Introduction

The acid–base concept of organic compounds have been considered from theoretical and experimental aspects [1,2,3,4,5]. The basicity of carbenes [6,7,8,9], silylenes [10,11,12], and germylenes [13,14,15,16] have been studied at theoretical levels. Silylenes mostly prefer to exist as singlets and can interact as a Lewis acid or base. The basicity of silylenes may be triggered by their nucleophilicity or proton affinity. Theoretical studies have been reported on the geometrical and thermodynamic properties of two-, three-, and poly-cyclic hydrocarbons including biphenyl, indene, azulene, naphthalene, anthracene, phenanthrene and pyrene, at various theoretical levels [17,18,19,20,21]. Furthermore, stability, electron and proton affinities, basicity and nucleophilicity have been assessed for heterocyclic [22,23,24,25] and polycyclic hydrocarbons [26, 27] and silylenes [28]. The stability of a conjugated base has a significant effect on the acidity of the corresponding acid. Conjugation, electronegativity, hybridization, and solvation appear as the most prominent contributing factors to the stability of the conjugated base and consequently the higher strength of its acid [29,30,31]. Also, the hard and soft acids and bases (HSAB) are widely used to determine the stability of many compounds [32]. Effects of aromatic rings on stabilization of carbenes are probed, at B3LYP/6-311++G** level. As the extent of conjugation increases, the stability of carbene improves [33,34,35]. In some cases, higher aromaticity may lead to more kinetic stability. For instance, di(triptycyl)carbene appears as a persistent triplet dialkyl carbene [36]. The stability of NHX (X = C, Si, and Ge) type polycyclic structures are also scrutinized, at theoretical levels [37]. Theoretical studies are reported on some cyclohepta-2,4,6-trienylidenes fused with benzene rings, which may be considered analogs of our novel one-, two-, three-, and four-cyclic conjugated silylenes (19, Scheme 1). The singlet carbene analogs of 1, 2, and 4 were found not to be minima for showing two imaginary frequencies [38]. In another survey, during the rearrangement of benzofulvene to naphthalene, the six-membered carbenic analogs of 2–4 were characterized as intermediates, through flash vacuum pyrolysis (FVP) [39]. Thermodynamic parameters of the carbenic analog of 5 has been scrutinized [40]. The most similar report includes the MNDO study on the carbenic analogs of 1–5 that we reported earlier [41]. Several methods are used to calculate the acidity of organic and inorganic compounds in either gas or solution phase [42,43,44,45,46,47,48,49,50]. The gas-phase acidity is mainly influenced by ion-pairing, aggregation effects, and entropic terms. Theoretical pKa values are usually obtained considering thermodynamic cycles which combine gas-phase and liquid-phase calculations, using different solvents and solvation methods, including CPCM [51,52,53,54,55,56]. Following up on our quest for stable group 14, here we have scrutinized thermodynamic and structural parameters of novel one-, two-, three-, and four-cyclic conjugated silylenes (19, Scheme 1). To this end, basicity of 19 are estimated by considering their corresponding protonated structures (1H9H, Scheme 2).

Scheme 1
scheme 1

2,4,6-Cycloheptatrienesilylene (1) with one, two, and three fused benzene rings (29), scrutinized at B3LYP/6-311++G** level

Scheme 2
scheme 2

The protonated structures of 2,4,6-cycloheptatrienesilylene (1H), with one, two, and three fused benzene rings (2H9H), scrutinized at B3LYP/6-311++G** level

Computational methods

All structures are optimized using Gaussian 98 via B3LYP/6-311++G** Pople’s basis set. Harmonic vibrational frequencies (\({\nu_{\min}}\)), ΔEH–L, ΔEs–t, the changes of enthalpies (ΔHs–t), and Gibbs free energies (ΔGs–t) are also calculated at 298.15 K and 1.0 atm. In order to determinate the acidity (pKa) of 19, ΔGs–t are calculated in gas and solution phases using DMSO as solvent. To predict the Gibbs free energy of solvation, the CPCM [57,58,59,60], which is based on the polarized continuum model (PCM), is applied. The nucleophilicity index [61] (N = EHOMO(Nu) − EHOMO(TCNE)) and the electrophilicity index [62] (ω = μ2/2η) are provided where μ is the chemical potential [63] (μ = (EHOMO − ELUMO)/2), and η is the chemical hardness [64] (η = ELUMO − EHOMO). In addition, the gas-phase proton affinities (PA = − ΔHProtonation) [65, 66], ionization potential (IP) [67,68,69], electron affinity (EA) [69,70,71], Mulliken electronegativity (\({\mathcal{X}}\) = (IP + EA)/2), and absolute hardness (ηabs = (IP − EA)/2) are calculated [72,73,74,75]. Furthermore, dipole moments (D) and structural parameters such as bond lengths (Å), divalent and dihedral angles (\({\hat{\text{A}}}\) and \({\hat{\text{D}}}\), respectively), and symmetries are determined. The NICS [76,77,78,79,80] values and the natural bond orbital (NBO) [81,82,83,84,85] charges are computed to give clear evidences of stabilization and basicity.

