Next Article in Journal
Chiral, Topological, and Knotted Colloids in Liquid Crystals
Previous Article in Journal
Advancements in and Applications of Crystal Plasticity Modelling of Metallic Materials
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

X-ray Structure of Eleven New N,N′-Substituted Guanidines: Effect of Substituents on Tautomer Structure in the Solid State

by
Vijayaragavan Elumalai
1,†,
Vaclav Eigner
2,
Nicholas Alexander Janjua
1,
Per-Olof Åstrand
1,
Torkild Visnes
3,
Eirik Sundby
4 and
Bård Helge Hoff
1,*
1
Department of Chemistry, Norwegian University of Science and Technology (NTNU), N-7491 Trondheim, Norway
2
Department of Structure Analysis, FZU—Institute of Physics of the Czech Academy of Sciences, Na Slovance 1999/2, 182 00 Prague 8, Czech Republic
3
Department of Biotechnology and Nanomedicine, SINTEF Industry, N-7034 Trondheim, Norway
4
Department of Materials Science and Engineering, Norwegian University of Science and Technology (NTNU), N-7491 Trondheim, Norway
*
Author to whom correspondence should be addressed.
Current address: Chiron AS, Arkitet Ebbells veg 26, N-7041 Trondheim, Norway.
Crystals 2024, 14(10), 884; https://doi.org/10.3390/cryst14100884
Submission received: 5 September 2024 / Revised: 7 October 2024 / Accepted: 9 October 2024 / Published: 11 October 2024
(This article belongs to the Section Organic Crystalline Materials)

Abstract

:
Guanidine-containing molecules are an interesting class of compounds within both medicinal and material sciences. Having knowledge of their tautomerism is key in designing guanidines that interact with biological and chemical receptors. However, there are limited data on the solid-phase structure of N,N′-substituted guanidines. Thus, eleven guanidines bearing a 4,6-dimethylpyrimidyl at one end and substituents of varying sizes and electronic properties at the other side, were synthesised, crystallised, and analysed by X-ray crystallography. Calculations of isolated molecules of tautomer energies and bond lengths were performed for comparison. One class of guanidines crystallised as a cis–trans tautomer with the shorter bond directed towards the pyrimidyl unit. When more electron-deficient aniline substituents were inserted, the crystallised tautomer changed to a cis–cis form where the shorter bond was directed towards the aniline. The switch in the tautomer structure is concluded to be due to both the electronic properties of the substituents and the intermolecular hydrogen bonding in the crystal lattice.

1. Introduction

Guanidines are of interest as sweeteners [1], organo-catalysts [2], catalysts for asymmetric synthesis [3], bases [4], metal ligands [5], sensor devices [6], drug candidates [7], and polyelectrolyte materials [8]. In all of these applications, intra- and intermolecular hydrogen bonding involving guanidines are crucial for the performance of the molecule or material. In this respect, having knowledge of the preferred tautomer structure is of importance. Generally, the most stable tautomer can be deduced from the relative acidity of the protonated version of the molecule [9]. While guanidine itself is highly basic (pKa = 13.6), N-substitution with aromatic groups lowers the pKa [10,11]. For instance 1,3-diphenyl guanidine has a pKa of 9.1 in dimethylformamide [12]. Very few X-ray structures have been reported for N,N′-substituted guanidines, but some are shown in Figure 1. Paixao et al. reported an X-ray structure of 1,3-diphenylguanidine (I) [13], Zheng et al. analysed structures II and III [14], while Carpy et al. [15] provided structure IV. These four structures were found to have one short bond (1.271–1.308 Å) corresponding to that typical for C=N bonds [14]. The tautomer structures of IIV follow that expected from the pKa. In contrast, Brown et al. [16], for structure V, reasoned that all three hydrogens were located at the central nitrogen atom.
Hydrogen bonding can also increase the stability of a specific tautomer form [17,18,19]. Some of these intramolecular hydrogen-bonded structures can be regarded as a semi-aromatic six-membered ring with a bond length deviating from that of a classical-type tautomer [20]. This type of bonding has been termed resonance-assisted hydrogen bonding. Additionally, in the solid state, intermolecular interactions are also factors strongly affecting the structure. We have previously published the X-ray structure of 1-(3-chlorophenyl)-2-(4,6-dimethylpyrimidin-2-yl)guanidine [21]. Herein, we have included data for eleven additional guanidine X-ray crystals. These contain the same 4,6-dimethylpyrimidyl unit at one side, which are decorated with amines with variance in basicity, and steric bulk at the other side. We show that the guanidine tautomer structure is dictated by the electronic properties of the guanidine substituents and by intermolecular hydrogen bonding. Hopefully, information on guanidine tautomerism will be useful in the design of biologically active molecules.

2. Materials and Methods

2.1. Chemicals, Equipment, and Analysis

N-(4,6-Dimethylpyrimidin-2-yl)cyanamide was prepared as previously described [21]. All other reagents and solvents were purchased from Merck and used directly without any further purification. NMR spectra were recorded at an ambient temperature at a frequency of 400 and 600 MHz for 1H and 101 and 151 MHz for 13C, respectively. The chemical shift values are reported in ppm from tetramethylsilane as the internal standard, calibrated with a residual signal of DMSO-d5 for proton (δ = 2.50 ppm), with a signal from DMSO-d6 for carbon (δ = 39.52 ppm), or with a signal from AcOH-d3 for proton (δ = 2.04 and 11.65 ppm), with a signal from AcOH-d4 for carbon (δ = 20 and 178.99 ppm).
19F NMR was recorded with hexafluorobenzene as the internal standard. The splitting patterns are reported as singlet (s), broad singlet (br s), doublet (d), triplet (t), quartet (q), doublet of doublet (dd), doublet of triplet (dt), triplet of doublet (td), doublet of doublet of doublets (ddd), and multiplet (m). The coupling constants (J) are reported in Hertz (Hz). High-resolution mass spectra (HRMS) were recorded on a water ‘Synapt G2-S’ Q-TOF instrument with electrospray (ES) positive mode ionisation. The IR spectra were obtained using a Bruker Alpha Eco-ATR FTIR spectrometer (Bruker, Billerica, MA, USA). The microwave reactions were performed using a Biotage initiator+ Microwave system (Biotage, Uppsala, Sweeden).

2.2. General Procedure A: Classical Heating

N-(4,6-Dimethylpyrimidin-2-yl)cyanamide (1 equiv.) and the amine derivative (1 equiv.) in dioxane (20 mL) were refluxed for 24 h and then cooled to 22 °C. The reaction mixture was then added to a basic solution (2% NH3 in 50 mL distilled water). The precipitate that formed (pH: 11) was isolated by filtration, washed with water, and dried under vacuum to obtain the target molecules. The attained solid was further purified by recrystallisation from 2-PrOH, and the obtained crystals were used for X-ray crystallographic analysis.

2.3. General Procedure B: Microwave

In a microwave vial (20 mL), N-(4,6-dimethylpyrimidin-2-yl)cyanamide (1 equiv.) and an amine derivative (1 equiv.) in dioxane (15 mL) were mixed and heated under microwave irradiation at 150 °C for 1 h. The reaction mixture was cooled to 22 °C and added to a basic solution (2% NH3 in 50 mL water). The precipitate that formed (pH: 11) was isolated by filtration, washed with water, and dried under vacuum to obtain the target compounds. For some compounds, the attained solid was further purified by recrystallisation from 2-PrOH, and the obtained crystals were used for X-ray crystallographic analysis.

