Next Article in Journal
Impact Response of Preplaced Aggregate Fibrous Concrete Hammerhead Pier Beam Designed with Topology Optimization
Previous Article in Journal
Transformation of SnS Nanocompisites to Sn and S Nanoparticles during Lithiation
Previous Article in Special Issue
Intramolecular Hydrogen Bond Energy and Its Decomposition—O–H∙∙∙O Interactions
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Weak Interactions in the Structures of Newly Synthesized (–)-Cytisine Amino Acid Derivatives

by
Anna K. Przybył
*,
Anita M. Grzeskiewicz
and
Maciej Kubicki
*
Faculty of Chemistry, A. Mickiewicz University in Poznań, Uniwersytetu Poznańskiego 8, 61-614 Poznań, Poland
*
Authors to whom correspondence should be addressed.
Crystals 2021, 11(2), 146; https://doi.org/10.3390/cryst11020146
Submission received: 15 January 2021 / Revised: 24 January 2021 / Accepted: 26 January 2021 / Published: 30 January 2021
(This article belongs to the Special Issue Hydrogen Bonds in Crystals)

Abstract

:
Eight new (–)-(N-[(AA)-(N-phtaloyl)]cytisines (where AA is amino acid: glycine, β-alanine, D,L-valine, L-valine, L-isoleucine, L-leucine, D-leucine and D,L-phenyloalanine), were synthesized and fully spectroscopically characterized (NMR, FTIR and MS). For two of these compounds, N-[glycine-(N-phtaloyl)]cytisine and N-[L-isoleucine-(N-phtaloyl)]cytisine, X-ray crystal structures were obtained and used as the basis for an in-depth analysis of intermolecular interactions and packing energies. The structural geometrical data (weak hydrogen bonds, π···π interactions, etc.) were compared with the energies of interactions and the topological characteristics (electron density, Laplacian at the appropriate critical point) based on the atoms-in-molecules theory. The results suggest that there is no straightforward connection between the geometry of point-to-point interactions and the molecule-to-molecule energies. Additionally, the usefulness of the transfer of multipolar parameters in estimating of critical points’ characteristics have been confirmed.

1. Introduction

Intermolecular interactions are responsible for the aggregation of molecules into molecular crystals as well as into other larger moieties and in general into supramolecules [1]. This is the core of supramolecular chemistry, one of the most swiftly developing branches of chemistry, which found numerous applications in such different topics as molecular self-assembly, nanotechnology, catalysis, drug delivery and molecular switches (cf. e.g., [2]).
Therefore, systematical studies of intermolecular interactions are relevant for general chemistry, as well as for biology, material science, etc. Starting from “classical” hydrogen bonds of, e.g., O–H···O or O–H···N type, well-known for more than a century, the realm of such interactions expanded steadily to become a huge conglomerate of very different members, differing in their energies, geometries, nature, importance, etc. These include non-classical hydrogen bonds, for instance with C–H donors (which were the subject of fiery debate in the 1970s, e.g. [3]), with π–electron acceptors, etc., as well as halogen bonding [4], chalcogen bonding [5], pnicogen bonding [6], tetrel bonding [7], π···π interactions [8], cation···π interactions [9], anion···π interactions [10], etc. The huge variety of (more or less) real interactions makes it necessary to investigate their mutual importance, hierarchies and characteristics.
The aims for such studies include geometrical characteristics (distances, angles), most conveniently obtained by means of X-ray diffraction, energetic characteristics (interaction energies), which can be obtained by quantum-chemical calculations, topological characteristics (atoms-in-molecules description), etc. Here we present the results of such wide studies of intermolecular interactions in the crystal structures of two (–)-cytisine derivatives. A number of such derivatives was synthesized (Figure 1) and spectroscopically characterized, although for the sake of this study, two derivatives were picked (N-[glycine-(N-phtaloyl)]cytisine (3A) and N-[L-isoleucine-(N-phtaloyl)]cytisine (3E)), because for these compounds we were able to obtain the crystals of quality appropriate for X-ray structural analysis. Additionally, we wanted to check the importance of the quality of the electron density model in estimating the critical points’ characteristics, e.g., density at the critical point or the Laplacian value by comparing the values obtained for the standard X-ray determination with/without modifications, such as hydrogen position correction or multipolar parameters transfer with the results of high-resolution diffraction studies of similar compounds.
(–)-Cytisine is a natural product, extracted from the plants of the Fabaceae (Leguminosae) family, and its importance stems from the fact that it has been found to have an affinity towards the specific subunits of nicotine acetylcholine receptors (nAChRs). The strongest nAChR agonists (receptor activators) are nicotine and epibatidine, which, however, due to their high toxicity, cannot be used in medical therapy [11]. Hence the increased importance of (–)-cytisine, which is a more selective ligand towards main nAChR receptors and less toxic than nicotine. Because of the similar mode of action and lower toxicity, (–)-cytisine (1) has been applied as a nicotine substitute in antinicotine therapy. (–)-Cytisine as a partial agonist of nAChR, and moderately increases the concentration of dopamine and thus alleviates the symptoms of nicotine withdrawal, the so-called nicotine craving. When administered together with nicotine, (–)-cytisine acts antagonistically towards nicotine, which reduces the receptor’s response to the latter [12].
In the search for new ligands interacting with nicotine receptors, a number of (–)-cytisine derivatives have been obtained. The synthetic analogues of (1) have been mainly obtained by substitution of the nitrogen atom N–12 in ring C (ring names as in Scheme 1) and modification of the quasi-aromatic ring A [13]. In general, substitution at N-12 is considered as non-conducive to increasing affinity to receptors, but it is the simplest method to enhance the lipophilicity of (–)-cytisine, which increases the chances of its overcoming the blood–brain barrier. Due to the broad spectrum of effects, (1) is considered as a pharmacophoric block to conjugate with other synthetic or natural compounds. It has been proven that the linkers of different lengths and bioactivities are important factors in the effective design of new bioactive compounds. Additionally, it has been shown that the combination of two biological agents including cytisine and camphecene make a promising novel class of potential antiviral compounds [14].

