Vjekoslav
Štrukil
a,
Ivica
Đilović
b,
Dubravka
Matković-Čalogović
b,
Jaan
Saame
c,
Ivo
Leito
c,
Primož
Šket
de,
Janez
Plavec
def and
Mirjana
Eckert-Maksić
*a
aRuđer Bošković Institute, Bijenička cesta 54, 10000 Zagreb, Croatia. E-mail: mmaksic@emma.irb.hr; Fax: +385-1-4680-195; Tel: +385-1-4680-197
bFaculty of Science, Department of Chemistry, Horvatovac 102A, 10000 Zagreb, Croatia
cUniversity of Tartu, Institute of Chemistry, Ravila 14a, 50411 Tartu, Estonia
dSlovenian NMR Centre, National Institute of Chemistry, Hajdrihova 19, SI-1000 Ljubljana, Slovenia
eEN-FIST Centre of Excellence, Dunajska 156, SI-1000 Ljubljana, Slovenia
fFaculty of Chemistry and Chemical Technology, University of Ljubljana, Askerceva cesta 5, SI-1000 Ljubljana, Slovenia
First published on 25th October 2011
It is shown that guanidine and its N,N-dimethyl-derivative react with substituted carbodiimides, affording hitherto unknown 1,2-dihydro-1,2,3-triazine derivatives. The structures of three novel compounds of this type and their perchlorate salts were elucidated by spectroscopic (IR, 1H and 13C NMR and 15N solid-state NMR) and X-ray diffraction methods. The acid/base properties were also determined experimentally and by using DFT calculations with the B3LYP functional. The most basic compound was found to be dihydrotriazine 3, the basicity of which with the pKa value of 23.3 is of the same order of magnitude as that of tetramethylguanidine. Acidity measurements revealed that all the compounds studied are very weak acids with pKa values in the range of 25.8–30.8 pKa units in acetonitrile.
Scheme 1 The structure of cycloguanil, an antimalarial1,2-dihydrotriazine inhibitor of dihydrofolate reductase (Ar = 4-ClC6H4).3 |
Other successful applications include use as herbicides,6insecticides7 and corrosion inhibitors.8 Consequently, their syntheses and the evaluation of their biological activity continue to attract considerable attention.9 In the course of our ongoing project on the design and reactivity of strong guanidine bases,10 we became interested in exploring the possibility of preparing some specifically substituted dihydrotriazines using guanidine derivatives and carbodiimides as reactants. This approach, as well as the reaction of biguanide derivatives and carbodiimides, has been extensively explored in the past. In all the reports published so far, only the formation of melamine or isomelamine derivatives was reported.11–13 The reaction was proposed to proceed through the primary addition of the reactants, followed by the cyclization of the resulting intermediate triguanide, with the loss of amine, to the heterocyclic end-product.11 However, although this approach to melamine and isomelamine derivatives has been known for a long time, no thorough study of a reaction mechanism and a structure of the so obtained products has been published as yet.11,12 Therefore, we considered it of interest to address some of these questions in the present work. Furthermore, an additional focus of our interest in the properties of dihydrotriazines concerned their acid–base properties. Namely, due to the presence of guanidine-type subunits in their molecular framework, they are expected to exhibit basicity of a similar order of magnitude as guanidine or biguanide derivatives.14
Herewith, we report on the synthesis and structural features of three novel 1,2-dihydro-1,3,5-triazine derivatives, namely 1–3 (Scheme 2). It should be noted that they differ from previously reported members of this family of compounds by the presence of the exo-cyclic imino group at ring position 2. We were particularly interested in establishing the structure of compound 1, for which the isomelamine geometry (see Scheme 5) from UV spectral analysis was previously proposed.11,12 Our second aim in this work was to evaluate their basicity and acidity experimentally by employing quantum chemical calculations. This is of considerable importance since due to the presence of exo-cyclic imino- and amino-type nitrogen atoms, they are expected to have the ability either to donate or accept a proton, i.e., to act as either an acid or a base (Scheme 2). In addition, the results of a computational study of the effect of p-substituents in aryl-substituted species on the protonation energies, with the goal of designing potentially stronger acids/bases belonging to this class of compounds, will be presented. To the best of our knowledge, this has been the first comprehensive study of the structural features and acid/base properties of 1,2-dihydro-1,3,5-triazine derivatives so far.