2.4. Synthesised Guanidines

2.4.1. 1-Benzyl-2-(4,6-dimethylpyrimidin-2-yl)guanidine (5)

N-(4,6-Dimethylpyrimidin-2-yl)cyanamide (578 mg, 3.90 mmol) and benzylamine (0.42 mL, 3.89 mmol) were reacted following General Procedure A. The resulting precipitate was isolated as an off-white solid (0.565 g, 57%). This material was further purified by recrystallisation using 2-PrOH (13 mL) giving 301 mg (1.18 mmol, 30%) of an off-white crystalline solid, as follows: mp. 207–208 °C; 1H NMR (400 MHz, DMSO-d6) δ 7.68 (s, 2H, br, NH2), 7.34 (m, 4H), 7.30–7.19 (m, 1H), 6.48 (s, 1H), 4.50 (s, 2H), 2.22 (s, 6H); 13C NMR (101 MHz, DMSO-d6) δ 166.5, 166.2 (2C), 158.2, 140.7, 128.7 (2C), 127.5 (2C), 127.1, 110.6, 44.0, 24.0 (2C); HRMS (TOF, ES+): found 256.1565, calcd for C14H18N5 (M + H)+, 256.1562; IR (neat, cm-1): 3272, 3109, 3024, 2980, 2918, 1566, 1520, 1411, 1373, 1328, 1173, 805, 732.

2.4.2. 1-Benzyl-2-(4,6-dimethylpyrimidin-2-yl)guanidine hydrochloride (5-HCl)

Compound 5 was mixed with an EtOH solution of HCl and stirred for 1 h. Upon evaporation and drying, a solid formed, as follows: 1H NMR (400 MHz, DMSO-d6) δ 11.13 (br s, 1H), 10.16 (br s, 1H), 9.16 (br s, 2H), 7.46–7.34 (m, 4H), 7.32 (m, 1H), 7.05 (s, 1H), 4.69 (d, J = 6.0 Hz, 2H), 2.38 (s, 6H).

2.4.3. 2-(4,6-Dimethylpyrimidin-2-yl)-1-(2-methylbenzyl)guanidine (6)

N-(4,6-Dimethylpyrimidin-2-yl)cyanamide (0.578 g, 3.90 mmol) and 2-methylbenzylamine (0.47 g, 0.48 mL, 3.89 mmol) were reacted following General Procedure A. The resulting precipitate was isolated as an off-white solid (0.60 g, 57%). The material was further purified by recrystallisation using refluxing 2-PrOH (200 mL) to give 210 mg (0.78 mmol, 20%) of an off-white crystalline solid, as follows: mp. 184–185 °C; 1H NMR (400 MHz, DMSO-d6) δ 7.58 (br s, 2H), 7.31–7.22 (m, 1H), 7.22–7.11 (m, 2H), 6.47 (s, 1H), 4.43 (s, 2H), 2.31 (s, 2H), 2.21 (s, 5H). The proton spectrum matched well that reported in [22], as follows: 1H NMR (400 MHz, CD3CO2D) δ 7.47–7.33 (m, 1H), 7.36–7.15 (m, 3H), 6.98 (s, 1H), 4.69 (s, 2H), 2.42 (s, 6H), 2.41 (s, 3H); 13C NMR (101 MHz, CD3CO2D) δ 170.6 (2C), 158.7, 155.8, 138.0, 134.9, 132.4, 130.0, 129.0, 128.1, 117.8, 45.2, 24.1 (2C), 19.8; HRMS (TOF, ES+): found 270.1723, calcd for C15H20N5 (M + H)+, 270.1719; IR (neat, cm−1): 3266, 3097, 2956, 1519, 1407, 1373, 1331, 1150, 805, 735, 557, 472.

2.4.4. (S)-1-(4,6-Dimethylpyrimidin-2-yl)-3-(1-phenylethyl)guanidine (7)

N-(4,6-Dimethylpyrimidin-2-yl)cyanamide (576 mg, 3.89 mmol) and (S)-1-phenylethan-1-amine (565 mg, 4.67 mmol) were reacted following General Procedure B. After cooling to rt (22 °C), the mixture was quenched in an aqueous solution of ammonia (50 mL, 2%). The resulting solid was isolated by filtration and washed with water. This gave, after drying, 344 mg (1.28 mmol, 33%) of a light beige solid, as follows: mp. 154.6–157.5 °C, [α ] 20 D = 56.0 (c 1.00, MeOH); 1H NMR (400 MHz, DMSO-d6) δ 7.54 (br s, 2H), 7.41–7.30 (m, 4H), 7.28–7.19 (m, 1H), 6.48 (s, 1H), 5.14 (s, 1H), 2.22 (s, 6H), 1.41 (d, J = 6.9 Hz, 3H); 1H NMR (600 MHz, CD3CO2D) δ 7.51–7.47 (m, 2H), 7.44 (t, J = 7.7 Hz, 2H), 7.39–7.34 (m, 1H), 6.99 (s, 1H), 5.06 (d, J = 9.0 Hz, 1H), 2.47 (s, 6H), 1.70 (d, J = 6.7 Hz, 3H); 13C NMR (151 MHz, CD3CO2D) δ 168.8 (2C), 157.0, 153.3, 141.1, 129.0 (2C), 128.1, 125.7 (2C), 116.0, 51.9, 22.5 (2C), 22.4; HR-MS (TOF MS ES+): found 270.1716, calcd for C15H20N5, (M + H)+, 270.1719; IR (neat, cm−1): 3269, 3112, 2960, 2923, 2852, 1611, 1572, 1519, 1410, 1376, 1335, 1254, 1211, 1120, 806, 700, 559.

2.4.5. (1-(4,6-Dimethylpyrimidin-2-yl)-3-(2,2,2-trifluoro-1-phenylethyl)guanidine (8)

To a solution of N-(4,6-dimethylpyrimidin-2-yl)cyanamide (230 mg, 1.55 mmol) in dioxane (10 mL), 2,2,2-trifluoro-1-phenylethan-1-amine (326 mg, 1.86 mmol) and HCl (0.13 mL, 1.2 equiv.) were added. The mixture was treated by microwave irradiation for 1 h at 150 °C (6 bar), before quenching it with NaOH (2M, 0.5 mL) until reaching pH = 8. Upon the addition of water (5 mL), a precipitate was formed, which was isolated by filtration and washed with MeCN (5 mL) to afford 224 mg (0.69 mmol, 45%) of white powder. A sample (129 mg) was recrystallised from 2-PrOH (10 mL) to yield 47 mg (0.15 mmol, 36%) of translucent starshaped crystals, as follows: mp. 186.7–188.8 °C; 1H NMR (400 MHz, DMSO-d6): 8.10-7.30 (m, 8H), 6.57 (s, 1H), 6.15 (br s, 1H), 2.26 (s, 6H); 1H NMR (400 MHz, CD3CO2D) δ 7.67–7.60 (m, 2H), 7.53 (dd, J = 5.1, 2.0 Hz, 3H), 7.04 (d, J = 2.1 Hz, 1H), 5.94 (q, J = 6.9 Hz, 1H), 2.52 (s, 6H). 13C NMR (101 MHz, CD3CO2D) δ 168.8 (2C), 156.9, 154.7, 130.5, 130.1, 129.2 (2C), 128.4–112.0 (q, J = 282.8 Hz), 128.0 (2C), 116.3, 57.3–56.3 (q, J = 32.3 Hz), 22.5 (2C); 19F NMR (565 MHz, DMSO-d6) δ −72.9; HRMS (TOF MS ES+): found 324.1440, calcd for C15H17N5F3, (M + H)+, 324.1436; IR (neat, cm−1): 3269, 3109, 2974, 2927, 1612, 1576, 1514, 1407, 1377, 1337 1252, 1210, 1166, 1118, 700, 637.