2. Materials and Methods

Flash chromatography was carried out on silica gel 60 G F254 (Merck). Melting points were determined on a Boetius apparatus (PHMK 05 VEB Wagetechnik Rapido, Radebeul). EIMS mass spectra were recorded using a model 402 two-sector mass spectrometer (AMD Intectra GmbH, Harpsted, Germany; ionizing voltage 70 eV, accelerating voltage 8 kV, resolution 1000 for low-resolution and 10 000 for high-resolution mass spectra). IR spectra were obtained on a FT-IR Bruker IFS 113v spectrometer (KBr pellets technique). NMR spectra were recorded on a Bruker AVANCE II 400 (400.13 MHz) spectrometer, with a 5 mm broadband probe head with actively shielded z gradient coil (90° 1H pulse width 10.8 s, 13C pulse width 12 μs.

2.1. Synthesis

General procedure for the synthesis of cytisine amino-acid derivatives (Figure 1 and Scheme 1):
A suitable amino acid with phthalic anhydride in a mortar in a 1:1 ratio were triturated. The ground reagents were transferred to a flask. For 1 hour, the substrates were melted in an oil bath at 150–160 °C under an argon atmosphere [15]. The resulting precipitate or oil was allowed to cool. The course of the reaction was monitored by TLC and developed in solution with ninhydrin to detect unreacted amino acid residues.
Blocked amino acids were treated with excess thionyl chloride (eq. 1:6.5). The mixture was cooled for about 30 minutes in an ice bath. Then the reagents were heated for 3 h [16]. The excess of thionyl chloride was removed on a rotary evaporator. The dissolved solid (in the case of compounds 2A, 2B and 2H—the details in SI) or the yellow oil (in the case of compounds 2C2G—the details in SI) was dissolved in toluene to separate the product from insoluble impurities. The solvent was evaporated, and the product obtained was washed with warm hexane and left to dry [17]. Compound 2 was dissolved in a solution of CH2Cl2/toluene (2:1) and cooled in an ice bath. Then cytisine dissolved in the same solvent mixture was added (molar ratio 1:1). Next, 0.5 mL of triethylamine was added, and the reaction mixture was stirred overnight at room temperature [16]. The solvent was evaporated and the product purified on an Al2O3 column and washed out with CH2Cl2. The yield of the obtained compounds:
3A (N-[glycine-(N-phtaloyl)]cytisine), 97%; 3B (N-[β-alanine-(N-phtaloyl)]cytisine), 54%; 3C (N-[D,L-valine-(N-phtaloyl)]cytisine) 60%; 3D (N-[L-valine-(N-phtaloyl)]cytisine), 69%; 3E (N-[L-isoleucine-(N-phtaloyl)]cytisine), 58%; 3F (N-[L-leucine-(N-phtaloyl)]cytisine), 89%; 3G (N-[D-leucine-(N-phtaloyl)]cytisine), 60%; 3H (N-[D,L-phenyloalanine-(N-phtaloyl)]cytisine), 47%.
Details of spectral analysis are submitted as Supplementary Information.
Mass Spectrometry: The degree of contamination of the products was evaluated on the basis of GC–MS, although this method did not permit detection of triethylamine and its salt in the samples studied. Analysis of chromatograms provided the first information on the retention times (in Supplementary Information).
L-isomers were found to be characterized by longer Rt than their D-isomers, for example, for D-valine (3C) Rt = 43.336, while for L-valine (3D) Rt = 47.642 and in the case of D-leucine (3G) Rt = 51.099, and L-leucine (3F) Rt = 53.591 minutes. The obtained amino acid derivatives of cytisine were subjected to EI–MS examination [18,19], showing that molecular ions of the compounds studied, formed upon EI, undergo mass fragmentation in which the bonds at both sides of the carbonyl group break up, leading to formation of even-electron fragment ions i and g (Table 1, Scheme 2). Breaking up of a radical of the ion g structure from the molecular ion gives the ion f, which gives the ion i as a result of elimination of a CO molecule. The even-electron fragment ion g is the parent ion for two subsequent ions j and m. For the derivatives L-isoleucine (3E) and L-leucine (3F), we found that it is possible to distinguish between the two regio-isomers, as the spectrum of L-leucine (3F) differs from that of L-isoleucine (3E) by a higher relative abundance of the ions d and l (Table 1, Scheme 2).
The molecular ions (M+) obtained for the compounds containing fragments of valine, leucine or isoleucine also undergoes decomposition involving elimination of CH2=C(CH3)2 molecule, which leads to the ion at m/z 377 (Table 1, Scheme 2).
A detail analysis of EI–MS spectra (SI) permits distinction of stereoisomers D and L, present in the series of the analyzed compounds 3A–H, just on the basis of relative abundances of selected fragmentation ions. In the spectra of stereoisomers D (3C-D; 3G; 3H-D, cytisine derivatives with valine or leucine) the relative intensity of the molecular ion (M+) is higher and the fragmentation ion d is lower than the corresponding ones in the spectra of stereoisomers L (3D, 3F, 3HL).
NMR spectra analysis. Amide conformers of cytisine derivatives occur in solutions in dynamic equilibrium [20,21]. Additionally, in the case of amino acid derivatives of cytisine, the presence of two cis and trans conformers is observed in 1H NMR and 13C NMR spectra (SI), which is recorded as a double set of signals. Each set corresponds to one of the two conformations of the analyzed compounds (Table 2).
X-ray diffraction. Diffraction data were collected by the ω-scan technique at room temperature on two Rigaku four-circle diffractometers: for 1 on XCalibur diffractometer with Sapphire2 CCD detector and graphite-monochromated MoKα radiation (λ = 0.71069 Å), and for 2 on SuperNova with Atlas detector and mirror-monochromated CuKα radiation (λ = 1.54178 Å). The data were corrected for Lorentz-polarization as well as for absorption effects [22]. Precise unit-cell parameters were determined by a least-squares fit of 2190 (1), and 11309 (2) reflections of the highest intensity, chosen from the whole experiment. The structures were solved with SHELXT-2013 [23] and refined with the full-matrix least-squares procedure on F2 by SHELXL-2013 [24]. All non-hydrogen atoms were refined anisotropically, and hydrogen atoms were placed in idealized positions and refined as a "riding model" with isotropic displacement parameters set at 1.2 (1.5 for methyl groups) times Ueq of appropriate carrier atoms.