Scheme 2 The 1,2-dihydro-1,3,5-triazine derivatives 1–3 studied in this work. |
Scheme 3 Synthesis of triazines 1–3. |
Scheme 4 Possible mechanisms of the formation of 1,2-dihydro-1,3,5-triazine derivatives 1–3 (R1 = H or Me, R2 = Ph or i-Pr). |
This confirms the previous proposition11,12 that an intermediately formed triguanide derivative undergoes spontaneous cyclization accompanied by the elimination of amine R2NH2 (aniline in the reaction of N,N′-diphenylcarbodiimide (6) or isopropylamine when N,N′-diisopropylcarbodiimide (7) is used). From a mechanistic point of view, two modes of cyclization to 1,2-dihydrotriazines could be envisaged (Scheme 4): one starting directly from the triguanide moiety 10 (path A) or from its ring tautomer intermediate 11, which is formed by a nucleophilic attack of the lone pair at the imino nitrogen N1 to the carbon atom C6 in triguanide 10, followed by the elimination of the R2NH2 molecule (path B). This type of tautomerism is well known from, e.g., sugar chemistry.15 More recently, chain–ring tautomerism, which, in fact, corresponds to thermally allowed 6π-electrocyclization, has also been invoked to explain the mechanism of the cyclization of some unsaturated hetero chains of the oligonitrile type into 1,2-dihydro-1,3,5-triazine derivatives.16
However, in this work, we could not confirm the presence of these intermediates experimentally due to their obvious instability. Therefore, we calculated the energies of both types of tautomers for the model compounds in which R1 and R2groups were replaced by methyl groups, using the B3LYP/6-311+G(d,p)//B3LYP/6-31G(d,p) method. In addition to the gas phase, calculations were carried out in THF as the solvent. The geometries of the fully optimized species are shown in Fig. S3 and Table S1 (see ESI†) and the computed energies are listed in Table 1. A glance at the calculated energies in Table 1 clearly shows that the open-chain and cyclic tautomers 10 and 11 are of similar stability, with the latter being slightly more stable in THF (by 1.07 kcal mol−1). Thus, based on these data, albeit qualitative, we presume that path B would be favored.
Structurea | E el/a.u. | E ZPV/a.u. | E tot/a.u. | E rel/kcal mol−1 |
---|---|---|---|---|
a R1 = R2 = CH3, see Scheme 4. TS corresponds to the transition structure for chain-tautomer cyclization. | ||||
Gas phase | ||||
10 | −739.06485 | 0.32223 | −738.74262 | 0.00 |
TS | −739.04652 | 0.32210 | −738.72442 | 11.42 |
11 | −739.06342 | 0.32486 | −738.73856 | 2.55 |
THF solution | ||||
ΔGsolv/kcal mol−1 | E solvrel/kcal mol−1 | |||
10 | 11.61 | 0.00 | ||
TS | 8.81 | 8.62 | ||
11 | 7.99 | −1.07 |
The 1,2-dihydrotriazine products were fully characterized by elemental analysis and spectroscopic methods (IR, HRMS, 1H and 13C solution and 1H, 13C and 15N solid-state NMR). The structures of compounds 1 and 2 were ultimately confirmed by X-ray diffraction analysis. In the mass spectra, all three compounds gave the expected molecular ion [MH+] along with some fragment ions. The proposed structures were supported by analysis of the solution and solid-state NMR spectra. For instance, taking the 1H NMR spectra in the solution of compound 3 as an example, we observe a broad doublet at 6.15 for the C–NHproton, a singlet for the protons of the N–CH3 group at 2.96 ppm, three multiplets for the N–CHprotons in the range of 3.96–4.76 ppm and three doublets for the CH–CH3 protons in the range of 0.96–1.41 ppm. The corresponding 13C NMR spectrum in solution displays three signals between 159.2–148.2, three signals between 45.3–42.8 ppm, one signal at 35.5 ppm and three signals between 24.5–19.4 ppm assigned to the ring carbon atoms, CH carbon atoms of three chemically different isopropyl groups, N–CH3 and C–CH3 carbon atoms, respectively.
Similar chemical shifts for carbon atoms have been observed in the solid state as well (Fig. 1). The 1H and 13C NMR spectra of triphenyl derivatives 1 and 2 (see the Experimental section) in solution, as well as in the solid state (Fig. S4–S7 and S19–S22 in ESI†), are also fully consistent with the 1,2-dihydro-1,3,5-dihydrotriazine structures. Notably, the signals for the NHproton in the 1H NMR spectra of 1 and 2, taken in DMSO, appear in the same region (7.61 ppm in 1 and 7.75 ppm in 2 in DMSO), thereby confirming that the imino proton is in the same chemical environment in both compounds (Fig. S4 and S6 in ESI†).
