5. Energies (kJ·mol–1). The minimum energy conformation in black; the crystallographic structure, TUYZEU, in italics.
CompoundConformationRelative Energy, Conformations Relative Energy, Isomers
333ddd37.1
udd11.2
uud0.00.0
uuu2.4
TUYZEUuud0.0
335ddd37.439.1
ddu15.116.8
duu17.319.0
udd12.614.2
uud0.01.7
uuu12.113.8
355ddd23.241.0
ddu0.017.9
duu2.520.4
udd4.122.0
uud9.927.8
uuu10.728.6
555ddd19.343.1
udd0.023.8
uud5.929.7
uuu18.542.3
Table 6. Torsions (°). The minimum energy conformation in black; the crystallographic structure, TUYZEU, in italics.
Table 6. Torsions (°). The minimum energy conformation in black; the crystallographic structure, TUYZEU, in italics.
CompoundConformation8-7-1-2313-12-1-233-2-1-23
333ddd145.3145.3145.3
udd−5.5177.7159.1
uud−32.9−28.1174.4
uuu−40.5−40.5−40.5
TUYZEUuud−32.9−28.2174.4
X-rayuud−27.97−24.48170.87
335ddd−143.4−149.9−136.1
ddu−173.0−85.712.3
duu145.3−38.6−26.9
udd−170.14.7−151.7
uud−31.7−26.7174.9
uuu−52.1−41.1−29.0
X-rayuud−39.0(6)−17.4(7)168.2(3)
355ddd−136.2−147.6−139.6
ddu−173.0−85.712.3
duu145.3−38.6−26.9
udd−170.14.7−151.7
uud−57.9162.3−22.9
uuu−43.3−43.1−34.4
555ddd139.2139.2139.2
udd−18.4129.3151.6
uud−24.2−50.0156.1
uuu−40.2−40.2−40.2

Share and Cite

MDPI and ACS Style

Silva, V.L.M.; Silva, A.M.S.; Claramunt, R.M.; Sanz, D.; Infantes, L.; Martínez-López, Á.; Reviriego, F.; Alkorta, I.; Elguero, J. An Example of Polynomial Expansion: The Reaction of 3(5)-Methyl-1H-Pyrazole with Chloroform and Characterization of the Four Isomers. Molecules 2019, 24, 568. https://doi.org/10.3390/molecules24030568

AMA Style

Silva VLM, Silva AMS, Claramunt RM, Sanz D, Infantes L, Martínez-López Á, Reviriego F, Alkorta I, Elguero J. An Example of Polynomial Expansion: The Reaction of 3(5)-Methyl-1H-Pyrazole with Chloroform and Characterization of the Four Isomers. Molecules. 2019; 24(3):568. https://doi.org/10.3390/molecules24030568

Chicago/Turabian Style

Silva, Vera L. M., Artur M. S. Silva, Rosa M. Claramunt, Dionisia Sanz, Lourdes Infantes, Ángela Martínez-López, Felipe Reviriego, Ibon Alkorta, and José Elguero. 2019. "An Example of Polynomial Expansion: The Reaction of 3(5)-Methyl-1H-Pyrazole with Chloroform and Characterization of the Four Isomers" Molecules 24, no. 3: 568. https://doi.org/10.3390/molecules24030568

Article Metrics

Back to TopTop