2.4.6. (1-(4,6-Dimethylpyrimidin-2-yl)-3-(2,2,2-trifluoro-1-phenylethyl)guanidine hydrochloride (8-HCl)

A small sample of compound 8 was treated with an EtOH solution of HCl and evaporated to dryness. 1H NMR (400 MHz, DMSO-d6): 11.14 (br s, 1H), 10.12 (br s, 1H), 9.95 (br s, 1H), 7.73–7.70 (m, 2H), 7.53–7.45 (m, 3H); 7.09 (s, 1H), 6.67 (br s, 1H), 2.42 (s, 6H).

2.4.7. 1-(Cyclohexylmethyl)-2-(4,6-dimethylpyrimidin-2-yl)guanidine (9)

N-(4,6-Dimethylpyrimidin-2-yl)cyanamide (576 mg, 3.89 mmol) and cyclohexyl methylamine (528 mg, 4.67 mmol) were reacted following General Procedure B. The precipitate was isolated and recrystallised from 2-PrOH (56 mL) to give 676 mg (2.59 mmol, 67%) as a white solid, as follows: mp. 231.6–232.4 °C (lit. [23] 236–238 °C; 1H NMR (400 MHz, DMSO-d6) δ 7.36 (s, 2H), 6.45 (s, 1H), 3.06 (d, J = 6.7 Hz, 2H), 2.22 (s, 6H), 1.74–1.60 (m, 5H), 1.25–1.10 (m, 3H), 0.94 (m, 2H); 1H NMR (400 MHz, CD3CO2D) δ 7.00 (s, 1H), 3.32 (d, J = 6.7 Hz, 2H), 2.48 (s, 6H), 1.93–1.70 (m, 6H), 1.43–1.05 (m, 5H); 13C NMR (101 MHz, CD3CO2D) δ 168.79 (2C), 156.97, 153.98, 115.92, 47.52, 36.87, 30.28 (2C), 25.97, 25.44 (2C), 22.48 (2C); HRMS (TOF MS ES+): found 262.2036, calcd for C14H24N5, (M + H)+, 262.2032; IR (neat, cm−1): 3259, 3098, 2920, 2848, 1519, 1412, 1370, 1331, 1176, 806.

2.4.8. 1-Butyl-2-(4,6-dimethylpyrimidin-2-yl)guanidine (10)

N-(4,6-Dimethylpyrimidin-2-yl)cyanamide (0.578 g, 3.90 mmol) and butylamine (0.38 mL, 3.89 mmol) were reacted following General Procedure A. The resulting precipitate was isolated as an off-white solid (0.657 g, 76%). The material was further purified by recrystallisation from 2-PrOH (45 mL, at 82 °C), giving 495 mg (2.24 mmol, 57%) of tiny off-white crystals, as follows: mp. 188–189 °C; 1H NMR (600 MHz, DMSO-d6) δ 7.47 (s, 2H, br, NH2), 6.44 (s, 1H), 3.19 (t, J = 7.0 Hz, 2H), 2.21 (s, 6H), 1.50–1.42 (m, 2H), 1.38–1.29 (m, 2H), 0.90 (t, J = 7.4 Hz, 3H); 13C NMR (151 MHz, DMSO-d6) δ 166.1, 165.5 (2C), 157.9, 109.7, 40.1, 31.4, 23.6 (2C), 19.6, 13.7; HRMS (TOF, ES+): found 222.1723, calcd for C11H20N5, (M + H)+, 222.1719; IR (neat, cm−1): 3264, 3099, 2957, 2923, 2870, 1525, 1410, 1374, 1330, 1176, 806, 733.

2.4.9. N′-(4,6-Dimethylpyrimidin-2-yl) pyrrolidine-1-carboximidamide (11)

N-(4,6-Dimethylpyrimidin-2-yl)cyanamide (0.578 g, 3.90 mmol) and pyrrolidine (0.44 mL, 3.89 mmol) were reacted following General Procedure B. The resulting homogeneous solution was kept at 22 °C overnight, and the desired product was crystallised slowly. The crystals were isolated by filtration, washed with water, and then dried under vacuum to give 715 mg (3.26 mmol, 84%) of an off-white crystalline solid, as follows: mp. 210–211 °C; 1H NMR (400 MHz, DMSO-d6) δ 8.06 (s, 2H, br, NH2), 6.43 (s, 1H), 3.38 (m, 4H), 2.21 (s, 6H), 1.91–1.79 (m, 4H); 13C NMR (101 MHz, DMSO-d6) δ 166.0, 165.7 (2C), 156.3, 109.8, 46.2 (2C), 25.1 (2C), 23.8 (2C); HRMS (TOF, ES+): found 220.1565, calcd for C11H18N5, (M + H)+, 220.1562; IR (neat, cm−1): 3316, 3153, 2947, 2880, 1616, 1531, 1411, 1334, 1112, 802.

2.4.10. N′-(4,6-Dimethylpyrimidin-2-yl)-1,3,4,5-tetrahydro-2H-pyrido[4,3-b]indole-2-carboximidamide (12)

N-(4,6-Dimethylpyrimidin-2-yl)cyanamide (578 mg, 3.90 mmol) and 1,2,3,4-tetrahydro-9H-pyrido[3,4-b]indole (670 mg, 3.89 mmol) were reacted following General Procedure B. The resulting homogeneous solution was kept at 22 °C, and the product was precipitated slowly. The precipitate was isolated by filtration, washed with water, and dried under vacuum to give a yellow solid (1.03 g, 82%). The obtained solid was crystallised from 2-PrOH (50 mL) to give 880 mg (2.75 mmol, 71%) of a yellow crystalline solid, as follows: mp. 210–211 °C, 1H NMR (400 MHz, DMSO-d6) δ 10.91 (s, 1H, NH-indole), 8.53 (s, 2H, NH2), 7.41 (d, J = 7.7 Hz, 1H), 7.32 (d, J = 8.1 Hz, 1H), 7.04 (ddd, J = 8.2, 7.0, 1.3 Hz, 1H), 6.97 (td, J = 7.5, 1.1 Hz, 1H), 6.51 (s, 1H), 4.87 (s, 2H), 3.88 (t, J = 5.6 Hz, 2H), 2.76 (t, J = 5.7 Hz, 2H), 2.25 (s, 6H); 13C NMR (151 MHz, DMSO-d6) δ 166.2 (2C), 166.1, 157.9, 136.5, 132.7, 127.1, 121.1, 118.9, 117.9, 111.5, 110.8, 107.4, 42.8, 42.8, 24.0 (2C), 21.4; HRMS (ES): found 321.1831, calcd for C18H21N6, (M + H)+, 321.1828; IR (neat, cm−1): 3349, 3254, 3162, 2998, 2957, 2916, 2847, 1618, 1577, 1520, 1490, 1431, 1412, 1335, 1109, 1010, 893, 736.