2.2. Crystal Data

1: C21H19N3O4, Mr = 377.39, orthorhombic, P212121, a = 6.5691(6) Å, b = 12.2108(11) Å, c = 22.0780(16) Å, V = 1771.0(3) Å, Z = 4, dx = 1.415 g·cm−3, F(000) = 792 μ(MoKα) = 0.100 mm−1, 9310 reflections measured up to Θ = 26.91, 3505 symmetry-independent (Rint = 2.99%), 2740 with = I > 2σ(I). Final R[I > 2σ(I)] = 4.10%, wR2[I > 2σ(I)] = 8.12%, R[all refl.] = 6.17%, wR2[all refl.] = 9.07%, S = 1.003, Δρmax/Δρmin = 0.12/–0.16e·Å−3.
2: C25H27N3O4·CH3CN, Mr = 474.55, orthorhombic, P212121, a= 7.26005(11) Å, b = 8.55757(14) Å, c = 39.3690(7) Å, V = 2445.93(7) Å, Z = 4, dx = 1.289 g·cm−3, F(000) = 1008 μ(CuKα) = 0.712 mm−1, 22727 reflections measured up to Θ = 75.55, 4966 symmetry-independent (Rint = 5.95%), 4726 with I > 2σ(I). Final R[I > 2σ(I)] = 5.36%, wR2[I > 2σ(I)] = 15.21%, R[all refl.] = 5.57%, wR2[all refl.] = 15.38%, S=1.063, Δρmax/Δρmin = 0.31/−0.16e·Å−3.
Energy calculations. The calculations of interaction energies between pairs of molecules and packing energies were performed with two methods:
(a)
Using wavefunctions at B3LYP/6-31G(d,p) level (hereinafter: B3LYP), the energy of interaction was calculated within the CrystalExplorer software [25] in terms of four key components: electrostatic, polarization, dispersion and exchange–repulsion Etot = keleEele + kpolEpol + kdisEdis +k repErep
(b)
and with the PIXEL method [26,27], as included in Mercury program [28].
In both cases the hydrogen atoms were moved to the average geometry as determined by neutron diffraction.
AIM topological analysis. The topology (atoms-in-molecules, [29]) of electron density distribution has been calculated using MoPro software [30].

3. Results and Discussion

Eight new (–)-(N-[AA-(N-phtaloyl)]cytisines were synthesized (cf. Figure 1). For two of them (3A (hereinafter 1) and 3E (2)) the crystal structures were determined, and the wide analysis of the structures and intermolecular interactions were performed. The comparison of the molecules shows that the overall conformations are radically different. Of course, in both two cases the cytisine moieties are rigid and the variations in their geometries are negligible, but the mutual orientation of the cytisine and phthalimide fragments is totally different. Figure 2 shows the comparison of two molecules with cytisine fragments fitted one onto another.
In 3A the ring A and phthalimide planes make an angle of 68.7°, while in 3E these two planes are almost parallel, with dihedral angle of 14.0°. In this latter case intramolecular π-stacking (distance between the planes is ca. 3.7Å) might afford the additional factor stabilizing this conformation.
Figure 3 and Figure 4 show the perspective views of the molecules 3A and 3E, respectively, and Table 3 lists the relevant geometrical parameters.
It might be noted that in nine N-carbonyl cytisine derivatives deposited in the CSD [31], only in one case (carbamide, 6-oxo-7,11-diazatricyclo[7.3.1.02,7]trideca-2,4-diene-11-carboxamide [32]) the conformation of C=O bond is anti with respect to C11–N12 bond (as in 1), while in the other eight cases the conformation is syn, like in 2, with the mean value of C11–N12–C14–O14 torsion angle at 2.2(19)°.

3.1. The Importance of the Quality of the Electron Density Model

Three different models of electron density distribution were tested:
(a) original X-ray independent atom model (IAM-standard model of X-ray structure determination);
(b) IAM model with hydrogen atoms moved to the average geometries determined by means of neutron diffraction; and
(c) The model with geometry as in (b) but upgraded to the multipolar level [33] by transferring the coefficients of multipolar expansion from the ELMAM2 database [34], (except for the cyan group, for which the parameters were transferred from the published molecule of [35] due to the absence in ELMAM2).
It turned out that the results obtained in the latter two approaches for intermolecular contacts were quite similar in terms of both critical points positions and their characteristics, but the results for intramolecular bonds showed obvious advantages of transferring procedure, especially for polarized C=O bonds.
To have deeper insight into this, we compared the topological parameters of the covalent bonds with those obtained experimentally, by means of high-resolution diffraction studies, for similar compounds: (–)-cytisine and N-methyl-cytisine [36]. The results of such comparison for selected bonds are presented in Figure 5 for electron density value at the critical point and for Laplacian value. The results for all bonds between non–H atoms are submitted as supplementary materials.
The results show clearly that the transfer of multipolar parameters improves the quality of the electron density maps. It is especially obvious for Laplacian values, where for most of the bonds even the sign is incorrect. It might be noted also that the distances between atoms, and in general geometries of the cytisine skeletons were almost identical in all studied cases.