Results and discussion

To scrutinize the effects of fused benzene ring(s) on 2,4,6-cycloheptatrienesilylene (1), we characterized singlet (s) and triplet (t) states of 1 fused with zero-, one-, two-, and three-conjugated phenyl rings (1, 24, 58, and 9, respectively, Scheme 1 and Table 1). The NBO charges, structural features, and thermodynamic parameters including ΔEs–t, ΔHf, ΔGf, and ΔEH–L, N, ω, μ, and also, NICS (1) data are obtained (Tables 1, 2).

Table 1 The NBO charges, and the structural parameters including bond lengths (Å), divalent angles (C–Si–C, \({\hat{\text{A}}}\), degree), dihedral angles (C–Si–C–C, \({\hat{\text{D}}}\), degree), symmetry, and dipole moment (D, Debye) for singlet silylenes 19, at B3LYP/6-311++G** level
Table 2 The singlet–triplet energy gap (ΔEs–t, kcal/mol), changes of enthalpy (ΔHs–t, kcal/mol) and Gibbs free energy (ΔGs–t, kcal/mol), band gap (ΔEH–L, eV), nucleophilicity (N, eV), electrophilicity (ω, eV), chemical potential (μ, eV), and the nucleus independent chemical shifts (NICS, ppm), calculated at 1 Å above the ring centre, are calculated for one-, two-, three-, and four-cyclic conjugated silylenes (19), at B3LYP/6-311++G** level

Silylenes 26 are planar while 1 and 79 appear non-planar (Table 1). The highest \({\hat{\text{D}}}\) is displayed by 9 (42.22°) that exhibits the lowest dipole moment (D, 1.82 Debye). In addition, 9 carries nearly the highest NBO positive charge on its Si atom (1.003). This appears consistent with its NICS (1) value (− 0.65 ppm, Table 2) and indicates the lower ring current of 9 compared to 18. Also, 7 shows nearly the lowest NBO positive charge on its Si atom (0.774) which is consistent with its highest N (4.01 eV).

The more negative values of ΔEs–t for 19 indicate the higher stability of singlet states compared to their corresponding triplets. The lower absolute value of ΔEs–t reflects the lower stability, and higher chemical activity. The results show that all our silylenes are rather nucleophilic. For instance, 1 and 3 show 3.80 and 3.85 eV for N and 2.71 and 3.62 eV for ω, respectively (Table 2). Structure 7 appears as our most nucleophilic (N = 4.01 eV) silylene. Also, 7 has the lowest ΔEs–t (− 15.71 eV), ΔHs–t (− 15.71 eV), ΔGs–t (− 14.75 eV), and ΔEH–L (− 2.54 eV). In contrast, 9 shows the lowest N (3.55 eV), and the highest ΔEH–L (− 3.32 eV). The overall order of silylenes ΔEs–t, ΔHf, and ΔGf is: 5 > 2 > 9 > 4 > 8 > 6 > 3 > 1 > 7. The overall trend for N shows 7 > 6 > 3 > 1 > 2 > 4 > 5 > 8 > 9. For all structures, we diagnosed the aromaticity by NICS at 1 Å above the center of aromatic system (NICS (1)). The number of benzene rings make a difference in the diatropic current at the center of the aromatic system. All scrutinized structures obey Hückel rule (4n + 2; 1: 6πe, 24: 10πe, 58: 14πe, and 9: 18πe). Evidently, planarity plays a dominant role in aromaticity of 19. Two-cyclic species (24) show higher aromaticity with respect to the remaining structures for being more planar. On the other extreme, 9 is the most non-planar silylene with the least NICS (1) value (− 0.65 ppm). Silylenes with one benzene ring (24) appear more planar and aromatic than those with two benzene rings (58), which turn out more planar and aromatic than the silylene with three benzene rings (9). Hence, the overall trend of NICS (1) values is: 3 > 4 > 2 > 7 > 6 > 5 > 1 > 8 > 9. Evidently, non-planar 1 that has no stabilizing benzo-substituent falls between 5 and 8. The N, PA2 and NICS (1) values show apparent correlation with the corresponding electrostatic potential (ESP) maps (Fig. 1). For instance, silylenes 24 illustrate the higher total electron densities at the centre of the aromatic systems that are consistent with NICS (1) values.