2.5. X-ray Crystal Determination

The crystalline material was mounted on glass fibre and measured using either Rigaku OD Gemini, utilising an Atlas S2 CCD camera and mirror-collimated Cu-Kα radiation from a sealed X-ray tube (comp. 5, 6, and 11), or Rigaku OD Supernova, utilising an Atlas S2 CCD camera and Cu-Kα radiation from a micro-focused sealed X-ray tube (1, 2, 3, 4, 7, 8, 9, 10, and 12). Both diffractometers and their detectors were supplied by Rigaku Europe SE, Neu-Isenburg, Germany In both cases, the samples were cooled during measurement using an open flow N2 system Cryojet 5 from Oxford Cryosystems to 180 K for 5, 6, and 7 and Cryostream 800 from Oxford Cryosystems (Oxford, United Kingdom) to 95 K for compounds 1, 2, 3, 4, 7, 8, 9, 10, and 12. The data integration, reduction, scaling, and absorption correction were handled in CrysAlis PRO [24]. The phase problem was solved using charge flipping methods in Superflip [25]. The structure models were refined using Crystals [26], except for 10, which was solved in Jana2020 [27]. Samples 9 and 10 were found to be twinned with twin volume ratios [28] of 764(4):236(4) and 5592(12):4408(12), respectively. Structure 10 had a significant number of overlapping reflections and benefited significantly from hklf5 refinement. The aromatic substituent in 4 was found to be disordered over two positions and refined with restrained geometry and a constrained sum of occupancies to 1, resulting in an occupancy ratio of 917(3):83(3). The hydrogen atoms bonded to heteroatoms were refined with restrained geometry, while the hydrogen atoms bonded to carbon atoms were placed in calculated positions and refined with riding constraints. The residual electron density maps were visualised using MCE 4.2.00 [29]. The crystal structures were visualised using Diamond 3.0c [30]. For further information on data collection, data reduction, and refinement, see Supplementary Materials, Tables S1–S3.

3. Results and Discussion

3.1. Synthesis

We selected 4,6-dimethylpyrimidine for one side of the guanidines, while the other side was varied with amine substituents having differences in basicity and steric bulk (see Scheme 1). Details on the synthesis of the aryl-containing guanidines 14 using HCl in 2-PrOH have recently been reported [21]. Due to its low basicity, trifluoro-containing guanidine 8 was also prepared by this method. In contrast, the guanidines containing aliphatic and benzylic amines (57 and 912) were more efficiently prepared from N-(4,6-dimethylpyrimidin-2-yl)cyanamide without acid by thermal or microwave heating in 1,4-dioxane. After the reaction, the mixture was quenched with aqueous ammonia to ensure that the guanidines were in the neutral form. Most of the compounds were recrystallised from 2-PrOH, giving yields in the range of 20–84%.
Guanidines 78 and 1112 are new chemical entities, while NMR spectroscopic data were lacking for benzyl derivative 5 and n-butyl analogue 10. Compounds 6 and 9 were previously well characterised and have been investigated for their biological activity as inhibitors of FOXO3-induced gene transcription [22] and Rho GTPase Rac1 inhibitors [23], respectively.
1H NMR spectroscopy in DMSO-d6 of compounds 512 showed a broad two-proton singlet from the guanidine residing at 7.35–8.35 ppm, while one of the guanidine protons underwent rapid exchange and was not observed. A diagnostic shift confirming that the guanidines are in neutral form is the one proton singlet at C-5 of the pyrimidine moiety, which resonates at 6.43–6.57 ppm, depending on the structure. For the protonated version of the molecule, and when analysed in deuterated acetic acid, these signals shifted 0.5 ppm towards the lower field. 1H NMR of the hydrochloride salt of 5 (5-HCl) in DMSO-d6 showed that proton exchange in the guanidine unit is slower than that observed in the case of neutral 5. Thus, all four N-H protons could be observed by 1H NMR and coupling between the benzylic protons, and the nearby N-H group was detected. Not all of these neutral guanidines have sufficient solubility in DMSO-d6 to be analysed by 13C NMR spectroscopy. However, their protonated versions have good solubility, and deuterated acetic acid was, therefore, applied as a solvent for some 13C NMR spectra. The NMR spectra are shown in Supplementary Materials, Figures S20–S40. The IR spectra of compounds 112 are presented in Tables S13 and S14 and in Figures S41–S52. Compounds 24, which crystallised in a T2 cis–cis tautomer form, have strong NH stretching in the range of 3420–3504 cm−1. In contrast, structures 1 and 512 have broader bands in the range of 3220–3282 cm−1, indicative of the other form of hydrogen bonding. The region 1700–1500 cm−1 is crowded with bands from C=N and aromatic bonds. Compounds 14 appear to have C=N stretching in the range of 1650–1656 cm−1, while the corresponding absorbance for 512 lies in the range of 1610–1621 cm−1.

3.2. X-ray Crystal Structures

Table 1 provides an overview of the X-ray structures, their space group, and the number of molecules in the asymmetric unit. The aniline-substituted structures 1, 2, and 4, alongside benzylic compounds 5 and 6, all have one molecule in the asymmetric unit.
Two molecules in the asymmetric units are seen for the pentafluoro derivative 3, and the guanidines have aliphatic primary amines as substituents (comp. 9 and 10). The presence of two molecules in the asymmetric unit allows for better adaption for steric demands in the structure. The main part of the scaffold remains almost planar; however, the twisting is more pronounced than that observed in the former compounds. In the case of guanidines substituted with chiral (7 and 8) and secondary amines (11 and 12), the added steric bulk leads to three molecules in the asymmetric unit and significant twisting in the main part of the molecule.