3.2. Crystal Packing, Intermolecular Interactions

Due to the absence of strong hydrogen bond donors, charged species and halogen atoms, the crystal architecture is determined by a number of weak interatomic interactions. Here we are going to show the comparison between different approaches: geometrical, energy calculations and topological analysis.
The most important interactions can be conveniently identified by means of Hirshfeld surface analysis [37], which allows one to visualize the regions of the molecule under consideration with the closest contacts (in the terms of van der Waals radii) with neighboring molecules.
In such a way, the closest (as compared to sums of van der Waals radii) intermolecular contacts have been found and then each of them analyzed (Figure 6).
In N-[glycine-(N-phtaloyl)]cytisine (1), the most pronounced contact is related with the C9–H9···O14 (−1 + x,y,z) weak hydrogen bond, with the shortest H···O distance (2.45 Å), which connects the molecules into infinite chains along the x direction. It is also the interaction which defines two molecules with the highest interaction energy, calculated at −51.8 kJ/mol by B3LYP and at −55.3kJ/mol by the PIXEL method (cf. experimental part). The critical point found for this interaction is also described by the highest values of electron density (0.092/0.072 e/Å3; hereinafter the first value is obtained for the IAM model with elongated C–H bonds, and the second one with the model with transferred multipole parameters; cf. Experimental part) and Laplacian (1.099/1.090 e/Å5) at the CP. The second shortest H···O distance was found for H15A···O24 (−1/2 + x, 3/2−y, −z), and for these two molecules also the second-highest energies (−43.0/−40.9 kJ/mol) and the critical point of second-highest electron density (0.078/0.065) and Laplacian (0.939/0.771) were calculated. More details can be found in Table 3; here, we would like to show examples for which differences between DFT (B3LYP) and PIXEL methods are significant. For two molecules connected by weak H21···O2 (3/2 + x,3/2 − y, −z), with the H···O distance of 2.59Å, the interaction energy calculated with the first method is −17.8 kJ/mol, while with the second one it is −9.2 kJ/mol. For this interaction, well-defined critical points with one of the highest characteristics (0.076/0.059 and 0.910/0.718) were found.
Additionally, in N-[L-isoleucine-(N-phtaloyl)]cytisine (2), the shortest C–H···O contact (H5·· O14, 2.38 Å), again between the molecules related by unique translation along x, is connected with the highest interaction energy (−44.2/−45.8 kJ/mol). The characteristics of the appropriate bond critical point (0.119/0.093, 1.363/1.169) indicate the strongest interaction. Energetically comparable is the interaction between molecules related by y unit-cell translation, although related with longer H13A···O2 contact of 2.65 Å. The energies calculated for this pair are −41.7 and −35.6 kJ/mol for DFT (B3LYP) and PIXEL methods, respectively. This energy is significantly larger than that calculated for apparently shorter contact, H3A2···O2 (2.45 Å): −26.8/−9.3 kJ/mol. Interestingly enough, both the electron density and Laplacian values at the appropriate critical points have the higher values for the latter contact (0.096/0.075 and 1.127/0.909) than for the former one (0.061/0.042 and 0.757/0.731).
The explanation of such discrepancies lies probably in the number of interactions (or contacts) between the molecules, as the overall interaction energies consider all these interactions, even very weak ones. For instance, taking the strongest intermolecular interactions in 1, this between molecules at (x,y,z) and (−1 + x,y,z), there are eight contacts for which critical points were found. The results for this contact are presented in Figure 7 and Table 4. The full set of such comparisons is submitted as supplementary information.
Similarly, for 2, the highest energy was found for the pair of molecules, for which as many as 10 contact critical points were found (Figure 7 and Figure 8, Table 5).

4. Conclusions

Eight new derivatives of (–)-cytisine, an alkaloid of biological importance, were synthesized and fully spectroscopically analyzed (NMR, FTIR). The general formula of the new compounds is (–)-(N-[(amino acid)-(N-phtaloyl)]cytisine. Mass spectrometry was used for proposing the fragmentation routes of new compounds. For two derivatives, namely glycine and L-isoleucine derivatives, the X-ray crystal structures were determined and used for in-depth analysis of intermolecular interactions, based on geometries of the structures, quantum chemical calculations of interaction energies, and characterization of the critical points of the electron density distribution based on atoms-in-molecules theory. No direct and strict correlation between the geometrical characteristics (distances, even angles) of the interaction and the energetical or topological parameters could be found. Instead, the results suggest that the interaction energies are correlated with the number of contacts and of critical points between molecules, rather in line with the reasoning of Dunitz and Gavezzotti [38]. Additionally, by comparing the results of the topological analysis with those obtained for cytisine and N-methyl cytisine by means of experimental method (high-resolution diffraction), the importance of the transfer of multipolar parameters was shown—only such a model was able to give results similar to the experimental ones.

Supplementary Materials

The following are available online at https://www.mdpi.com/2073-4352/11/2/146/s1: Table S1: FTIR–characteristic bands for compounds 3 [cm−1]; Figures. S1–S39: GC-MS, EI-MS, FTIR 13V NMR spectra for compounds 3A3F. Tables S2 and S3: Intermolecular interaction details for 3A and 3E.

Author Contributions

Conceptualization, A.K.P. and M.K.; methodology, A.K.P., A.M.G., M.K.; software, A.M.G., M.K.; validation, A.M.G., M.K.; investigation, A.K.P., A.M.G., M.K.; resources, M.K.; writing—original draft preparation, A.K.P., M.K.; writing—review and editing, A.K.P., M.K.; visualization, A.K.P., A.M.G., M.K.; supervision, A.K.P., M.K..; project administration, M.K.; funding acquisition, M.K. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by National Science Center (Poland) project (Grant No. 2013/11/B/ST5/01681).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Crystallographic data for the structural analysis has been deposited with the Cambridge Crystallographic Data Centre, Nos. CCDC–1851176 (1) and 1851175 (2). Copies of this information may be obtained free of charge from: The Director, CCDC, 12 Union Road, Cambridge, CB2 1EZ, UK. Fax: +44(1223)336-033, e-mail: [email protected], or www: www.ccdc.cam.ac.uk.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript, or in the decision to publish the results.