Fig. 1 13C CP-MAS NMR spectra of compounds 1 (a), 2 (b) and 3 (c). |
Furthermore, in the 15N CP-MAS NMR spectra, the most upfield signals between −290 and −300 ppm indicate the presence of the amino/dimethylamino group attached to the C4 atom (Fig. 2).
Fig. 2 15N CP-MAS NMR spectra of compounds 1 (a) and 2 (b). |
Even more importantly, the signals for N3 and N5 atoms have different chemical shifts. In the case of 1, this excludes the existence of the quinoid and isomelamine structures 1T2 and 1T3 (Scheme 5), in which we would expect to observe a single resonance for both atoms due to the symmetry of the molecule. In order to obtain insights into the intrinsic stabilities of tautomers 1T1–1T3 of compound 1, their energies were calculated using the B3LYP/6-311+G(d,p)//B3LYP/6-31G(d) method. At this level of theory, tautomer 1T1 was found to be more stable than 1T2 and 1T3 by 15.2 and 11.6 kcal mol−1, thus being in accordance with NMR data (see Table S3 for energies in ESI†).
Scheme 5 Possible tautomeric structures of molecule 1. 1T1-imino, 1T2-quinoid and 1T3-isomelamine structures. |
Finally, an interesting feature observed in the 15N CP-MAS NMR spectrum of compound 1 concerns the presence of a larger number of signals than in compound 2, indicating that two molecules are present in the asymmetric unit of compound 1, while a single molecule is present in the asymmetric unit of compound 2 in the solid state. This, as will be shown below, was confirmed by X-ray diffraction data.
Fig. 3 ORTEP drawing of 1·0.5 MeOH. Thermal ellipsoids are at the 30% probability level. |
Fig. 4 ORTEP drawing of 2. Thermal ellipsoids are at the 30% probability level. |
Bond | 1a | 1b | 2 |
---|---|---|---|
a For numeration of atoms see Fig. 3 and 4. | |||
N1–C2 | 1.416(2) | 1.419(2) | 1.433(2) |
N1–C6 | 1.374(2) | 1.372(2) | 1.371(2) |
C2–N3 | 1.360(2) | 1.370(2) | 1.354(2) |
N3–C4 | 1.331(2) | 1.329(2) | 1.328(2) |
C4–N9 | 1.325(2) | 1.333(2) | 1.346(2) |
C4–N5 | 1.367(2) | 1.366(2) | 1.362(2) |
N5–C6 | 1.302(2) | 1.317(2) | 1.308(2) |
C2–N7 | 1.292(2) | 1.287(2) | 1.292(2) |
C6–N8 | 1.359(2) | 1.343(2) | 1.361(2) |
N1–Ph | 1.447(2) | 1.444(2) | 1.444(2) |
N7–Ph | 1.423(2) | 1.416(2) | 1.411(2) |
N8–Ph | 1.418(2) | 1.433(2) | 1.424(2) |
1·0.5 MeOH | 2 | 2·HClO4 | 3·HClO4 | |
---|---|---|---|---|
a R = ∑||Fo| − |Fc||/∑Fo, w = 1/[σ2(F2o) + (g1P)2 + g2P] where P = (F2o + 2F2c)/3, S = Σ[w(F2o – F2c)2/(Nobs − Nparam)]1/2. b wR = [Σ(F2o − F2c)2/Σ(F2o)2]1/2. | ||||
Empirical formula | 2(C21H18N6), CH4O | C23H22N6 | C23H23N6, ClO4 | C14H29N6, ClO4 |
Formula weight | 740.87 | 382.47 | 482.92 | 380.88 |
Crystal system, space group | Monoclinic, P 21/c | Orthorhombic, P bca | Orthorhombic, P 212121 | Monoclinic, P 21/c |
Unit cell dimensions (Å, °) | ||||
a | 14.1696(2) | 19.0976(8) | 6.9133(5) | 9.4056(10) |
B | 17.3971(4) | 9.1823(3) | 15.5643(12) | 9.3787(9) |
c | 15.0518(3) | 21.9471(8) | 20.7086(14) | 22.2736(18) |
α | 90.00 | 90.00 | 90.000 | 90.00 |
β | 91.7700(20) | 90.00 | 90.000 | 98.682(7) |
γ | 90.00 | 90.00 | 90.000 | 90.00 |
Volume/Å3 | 3708.65(12) | 3848.6(2) | 2228.3(3) | 1942.3(3) |
Z | 4 | 8 | 4 | 4 |
D calc/g cm−3 | 1.327 | 1.320 | 1.440 | 1.303 |
T/K | 100(2) | 100(2) | 150(2) | 110(2) |
Reflections observed/independent (Rint) | 35085/6866 (0.0366) | 12539/3363 (0.0426) | 5090/3659 (0.0297) | 15999/3799 (0.0317) |