3.3. Calculations

We performed gas-phase calculations (B3LYP [31]/cc-pVDZ [32,33]) with the PySCF software [34,35] of the stability of the different tautomers and their bond lengths (see Figure 2 and Table S3). Obviously, gas-phase calculation will not be able to predict the solid-phase structure but is otherwise a useful reference point for bond lengths. Tautomer T1 has a shorter bond (double-bond character) between atoms N7 and C8, tautomer T3 has double bond to the central amine, and tautomer T2 has double-bond character between atoms C8 and N10. The guanidines can also be in a different conformation, with respect to the position of the R-group (cis or trans in relation to the central nitrogen atom).
Overall, gas-phase calculation indicated tautomer T2 with a cis–cis geometry to be most stable for compounds 19, while n-butyl derivative 10 appeared most stable as tautomer T1-cis–trans with a low margin. Derivatives 1112, having cyclic secondary amine substituents, are also indicated to be most stable as tautomer T1. Note that the T2 structure is prohibited by the lack of a proton at this side. Tautomer T3 was found to be 15–17 KJ/mol higher in energy. Tautomer T1-cis–trans was relatively stable for the non-aniline structures 512 (see Supplementary Materials, Table S7), but high in energy for the aniline structures. Additionally, the tautomer stability of five guanidines (comp. 1, 3, 4, 5, and 8) was also computationally evaluated, with other basis sets and density functionals showing similar trends (Supplementary Materials, Table S8). The average of the calculated bond lengths is shown in Figure 2. No significant difference in bond length was seen between the cis–cis and cis–trans forms of the tautomers (Supplementary Materials, Table S9). Pyrimidines are potential hydrogen bond acceptors. The distance between the pyrimidine atom N3 (hydrogen bond acceptor) and N9 (hydrogen bond donor) was indicated to be within distance for medium strength hydrogen bonding (Supplementary Materials, Table S9). The average distance between N3 and N9 was 2.67 Å, 2.76 Å, and 2.91 Å for T1, T2, and T3 structures, respectively. This could indicate stronger hydrogen bonds for the T1 tautomers, as compared to the others. The Moltaut program [36] predicted tautomer T1 to be most stable in 10 of the 11 cases tested (Supplementary Materials, Table S10). The program was not able to handle the larger molecule 12. Furthermore, to indicate the effect of electronic properties of the different amine substituents, the proton affinities of the guanidines were calculated using B3LYP/cc-pVDZ (Table S11). The correlation with the experimental pKa values was reasonably good (R2 = 0.952, Figure S19).

3.4. Description of the Solid-Phase Guanidine Structures

The solid-phase structures of guanidines 112 depend on the amine substituents and can broadly be classified in two groups. The bond length data are shown in Table 2, while representative X-ray structures are displayed in Figure 3 and in Supplementary Materials, Figures S1–S19.
For meta-chloroaniline-substituted 1 and benzylic/aliphatic-substituted 510, the bond lengths from the central carbon (C8) to the three connected nitrogen atoms are very similar, with average values of N7-C8: 1.333 ± 0.003 Å, C8-N9: 1.345 ± 0.005 Å, and C8-N10: 1.347 ± 0.008 Å. The N7-C8 bond is longer than that seen in the analogues IIV [13,14,15] (1.294 ± 0.0148 Å, structures in Figure 1) and that calculated in the gas phase. As a note, the N10-C11 bond in compound 1 is shorter than that observed for the aliphatic and benzylic amines, which is typical for anilines [37]. The measured bond lengths of guanidines 1 and 510 are consistent with a T1 tautomer structure, and they all are substituted with amines with a pKa > 3.8. The solid-phase structure of pyrrolidine-substituted 11 has N7-C8, C8-N9, and C8-N10 bonds, which are even more equal (1.341–1.344 Å). The lengths for the corresponding bonds in compound 12, however, mimic those seen for compound 1. A larger variance in bond length between the central atoms was seen for compounds 24. The average values were N7-C8: 1.384 ± 0.003 Å, C8-N9: 1.338 ± 0.003 Å, and C8-N10 1.303 ± 0.007 Å. Thus, the N7-C8 bond and the C8-N10 bonds are more in line with that estimated by DFT calculations in the gas phase (Supplementary Materials, Table S9).
These bonds are approaching single- and double-bond character, respectively, and the visualisation of tautomer T2 with a cis–cis geometry, as shown in Figure 2, is descriptive. The anilines used in the preparation of these three molecules have pKa < 3.2. Thus, the electronic properties of the substituents appear to play a major role in determining the tautomer type. The bond and torsion angels for the central part of the molecules are visualised in Figure 4 and in Supplementary Materials, Tables S5 and S6. For compounds 1 and 512, the following average bond angles were observed: N7-C8-N9: 126.1 ± 0.6°, N7-C8-N10: 118.2 ± 0.8°, and N9-C8-N10: 115.7 ± 1.3°. The bond angles for guanidines 24 are very similar, but they are arranged in an opposite way, being N7-C8-N9: 118.0 ± 0.2°, N7-C8-N10: 115.6 ± 0.4°, and N9-C8-N10: 126.4 ± 0.3°.
Guanidines 1 and 512 have N7-C8-N10-C11 torsion angle between 28.1° and −16.9°, with an average absolute value of 8.6°. The N9-C8-N10-C11 torsion angle has an average absolute value of 173.1°. Thus, in these compounds, the substituent on N10 is trans in relation to N9. An outlier in this series is m-chloroaniline derivative 1, which, due to its bulk and shorter C-N bond, requires some reorientation of the structure. For guanidines 24, the situation is reversed, with the N7-C8-N10-C11 torsion angle being close to 180°, with an average absolute value of 176.6° and an average N9-C8-N10-C11 torsion angle of 3.9°. Thus, these structures are crystallised with a cis–cis structure. The deviation from planarity within the central part of the molecules defined by atoms N3-N7-C8-N9-N10 is only minor. For the cis–cis tautomer structures, atomN7 shifts out of this plane by an average value of 0.105 Å, while the value for cis–trans is 0.051 Å.

3.5. Hydrogen Bonding and the Effect on Solid-Phase Structure

One possible explanation for the bond length averaging especially seen for compounds 1 and 512 could be intramolecular hydrogen bonding leading to semi-aromatic substructures. Such a phenomenon has been explained by resonance-assisted hydrogen bonding [17,18,19]. In all of the X-ray structures, intramolecular hydrogen bonding involving the central N9 atom as a donor and pyrimidine N3 as an acceptor is indicated. The distance between N3 and N9 is in the range expected for hydrogen bonds (average value for all compounds, 2.68 Å), but the interatomic distances (2.66–2.71 Å) are not shorter for these molecules in comparison with guanidines 24. Thus, resonance-assisted hydrogen bonding is not likely to explain the X-ray tautomer forms for 1 and 512. However, the intermolecular hydrogen bonding in the crystals differ. The cis–trans geometry of guanidines 1 and 510 dictate that two hydrogens (on N9 and N10) are heading in the same direction, providing two interaction sites. Atoms N1 and N7 in a neighbouring molecule are in perfect distance to be acceptors of these hydrogen bonds. This leads to the formation of infinite chains in the structure connected by hydrogen bonds (see Figure 5).
For guanidines 24, the two hydrogens are on opposite sides, thus not allowing for the same intermolecular hydrogen bonding. However, another preorganised block emerges, allowing for the formation of dimers via N7-H···N1 and N7-H ···N10 hydrogen bonds. This is probably the distinct IR absorptions seen for 24 in the range of 3400–3500 cm−1 (Supplementary Materials, Table S13). Fluorine can also engage in weak hydrogen bonds, with bond distances and angles deviating from those of more classical hydrogen bonds [38]. Pentafluoro derivative 3 has a N9-F distance of 2.98 Å at an angle of 101.5°, while trifluoromethyl analogue 8 has a N10-F distance of 2.80 Å at 95°. Both are within length–angle boundaries expected for a weak N-H···F interaction [38]; however, they are not expected to significantly affect the structure in the solid state. For 2,5-dimethoxy derivative 4, the N9-O distance is 3.33 Å at an angle of 95°. Thus, hydrogen bonding for this pair is not likely. Compounds 11 and 12 have disubstituted N10, which does not allow for the pair of intermolecular hydrogen bonds seen in other structures. For 11, a water molecule donates a hydrogen bond to N7 (3.0 Å at 166.6°). This same water molecule might also interact weakly with the pyrimidine N1 (3.4 Å, 126.8°). Compound 12 instead forms a supramolecular block consisting of two molecules of 12 and one 1,4-dioxane molecule. The pyrrole unit of the N10 substituent is indicated to be an acceptor of a T-shaped hydrogen bond from the CH unit of the 1,4-dioxane, in addition to other van der Waal contacts. To summarise, the solid-phase structure of the guanidines investigated is in part dictated by the electronic properties of the amine substituent and in part by intermolecular hydrogen bonding. The guanidines substituted with amines with a pKa > 3.8 crystallised in a cis–trans T1 form, while those substituted with anilines with pKa < 3.2 had a cis–cis T2 solid-phase structure.