References

  1. Dunitz, J.D. Phase transitions in molecular crystals from a chemical viewpoint. Pure Appl. Chem. 1991, 63, 177–185. [Google Scholar] [CrossRef] [Green Version]
  2. Lehn, J.-M. From supramolecular chemistry towards constitutional dynamic chemistry and adaptive chemistry. Chem. Soc. Rev. 2007, 36, 151–160. [Google Scholar] [CrossRef] [PubMed]
  3. Desiraju, G.R. The C-H···O hydrogen bond: Structural implications and supramolecular design. Acc. Chem. Res. 1996, 29, 441–449. [Google Scholar] [CrossRef] [PubMed]
  4. Cavallo, G.; Metrangolo, P.; Milani, R.; Pilati, T.; Priimagi, A.; Resnati, G.; Terraneo, G. The halogen bond. Chem. Rev. 2016, 116, 2478–2601. [Google Scholar] [CrossRef] [Green Version]
  5. Scilabra, P.; Terraneo, G.; Resnati, G. The chalcogen bond in crystalline solids: A world parallel to halogen bond. Acc. Chem. Res. 2019, 52, 1313–1324. [Google Scholar] [CrossRef] [PubMed]
  6. Scheiner, S. The pnicogen bond: Its relation to hydrogen, halogen, and other noncovalent bonds. Acc. Chem. Res. 2020, 46, 280–288. [Google Scholar] [CrossRef]
  7. Bauza, A.; Mooibroek, T.J.; Frontera, A. Tetrel-Bonding interacion: Rediscovered supramolecular force? Angew. Chem. Int. Ed. 2013, 52, 12317–12321. [Google Scholar] [CrossRef]
  8. Thakuria, R.; Nath, N.K.; Saha, B.K. The nature and applications of π···π interactions: A perspective. Cryst. Growth Des. 2019, 19, 5230528. [Google Scholar] [CrossRef]
  9. Mahadevi, A.S.; Sastry, N. Cation-π interaction: Its role and relevance in chemistry, biology, and material science. Chem. Rev. 2013, 113, 2100–2138. [Google Scholar] [CrossRef]
  10. Wang, D.-X.; Wang, M.-X. Anion-π interactions: Generality, binding strength, and structure. J. Am. Chem. Soc. 2013, 135, 892–897. [Google Scholar] [CrossRef]
  11. Tutka, P.; Zatoński, W. Cytisine for the treatment of nicotine addiction: From a molecule to therapeutic efficiacy. Pharmacol. Rep. 2005, 58, 777–798. [Google Scholar]
  12. Rouden, J.; Lasne, M.-C.; Blanchet, J.; Baudoux, J. (−)-Cytisine and derivatives: Synthesis, reactivity, and applications. Chem. Rev. 2014, 114, 712–778. [Google Scholar] [CrossRef] [PubMed]
  13. Kozikowski, A.P.; Chellappan, S.K.; Xiao, Y.; Bajjuri, K.M.; Yuan, H.; Kellar, K.J.; Petukhov, P.A. Chemical medicine: Novel 10-substituted cytisine derivatives with increased selectivity for α4β2 nicotinic acetylcholine receptors. ChemMedChem 2007, 2, 1157–1161. [Google Scholar] [CrossRef] [PubMed]
  14. Artyushin, O.I.; Moiseeva, A.A.; Zarubaev, V.V.; Slita, A.V.; Galochkina, A.V.; Muryleva, A.A.; Borisevich, S.S.; Yarovaya, O.I.; Salakhutdinov, N.F.; Brel, V.K. Synthesis of camphecene and cytisine conjugates using click chemistry methodology and study of their antiviral activity. Chem. Biodivers. 2019, 16, e1900340. [Google Scholar] [CrossRef] [PubMed]
  15. Bhat, S.V.; Nagasampagi, B.A.; Sivakumar, M. Chemistry of Natural Products; Springer: Berlin/Heidelberg, Germany, 2005; pp. 237–315. [Google Scholar]
  16. Juaristi, E.; Beck, A.K.; Hansen, J.; Matt, T.; Mukhopadhyay, T.; Simson, M.; Seebach, D. Enantioselective aldol and Michael additions of achiral enolates in the presence of chiral lithium amides and amines. Synthesis 1993, 1993, 1271–1290. [Google Scholar] [CrossRef]
  17. Kánai, K.; Podányi, B.; Bokotey, S.; Hajdú, F.; Hermecz, I. Stereoselective sulfoxide formation from a thioproline derivative. Tetrahedron Asymmetry 2002, 13, 491–496. [Google Scholar] [CrossRef]
  18. Przybył, A.K.; Prukała, W.; Kikut-Ligaj, D. EI-MS study of selected N-amide and N-alkyl derivatives of cytisine. Rapid Commun. Mass Spectrom. 2007, 21, 1409–1413. [Google Scholar] [CrossRef]
  19. Przybył, A.K.; Nowakowska, Z. EIMS study of Halogenated N-Acetyl- and N-Propionylcytisines. Rapid Commun. Mass Spectrom. 2011, 25, 1193–1197. [Google Scholar] [CrossRef]
  20. Przybył, A.K.; Kubicki, M. A comparative study of dynamic NMR spectroscopy in analysis of selected N-alkyl-, N-acyl-, and halogenated cytisine derivatives. J. Mol. Struct. 2011, 985, 157–166. [Google Scholar] [CrossRef]