Fig. 1
figure 1

Correlation between N, PA1, NICS (1), and the corresponding ESP map for silylenes (19), at B3LYP/6-311++G** level

To determine the ηabs and \({\mathcal{X}}\), the EA and IP are computed (Table 3). The accuracy of \({\mathcal{X}}\) is dominated by IP which has higher values than EA. The positive and negative electron affinities suggest that the addition of an electron is an endothermic and exothermic processes, respectively. Structures 3 and 7 show the lowest and the highest \({\mathcal{X}}\) values (2.90 and 4.34 eV, respectively). The latter data on 7 appears consistent with its highest value of N (4.01 eV), the lowest ηabs (0.10 eV) and IP (6.76 eV). Therefore, 7 can easily donate its valence electrons and turn out as the most basic silylene. The proton affinities in gas (PA1) and solution (PA2) phases show different values for 19. The PA2 values indicate that silylenes have less tendency to react with proton because the solvation effects [dielectric constant (ε) of DMSO = 47 F/m] [86] that stabilize the initial reactants (H+ and M, M = 19) with respect to that in gas phase. For instance, 5 shows 243.89 and 178.72 kcal/mol for PA1 and PA2, respectively. The overall trend of PA2 appears consistent with N: 7 > 6 > 3 > 1 > 2 > 4 > 5 > 8 > 9.

Table 3 The electron affinity (EA, eV), ionization potential (IP, eV), proton affinities in gas and solution phases (PA1 and PA2, kcal/mol, respectively), absolute chemical hardness (ηabs, eV), Mulliken electronegativity (\({\mathcal{X}}\) eV), and pKa (in DMSO, for 1H9H) are calculated, at B3LYP/6-311++G** level

A thermodynamic cycle is employed to account for deprotonation of HA+ (HA+ = 1H9H) in gas and solution phases (Scheme 3).

Scheme 3
scheme 3

Thermodynamic cycles in gas and solution phases for determining the pKa of protonated silylenes (HA+ = 1H9H) and basicity of A = 19, at B3LYP/6-311++G** level

Our silylenes 19 are represented by A. Their changes of Gibbs free energy for deprotonation in gas, solution, and solvation in DMSO are shown by \(\Delta G^{\circ}_{\text{gas}} , \Delta G^{\circ}_{\text{sol}} ,\;{\text{and}}\;\Delta G^{\circ}_{\text{s}}\), respectively.

The criteria for acidity is given by

$$pK_{\text{a}} = - \log K_{\text{a}} \;{\text{and}}\;\Delta G^{\circ}_{\text{sol}} = - 2.303 \, RT \, \log K_{\text{a}}$$
(1)

where

$$\Delta G^{\circ}_{\text{sol}} = G^{\circ} \left( {A_{\text{sol}} } \right) + G^{\circ} \left( {H^{ + }_{\text{sol}} } \right) - G^{\circ} \left( {HA^{ + }_{\text{sol}} } \right) = G^{\circ} \left( {A_{\text{gas}} } \right) + \Delta G^{\circ}_{\text{s}} \left( A \right) + G^{\circ} \left( {H^{ + }_{\text{gas}} } \right)\quad +\, \Delta G^{\circ}_{\text{s}} \left( {H^{ + } } \right) - G^{\circ} \left( {HA^{ + }_{\text{gas}} } \right) - \Delta G^{\circ}_{\text{s}} \left( {HA^{ + } } \right)$$
(2)