4. Conclusions

This series of N,N′-substituted guanidines was prepared from N-(4,6-dimethylpyrimidin-2-yl)cyanamide. The conditions used depend on the pKa of the amine nucleophile. The compounds were crystallised and analysed by X-ray crystallography. Two main types of structures were obtained. Guanidines substituted with amines with pKa > 3.8 had a cis–trans geometry, with the shortest bond between atoms N7 and C8. When the amine used to form the guanidines had pKa below 3.2, the solid-phase structure had cis–cis geometry with a short bond on the opposite side of the guanidine. The bond lengths differ from that expected from single and double bonds, due to the conjugate nature of these guanidines alongside the effects from inter- and intramolecular hydrogen bonding.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/cryst14100884/s1. The file contains X-ray data and images; calculations of tautomer energies, angles, and bond lengths; reference to pKa of amines [39,40,41,42,43,44]; and NMR, IR, and high-resolution mass spectra of the guanidines. For the DFT calculations, files in xyz-formate can be found at DataverseNO: https://doi.org/10.18710/QW76UM.

Author Contributions

Conceptualisation, B.H.H. and V.E. (Vaclav Eigner); methodology, V.E. (Vijayaragavan Elumalai); formal analysis, V.E. (Vaclav Eigner); investigation, V.E. (Vijayaragavan Elumalai), N.A.J., P.-O.Å. and V.E. (Vaclav Eigner); writing—original draft preparation, B.H.H.; writing—review and editing, B.H.H.; visualisation, V.E. (Vaclav Eigner) and B.H.H.; supervision, B.H.H. and E.S.; funding acquisition, T.V. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Research Council of Norway, grant number NFR 303369.

Data Availability Statement

Structures 212 have been deposited into the Cambridge Structural Database. The CCDC numbers are shown in Supplementary Materials, Table S1.

Acknowledgments

The support from the Research Council of Norway to the project (grant number: NFR 303369) and the Norwegian NMR Platform (project number 226244/F50) is highly appreciated, as is the help from the Mass Spectrometry Lab at the NV Faculty at NTNU. Roger Aarvik is thanked for technical support. The X-ray single crystal analysis was supported by the CzechNanoLab project LM2023051, funded by MEYS. CR is gratefully acknowledged for the financial support of the measurements/sample fabrication at LNSM Research Infrastructure.

Conflicts of Interest

There are no conflicts of interest to declare.