  21. Przybył, A.K.; Kubicki, M.; Hoffmann, M. The amide protonation of (-)-N-benzoyl-cytisine in its perchlorate salts. Spectrochim. Acta A Mol. Biomol. Spectr. 2014, 129, 1–6. [Google Scholar] [CrossRef]
  22. CrysAlisPro; Rigaku Oxford Diffraction LTd.: Oxford, UK, 2013.
  23. Sheldrick, G.M. SHELXT—Integrated space-group and crystal-structure determination. Acta Cryst. Part A 2015, 71, 3–8. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Sheldrick, G.M. Crystal structure refinement with SHELXL. Acta Cryst. Part C 2015, 71, 3–8. [Google Scholar] [CrossRef] [PubMed]
  25. Turner, M.J.; McKinnon, J.J.; Wolff, S.K.; Grimwood, D.J.; Spackman, P.R.; Jayatilaka, D.; Spackman, M.A. CrystalExplorer17. University of Western Australia. 2017. Available online: http://crystalexplorer.scb.uwa.edu.au/ (accessed on 29 January 2021).
  26. Gavezzotti, A. Are crystal structures predictable? Acc. Chem. Res. 1994, 27, 309–314. [Google Scholar] [CrossRef]
  27. Gavezzotti, A.; Fillippini, G. Geometry of the intermolecular X-H···Y (X, Y = N, O) hydrogen bond and the calibration of empirical hydrogen-bond potentials. J. Phys. Chem. 1994, 98, 4831–4837. [Google Scholar] [CrossRef]
  28. Macrae, C.F.; Sovago, I.; Cottrell, S.J.; Galek, P.T.A.; McCabe, P.; Pidcock, E.; Platings, M.; Shields, G.P.; Stevens, J.S.; Towler, M.; et al. Mercury 4.0: From visualization to analysis, design and prediction. J. Appl. Cryst. 2020, 53, 226–235. [Google Scholar] [CrossRef] [Green Version]
  29. Bader, R.F.W. Atoms in Molecules: A Quantum Theory; Clarendon Press: Oxford, UK, 1990. [Google Scholar]
  30. Guillot, B.; Viry, L.; Guillot, R.; Lecomte, C.; Jelsch, C. Refinement of proteins at subatomic resolution with MOPRO. J. Appl. Cryst. 2001, 34, 214–223. [Google Scholar] [CrossRef] [Green Version]
  31. Groom, C.R.; Bruno, I.J.; Lightfoot, M.P.; Ward, S.C. The cambridge structural database. Acta Cryst. Part B 2016, 72, 171–179. [Google Scholar] [CrossRef]
  32. Suponitskii, K.Y.; Tsypysheva, I.P.; Kovalskaya, A.V. Molecular and crystal structure of (1R,5S)-8-oxo-1,5,6,8-tetrahydro-2H-1,5-methanopyrido[1,2-a][1,5]diazocine-3(4H)-carboxamide and (1R,5S)-8-OXO-1,5,6,8-tetrahydro-2H-1,5-methanopyrido[1,2-a][1,5]diazocine-3(4H)-thiocarbo-xamide. J. Struct. Chem. 2015, 56, 188–190. [Google Scholar] [CrossRef]
  33. Hansen, N.; Coppens, P. Testing aspherical atom refinements on small-molecule data sets. Acta Cryst. Part A 1978, 71, 3–8. [Google Scholar] [CrossRef]
  34. Domagala, S.; Fournier, B.; Liebschner, D.; Guillot, B.; Jelsch, C. An improved experimental databank of transferable multipolar atom models—ELMAM2. Construction details and applications. Acta Cryst. Part A 2012, 68, 337–351. [Google Scholar] [CrossRef] [Green Version]
  35. Paul, A.; Kubicki, M.; Kubas, A.; Jelsch, C.; Fink, K.; Lecomte, C. Charge density analysis of 2-methyl-4-nitro-1-phenyl-1H-imidazole-5-carbonitrile: An experimental and theoretical study of C≡N ···C≡N interactions. J. Phys. Chem. A 2011, 115, 12941–12952. [Google Scholar] [CrossRef] [PubMed]
  36. Owczarzak, A.; Grześkiewicz, A.M.; Kubicki, M. Experimental studies of charge density distribution in the crystals of cytisine and N-methylcytisine. Inside the fake tobacco. Struct. Chem. 2017, 28, 1359–1367. [Google Scholar] [CrossRef] [Green Version]
  37. McKinnon, J.J.; Spackman, M.A.; Mitchell, A.S. Hirshfeld surfaces: A new tool for visualizing and exploring molecular crystals. Chem. Eur. J. 1998, 4, 2136–2141. [Google Scholar] [CrossRef]
  38. Dunitz, J.D.; Gavezzotti, A. Molecular recognition in organic crystals. Directed intermolecular bonds or nonlocalized bonding? Angew. Chem. Int. Ed. 2005, 44, 1766–1787. [Google Scholar] [CrossRef]
Figure 1. The structures of studied compounds. R = –CH2– (3A, N-[glycine-(N-phtaloyl)]cytisine), –CH2–CH2– (3B, N-[β-alanine-(N-phtaloyl)]cytisine), –CH(CH3)2– (3C, N-[D,L-valine-(N-phtaloyl)]cytisine), –CH(CH3)2 (3D, N-[L-valine-(N-phtaloyl)]cytisine), –CH((CH3)CH2CH3 (3E, N-[L-isoleucine-(N-phtaloyl)]cytisine), –CH2CH(CH3)2 (3F, N-[L-leucine-(N-phtaloyl)]cytisine), –CH2CH(CH3)2 (3G, N-[D-leucine-(N-phtaloyl)]cytisine), –CH2C6H5 (3H, N-[D,L-phenyloalanine-(N-phtaloyl)]cytisine).