Observed reflections [I > 2σ(I)] | 4991 | 2698 | 2760 | 3083 |
Goodness-of-fit on F2 | 0.979 | 1.072 | 1.033 | 1.121 |
R/wR [I > 2σ(I)]a | 0.0443/0.1067 | 0.0454/0.0892 | 0.0632/0.1065 | 0.0488/0.1235 |
R/wR (all data)b | 0.0642/0.1172 | 0.0642/0.0958 | 0.0910/0.1167 | 0.0624/0.1317 |
Flack parameter | — | — | −0.04(12) | — |
It is important to note that the NH2 group (NMe2 in 2) is practically coplanar with the triazine moiety, thus enabling effective delocalization of the nitrogen lone pair in the triazine ring. Another notable feature concerns the difference in the length of the exocyclic C2–N7 and C6–N8 bonds. While the length of the C2–N7 bond (1.292(2) Å and 1.287(2), 1.292(2) Å in 1a, 1b, and 2, respectively) corresponds to a typical CN double bond, the C6–N8 bond (ranging from 1.343(2) to 1.361(2) Å) has a value closer to a typical Nsp2–Csp2 single bond, such as in ref. 18.
In the crystal structure of 1·0.5 MeOH, the two conformers are connected by hydrogen bonds of the type N–H⋯N (2.912(2) and 3.053(2) Å), forming a ring that can be described by the graph-set notation R22(8).
The methanol molecule is a donor in the hydrogen bond O–H⋯N (2.915(2) Å) to molecule 1a and is an acceptor of a weak hydrogen bond C–H⋯O from a neighboring 1a molecule (Fig. 5). Such supramolecular assemblies are interconnected in the crystal structure by weak interactions of the type C–H⋯O, C–H⋯N, C–H⋯π and N–H⋯π.
Fig. 5 A supramolecular assembly in the structure of 1·0.5 MeOH showing molecules interconnected by hydrogen bonds (blue and red lines). Symmetry code * = 1 − x, −y, −z. |
In contrast, packing in the crystal structure of 2 is achieved through weak interactions, mostly of the C–H⋯π type. The protonated nitrogen atom N8 is sterically hindered and not involved in hydrogen bonding.
Fig. 6 ORTEP drawings of (a) 2·HClO4 and (b) 3·HClO4. The thermal ellipsoids are at the 30% probability level. |
Bond | d/Å | ||||
---|---|---|---|---|---|
Theory | Measured | Theory | Measured | Theory | |
1H++ | 2·HClO4 | 2H++ | 3·HClO4 | 3H++ | |
a For numbering of the atoms see Scheme 2 and Fig. 6. | |||||
N1–C2 | 1.401 | 1.397(5) | 1.401 | 1.393(2) | 1.400 |
N1–C6 | 1.402 | 1.401(6) | 1.401 | 1.389(2) | 1.401 |
C2–N3 | 1.317 | 1.304(5) | 1.311 | 1.320(3) | 1.317 |
N3–C4 | 1.349 | 1.376(6) | 1.357 | 1.347(2) | 1.354 |
C4–N9 | 1.336 | 1.318(5) | 1.340 | 1.336(2) | 1.341 |
C4–N5 | 1.349 | 1.350(5) | 1.357 | 1.349(3) | 1.355 |
N5–C6 | 1.317 | 1.307(5) | 1.311 | 1.322(2) | 1.316 |
C2–N7 | 1.343 | 1.352(6) | 1.347 | 1.330(3) | 1.340 |
C6–N8 | 1.343 | 1.333(5) | 1.347 | 1.333(3) | 1.344 |
N1–Ph (i-Pr) | 1.449 | 1.446(5) | 1.448 | 1.503(2) | 1.505 |
N7–Ph (i-Pr) | 1.432 | 1.437(5) | 1.432 | 1.478(3) | 1.483 |
N8–Ph (i-Pr) | 1.432 | 1.429(6) | 1.432 | 1.479(3) | 1.484 |
Before analyzing the effect of protonation on 1–3 in detail, we note that the trend of the changes in the calculated structural parameters in 2H++ and 3H++ closely resembles those found in the X-ray structures of their perchlorate salts. The same holds for the comparison of the experimental and calculated structures of 1 and 2 and their protonated forms with those of the corresponding perchlorate salts. This, in turn, lends credence to the structural parameters calculated for the other species studied in this work.