References

  1. Glaser, D. Specialization and phyletic trends of sweetness reception in animals. Pure App. Chem. 2002, 74, 1153–1158. [Google Scholar] [CrossRef]
  2. Leow, D.; Tan, C.-H. Chiral Bicyclic Guanidine, Bis-Guanidinium, Pentanidium and Related Organocatalysts. Top. Heterocycl. Chem. 2017, 50, 129–156. [Google Scholar] [CrossRef]
  3. Dong, S.; Feng, X.; Liu, X. Chiral guanidines and their derivatives in asymmetric synthesis. Chem. Soc. Rev. 2018, 47, 8525–8540. [Google Scholar] [CrossRef]
  4. Geiselhart, C.M.; Schmitt, C.W.; Jöckle, P.; Mutlu, H.; Barner-Kowollik, C. A Guanidine-Based Superbase as Efficient Chemiluminescence Booster. Sci. Rep. 2019, 9, 14519. [Google Scholar] [CrossRef]
  5. Bailey, P.J.; Pace, S. The coordination chemistry of guanidines and guanidinates. Coord. Chem. Rev. 2001, 214, 91–141. [Google Scholar] [CrossRef]
  6. Manna, S.; Truong, J.; Hammond, M.C. Guanidine Biosensors Enable Comparison of Cellular Turn-on Kinetics of Riboswitch-Based Biosensor and Reporter. ACS Synth. Biol. 2021, 10, 566–578. [Google Scholar] [CrossRef]
  7. Gomes, A.R.; Varela, C.L.; Pires, A.S.; Tavares-da-Silva, E.J.; Roleira, F.M.F. Synthetic and natural guanidine derivatives as antitumor and antimicrobial agents: A review. Bioorg. Chem. 2023, 138, 106600. [Google Scholar] [CrossRef]
  8. Gorbunova, M.; Lemkina, L.; Borisova, I. New guanidine-containing polyelectrolytes as advanced antibacterial materials. Eur. Polym. J. 2018, 105, 426–433. [Google Scholar] [CrossRef]
  9. Cunningham, I.D.; Wan, N.C.; Cox, B.G. 1H and 15N NMR studies of N-substituted-N′-cyanoguanidines. J. Chem. Soc. Perkin Trans. 2 1994, 1849–1853. [Google Scholar] [CrossRef]
  10. Caine, B.A.; Dardonville, C.; Popelier, P.L.A. Prediction of Aqueous pKa Values for Guanidine-Containing Compounds Using Ab Initio Gas-Phase Equilibrium Bond Lengths. ACS Omega 2018, 3, 3835–3850. [Google Scholar] [CrossRef]
  11. Dardonville, C.; Caine, B.A.; Navarro de la Fuente, M.; Martín Herranz, G.; Corrales Mariblanca, B.; Popelier, P.L.A. Substituent effects on the basicity (pKa) of aryl guanidines and 2-(arylimino)imidazolidines: Correlations of pH-metric and UV-metric values with predictions from gas-phase ab initio bond lengths. New J. Chem. 2017, 41, 11016–11028. [Google Scholar] [CrossRef]
  12. Tanganov, B.B.; Alekseeva, I.A. Acid-base equilibria in solutions of polyacid bases (model and experiment): I. Thermodynamic dissociation constants of protonated mono- and diacid bases. Russ. J. Gen. Chem. 2006, 76, 1724–1728. [Google Scholar] [CrossRef]
  13. Paixao, J.A.; Matos Beja, A.; Pereira Silva, P.S.; Ramos Silva, M.; Alte da Veiga, L. A new orthorhombic phase of N,N’-diphenylguanidine. Acta Crystallogr. Sect. C 1999, 55, 1037–1040. [Google Scholar] [CrossRef]
  14. Zeng, C.-J.; Chen, C.-J.; Chang, C.-W.; Chen, H.-T.; Chien, T.-C. Copper(I) Iodide-Catalyzed Synthesis of N,N′-Disubstituted Guanidines from N-Substituted Cyanamides. Aust. J. Chem. 2014, 67, 1134–1137. [Google Scholar] [CrossRef]
  15. Carpy, A.; Leger, J.-M.; Wermuth, C.-G.; Leclerc, G. Structure de la (dichloro-2,6 phenyl)-2 methyl-1 guanidine base et chlorhydrate; un analogue ouvert de la clonidine. Acta Cryst. 1981, B37, 885–889. [Google Scholar] [CrossRef]
  16. Brown, C.J.; Gash, D.J. N,N’-Bis(2-methylphenyl)guanidine, C15H17N3. Acta Crystallogr. Sect. C 1984, 40, 562–564. [Google Scholar] [CrossRef]
  17. Pareras, G.; Palusiak, M.; Duran, M.; Solà, M.; Simon, S. Tuning the Strength of the Resonance-Assisted Hydrogen Bond in o-Hydroxybenzaldehyde by Substitution in the Aromatic Ring1. J. Phys. Chem. A 2018, 122, 2279–2287. [Google Scholar] [CrossRef] [PubMed]
  18. Janusz Grabowski, S. π-Electron delocalisation for intramolecular resonance assisted hydrogen bonds. J. Phys. Org. Chem. 2003, 16, 797–802. [Google Scholar] [CrossRef]
  19. Rusinska-Roszak, D. Energy of Intramolecular Hydrogen Bonding in ortho-Hydroxybenzaldehydes, Phenones and Quinones. Transfer of Aromaticity from ipso-Benzene Ring to the Enol System(s). Molecules 2017, 22, 481. [Google Scholar] [CrossRef]
  20. Gilli, G.; Bellucci, F.; Ferretti, V.; Bertolasi, V. Evidence for resonance-assisted hydrogen bonding from crystal-structure correlations on the enol form of the .beta.-diketone fragment. J. Am. Chem. Soc. 1989, 111, 1023–1028. [Google Scholar] [CrossRef]
  21. Elumalai, V.; Vaclav, E.; Visnes, T.; Sundby, E.; Hoff, B.H. Improved Synthetic Methodology, Substrate Scope and X-ray Crystal Structure for N, N’-disubstituted Guanidines. ChemistrySelect 2024, 9, e202304381. [Google Scholar] [CrossRef]
  22. Kohoutova, K.; Dočekal, V.; Ausserlechner, M.J.; Kaiser, N.; Tekel, A.; Mandal, R.; Horvath, M.; Obsilova, V.; Vesely, J.; Hagenbuchner, J.; et al. Lengthening the Guanidine–Aryl Linker of Phenylpyrimidinylguanidines Increases Their Potency as Inhibitors of FOXO3-Induced Gene Transcription. ACS Omega 2022, 7, 34632–34646. [Google Scholar] [CrossRef] [PubMed]
  23. Ciarlantini, M.S.; Barquero, A.; Bayo, J.; Wetzler, D.; Dodes Traian, M.M.; Bucci, H.A.; Fiore, E.J.; Gandolfi Donadio, L.; Defelipe, L.; Turjanski, A.; et al. Development of an Improved Guanidine-Based Rac1 Inhibitor with in vivo Activity against Non-Small Cell Lung Cancer. ChemMedChem 2021, 16, 1011–1021. [Google Scholar] [CrossRef] [PubMed]
  24. Rigaku, O.D. CrysAlis PRO; Rigaku Oxford Diffraction Ltd.: Yarnton, UK, 2020. [Google Scholar]
  25. Palatinus, L.; Chapuis, G. SUPERFLIP—A computer program for the solution of crystal structures by charge flipping in arbitrary dimensions. J. Appl. Cryst. 2007, 40, 786–790. [Google Scholar] [CrossRef]
  26. Betteridge, P.W.; Carruthers, J.R.; Cooper, R.I.; Prout, K.; Watkin, D.J. CRYSTALS version 12: Software for guided crystal structure analysis. J. Appl. Crystallogr. 2003, 36, 1487. [Google Scholar] [CrossRef]
  27. Petříček, V.; Palatinus, L.; Plášil, J.; Dušek, M. Jana2020—A new version of the crystallographic computing system Jana. Z. Für Krist. —Cryst. Mater. 2023, 238, 271–282. [Google Scholar] [CrossRef]
  28. Abou-Shehada, S.; Teasdale, M.C.; Bull, S.D.; Wade, C.E.; Williams, J.M.J. Lewis Acid Activation of Pyridines for Nucleophilic Aromatic Substitution and Conjugate Addition. ChemSusChem 2015, 8, 1083–1087. [Google Scholar] [CrossRef]
  29. Rohlicek, J.; Husak, M. MCE2005—A new version of a program for fast interactive visualization of electron and similar density maps optimized for small molecules. J. Appl. Cryst. 2007, 40, 600–601. [Google Scholar] [CrossRef]
  30. Brandenburg, K. DIAMOND; Crystal Impact GbR: Bonn, Germany, 1999. [Google Scholar]
  31. Stephens, P.J.; Devlin, F.J.; Chabalowski, C.F.; Frisch, M.J. Ab Initio Calculation of Vibrational Absorption and Circular Dichroism Spectra Using Density Functional Force Fields. J. Phys. Chem. 1994, 98, 11623–11627. [Google Scholar] [CrossRef]
  32. Dunning, T.H., Jr. Gaussian basis sets for use in correlated molecular calculations. I. The atoms boron through neon and hydrogen. J. Chem. Phys. 1989, 90, 1007–1023. [Google Scholar] [CrossRef]