Figure 1. The structures of studied compounds. R = –CH2– (3A, N-[glycine-(N-phtaloyl)]cytisine), –CH2–CH2– (3B, N-[β-alanine-(N-phtaloyl)]cytisine), –CH(CH3)2– (3C, N-[D,L-valine-(N-phtaloyl)]cytisine), –CH(CH3)2 (3D, N-[L-valine-(N-phtaloyl)]cytisine), –CH((CH3)CH2CH3 (3E, N-[L-isoleucine-(N-phtaloyl)]cytisine), –CH2CH(CH3)2 (3F, N-[L-leucine-(N-phtaloyl)]cytisine), –CH2CH(CH3)2 (3G, N-[D-leucine-(N-phtaloyl)]cytisine), –CH2C6H5 (3H, N-[D,L-phenyloalanine-(N-phtaloyl)]cytisine).
Crystals 11 00146 g001
Scheme 1. Synthesis of new derivatives of cytisine (N-[amino acid-(N-phtaloyl)]cytisines, 3A–J).
Scheme 1. Synthesis of new derivatives of cytisine (N-[amino acid-(N-phtaloyl)]cytisines, 3A–J).
Crystals 11 00146 sch001
Scheme 2. Principal EI mass spectral fragmentation routes of compounds 3A–3H.
Scheme 2. Principal EI mass spectral fragmentation routes of compounds 3A–3H.
Crystals 11 00146 sch002
Figure 2. A comparison of conformations of 3A (red) and 3E (blue). The cytisine fragments were fitted one onto another.
Figure 2. A comparison of conformations of 3A (red) and 3E (blue). The cytisine fragments were fitted one onto another.
Crystals 11 00146 g002
Figure 3. A perspective view of the molecule 3A. Ellipsoids are drawn at the 50% probability level, hydrogen atoms are represented by spheres of arbitrary radii.
Figure 3. A perspective view of the molecule 3A. Ellipsoids are drawn at the 50% probability level, hydrogen atoms are represented by spheres of arbitrary radii.
Crystals 11 00146 g003
Figure 4. A perspective view of the molecule 3E. Ellipsoids are drawn at the 50% probability level, hydrogen atoms are represented by spheres of arbitrary radii.
Figure 4. A perspective view of the molecule 3E. Ellipsoids are drawn at the 50% probability level, hydrogen atoms are represented by spheres of arbitrary radii.
Crystals 11 00146 g004
Figure 5. A comparison of the values of electron density at bond critical point (top) and Laplacian (bottom) at this point. The codes are: CYT1, CYT2–two symmetry independent molecules of (-)cytisine [19], m-CYT—N-methylcytisine [19], 3A-T, 3E-T: molecules 3A and 3E with transferred multipolar coefficients; 3A–H, 3E–H: 3A and 3E at IAM level with H atoms moved to neutron positions.
Figure 5. A comparison of the values of electron density at bond critical point (top) and Laplacian (bottom) at this point. The codes are: CYT1, CYT2–two symmetry independent molecules of (-)cytisine [19], m-CYT—N-methylcytisine [19], 3A-T, 3E-T: molecules 3A and 3E with transferred multipolar coefficients; 3A–H, 3E–H: 3A and 3E at IAM level with H atoms moved to neutron positions.
Crystals 11 00146 g005
Figure 6. Hirshfeld surfaces of molecules 3A (left) and 3E (right) with the closest neighbors.
Figure 6. Hirshfeld surfaces of molecules 3A (left) and 3E (right) with the closest neighbors.
Crystals 11 00146 g006
Figure 7. An example of a pair of molecules (related by unit translation along the x-direction) and connected by weak, numerous interactions, giving rise to relatively high interaction energy: Eint: −51.8 kJ/mol (B3LYP); −55.3 kJ/mol (PIXEL).
Figure 7. An example of a pair of molecules (related by unit translation along the x-direction) and connected by weak, numerous interactions, giving rise to relatively high interaction energy: Eint: −51.8 kJ/mol (B3LYP); −55.3 kJ/mol (PIXEL).
Crystals 11 00146 g007
Figure 8. An example of a pair of molecules (related by unit translation along x-direction) and connected by weak, numerous interactions, giving rise to relatively high interaction energy: Eint: −44.2 kJ/mol (B3LYP); −43.3 kJ/mol (PIXEL).
Figure 8. An example of a pair of molecules (related by unit translation along x-direction) and connected by weak, numerous interactions, giving rise to relatively high interaction energy: Eint: −44.2 kJ/mol (B3LYP); −43.3 kJ/mol (PIXEL).
Crystals 11 00146 g008
Table 1. Elemental compositions and relative abundances of the ions in the EI spectra of compounds 3A3H.
Table 1. Elemental compositions and relative abundances of the ions in the EI spectra of compounds 3A3H.
IonElemental Compositionm/z3A3B3C
(D)
3D
(L)
3E
(L)
3G
(D)
3F
(L)
3H
(D)
3H
(L)
M+• aC21H19N3O4
C22H21N3O4
C24H25N3O4
C25H27N3O4
C28H25N3O4
 