The most striking feature of the experimental structures of the protonated species concerns the elongation of the C2–N7 bond relative to that in the neutral molecule, confirming that protonation occurs at the N7 atom. This results in a near equalization of the exocyclic C2–N7 and C6–N8 bonds, accompanied by changes in the bond lengths within the triazine ring, which assumes a paraquinonoid character with the C2–N3 and N5–C6 bond lengths being close to typical Csp2–Nsp2 bonds.
Compd. (B) | Reference base (Rb) | pKa (Rb) | ΔpKaa | pKa (B) | Assigned pKa(B) |
---|---|---|---|---|---|
a pKa(Rb) − pKa(B). b The standard uncertainties of the pKa values relative to the acetonitrile pKa scale are estimated as 0.05 pKa units (see ref. 20e). | |||||
1 | 2-Cl–C6H4P1(pyrr) | 20.17 | 1.69 | 18.48 | 18.51 |
2-Cl–C6H4P1(dma) | 19.07 | 0.59 | 18.48 | ||
2,6-Cl2–C6H3P1(pyrr) | 18.56 | 0.00 | 18.56 | ||
2 | PhP1(dma)2Me | 21.03 | 1.89 | 19.14 | 19.14 |
2-Cl–C6H4P1(pyrr) | 20.17 | 1.01 | 19.16 | ||
2-Cl–C6H4P1(dma) | 19.07 | −0.05 | 19.12 | ||
3 | 2-Cl–C6H4P2(dma) | 24.23 | 1.23 | 23.00 | 23.02 |
PhP1(pyrr) | 22.34 | −0.72 | 23.06 |
Compd. (A) | Reference acid (Ra) | pKa (Ra) | ΔpKaa | pKa (A) | Assigned pKab (A) |
---|---|---|---|---|---|
a pKa(Ra) − pKa(A). b The standard uncertainties of the pKa values relative to the acetonitrile pKa scale are estimated as 0.06 pKa units for 1 and 0.2 pKa units for 2 and 3 (see ref. 20e). | |||||
1 | (C6F5)(C6H5)CHCN | 26.14 | 0.31 | 25.83 | 25.83 |
(C5F4N)(C6H5)NH | 26.34 | 0.51 | 25.83 | ||
2 | (C6F5)(C6H5)CHCN | 26.14 | −0.47 | 26.61 | 26.6 |
(C5F4N)(C6H5)NH | 26.34 | −0.32 | 26.66 | ||
(4-Me-C6F4)(C6F5)NH | 24.94 | −1.48 | 26.42 | ||
3 | 9-C6F5-Fluorene | 28.11 | −2.7 | 30.8 | 30.8 |
2,3,4,5,6-(CF3)5-Toluene | 28.7 | −2.1 | 30.8 |
Analysis of the results for the basicity measurements of 1–3 (Table 5) reveals that the basicity of 3 is the highest among the studied dihydrotriazines. Its pKa value lies between those for the 2-Cl-C6H4P2(dma) and PhP1(pyrr) reference bases. In relation to our current interest in the basicity of guanidine compounds,19 which, similarly to the present compounds, undergo protonation at the imino group, we note that this pKa value is of the same order of magnitude as that of tetramethylguanidine (23.3 pKa units)20f and 3.9 pKa units smaller than the pKa value of N,N′,N′′-tris(3-dimethylaminopropyl)guanidine, which is the most basic acyclic guanidine derivative studied so far.19 It should be, however, emphasized that the high basicity of the latter compound is to a large extent due to the presence of cooperative intramolecular hydrogen bonds and their impact on the stabilities of the base and its conjugated acid,19 which are not present in dihydrotriazines studied in this work. Replacement of the isopropyl groups with the electron-withdrawing phenyl rings, leading to 2 and 1, leads to a decrease in basicity of 3.9 and 4.5 pKa units (in acetonitrile), respectively. It is noteworthy that replacement of the amino group with the dimethylamino moiety in 1 leading to 2 causes partial cancellation of the electron-withdrawing effect of the phenyl rings, resulting in a slight enhancement of basicity.
Comparison of the measured acidity data (Table 6) with those encompassed by the previously published acidity scale of the neutral compounds indicates that the acidity of the NH proton in 1–3 has been among the lowest measured in acetonitrile so far. Thus, the acidity of 3 is ca. 2 pKa units lower than that of 9-C6F5-fluorene (28.11), which has the lowest acidity in the previously reported acidity scale20a and thus 3 extends the previously established pKa scale of acidity in acetonitrile. The acidity of the respective protons in 1 and 2 is 5.0 and 4.2 pKa units higher relative to 3. The observed enhancement in acidity can be understood by comparing the calculated charge density distribution in the molecules considered. Due to the delocalization of electronic density on the deprotonation site into the neighboring phenyl ring in molecules 1 and 2, the loss of the proton from this site becomes easier and the resulting anion more stabilized than in the case of 3, where such delocalization is not possible (Tables S8–S10 in the ESI†).
Having measured the pKa values of guanidines 1–3, we decided to test the applicability of such correlations for calculating the pKa values of these types of bases. For this purpose, we used a slightly modified approach as employed in ref. 25. Specifically, the IEF-PCM model was used for calculating solvation energies instead of the IPCM method, due to a problem in the convergence of the isodensity surfaces for several structures. The same approach was subsequently used to evaluate the effect of a series of selected substituents in the para position of the phenyl rings, using molecule 2 as an example. All the optimized geometries were verified to be minima by vibrational analysis at the same level of theory. The resulting internal coordinates of all the optimized species are shown in Table S11 in the ESI.† The electronic energies and Gibbs energy corrections were calculated by the B3LYP/6-311+G(d,p) and B3LYP/6-31G(d) methods, respectively. Zero point vibrational energies were used unscaled. pKa values for the examined dihydrotriazines were calculated using linear eqn (1).26
pKa(calc) = 0.617 × GB′(B)AN − 155.585 | (1) |
GB′(AN) = GB′(B)gas + ΔG(BH+)AN − ΔG(B)AN | (2) |
The parameters for eqn (1) were evaluated from the linear relationship calculated for the test set of the 57 different nitrogen bases also used in our previous calculations, which span a range of ca. 40 pKa units. The calculated GB′(B)AN and pKa(calc) values of the compounds considered (as bases) are summarized in Table 7.
Molecule | GB(B)gas | GB′(B)gas | ΔG(X)AN/kcal mol−1 | GB′(B)AN/kcal mol−1 | pKa(calc) | pKa (exp) | Δ(pKa) | |
---|---|---|---|---|---|---|---|---|
(kcal mol−1) | X = B | X = BH+ | ||||||
a For numeration of atoms see Scheme 2. | ||||||||
1 | 247.35 | 253.63 | 3.80 | −25.18 | 282.61 | 18.78 | 18.51 | −0.27 |
2 | 249.49 | 255.77 | 9.30 | −18.36 | 283.43 | 19.29 | 19.14 | −0.15 |
3 | 250.42 | 256.70 | 14.66 | −18.50 | 289.86 | 23.26 | 23.02 | −0.24 |
2-NO2 | 226.77 | 233.05 | −0.26 | −42.10 | 274.89 | 14.02 | ||
2-CN | 230.09 | 236.37 | 1.10 | −39.37 | 276.84 | 15.23 | ||
2-Cl | 242.59 | 248.87 | 10.82 | −22.57 | 282.26 | 18.57 | ||
2-F | 243.97 | 250.25 | 10.41 | −22.53 | 283.19 | 19.14 | ||
2-Me | 253.53 | 259.81 | 13.42 | −12.68 | 285.91 | 20.82 | ||
2-OMe | 255.71 | 261.99 | 8.67 | −16.43 | 287.09 | 21.55 | ||
2-NH2 | 260.28 | 266.56 | 0.27 | −22.49 | 289.32 | 22.93 | ||
2-NMe2 | 264.77 | 271.05 | 13.68 | −6.46 | 291.19 | 24.08 | ||
2-N7PhNMe2 | 255.46 | 261.74 | 10.60 | −14.52 | 286.86 | 21.41 | ||
2-N7,N8PhNMe2 | 259.37 | 265.65 | 11.92 | −10.82 | 288.39 | 22.35 |
Before considering the effect of substituents on the basicity of compound 2, we note that the calculated pKa values for bases 1–3 are in excellent agreement with the experimentally measured values. Analysis of the data calculated for compound 2 substituted in the para position of the phenyl rings shows the strong impact of the electronic properties of the substituents on basicity. As expected, the effect of strong electron-donating dimethylamino groups is the most profound, causing an increase in basicity by ca. 5 pKa units. It is interesting to note that the effect of the dimethylamino group is additive; the largest contribution arises from the dimethylamino group at the phenyl ring attached to the N7 (2.12 pKa units), at which protonation takes place. This is followed by the contribution of the dimethylamino group at the N1 (1.73 pKa units) and N8 (0.94 pKa units) (Table 7). A somewhat weaker effect is observed for the amino substituted species, while methoxy and methyl groups increase basicity only moderately. On the other hand, substitution by the electron-accepting groups lowers basicity, with the largest effect encountered for the nitro-substituted species. To summarize, the chosen set of substituents facilitates fine-tuning of the basicity of the examined compound in the range of ca. 10 pKa units. It is quite plausible that all of these species can be easily prepared. Their use in further studies could lead to fine refinement of the existing scales of basicity in the range of ca. 14 to 24 pKa values.
NMR spectra of solid samples were recorded on a Varian NMR System 600 MHz NMR spectrometer equipped with a 3.2 mm NB Double Resonance HX MAS Solids Probe. The Larmor frequencies of protons, carbon and nitrogen nuclei were 599.77 MHz, 150.83 MHz and 60.78 MHz, respectively. The 1H MAS NMR spectra were externally referenced using adamantane. The 13C CP-MAS NMR spectra were externally referenced using hexamethylbenzene (HMB). The 15N CP-MAS NMR spectra were externally referenced using ammonium sulfate (δ −355.7 ppm regarding nitromethane at δ 0.0 ppm). Samples were spun at the magic angle at 20 kHz during 1H measurement and 10 kHz during 13C and 15N measurements. The 1H spectra were acquired using an echo pulse sequence. The repetition delay was 5 s and the number of scans was 16. The pulse sequences used for acquiring the 13C and 15N spectra were standard cross-polarization MAS pulse sequences with high-power proton decoupling during acquisition. The repetition delay was 5 s. The numbers of scans were between 600 and 2200 for the 13C measurements and 22000 and 27000 for 15N measurements.
HPLC analyses were performed on a Varian ProStar Instrument supplied with a UV/Vis detector using a Restek UltraIBD C18 (reversed phase) 5 μm 250 × 4.6 mm column operated at room temperature with a flow rate of 1 mL min−1; gradient of 2% acetic acid (solvent A) and methanol (solvent B): 95% A + 5% B, 0–5 min; 85% A + 15% B, 5–45 min, 35% A + 65% B, 45–55 min, 5% A + 95% B, 55 min. Melting points were determined on a Kofler hot-stage apparatus and are uncorrected. Elemental analyses were performed on a Perkin Elmer 2400 Series II CHNS/O Analyzer.
υ max/cm−1 3352, 1638, 1430, 1070. δH(600 MHz; d6-DMSO; Me4Si) 2.74 (6 H, s, CH3), 4.30–4.70 (3 H, br s, NH). δC(150 MHz, d6-DMSO; Me4Si) 37.5, 160.0.
Evaporation of the filtrate afforded a pale yellowish residue as a mixture of N,N′-diphenylbiguanide (∼8%) and N,N′,N′′-triphenylguanidine (8) (∼92%, by HPLC). An analytically pure sample of guanidine by-product 8 was obtained by the recrystallisation of the residue from ethanol.
A single crystal suitable for X-ray crystallographic measurement was obtained by dissolving a sample of 1 (0.10 g) in an ethanol–methanol 2:1 mixture (15 cm3). After 4 days of slow evaporation, the product was crystallized in the form of colorless prisms (65 mg, 65%).
mp 277–278 °C (from EtOH–MeOH 2:1 mixture) (lit.,1 270–271 °C). Found: C, 70.0%; H, 5.4%; N, 23.0% C21H18N6·0.5CH3OH requires C, 69.7%; H, 5.45%; N, 22.7%. υmax/cm−1 3393, 3154, 1637, 1588, 1534, 1463, 1362, 1295, 1143, 772, 743, 700. δH(600 MHz; d6-DMSO; Me4Si) 6.40–7.40 (12 H, m, overlapped Ph and NH2 protons), 7.44–7.46 (3 H, m, Ph), 7.52–7.55 (2 H, m, Ph), 7.61 (1 H, s, NH). δC(150 MHz, d6-DMSO; Me4Si) 119.5–121, 123–125 (br), 127.9, 128.4, 129.62, 129.64, 136.5, 162.4. HRMS-MALDI found: 355.1683; calc. for C21H19N6 (MH+): 355.1666.