Thermodynamic parameters such as PA and N correlate with pKa values. Relative basicity of 19 can indicate the relative stability of divalent structures with respect to their corresponding protonated forms. In other words, a stronger acid generates a weaker conjugated base [87]. The results show high basicity, PA, and pKa for 19 that are similar to those reported in literature for super bases or proton sponges [88,89,90,91]. The proton sponge definition that was first introduced by Alder et al. has been attributed to the compounds with intramolecular hydrogen bonding after protonation such as 1,8-bis(dimethylamino)naphthalene [92]. Therefore, most of our scrutinized 19 may be considered as super bases or proton sponges for showing PA, intrinsic gas phase basicity (GB), and pKa approximately above 245.3 kcal/mol, 239 kcal/mol, and 7.5 (in DMSO), respectively [93, 94]. To determine the pka values, we calculate the absolute Gibbs free energies in gas and solution phases (\(G^{\circ}_{\text{Gas}}\), \(G^{\circ}_{\text{DMSO}}\), respectively) and \(\Delta G^{\circ}_{\text{S}}\), for 19 and their protonated structures (1H9H, Schemes 1, 2, 3, Tables 4 and 5).

Table 4 Absolute values of Gibbs free energies in gas and solution phases (\(G^{\circ}_{\text{Gas}}\) and \(G^{\circ}_{\text{DMSO}}\), a.u.) and the changes of Gibbs free energies for solvation (\(\Delta G^{\circ}_{\text{S}}\), kcal/mol) are calculated for silylenes (19), at B3LYP/6-311++G** level
Table 5 Absolute values of Gibbs free energies in gas and solution phases (\(G^{\circ}_{\text{Gas}}\) and \(G^{\circ}_{\text{DMSO}}\), a.u.) and the changes of Gibbs free energies for solvation (\(\Delta G^{\circ}_{\text{S}}\), kcal/mol) are calculated for protonated silylenes (1H9H), at B3LYP/6-311++G** level

The \(\Delta G^{\circ}_{\text{S}}\) for 1H9H show higher values than their corresponding 19. This may be attributed to the solvent stabilization of the positive charge on the former. Using Eqs. 1 and 2, we may get rather accurate pka values in DMSO. A higher pka represents lower acidity and indicates higher basicity of 19. Structures 1H9H benefit from aromatization that disperse the positive charge on Si atom. This aromatization and solvent effects, makes 1H9H more stable than 19. The non-planarity caused by the three fused benzene rings in 9 diminishes the possibility of their electron delocalization. The overall trend of pka values with respect to N, and PA2 is: 7 > 6 > 3 > 1 > 2 > 4 > 5 > 8 > 9.

A higher pka may imply a higher tendency for a silylene to react as a nucleophile and/or a base. Structure 7 is the most basic structure among our scrutinized silylenes (pka = 12.19). Evidently, PA2 shows a direct relationship with pKa [95]. An excellent correlation between pKa and PA2 (R2 = 0.9916, Fig. 2) indicates that the basicity of a silylene may enhance as its PA2 increases. Almost all probed silylenes (except 8 and 9) represent super basicity for showing relatively high PA and pKa (approximately 712).

Fig. 2
figure 2

Correlation between pKa and PA2 for silylenes 19, at B3LYP/6-311++G** level

Conclusions

In this study, effects of fused benzene rings on 2,4,6-cycloheptatrienesilylene (1) basicity is probed. To this end, we have scrutinized the novel one-, two-, three-, and four-cyclic conjugated silylenes (19) and determined their basicity using the corresponding protonated structures (1H9H), at B3LYP/6-311++G** level. Most of the silylenes show relatively high basicity that can be fall in the category of super bases and proton sponge compounds. All of the scrutinized structures appear as minima on their energy surfaces for showing no imaginary frequencies. The structural and thermodynamic parameters for 19 are also calculated. The overall order of silylenes ΔEs–t, ΔHf, and ΔGf is: 5 > 2 > 9 > 4 > 8 > 6 > 3 > 1 > 7. To suggest the basicity trend of 19, the CPCM is applied in DMSO as a solvent. The thermodynamic cycles of Gibbs free energies give the relative pka values for protonated silylenes (1H9H). Aromaticity, planarity, N, PA, steric and solvation effects appear as significant factors in our study of basicity. The overall trend of basicity, PA2, and N is: 7 > 6 > 3 > 1 > 2 > 4 > 5 > 8 > 9. Among our scrutinized silylenes, 7 shows the highest basicity, N, PA, \({\mathcal{X}}\); and the lowest ΔEH–L, ΔEs–t, \({\hat{\text{A}}}\), Si–C bond length, and ηabs, while 9 shows the lowest N, PA, NICS (1), dipole moment, pKa and the highest ΔEH–L, and \({\hat{\text{D}}}\). The results suggest reaching at excellent basicity values for 19 silylenes with high pka and PAs values in gas phase and DMSO.