  33. Woon, D.E.; Dunning, T.H., Jr. Gaussian basis sets for use in correlated molecular calculations. III. The atoms aluminum through argon. J. Phys. Chem. 1993, 98, 1358–1371. [Google Scholar] [CrossRef]
  34. Sun, Q.; Berkelbach, T.C.; Blunt, N.S.; Booth, G.H.; Guo, S.; Li, Z.; Liu, J.; McClain, J.D.; Sayfutyarova, E.R.; Sharma, S.; et al. PySCF: The Python-based simulations of chemistry framework. WIREs Comput. Mol. Sci. 2018, 8, e1340. [Google Scholar] [CrossRef]
  35. Sun, Q.; Zhang, X.; Banerjee, S.; Bao, P.; Barbry, M.; Blunt, N.S.; Bogdanov, N.A.; Booth, G.H.; Chen, J.; Cui, Z.-H.; et al. Recent developments in the PySCF program package. J. Chem. Phys. 2020, 153, 024109. [Google Scholar] [CrossRef] [PubMed]
  36. Pan, X.; Zhao, F.; Zhang, Y.; Wang, X.; Xiao, X.; Zhang, J.Z.H.; Ji, C. MolTaut: A Tool for the Rapid Generation of Favorable Tautomer in Aqueous Solution. J. Chem. Inf. Model. 2023, 63, 1833–1840. [Google Scholar] [CrossRef] [PubMed]
  37. Allen, F.H.; Kennard, O.; Watson, D.G.; Brammer, L.; Orpen, A.G.; Taylor, R. Tables of bond lengths determined by X-ray and neutron diffraction. Part 1. Bond lengths in organic compounds. J. Chem. Soc. Perkin Trans. 1987, 2, S1–S19. [Google Scholar] [CrossRef]
  38. Pietruś, W.; Kafel, R.; Bojarski, A.J.; Kurczab, R. Hydrogen Bonds with Fluorine in Ligand-Protein Complexes-the PDB Analysis and Energy Calculations. Molecules 2022, 27, 1005. [Google Scholar] [CrossRef]
  39. Gross, K.C.; Seybold, P.G.; Peralta-Inga, Z.; Murray, J.S.; Politzer, P. Comparison of Quantum Chemical Parameters and Hammett Constants in Correlating pKa Values of Substituted Anilines. J. Org. Chem. 2001, 66, 6919–6925. [Google Scholar] [CrossRef]
  40. Tehan, B.G.; Lloyd, E.J.; Wong, M.G.; Pitt, W.R.; Gancia, E.; Manallack, D.T. Estimation of pKa Using Semiempirical Molecular Orbital Methods. Part 2: Application to Amines, Anilines and Various Nitrogen Containing Heterocyclic Compounds. Quant. Struct.-Act. Relat. 2002, 21, 473–485. [Google Scholar] [CrossRef]
  41. Hall, H.K., Jr. Correlation of the Base Strengths of Amines1. J. Am. Chem. Soc. 1957, 79, 5441–5444. [Google Scholar] [CrossRef]
  42. Smith, P.J.; Noble, A. A Primary Hydrogen-Deuterium Isotope Effect Study on the Carbonyl Elimination Reaction of 9-Fluorenyl Nitrate with Various Bases. Can. J. Chem. 1975, 53, 263–268. [Google Scholar] [CrossRef]
  43. Juranic, I. Simple Method for the Estimation of pKa of Amines. Croat. Chem. Acta 2014, 87, 343–347. [Google Scholar] [CrossRef]
  44. CRC Handbook of Chemistry and Physics; CRC Press: Boca Raton, FL, USA, 2016; Volume 97.
Figure 1. Structure of N,N′-substituted guanidines previously analysed by X-ray.
Figure 1. Structure of N,N′-substituted guanidines previously analysed by X-ray.
Crystals 14 00884 g001
Scheme 1. Structure of the N,N′-substituted guanidines 112 included in the study.
Scheme 1. Structure of the N,N′-substituted guanidines 112 included in the study.
Crystals 14 00884 sch001
Figure 2. Possible tautomers of the guanidine structures. The most important results from gas-phase calculations are shown alongside the atom numbering system used in the discussion.
Figure 2. Possible tautomers of the guanidine structures. The most important results from gas-phase calculations are shown alongside the atom numbering system used in the discussion.
Crystals 14 00884 g002
Figure 3. Representative X-ray structures for the compound collection: compounds 1 and 9 have a cis–trans tautomer-T1-like structure, compound 11 crystallised with two molecules of water, and compound 2 has a T2-cis–cis tautomer type.
Figure 3. Representative X-ray structures for the compound collection: compounds 1 and 9 have a cis–trans tautomer-T1-like structure, compound 11 crystallised with two molecules of water, and compound 2 has a T2-cis–cis tautomer type.
Crystals 14 00884 g003
Figure 4. Average bond and torsion angles for the guanidines.
Figure 4. Average bond and torsion angles for the guanidines.
Crystals 14 00884 g004
Figure 5. Intermolecular hydrogen bonding in the two tautomeric classes of compounds.
Figure 5. Intermolecular hydrogen bonding in the two tautomeric classes of compounds.
Crystals 14 00884 g005
Table 1. Overview of the guanidines analysed by X-ray, the number of molecules in the asymmetric unit (AS), and their space group.
Table 1. Overview of the guanidines analysed by X-ray, the number of molecules in the asymmetric unit (AS), and their space group.
Comp.StructureAS 1Space Group
1Crystals 14 00884 i0011P21/c
2Crystals 14 00884 i0021P1
3Crystals 14 00884 i0032P21/n
4Crystals 14 00884 i0041P1
5Crystals 14 00884 i0051P21/c
6Crystals 14 00884 i0061P21/c
7Crystals 14 00884 i0073P21
8Crystals 14 00884 i0083Cc
9Crystals 14 00884 i0092P21/c
10Crystals 14 00884 i0102P21/c
11Crystals 14 00884 i0113P1
12Crystals 14 00884 i0123P21/c
1 Number of molecules in asymmetric unit.
Table 2. Bond length (Å) for the central part of the guanidine structures. The values are averaged for all of the molecules in the asymmetric unit.
Table 2. Bond length (Å) for the central part of the guanidine structures. The values are averaged for all of the molecules in the asymmetric unit.
Crystals 14 00884 i013
Comp.N3-C2C2-N1C2-N7N7-C8C8-N9C8-N10N10-C11
11.3461.3581.3751.3281.3401.3681.407
21.3411.3391.3841.3861.3431.3011.400
31.3421.3391.3871.3801.3351.3131.402
41.3341.3421.3791.3861.3361.2941.412
51.3521.3591.3671.3341.3411.3391.457
61.3491.3601.3691.3321.3421.3461.459
71.3521.3621.3671.3331.3471.3461.455
81.3511.3601.3681.3331.3431.3441.457
91.3531.3621.3681.3331.3471.3461.453
101.3491.3611.3621.3391.3551.3431.456
111.3521.3591.3711.3431.3411.3441.467; 1.462
121.3521.3571.3721.3281.3461.3681.470; 1.471
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Elumalai, V.; Eigner, V.; Janjua, N.A.; Åstrand, P.-O.; Visnes, T.; Sundby, E.; Hoff, B.H. X-ray Structure of Eleven New N,N′-Substituted Guanidines: Effect of Substituents on Tautomer Structure in the Solid State. Crystals 2024, 14, 884. https://doi.org/10.3390/cryst14100884

AMA Style

Elumalai V, Eigner V, Janjua NA, Åstrand P-O, Visnes T, Sundby E, Hoff BH. X-ray Structure of Eleven New N,N′-Substituted Guanidines: Effect of Substituents on Tautomer Structure in the Solid State. Crystals. 2024; 14(10):884. https://doi.org/10.3390/cryst14100884

Chicago/Turabian Style

Elumalai, Vijayaragavan, Vaclav Eigner, Nicholas Alexander Janjua, Per-Olof Åstrand, Torkild Visnes, Eirik Sundby, and Bård Helge Hoff. 2024. "X-ray Structure of Eleven New N,N′-Substituted Guanidines: Effect of Substituents on Tautomer Structure in the Solid State" Crystals 14, no. 10: 884. https://doi.org/10.3390/cryst14100884

APA Style

Elumalai, V., Eigner, V., Janjua, N. A., Åstrand, P. -O., Visnes, T., Sundby, E., & Hoff, B. H. (2024). X-ray Structure of Eleven New N,N′-Substituted Guanidines: Effect of Substituents on Tautomer Structure in the Solid State. Crystals, 14(10), 884. https://doi.org/10.3390/cryst14100884

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