Evaporation of the filtrate yielded a pale yellowish residue of N,N′,N′′-triphenylguanidine (8) (>95% by HPLC analysis, crude product yield 97%).
A single crystal suitable for X-ray crystallographic studies was obtained by dissolving a sample of 2 (0.10 g) in methanol (8.0 cm3). The product crystallized in the form of pale yellowish prisms (66 mg, 66%).
mp 197–198 °C (from MeOH). Found: C, 71.4%; H, 5.7%; N, 21.9% C23H22N6 requires C, 72.2%; H, 5.8%; N, 22.0%. υmax/cm−1 3390, 2923, 2853, 1629, 1611, 1584, 1534, 1487, 1467, 1401, 1218, 765, 694. δH(600 MHz; d6-DMSO; Me4Si) 2.97 (6 H, s, CH3), 6.60–7.60 (15 H, m, overlapped Ph protons), 7.75 (1 H, s, NH). δC(150 MHz, d6-DMSO; Me4Si) 35.7, 120, 123, 124, 127.7, 128.5, 129.58, 129.63, 136.3, 137, 150, 151, 155, 160.0. HRMS-MALDI found: 383.1971; calc. for C23H23N6 (MH+): 383.1979.
mp 203–205 °C (from H2O–EtOH mixture). υmax/cm−1 3402, 3063, 2930, 1657, 1622, 1589, 1543, 1462, 1411, 1225, 1120, 1095, 764, 745, 722, 692, 623, 561. δH(600 MHz; d6-DMSO; Me4Si) 3.01 (6 H, s, CH3), 7.22–7.27 (2 H, m, Ph), 7.32–7.41 (8 H, m, Ph), 7.70–7.84 (5 H, m, Ph), 8.81 (2 H, s, NH). δC(150 MHz, d6-DMSO; Me4Si) 36.1, 125.7, 126.1, 128.1, 129.4, 130.8, 131.62, 131.64, 136.3, 154.1, 159.9.
mp 42–43 °C. υmax/cm−1 3301, 3214, 2969, 2931, 2874, 1602, 1565, 1547, 1491, 1410, 1168, 778. δH(600 MHz; d6-DMSO; Me4Si) 0.96 (6 H, d, J 6.3, (CH3)2CH), 1.17 (6 H, d, J 6.6, (CH3)2CH), 1.41 (6 H, d, J 6.8, (CH3)2CH), 2.96 (6 H, s, CH3), 3.96–4.03 (1 H, m, (CH3)2CH), 4.16–4.24 (1 H, m, (CH3)2CH), 4.68–4.76 (1 H, m, (CH3)2CH), 6.15 (1 H, d, J 7.2, NH). δC (150 MHz, d6-DMSO; Me4Si) 19.4, 22.1, 24.5, 35.5, 42.8, 45.0, 45.3, 148.2, 155.1, 159.2. HRMS-MALDI found: 281.2437; calc. for C14H29N6 (MH+): 281.2448.
mp 178–179 °C (from MeOH). Found: C, 44.15%; H, 7.25%; N, 21.4% C14H29N6O4Cl requires C, 44.15%; H, 7.7%; N, 22.05%. υmax/cm−1 3215, 2969, 1607, 1559, 1453, 1411, 1166, 1140, 1118, 1087, 778, 627. δH(300 MHz; d6-DMSO; Me4Si) 1.24 (12 H, d, J 6.6, (CH3)2CH), 1.49 (6 H, d, J 7.0, (CH3)2CH), 3.12 (6 H, s, CH3), 4.27–4.40 (2 H, m, (CH3)2CH), 4.51–4.64 (1 H, m, (CH3)2CH), 7.13 (2 H, d, J 7.5, NH). δC(75 MHz, d6-DMSO; Me4Si) 19.1, 21.3, 36.1, 44.7, 48.7, 153.5, 159.3.
Footnote |
† Electronic supplementary information (ESI) available: Spectra, X-ray CIF files, tables of bond length, angles and Cartesian coordinates of the optimized structures. CCDC 830994–830998. For ESI and crystallographic data in CIF or other electronic format see DOI: 10.1039/c1nj20595a |
This journal is © The Royal Society of Chemistry and the Centre National de la Recherche Scientifique 2012 |