377 391
419
433
467
 
29
-
-
-
-
21
-
-
-
30
-
-
-
-
7
-
-
-
-
1
-
-
-
16
-
-
-
3
 
-
-
-
-
27
 
-
-
-
-
3
dC21H19N3O4377--412703446--
fC12H13N2O2217141861210613
gC9H6NO2
C13H14NO2
C10H8NO2
C12H12NO2
160
216
174
202
100
-
-
-
-
-
5
-
-
-
-
100
-
-
-
100
-
19
-
-
-
24
-
-
-
41
-
-
 
30
-
13
2
 
13
-
1
-
iC11H13N2O18935301827372321413
jC9H6NO21601007363671001001003013
lC8H5NO214721005664757236374
mC8H4NO2146878164585554363538
Table 2. Chemical shifts δ (ppm) of carbon atoms of cytisine (13C NMR in CDCl3).
Table 2. Chemical shifts δ (ppm) of carbon atoms of cytisine (13C NMR in CDCl3).
13A3B3D cis/trans3E cis/trans3F3H
At Cδ (ppm)
2162.26165.1/164.6163.6/163.5163.2/162.4163.3/162.4162.9/162.9162.9/162.9
3115.04118.3/117.7117.4/116.9117.3/116.9117.3/116.8117.9/117.3118.1/117.2
4138.64138.8/138.4139.5/139.1139.0/136.8139.0/136.9138.6/137.4137.0/136.8
5103.80105.6/104.7106.6/105.9105.8/104.0105.8/104.1105.1/104.3105.9/105.1
6152.34147.7/147.7148.4/148.2134.2/131.4148.1/148.1147.9/147.9134.1/133.9
734.6834.8/34.134.7/34.027.6/26.833.3/32.930.0/29.635.2/34.3
825.7826.1/25.925.71/25.620.8/20.026.1/25.924.9/24.826.1/25.9
927.1327.3/27.327.2/27.026.7/26.227.6/26.727.2/27.027.4/26.8
1049.3948.7/48.749.9/48.549.4/48.951.5/51.149.2/48.751.7/51.3
1152.5349.4/48.348.4/47.551.3/51.048.9/48.550.6/50.248.748.7
1353.4552.3/50.952.7/51.4355.8/55.354.7/54.351.6/51.651.7/51.3
Table 3. Relevant geometric parameters (Å, °) with s.u.’s in parentheses.
Table 3. Relevant geometric parameters (Å, °) with s.u.’s in parentheses.
3A3E 3A3E
N1–C21.400(3)1.418(4)N1–C61.380(3)1.369(4)
N1–C101.473(3)1.478(5)C2–O21.241(3)1.241(5)
C11–N121.464(3)1.463(4)N12–C131.462(3)1.461(4)
N12–C141.352(3)1.364(4)C14–O141.222(3)1.218(4)
C15–N161.437(3)1.470(4)N16–C171.395(3)1.397(4)
N16–C241.389(3)1.406(4)C17–O171.209(3)1.199(4)
C24–O241.203(3)
C2–N1–C6122.4(2)122.2(3)C2–N1–C10113.9(2)114.1(3)
C6–N1–C10123.7(2)123.6(3)C11–N12–C13115.9(2)114.6(3)
C11–N12–C14123.6(2)116.8(3)C13–N12–C14117.6(2)124.8(3)
C15–N16–C17124.0(2)122.3(2)C15–N16–C24124.4(2)124.6(2)
C17–N16–C24111.5(2)111.8(3)
C7–C6–N1–C103.2(3)−0.4(5)C6–C7–C8–C9−57.7(3)−61.8(4)
C7–C8–C9–C1068.3(3)63.6(4)C8–C9–C10–N1−43.3(3)−33.5(5)
C9–C10–N1–C67.5(3)1.4(5)C7–C8–C9–C11−59.0(3)−61.0(4)
C8–C9–C11–N1249.6(3)57.2(4)C9–C11–N12–C13−45.0(3)−53.0(5)
C11–N12–C13–C749.6(3)52.2(4)N12–C13–C7–C8−60.2(3)−56.3(4)
C13–C7–C8–C964.8(3)60.8(4)C9–C11–N12–C14154.7(2)147.9(4)
C7–C13–N12–C14−148.9(2)−150.6(3)C11–N12–C14–C15−13.1(3)177.9(3)
C13–N12–C14–C15−173.1(2)21.1(4)N12–C14–C15–N16−176.7(2)60.4(3)
C11–N12–C14–O14167.2(2)−2.6(5)C13–N12–C14–O147.2(4)−159.4(3)
C14–C15–N16–C17−77.6(3)99.2(3)C14–C15–N16–C2497.4(3)57.9(4)
Table 4. The details of the interactions for the pair of the highest interaction energy in 1 (i: −1+x, y, z: 8 contacts with critical points). Upper part: geometrical details; lower: topological; D12: distance between two atoms; Gcp: kinetic energy density (kJ/mol/Bohr3); Vcp: potential energy density (kJ/mol/Bohr3); Dcp1, Dcp2: distance from the first and the second atom to the critical point, respectively; Lap: laplacian [∇2bcp) (eÅ−5)]; Den: electron density [ρbcp (eÅ−3)].
Table 4. The details of the interactions for the pair of the highest interaction energy in 1 (i: −1+x, y, z: 8 contacts with critical points). Upper part: geometrical details; lower: topological; D12: distance between two atoms; Gcp: kinetic energy density (kJ/mol/Bohr3); Vcp: potential energy density (kJ/mol/Bohr3); Dcp1, Dcp2: distance from the first and the second atom to the critical point, respectively; Lap: laplacian [∇2bcp) (eÅ−5)]; Den: electron density [ρbcp (eÅ−3)].
DHAD–HH···AD···AD–H···A
C9H9O14i1.092.393.181129
C15H15BC22i1.092.733.896117
C11H11BO14i1.092.613.310122
C8H8BC5i1.093.313.854144
C10H10AC4i1.092.993.963150
C11H11AC23i1.093.143.896127
contacts
O17C20i 3.826
H8AH13Bi 2.90
Atom 1Atom 2GcpVcpD12Dcp1Dcp2DenLap
H9O14i23.65−17.622.37980.98481.39510.0721.09
H15BC22i15.46−11.132.72601.09671.66050.0520.727
H11BO14i14.63−9.442.60561.12241.50030.0390.727
H8BC5i7.76−5.042.98471.19591.80140.0270.385
H10AC4i5.61−3.912.98231.22981.76070.0260.269
H11AC23i6.2−3.913.13461.25371.92890.0220.312
O17C20i5.48−3.463.52351.72481.82220.0200.275
H8AH13Bi2.88−1.722.89971.43581.46470.0120.149
Table 5. The details of the interactions for the pair of the highest interaction energy in 2 (i 1+x, y, z: 10 contacts with critical points). Upper part: geometrical details; lower: topological; D12: distance between two atoms; Gcp: kinetic energy density (kJ/mol/Bohr3); Vcp: potential energy density (kJ/mol/Bohr3); Dcp1, Dcp2: distance from the first and the second atom to the critical point, respectively; Lap: laplacian [∇2bcp) (eÅ−5)]; Den: electron density [ρbcp (eÅ−3)].
Table 5. The details of the interactions for the pair of the highest interaction energy in 2 (i 1+x, y, z: 10 contacts with critical points). Upper part: geometrical details; lower: topological; D12: distance between two atoms; Gcp: kinetic energy density (kJ/mol/Bohr3); Vcp: potential energy density (kJ/mol/Bohr3); Dcp1, Dcp2: distance from the first and the second atom to the critical point, respectively; Lap: laplacian [∇2bcp) (eÅ−5)]; Den: electron density [ρbcp (eÅ−3)].
DHAD–HH···AD···AD–H···A
C5H5O14i1.092.233.257157
C7H7O14i1.092.653.798136
C25iH25iO171.092.723.537132
C4H4O24i1.092.813.551125
C5H5O24i1.092.863.566123
C28iH28BiO171.093.043.913138
contacts
C19O24i 3.526
H19H25 2.558
H19H28B 2.605
H7H11Ai 2.459
Atom 1Atom 2GcpVcpD12Dcp1Dcp2DenLap
H5O1427.12−22.42.23340.89431.33990.0931.169
H7O1411.61−7.792.64291.12011.52510.0380.566
O17H2510.21−6.682.71141.55971.1530.0330.505
H7H11A5.72−4.092.44751.23471.21510.0280.27
H4O247.83−4.762.81781.20651.61150.0230.4
H5O247.32−4.432.86051.22631.6360.0210.375
C19O245.83−3.563.52621.86991.66690.0190.297
H19H255.44−3.362.55441.26811.29190.0190.276
H19H28B4.47−2.712.61731.30341.31780.0160.229
O17H28B4.34−2.653.04641.72721.33380.0160.222
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Przybył, A.K.; Grzeskiewicz, A.M.; Kubicki, M. Weak Interactions in the Structures of Newly Synthesized (–)-Cytisine Amino Acid Derivatives. Crystals 2021, 11, 146. https://doi.org/10.3390/cryst11020146

AMA Style

Przybył AK, Grzeskiewicz AM, Kubicki M. Weak Interactions in the Structures of Newly Synthesized (–)-Cytisine Amino Acid Derivatives. Crystals. 2021; 11(2):146. https://doi.org/10.3390/cryst11020146

Chicago/Turabian Style

Przybył, Anna K., Anita M. Grzeskiewicz, and Maciej Kubicki. 2021. "Weak Interactions in the Structures of Newly Synthesized (–)-Cytisine Amino Acid Derivatives" Crystals 11, no. 2: 146. https://doi.org/10.3390/cryst11020146

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop