Skip to main content
Heliyon logoLink to Heliyon
. 2019 Oct 18;5(10):e02580. doi: 10.1016/j.heliyon.2019.e02580

Synthesis, crystal structures and electrochemical properties of ferrocenyl imidazole derivatives

Ayomide H Labulo 1, Bernard Omondi 1, Vincent O Nyamori 1,
PMCID: PMC6806399  PMID: 31692585

Abstract

Six ferrocenyl imidazole derivatives substituted with -Cl, -NO2 and -CH3 on the 2-position of the 1H-imidazole ring have been synthesized. Of the six compounds, the di-substituted ferrocenes, i.e. compounds 4 (1,1′-ferrocenylmethyl(2-chloroimidazole)), 5 (1,1′-ferrocenyl(2-nitroimidazole)), and 6 (1,1′-ferrocenylmethyl(2-methylimidazole)) are reported for the first time. The structure-property relationships of compounds 4, 5 and 6 were investigated by means of UV-visible, FTIR, 1H-NMR, 13C-NMR spectroscopy and electrochemical studies. UV-visible analysis in acetonitrile showed that the π -π* band of compounds 2 (1-ferrocenylmethyl(2-nitroimidazole)) and 5 appeared at longer wavelength compared to 1 (1-ferrocenylmethyl(2-chloroimidazole)), 3 (1-ferrocenylmethyl(2-methylimidazole)), 4 and 6. This phenomenon is due to the different electronics around the imidazole moieties. In cyclic voltammetry analysis, all compounds exhibited a quasi-reversible redox wave for the ferrocenyl and imidazole moieties. Density functional theoretical (DFT) calculations with the B3LYP/6-311+G(d) basis set were performed on compounds 16, and the calculated HUMO-LUMO band gap energies correlated with those obtained from electrochemical and spectroscopic data. The X-ray crystallographic analysis highlighted the effect of electron-withdrawing and electron-donating substituents on the conformation of the cyclopentadienyl rings attached to the ferrocenyl moiety.

Keywords: Materials chemistry, Theoretical chemistry, Ferrocenyl imidazole derivatives, UV-vis spectroscopy, Cyclic voltammetry, DFT calculations

1. Introduction

Ferrocenyl derivatives bearing heteroatoms, such as N, P and S, have been of great importance in organometallic chemistry. The synthesis of compounds containing the ferrocenyl unit with various acceptor units such as pyrrole [1], indole [2], thiourea [3], pyridine [4] and imidazole [5] have been previously reported. These particular ferrocenyl derivatives have potential applications in fields such as electrochemistry [6], medicinal science [7], nonlinear optics [8] and materials science [9, 10]. Ferrocenyl alkyl azoles have received considerable attention as a result of the unusual properties of the ferrocene system, and this is due to its unique steric protection, electron-donation properties and chemical stability [11]. The design and fabrication of series of donor-acceptor systems are increasing due to their use in the preparation of optoelectronics [12], liquid crystals [13], polymers [14], sensors [15] and magnetic materials [16]. The photonic and electronic properties of the donor-acceptor system can be fine-tuned by varying the acceptor, donor or spacer strength [17].

Several synthetic approaches have been exploited for the synthesis of ferrocenylalkyl azoles. Mono- and di-substituted ferrocenyl compounds with N-heterocyclic groups, such as ferrocenylmethylbenzimidazole, have been reported [18]. Ferrocenylmethyl benzimidazole has been prepared according to the classical method, i.e. from the reaction of thionyldibenzimidazole and ferrocenylmethanol in dichloromethane [19]. 1-Ferrocenylethyl benzotriazole has also been synthesized by using 1-ferrocenylethanol and benzotriazole in dichloromethane in the presence of aqueous fluoroboric acid at room temperature [20]. An alternative approach for the synthesis of ferrocenylalkyl imidazoles has been proposed with the use of N,N′-carbonyldiimidazole (CDI) [16]. For example, 1,1′-bis(imidazolylmethyl)ferrocenes have been synthesized by the reaction of diferrocenemethanol with CDI [21] or phosphorous tribromide [5]. Simenel et al. [22] reported a simple method for the synthesis of ferrocenyl(alkyl)imidazoles. This method involves the treatment of ferrocenylcarbinols with CDI in methylene dichloride. Additionally, N,N′-diferrocenylimidazol(in)ium salts have been synthesized by condensation of N,N′-diferrocenyldiazabutadiene with formaldehyde [23]. The electrochemical behaviour of a range of ferrocenylalkyl azoles has been studied and the compounds were found to have low reduction potentials [24, 25]. The oxidation potentials (Epa) for substituted ferrocenylalkyl azoles have been found to be markedly dependent on the degree of electron-accepting or electron-donating properties of the lateral substituents [26].

In terms of applications, a number of ferrocenyl imidazole salt derivatives ferrocenyl imidazole salts, have been utilized in various applications. For example, ferrocenylalkyl imidazole salts have been found to play a vital role in many biological systems such as metallo-proteins [27] and anticancer drugs [18]. Ferrocenyl compounds connected to an imidazole ring, such as ferrocenylethyl-N′-(trimethylphenyl)imidazole chloride have been found to show antimicrobial activity [28]. Substituted ferrocenyl imidazoles have also been employed as the catalyst for the synthesis of nitrogen-doped carbon nanotubes (N-CNTs) and N-CNT-metal nanocomposites [11, 29]. These compounds have produced carbon nanomaterials with improved thermal [30, 31], structural [32] and magnetic [33] properties.

In this study, we report the synthesis and characterization of -Cl, -NO2 and -CH3 substituted ferrocenyl imidazole compounds. The synthesis of 1-ferrocenylmethyl(2-chloroimidazole) (1), 1-ferrocenylmethyl(2-nitroimidazole) (2) and 1-ferrocenylmethyl(2-methylimidazole) (3) have been reported in our previous work [29], however, the synthesis of 1,1′-ferrocenylmethyl(2-chloroimidazole) (4), 1,1′-ferrocenyl(2-nitroimidazole) (5) and 1,1′-ferrocenylmethyl(2-methylimidazole) (6) are reported for the first time. The structure-property relationships of compounds 1–6 were investigated. The redox behaviour and electronic transitions of compounds 16 were investigated by both experimental (cyclic voltammetry and UV-visible spectroscopy, respectively) and theoretical (density functional theory (DFT) calculations) studies. The molecular structures of compounds 4 and 5 were determined by X-ray crystallography.

2. Results and discussion

2.1. Synthesis and characterization

Mono- (13) and di-substituted (4–6) ferrocenyl imidazole compounds were synthesized following a literature method, i.e. the reaction of ferrocenemethanol and 1,1′-ferrocenedimethanol, respectively, with 2-substituted (-Cl, -NO2 and -CH3) 1H-imidazoles in the presence of glacial acetic acid [34] (Scheme 1). Compound 4 is an orange solid, while compounds 5 and 6 are high melting compounds obtained as yellow solids. All compounds were obtained in relatively high isolated yields (75–83%), similar to those reported in the synthesis of ferrocenyl alkyl imidazole compounds, obtained from the reaction of ferrocenemethanol [35] and 1,1′-ferrocenedimethanol [5]. From 1H-NMR spectra, characteristic resonances for imidazole protons appeared at 6.77–7.66 ppm and at 124.87–132.43 ppm for C5/C6 in the case of 13C-NMR spectra (Table 1, Figs. 1 and 2). These values are similar to those reported for ferrocenyl(methyl)imidazole [36].

Scheme 1.

Scheme 1

Reaction schemes of (a) ferrocenemethanol and (b) 1,1′-diferrocenemethanol for the synthesis of compounds 1–6.

Table 1.

1H-/13C-NMR characteristic data for the imidazole moiety of compounds 16.

Compound 1H-NMR (ppm)
13C-NMR (ppm)
C3-H C5/C6-H C3 C4 C5/C6
1 4.75 6.78/6.81 81.81 120.65 127.93/131.22
2 5.34 6.92/7.01 80.15 144.47 124.87/128.04
3 4.73 6.76/6.84 82.78 143.94 128.70/130.87
4 4.76 6.85/6.89 82.87 131.22 128.15/130.87
5 5.38 7.15/7.66 81.57 124.97 128.23/128.25
6 4.76 6.77/6.82 83.66 143.92 128.79/132.46

Values taken from ref [29].

Fig. 1.

Fig. 1

1H-NMR of compounds (a) 4 (b) 5 and (c) 6.

Fig. 2.

Fig. 2

13C-NMR of compounds (a) 4 (b) 5 and (c) 6.

The resonance peaks of the unsubstituted ferrocenes and the substituted ferrocenes occur as a singlet between 4.12-4.23 ppm and as a triplet between 4.22-4.27 ppm, respectively [37]. The peak for C3-H was observed to be more downfield, i.e. 5.34 ppm for 2 and 5.38 ppm for 5, and this due to the electron-withdrawing inductive effect from the -NO2 substituent (Fig. 1 and Scheme 1).

The appearance of a FTIR absorption band between 1248-1360 cm−1 was observed in all compounds. This peak marked the appearance of the C-N absorption band while the O-H bands of the starting ferrrocenyl alcohols disappeared. The most important frequencies for the ferrocene moiety are found at 450–481 (νFe–Cp), 821–831 (πC–H), 1440 cm−1C–C) and 3382-3113 cm−1C–H) [37].

2.2. Crystal structures

Crystals of compounds 4 and 5 were grown in a dichloromethane-hexane solution. ORTEP drawings of these compounds together with the atom numbering schemes, are shown in Fig. 3. Selected bond angles and bond lengths for compounds 4 and 5 are listed in Table 2. The asymmetric unit of compound 4 contains one molecule while for compound 5 it contains half a molecule with the other half generated by the inversion centre on the Fe atom. The Cp rings in compound 5 adopt a perfect anti-conformation with a staggering angle of 35.82(4)°. In compound 4, the Cp rings of the ferrocene units are slightly twisted from the ideal eclipsed conformation with C-centroid-centroid-C torsion angles of 11.46(8)°. The Cp rings in compound 4 were slightly tilted by an average angle of 2.25(3)°.

Fig. 3.

Fig. 3

ORTEP diagrams, along with the atom-numbering scheme for compounds (a) 4 and (b) 5. The thermal ellipsoids are depicted at 50% probability.

Table 2.

Selected bond lengths (Å) and angles (°) for compounds 4 and 5.

4 5
C(1)-C(11) 1.503(2) C(1)-C(6) 1.496(2)
C(11)-N(1) 1.469(8) C(6)-N(1) 1.487(8)
N(2)-C(13)-N(1) 114.08(13) N(1)-C(7)-N(2) 121.76(14)
N(1)-C(11)-C(1) 112.28(12) C(7)-N(1)-C(6) 132.00(13)
C(11)-C(1)-C(6)-C(15) 79.42(1) C(6)-C(1)-C(1)-C(6) 180
C(4)-C(5)-C(1) 108.17(13) C(1)-C(2)-C(3) 108.13(14)
C(2)-Fe(1) 2.045(14) C(2)-Fe(1) 2.042(6)
C(5)-Fe(1) 2.034(14) C(5)-Fe(1) 2.036(15)

The N-C-N imidazole angles are 121.25° (N1-C13-N2) and 113.96° (N1-C7-N2) in compounds 4 and 5, respectively; and are larger than the values for imidazole salts (107.6–108.7°) reported by Gao et al. [22]. This increase in bond angle could be attributed to the electron-withdrawing effect of the -Cl and -NO2 substituents on the imidazole rings [38, 39]. In compound 4, the dihedral angles between the planes of the substituted Cp rings and the imidazole ring plane are 65.58(5)°and 67.28(2)°, respectively, while in compound 5, the dihedral angles are equal with a value of 71.46°(4). This slight change in the dihedral angle is probably due to the steric effect of the substituents on the imidazole moiety and the Cp rings [40]. The crystallographic data and structural refinement parameters for compounds 4 and 5 are shown in Table 3.

Table 3.

Crystallographic data and structural refinement parameters for compounds 4 and 5.

4 5
Empirical formula C18H16Cl2FeN4 C18H16FeN6O4
Formula weight 411.07 436.22
T(K) 173(2) 100(2)
λ (Å) 0.71073 0.71073
Crystal system Monoclinic Monoclinic
Space group P21/n C2/c
Unit cell dim. (Å/°)
β 102.9480(10)° 102.539(2)°
a 8.5123(3) 10.9972(3)
b 15.7406(6) 7.5371(2)
c, 13.1687(5) 21.7869(7)
V(Å3) 1719.59(11) 1762.78(9)
Z 4 4
σ (Mg/m3) 1.588 1.644
μ(mm−1) 1.195 0.897
F (000) 832 896
Crystal size (mm3) 0.326 × 0.319 × 0.272 0.220 × 0.160 × 0.080
θ range (°) 2.047 →27.461 1.915 → 28.321
Index ranges -10 ≤ h ≤ 10
-20 ≤ k ≤ 9
-16 ≤ l ≤ 16
-11 ≤ h ≤ 14
-9 ≤ k ≤ 9
-29 ≤ l ≤ 26
Reflections collected 22847 8361
Independent reflections 3288 2173
Completeness to θ %/(°) s 99.9 (25.242)
Data/restrs/para 3288/0/226 2173/0/133
GOOF on F2 1.073 1.035
Final R indices, R1, wR2 [I > 2sigma(I)] 0.0289, 0.0811 0.0299, 0.0734
R indices (all data), R1, wR2 0.0305, 0.0824 0.0400, 0.0793
Largest diff. peak & hole 0.946 and -0.288 e.Å−3 0.502 and -0.314 e.Å−3

2.3. UV-visible absorption spectroscopic studies

The absorption spectra for compounds 16 are shown in Fig. 4 and the relevant absorption characteristics are given in Table 4. The spectra obtained in acetonitrile show bands at approximately 328 and 438 nm. The bands at 328 nm represent π-π* transitions from the ferrocenyl to the imidazole moiety, while the bands at 438 nm are ascribed to d-d transitions, probably mixed with d-π* transitions [41]. There is a shift of the π-π* band of compounds 2 and 5, at 328 nm bathochromically to 331 nm, attributed to the stronger electron-withdrawing effect of the nitro group on the imidazole moiety [42, 43]. This shift could also be attributed to the strong electronic communication between the enriched imidazole and ferrocenyl moieties [44].

Fig. 4.

Fig. 4

UV-visible spectra of 0.005M solutions of 1-6 in acetonitrile.

Table 4.

UV-visible absorption characteristics of compounds 16.

Compounds λmax/nm (molar absorptivity/mol−1 dm3 cm−1)
d-d π-π*
1 438 (174) 328 (89)
2 none 331 (236)
3 438 (138) 328 (142)
4 433 (173) 328 (141)
5 none 331 (241)
6 437 (172) 328 (249)

In order to establish the solvatochromic behaviour of compounds 1, 3, 4 and 6, absorption studies were carried out in different solvents, that is, dimethylformamide (DMF) (polar aprotic), methanol (polar protic), acetonitrile (polar aprotic) and dichloromethane (DCM) (mid-polar aprotic). In all compounds, two bands (d-d transition and charge transfer (CT)) are observed. The data obtained are summarized in Table 5. CT transitions usually display a large solvent-dependent effect [45], resulting in negative solvatochromism as observed in methanol and DMF when compared with DCM (Fig. 5). This could be attributed to possible hydrogen-bonding between the compounds and the polar solvent molecules. It was found that the CT absorption band in compounds 1, 3, 4 and 6 is the same in all solvents; however, a red shift in the π-π* absorption band was observed upon a change in solvent polarity with compound 1 (showing the most significant shift from λmax = ∼320 nm methanol/DMF to DCM) (Fig. 5). Thus, the electron-withdrawing and electron-donating substituents influenced the dipole-dipole interactions between the compounds and solvents [46]. Similar behaviour has been reported for ferrocenyl imidazole derivatives [47].

Table 5.

Absorption wavelengths of ferrocenyl imidazole derivatives in DCM, methanol, DMF and acetonitrile.

Compounds λmax/nm
DCM Methanol DMF Acetonitrile
1 none 438 440 438
343 321 320 321
2 none none none none
328 327 327 331
3 436 429 436 428
334 320 322 321
4 446 434 442 427
331 326 328 328
5 none none none none
329 325 327 331
6 430 427 438 438
332 326 327 328

Fig. 5.

Fig. 5

UV-visible spectra of 0.005 M solutions of compound 1 in DMF, acetonitrile, methanol and DCM.

2.4. Electrochemical studies

Electrochemical studies of compounds 16 were performed in acetonitrile, with NaClO4 (0.5 M) as the supporting electrolyte, at a scan rate of 100 mV/s. Fig. 6 shows the cyclic voltammograms (CV), while the corresponding electrochemical data are presented in Table 6. The CV of the substituted ferrocenyl imidazole compounds exhibit the electrochemical oxidation wave of the ferrocene/ferrocenium ion, and this is coupled with a quasi-reversible peak. Compounds 16 show one fully or partially reversible one-electron oxidation and one or two quasi-reversible reductions for the imidazole acceptor moieties. The ferrocenyl moieties in compounds 16 show larger Epa values than free ferrocene. The Epa values follow the order: 5 > 4 > 2 > 3 > 1 > 6, which can be related to the effect of the substituents on the redox potential of the compounds and, subsequently, the electron transfer process. Compounds 1–6 gave ipa/ipc ratios of approximately 1. As expected, the introduction of electron-withdrawing -Cl and -NO2 substituents led to a more positive shift in the redox potential [48].

Fig. 6.

Fig. 6

Cyclic voltammogram of compounds (a) 13 and (b) 46 in 0.5 M NaClO4/acetonitrile at a platinum working electrode.

Table 6.

Cyclic voltammetry data for compounds 16.

Compound Epa (mV) Epc (mV) ΔEp (mV) E1/2 ​ (mV) ipa/ipc
Ferrocene [52] 470 413 57 442 1.170
1 446 373 73 410 1.129
2 456 385 71 421 1.279
3 452 376 76 414 1.168
4 546 475 71 511 1.111
5 560 484 76 522 1.014
6 426 353 73 390 1.075

Epa, anodic oxidation potential; Epc, cathodic reduction potential;  ​E1/2 ​, redox peak potential; (E1/2 ​ = (Epa + Epc)/2); ΔEp = Epa - Epc.

The CV curves of compounds 2 and 5 showed poorly resolved ferrocene-based waves (Fig. 6a and b). Probably the use of a coordinating solvent (i.e. acetonitrile) masked irregular features of the CV curves for compounds 2 and 5. The two observed waves for compounds 1, 3, 4 and 6 are associated with the typical symmetrical dimers which produce a mixed valent transition upon reduction or oxidation. Reductions potentials on the side groups of symmetrical complexes that generate mixed valent intermediates have been reported [49]. These systems either allow electron delocalization from direct metal-metal interaction or through the bridge path. The mixed valent intermediate group of the ferrocenyl and the ferrocenium group could be influenced by the electronegativities of the attached group.

E1/2 of mono-substituted derivatives, compounds 13, is between 410-421 mV, while the di-substituted derivatives, compounds 46, displayed E1/2 ​ values in the range of 390–522 mV. Compound 5 has the highest E1/2 ​ values, thus higher reduction propensity, while compound 6 gave the lowest E1/2 ​value indicating less favourable reduction of the Fe centre. From our results, compounds 2, 4 and 5 show relatively high E1/2 ​ values (421, 511 and 522 mV, respectively), corresponding to the oxidation of the ferrocenyl moiety, attributable to the electron-withdrawing effect of the -Cl and -NO2 substituents on the imidazole moiety, compared with compounds 1, 3 and 6 [5,50,51]. The Epa values for mono- and di-substituted ferrocenyl derivatives were shifted to more positive potentials (426–560 mV) compared with that of unsubstituted ferrocene (413 mV) (Table 6).

2.5. Theoretical calculation

To obtain a better insight into the electronic nature and electrochemical properties of compounds 16, DFT calculations were carried out with the B3LYP functional and M06-L with def2-TZVP. The optimized geometries of compounds 4 and 5 were in good agreement with the experimental ones obtained from single crystal XRD analysis (Fig. 3).

The molecular orbitals considered were the second highest (HOMO-1) and highest (HOMO) occupied once and the lowest (LUMO) and second-lowest (LUMO+1) unoccupied ones. Fig. 7 shows the frontier molecular orbitals (HOMO and LUMO) of compound 5. Generally, the systems are characterized by relatively low energies for LUMO and LUMO+1 (Table 7). It is evident from Fig. 7 that, in compound 5, the LUMO resides in the imidazole moiety while the HOMO is intense on the ferrocene ring. The high energies obtained for HOMO and HOMO-1 reflect the ability of electron donation [53]. Compound 3 has the highest HOMO and HOMO-1 values (-0.171 eV and -0.172 eV, respectively, Table 7), and this is due to the effect of the electron-donating group on the imidazole moiety, which increases the orbital energies of the ferrocenyl receptor [54]. These results indicate that more energy is involved during the electron transition in the HOMO relative to LUMO. The difference between ΕHOMO and ELUMO yields the band gap energy (ΔE), which is an important stability index [55]. A smaller band gap is a measure of greater prospect for transfer of electrons within a given system [56]. Compounds 2 and 5 gave the smallest band gap energies of 0.073 and 0.085 eV, respectively (Table 7). The HOMO–LUMO energy gaps of compounds 16 was found to approximately agree with the band gap energies calculated from the electrochemical (Table 7). As illustrated in Fig. 7, excitation from lower to lowest unoccupied orbital reflects a localized density on the functional groups while the electrons are highly occupied in the parent (ferrocene) moiety. The results obtained in Table 7 indicated that the ΔE corresponds to CT transitions in all the compounds.

Fig. 7.

Fig. 7

Lobes of the frontier molecular orbitals for the symmetric compound 5 obtained from the M1 theoretical model.

Table 7.

Frontier molecular orbital energies and quantum chemical descriptors of compounds 16 obtained from M1 level of theory.

Compound LUMO+1 (eV) LUMO (eV) HOMO (eV) HOMO-1 (eV) ΔE (eV) ΔE (CV) (eV)
1 -0.045 -0.046 -0.172 -0.173 0.127 0.073
2 -0.054 -0.107 -0.180 -0.181 0.073 0.081
3 -0.044 -0.045 -0.171 -0.172 0.126 0.070
4 -0.058 -0.058 -0.184 -0.185 0.126 0.121
5 -0.115 -0.115 -0.200 -0.200 0.085 0.070
6 -0.052 -0.053 -0.178 -0.179 0.126 0.160

ΔE - calculated band gap energy.

3. Conclusion

Novel ferrocene derivatives, namely of compounds 46, were synthesized and characterized. The single crystal X-ray crystallographic analysis of compounds 4 and 5 showed the preferred solid-state conformation and symmetry of the di-substituted ferrocenyl moiety. The electrochemical behaviour of the compounds was studied by cyclic voltammetry which shows one fully or partially reversible one-electron oxidation and one or two quasi-reversible reductions for the imidazole acceptor moieties with band gaps similar to those obtained from DFT calculations. The electron transfer process and redox behaviour of compounds 16 are affected by the electron-donating and withdrawing substituent on the imidazole moiety. Thus, the optical and electronic properties of the synthesized compounds can be fine-tuned by changing the substituent groups on the imidazole ring.

4. Experimental

2-Nitro-1H-imidazole (≥98.1%), 2-chloro-1H-imidazole (≥98.2%), 2-methyl-1H-imidazole (>98%), ferrocenemethanol (98%) and 1,1′-ferrocenedimethanol (98%) were supplied by Sigma Aldrich. Sodium borohydride (95%) and glacial acetic acid (99.8%) were supplied by Merck chemicals. All other solvents used were of analytical grade and were used as supplied by Sigma Aldrich, unless stated otherwise. TLC with varying solvent polarity was performed on silica gel 60 F254 plates. Silica gel with 0.063–0.2 mm mesh size was used for column chromatography to purify the products obtained. The melting point of each compound was determined using an open-end capillary tube in a Bibby Stuart Scientific model SMP3 apparatus. Infrared spectra were obtained with a PerkinElmer Spectrum 100 FTIR spectrometer fitted with a universal ATR accessory. 1H-NMR and 13C-NMR spectra were recorded on a 400 MHz Bruker AVANCE III spectrometer at room temperature using deuterated CDCl3 or DMSO as solvents. Mass spectra of the synthesized compounds were obtained from an Agilent Technologies, 1100 series, ion-trap mass spectrometer. Electronic spectra were recorded on a PerkinElmer Lambda 35 double-beam UV–visible spectrophotometer with 10 mm path length quartz cuvettes in acetonitrile, dichloromethane, and DMF. Cyclic voltammetry measurements were performed with a CHI660E electrochemical workstation (CH Instruments, USA) with a scan rate of 100 mV/s−1 in a three-electrode system consisting of a GCE (3.0 mm diameter) auxiliary electrode, Ag/AgCl reference electrode and a platinum-wire as the working electrode. The half-wave potential (E1/2)values were obtained using (Epa + Epc)/2, where Epa and Epc are the anodic and cathodic peak potentials, respectively. The band gap energies were estimated from the empirical relation ELUMO = [(Ered - E1/2(ferrocene)) + 4.8] eV or EHOMO = [(Eox - E1/2(ferrocene)) + 4.4] eV. Ferrocene was used as an external standard. It shows two peaks at 0.38 and 0.45 V hence the E1/2 (ferrocene) is equal to 0.42 V. The longest absorption wavelength λonset was used to calculate the optical gap energy, ΔE (Optical), according to the equation Eg = 1242/λonset.

4.1. General procedure for the synthesis of compounds 4–6

4.1.1. Representative procedure for 4–6

The prepared 1,1′-ferrocenedimethanol (240.9 mg, 1 mM) and 2-chloro-1H-imidazole (262.3 mg, 2.32 mM) in acetic acid (3 mL) were stirred for 6 hours at 60 °C. Product formation was monitored by using TLC plates with ethyl acetate/hexane (2:1) as eluent. The reaction mixture was extracted with ethyl acetate (50 mL), washed with 15 mL brine solution and dried over anhydrous sodium sulphate. Thereafter, the solvent was removed under reduced pressure to obtain the crude product which was purified by flash column chromatography on silica gel to give a yellow solid.

4.1.2. Synthesis of 1,1′-ferrocenylmethyl(2-chloroimidazole) (4)

The general procedure described in Section 4.1.1 was followed by using 1,1′-ferrocenedimethanol (246.0 mg, 1 mM) and 2-chloro-1H-imidazole (237.8 mg, 2.20 mM). The product was purified by column chromatography and concentrated under vacuum to obtain yellow crystals as the product. Yield: 60% (245.0 mg), Rf = 3.0, m.p = 170–174 °C. IR (cm−1): 439.01, 736.00, 820.00, 1036.00, 1334.37, 1465.70, 1702.04, 2923.79, 3100.43. 1H-NMR spectra (400 MHz, CDCl3, δ ppm): δ = 4.17–4.22 (8H, t, 2 C5H4), 4.76 (4H, s, 2 (CH2)), 6.85 (2H, s, imidazole CH=CH), 6.89 (2H, s, imidazole CH=CH) ppm. 13C-NMR spectra (400 MHz, CDCl3, δ ppm): δ = 14.11, 22.67, 29.64, 30.92, 45.87, 69.53, 69.71, 76.74, 77.06, 77.38, 82.63, 120.46, 128.15, 131.22, 206.98. HR-MS (C18H16Cl2FeN4). ES: [M]+ m/z calc. 414.00910, found [M + Na]+ 436.9902.

4.1.3. Synthesis of 1,1′-ferrocenyl(2-nitroimidazole) (5)

The general procedure described in Section 4.1.1 was followed by using 1,1′-ferrocenedimethanol (246.0 mg, 1 mM) and 2-nitro-1H-imidazole (262.3 mg, 2.32 mM). The product was purified by column chromatography and concentrated under vacuum to obtain orange crystals. Yield: 56% (242.2 mg), Rf = 0.59, m.p = 245 °C. IR (cm−1): 437.66, 831.62, 916.65, 1023.23, 1143.93, 1237.14, 1270.40, 1353.90, 1479.86, 1526.93, 1661.18, 2867.87, 3116.37, 3385.49. 1H-NMR spectra (400 MHz, CDCl3, δ ppm): δ = 4.23–4.25 (8H, t, 2 C5H4), 5.39 (4H, s, 2(CH2)), 7.15 (2H, s, imidazole CH=CH), 7.66 (2H, s, imidazole CH=CH) ppm. 13C-NMR spectra (400 MHz, CDCl3, δ ppm): δ = 30.89, 29.62, 39.83, 40.03, 49.28, 49.66, 60.22, 68.77, 69.86,70.26, 77.08, 77.33, 80.77, 81.57, 124.97, 128.02, 207.32. HR-MS (C18H16FeN6O4). ES: [M]+ m/z calc. 436.05800, found [M + Na]+ 459.04801.

4.1.4. Synthesis of 1,1′-ferrocenylmethyl(2-methylimidazole) (6)

The general procedure described in Section 4.1.1 was followed by using 1,1′-ferrocenedimethanol (246.0mg, 1 mM) and 2-methyl-1H-imidazole (190.5 mg, 2.32 mM). The product was purified by column chromatography and concentrated under vacuum to obtain orange solid. Yield: 64% (239.0 mg), Rf = 2.44, m.p = 258 °C; IR (cm−1): 487.66, 691.80, 831.62, 916.65, 1023.23, 1113.93, 1237.14, 1270.40, 1479.86, 1526.93, 1661.18, 2867.87, 3116.37, 3385.49. 1H-NMR spectra (400 MHz, CDCl3) 2.39 (6H, s, 2(CH3)), 4.14–4.23 (8H, t, 2 C5H4), 4.76 (4H, s, 2(CH2)), 6.77 (2H, s, imidazole), 6.83 (2H, s, imidazole CH=CH). 13C-NMR spectra (400 MHz, CDCl3): δ (ppm) = 10.96, 13.07, 14.05, 23.74, 28.92, 29.35, 30.35, 38.72, 45.42, 60.15, 61.58, 68.62, 69.04, 70.26, 72.66, 76.76, 77.07, 77.28, 89.00, 119.00, 126.32, 128.79, 130.89, 132.43, 143.93, 167.77. HR-MS (C20H22FeN4). ES: [M]+ m/z calc. 374.30012, found 374.30215.

4.2. X-ray crystal structures

Crystals of compounds 4 and 5 were attached onto the tip of a glass fibre with epoxy glue and centered in the X-ray beam under a video camera. Crystal evaluation and data were done and collected on a Bruker Smart APEX2 diffractometer with Mo Kα radiation (λ = 0.71073 Å) equipped with an Oxford Cryostream low-temperature apparatus. Omega scans of 0.5° width with an exposure time of 20 s per frame were used for data collection. The total number of images was based on results from the program COSMO [57] with the completeness of 100% out to 0.75 Å and expected redundancy was expected to be 4.0. The initial cell matrix was determined from three series of scans collected at intervals of 0.5° in a 6° range, with each of the series of scans collected at different starting angles and the APEXII [58] program suite employed for reflection indexing. Cell parameters were retrieved using APEXII [58] and refined by using SAINT [59] on all observed reflections. The structures were solved by the direct method using the SHELXS [60] program and refined. The visual crystal structures were presented using ORTEP-3 [61] and MERCURY [62]. Non-hydrogen atoms were first refined with anisotropic displacement parameter and then by anisotropic refinement with full-matrix least squares.

Ten different DFT models (B3LYP, [63, 64] B3LYP-D2, B3LYP-D3, [65] CAM-B3LYP, [66] BMK, M05-2X, M06, M06-2X, M06-HF and M06-L [67, 68]) with def2-TZVP [69] and 6-311+G(d) [70, 71] basis sets were examined. M06-L [68] plus either basis sets proved to be sufficient and more accurate than the rest of the DFT models studied. Calculations were executed within the Gaussian 09 RevD.01 software [72] and pre/post molecular analyses were carried out in GaussView [73].

4.2.1. Crystallographic data

Crystallographic data (excluding structure factors) have been deposited with the Cambridge Crystallographic Data Centre and can be obtained free of charge via www.ccdc.cam.ac.uk/data_request/cif. The supplementary crystallographic data of compounds 4 and 5 are contained in CCDC 1854200 and 1854185 publication number, respectively.

Declarations

Author contribution statement

Ayomide H. Labulo: Conceived and designed the experiments; Performed the experiments; Analyzed and interpreted the data; Wrote the paper.

Vincent O.Nyamori: Conceived and designed the experiments; Analyzed and interpreted the data; Contributed reagents, materials, analysis tools or data; Wrote the paper.

Bernard Omandi: Performed the experiments; Analyzed and interpreted the data; Contributed reagents, materials, analysis tools or data; Wrote the paper.

Funding statement

This work was supported by University of KwaZulu-Natal (UKZN), UKZN Nanotechnology Platform, National Research Foundation (NRF) SA, and Eskom Tertiary Education Support Programme (TESP) SA.

Competing interest statement

The authors declare no conflict of interest.

Additional information

No additional information is available for this paper.

References

  • 1.Yang W., Yin Z., Wang C.-H., Huang C., He J., Zhu X., Cheng J.-P. New redox anion receptors based on calix [4] pyrrole bearing ferrocene amide. Tetrahedron. 2008;64:9244–9252. [Google Scholar]
  • 2.Takase M., Inouye M. Ferrocene-modified bis (spiropyridopyran) s as synthetic signaling receptors for guanine–guanine dinucleoside derivatives. Chem. Commun. 2001;23:2432–2433. doi: 10.1039/b107306k. [DOI] [PubMed] [Google Scholar]
  • 3.Lorenzo Á., Aller E., Molina P. Iminophosphorane-based synthesis of multinuclear ferrocenyl urea, thiourea and guanidine derivatives and exploration of their anion sensing properties. Tetrahedron. 2009;65:1397–1401. [Google Scholar]
  • 4.Quintal S., Morais T.S., Matos C.P., Robalo M.P., Piedade M.F.M., de Brito M.J.V., Garcia M.H., Marques M., Maia C., Campino L. Synthesis, structural characterization and leishmanicidal activity evaluation of ferrocenyl N-heterocyclic compounds. J. Organomet. Chem. 2013;745:299–311. [Google Scholar]
  • 5.Niu H.-T., Yin Z., Su D., Niu D., Ao Y., He J., Cheng J.-P. Ferrocene-based imidazolium receptors for anions. Tetrahedron. 2008;64:6300–6306. [Google Scholar]
  • 6.Krejčík M., Daněk M., Hartl F. Simple construction of an infrared optically transparent thin-layer electrochemical cell: Applications to the redox reactions of ferrocene, Mn2(CO)10 and Mn(CO)3(3, 5-di-t-butyl-catecholate)−. J. Electroanal. Chem. Interfacial Electrochem. 1991;317:179–187. [Google Scholar]
  • 7.Biot C., Glorian G., Maciejewski L.A., Brocard J.S., Domarle O., Blampain G., Millet P., Georges A.J., Abessolo H., Dive D. Synthesis and antimalarial activity in vitro and in vivo of a new ferrocene− chloroquine analogue. J. Med. Chem. 1997;40:3715–3718. doi: 10.1021/jm970401y. [DOI] [PubMed] [Google Scholar]
  • 8.Rochford J., Rooney A.D., Pryce M.T. Redox control of meso-zinc (II) ferrocenylporphyrin based fluorescence switches. Inorg. Chem. 2007;46:7247–7249. doi: 10.1021/ic0703326. [DOI] [PubMed] [Google Scholar]
  • 9.Wang Z., Lu P., Chen S., Gao Z., Shen F., Zhang W., Xu Y., Kwok H.S., Ma Y. Phenanthro [9, 10-d] imidazole as a new building block for blue light emitting materials. J. Mater. Chem. 2011;21:5451–5456. [Google Scholar]
  • 10.Mortazavi-Derazkola S., Salavati-Niasari M., Amiri O., Abbasi A. Fabrication and characterization of Fe3O4@ SiO2@ TiO2@ Ho nanostructures as a novel and highly efficient photocatalyst for degradation of organic pollution. J. Energy Chem. 2017;26:17–23. [Google Scholar]
  • 11.Ombaka L.M., Ndungu P.G., Nyamori V.O. Tuning the nitrogen content and surface properties of nitrogen-doped carbon nanotubes synthesized using a nitrogen-containing ferrocenyl derivative and ethylbenzoate. J. Mater. Sci. 2015;50:1187–1200. [Google Scholar]
  • 12.Patil Y., Misra R. Tetracyanobutadiene bridged ferrocene and triphenylamine functionalized pyrazabole dimers. J. Organomet. Chem. 2017;840:23–29. [Google Scholar]
  • 13.Cheng H., Ma C., Chen Y., Ni H.L., Feng C., Wang B.Q., Zhao K.Q., Yu W.H., Hu P. Monosubstituted ferrocene liquid crystals containing click triazole with a wide nematic phase temperature range. Liquid Crystals. 2017;44:1450–1461. [Google Scholar]
  • 14.Manners I. John Wiley & Sons; 2006. Synthetic Metal-Containing Polymers. [Google Scholar]
  • 15.Rosa V., Gaspari A.P., Folgosa F., Cordas C.M., Tavares P., Santos-Silva T., Barroso S., Avilés T. Imine ligands based on ferrocene: synthesis, structural and Mössbauer characterization and evaluation as chromogenic and electrochemical sensors for Hg2+ New J. Chem. 2018;42:3334–3343. [Google Scholar]
  • 16.Zhang X., Chen L., Yun J., Kong J. Novel ferrocene-containing organosilicon polymers and uniform microspheres prepared by free radical copolymerization: Precursors for magnetic Si-C-Fe-(O) Mater. Design. 2018;144:86–97. [Google Scholar]
  • 17.He J., Guo F., Li X., Wu W., Yang J., Hua J. New bithiazole-based sensitizers for efficient and stable dye-sensitized solar cells. Chem. A Eur. J. 2012;18:7903–7915. doi: 10.1002/chem.201103702. [DOI] [PubMed] [Google Scholar]
  • 18.Lubov’V S., Simenel A.A., Nekrasov Y.S., Morozova E.A., Starikova Z.A., Peregudova S.M., Kuzmenko Y.V., Babin V.N., Ostrovskaya L.A., Bluchterova N.V. Synthesis, structure and redox potentials of biologically active ferrocenylalkyl azoles. J. Organomet. Chem. 2004;689:2473–2479. [Google Scholar]
  • 19.Boev V., Betankourt P., Popova L., Babin V. Alpha-ferrocenylalkylation of polyfunctional nucleophilic compounds. J. Gen. Chem USSR. 1991;61:1516–1526. [Google Scholar]
  • 20.Lubov’V S., Boev V.I., Nekrasov Y.S., Ilyin M.M., Davankov V.A., Starikova Z.A., Yanovsky A.I., Kolomiets A.F., Babin V.N. Synthesis and structure of biologically active ferrocenylalkyl polyfluoro benzimidazoles. J. Organomet. Chem. 1999;580:26–35. [Google Scholar]
  • 21.Gao Y., Twamley B., Shreeve J.N.M. The first (ferrocenylmethyl) imidazolium and (ferrocenylmethyl) triazolium room temperature ionic liquids, Inorg. Chem. 2004;43:3406–3412. doi: 10.1021/ic049961v. [DOI] [PubMed] [Google Scholar]
  • 22.Simenel A.A., Morozova E.A., Kuzmenko Y.V., Lubov'V S. Simple route to ferrocenyl (alkyl) imidazoles. J. Organomet. Chem. 2003;665:13–14. [Google Scholar]
  • 23.Bildstein B., Malaun M., Kopacka H., Wurst K., Mitterböck M., Ongania K.-H., Opromolla G., Zanello P. N,N‘-Diferrocenyl-N-heterocyclic carbenes and their derivatives. Organometallics. 1999;18:4325–4336. [Google Scholar]
  • 24.Wilkinson G., Stone F.G.A., Abel E.W. Vol. 4. Pergamon Press; 1982. (Comprehensive Organomet. Chem.). [Google Scholar]
  • 25.Alfonso M., Espinosa Ferao A., Tárraga A., Molina P. Electrochemical and fluorescent ferrocene-imidazole-based dyads as ion-pair receptors for divalent metal cations and oxoanions. Inorg. Chem. 2015;54:7461–7473. doi: 10.1021/acs.inorgchem.5b01071. [DOI] [PubMed] [Google Scholar]
  • 26.Alfonso M., Tárraga A., Molina P. Preparation and sensing properties of a nitrogen-rich ferrocene-imidazole-quinoxaline triad decorated with pyrrole rings. Dalton Trans. 2016;45:19269–19276. doi: 10.1039/c6dt03667h. [DOI] [PubMed] [Google Scholar]
  • 27.Patra M., Gasser G., Metzler-Nolte N. Small organometallic compounds as antibacterial agents. Dalton Trans. 2012;41:6350–6358. doi: 10.1039/c2dt12460b. [DOI] [PubMed] [Google Scholar]
  • 28.Çoban E.P., Fırıncı R., Biyik H., Günay M.E. Unsymmetrically substituted imidazolium salts: Synthesis, characterization and antimicrobial activity. Brazilian J. Pharm. Sci. 2017:53. [Google Scholar]
  • 29.Labulo A.H., Ngidi N.P., Omondi B., Nyamori V.O. Physicochemical properties of nitrogen-doped carbon nanotubes from metallocenes and ferrocenyl imidazolium compounds. J. Organomet. Chem. 2018;868:66–75. [Google Scholar]
  • 30.Salavati-Niasari M., Behfard Z., Amiri O., Khosravifard E., Hosseinpour-Mashkani S.M. Hydrothermal synthesis of bismuth sulfide (Bi2S3) nanorods: Bismuth (III) monosalicylate precursor in the presence of thioglycolic acid. J. Cluster Sci. 2013;24:349–363. [Google Scholar]
  • 31.Fleming E., Du F., Ou E., Dai L., Shi L. Thermal conductivity of carbon nanotubes grown by catalyst-free chemical vapor deposition in nanopores. Carbon. 2019;145:195–200. [Google Scholar]
  • 32.Wei Y., Yang Z., Zhao G., Yang W. Synthesis and catalytic properties of macroporous SiO2-coated CNT-sieve-composite-supported 12-tungstophosphoric acid catalysts with dual pore structure for the Baeyer–Villiger oxidation of cyclic ketones under microwave irradiation. J. Catal. 2019;371:196–206. [Google Scholar]
  • 33.Nasir A.N., Yahaya N., Zain N.N., Lim V., Kamaruzaman S., Saad B., Nishiyama N., Yoshida N., Hirota Y. Thiol-functionalized magnetic carbon nanotubes for magnetic micro-solid phase extraction of sulfonamide antibiotics from milks and commercial chicken meat products. Food Chem. 2019;276:458–466. doi: 10.1016/j.foodchem.2018.10.044. [DOI] [PubMed] [Google Scholar]
  • 34.He P., Du Y., Wang S., Cao C., Wang X., Pang G., Shi Y. Synthesis, structure, and reactivity of ferrocenyl-NHC palladium complexes. Z. Anorg. Allg. Chem. 2013;639:1004–1010. [Google Scholar]
  • 35.Nyamori V.O., Gumede M., Bala M.D. Synthesis, characterisation and properties of ferrocenylalkylimidazolium salts. J. Organomet. Chem. 2010;695:1126–1132. [Google Scholar]
  • 36.Taylor A.W., Licence P. X-ray photoelectron spectroscopy of ferrocenyl-and ferrocenium-based ionic liquids. ChemPhysChem. 2012;13:1917–1926. doi: 10.1002/cphc.201100829. [DOI] [PubMed] [Google Scholar]
  • 37.Kübler P., Sundermeyer J. Ferrocenyl-phosphonium ionic liquids–synthesis, characterisation and electrochemistry. Dalton Trans. 2014;43:3750–3766. doi: 10.1039/c3dt53402b. [DOI] [PubMed] [Google Scholar]
  • 38.Longevial J.-F.O., Langlois A., Buisson A., Devillers C.H., Clément S.B., Van Der Lee A., Harvey P.D., Richeter S.B. Synthesis, characterization, and electronic properties of porphyrins conjugated with N-heterocyclic carbene (NHC)–gold (I) complexes. Organometallics. 2016;35:663–672. [Google Scholar]
  • 39.Balasubramanian R., Wang W., Murray R.W. Redox ionic liquid phases: Ferrocenated imidazoliums. J. Am. Chem. Soc. 2006;128:9994–9995. doi: 10.1021/ja0625327. [DOI] [PubMed] [Google Scholar]
  • 40.Coleman K.S., Turberville S., Pascu S.I., Green M.L. Synthesis of a new bidentate ferrocenyl N-heterocyclic carbene ligand precursor and the palladium (II) complex trans-[PdCl2 (C∧ fc∧ C)], where (C∧ fc∧ C)= 1, 1′-di-tert-butyl-3, 3′-(1, 1′-dimethyleneferrocenyl)-diimidazol-2-ylidene. J. Organomet. Chem. 2005;690:653–658. [Google Scholar]
  • 41.Kowalski K., Szczupak Ł., Skiba J., Abdel-Rahman O.S., Winter R.F., Czerwieniec R., Therrien B. Synthesis, structure, and spectroelectrochemistry of ferrocenyl–meldrum’s acid donor–acceptor systems. Organometallics. 2014;33:4697–4705. [Google Scholar]
  • 42.Zhang Q., Song W.-L., Hossain A.S., Liu Z.-D., Hu G.-J., Tian Y.-P., Wu J.-Y., Jin B.-K., Zhou H.-P., Yang J.-X. Synthesis, crystal structure, electrochemistry and in situ FTIR spectroelectrochemistry of a bisferrocene pyrazole derivative. Dalton Trans. 2011;40:3510–3516. doi: 10.1039/c0dt01648a. [DOI] [PubMed] [Google Scholar]
  • 43.López-Mayorga B., Sandoval-Chávez C., Carreón-Castro P., Ugalde-Saldívar V., Cortés-Guzmán F., López-Cortés J., Ortega-Alfaro M. Ferrocene amphiphilic D–π–A dyes: synthesis, redox behavior and determination of band gaps. New J. Chem. 2018;42:6101–6113. [Google Scholar]
  • 44.Gélinas B., Forgie J.C., Rochefort D. Conductivity and electrochemistry of ferrocenyl-imidazolium redox ionic liquids with different alkyl chain lengths. J. Electrochem. Soc. 2014;161:161–165. [Google Scholar]
  • 45.Reid J.E., Bernardes C.E., Agapito F., Martins F., Shimizu S., da Piedade M.E.M., Walker A.J. Structure–property relationships in protic ionic liquids: a study of solvent–solvent and solvent–solute interactions. Phys. Chem. Chem. Phys. 2017;19:28133–28138. doi: 10.1039/c7cp05076c. [DOI] [PubMed] [Google Scholar]
  • 46.Dereka B., Vauthey E. Solute–solvent interactions and excited-state symmetry breaking: Beyond the dipole–dipole and the hydrogen-bond interactions. J. Phys. Chem. Lett. 2017;8:3927–3932. doi: 10.1021/acs.jpclett.7b01821. [DOI] [PubMed] [Google Scholar]
  • 47.Zhuo J.-B., Zhu X.-X., Lin C.-X., Bai S., Xie L.-L., Yuan Y.-F. Design, synthesis and anion recognition of ferrocene-based benzimidazolium receptors. J. Organomet. Chem. 2014;770:85–93. [Google Scholar]
  • 48.Özbek H.A., Aktaş P.S., Daran J.-C., Oskay M., Demirhan F., Çetinkaya B. Synthesis, structure, electrochemical and antimicrobial properties of N, N′-bis (ferrocenylmethyl) imidazolinium salts. Inorg. Chim. Acta. 2014;423:435–442. [Google Scholar]
  • 49.Du Plessis W.C., Erasmus J.J.C., Lamprecht G.J., Conradie J., Cameron T.S., Aquino M.A.S., Swarts J.C. Cyclic voltammetry of ferrocene-containing β-diketones as a tool to obtain group electronegativities. The structure of 3-ferrocenoyl-1, 1, 1-trifluoro-2-hydroxyprop-2-ene. Can. J. Chem. 1999;77:378. [Google Scholar]
  • 50.Zhou G., Baumgarten M., Müllen K. Arylamine-substituted oligo (ladder-type pentaphenylene) s: Electronic communication between bridged redox centers. J. Am. Chem. Soc. 2007;129:12211–12221. doi: 10.1021/ja073148s. [DOI] [PubMed] [Google Scholar]
  • 51.Shi Y., Xiao L., Wu D., Li F., Li D., Zhang J., Li S., Zhou H., Wu J., Tian Y. Synthesis, crystal structure, electrochemistry and third-order nonlinear optical properties of two novel ferrocene derivatives. J. Organomet. Chem. 2016;817:36–42. [Google Scholar]
  • 52.Njogu E.M., Omondi B., Nyamori V.O. Synthesis, physical and antimicrobial studies of ferrocenyl-N-(pyridinylmethylene) anilines and ferrocenyl-N-(pyridinylmethyl) anilines. S. Afr. J. Chem. 2016;69:51–66. [Google Scholar]
  • 53.Okoli C.P., Guo Q.J., Adewuyi G.O. Application of quantum descriptors for predicting adsorption performance of starch and cyclodextrin adsorbents. Carbohydr. Polym. 2014;101:40–49. doi: 10.1016/j.carbpol.2013.08.065. [DOI] [PubMed] [Google Scholar]
  • 54.Manivannan R., Elango K.P. Spectral and electrochemical studies on anion recognition by ferrocene based imidazoles possessing different electron acceptor moieties. J. Organomet. Chem. 2015;799:99–107. [Google Scholar]
  • 55.Satheshkumar A., Manivannan R., Elango K.P. Spectroscopic, electrochemical and theoretical studies on anion recognition by receptors with redox active ferrocene and quinone centers bridged by imidazole. J. Organomet. Chem. 2014;750:98–106. [Google Scholar]
  • 56.Deglmann P., Schafer A., Lennartz C. Application of quantum calculations in the chemical industry—An overview. Int. J. Quantum Chem. 2015;115:107–136. [Google Scholar]
  • 57.COSMO. Bruker AXS; Madison, WI: 2009. Bruker. [Google Scholar]
  • 58.APEXII. Bruker AXS; Madison, WI: 2009. Bruker. [Google Scholar]
  • 59.SAINT. Bruker AXS; Madison, WI, USA: 2009. Bruker. [Google Scholar]
  • 60.Sheldrick G.M. A short history of SHELX. Acta. Crystallogr. Sect. A. 2008;64:112. doi: 10.1107/S0108767307043930. [DOI] [PubMed] [Google Scholar]
  • 61.Farrugia L.J. WinGX and ORTEP for Windows: an update. J. Appl. Crystallogr. 2012;45:849. [Google Scholar]
  • 62.Macrae C.F., Bruno I.J., Chisholm J.A., Edgington P.R., McCabe P. e. Pidcock, L. Rodriguez-Monge, R. Taylor, J. van de Streek, P.A. Wood. Mercury CSD 2.0–new features for the visualization and investigation of crystal structures. J. Appl. Crystallogr. 2008;41:466. [Google Scholar]
  • 63.Lee C., Yang W., Parr R.G. Development of the Colle-Salvetti correlation-energy formula into a functional of the electron density. Phys Rev B. 1988;37:785. doi: 10.1103/physrevb.37.785. [DOI] [PubMed] [Google Scholar]
  • 64.Becke A.D. A new mixing of Hartree–Fock and local density-functional theories, J. Chem. Phys. 1993;98:1372–1377. [Google Scholar]
  • 65.Grimme S., Antony J., Ehrlich S., Krieg H. A consistent and accurate ab initio parametrization of density functional dispersion correction (DFT-D) for the 94 elements H-Pu, J. Chem. Phys. 2010;132(15) doi: 10.1063/1.3382344. [DOI] [PubMed] [Google Scholar]
  • 66.Yanai T., Tew D.P., Handy N.C. A new hybrid exchange–correlation functional using the Coulomb-attenuating method (CAM-B3LYP), Chem Phys Lett. 2004;393:51–57. [Google Scholar]
  • 67.Zhao Y., González-García N., Truhlar D.G. Benchmark database of barrier heights for heavy atom transfer, nucleophilic substitution, association, and unimolecular reactions and its use to test theoretical methods. J Phys. Chem. A. 2005;109:2012–2018. doi: 10.1021/jp045141s. [DOI] [PubMed] [Google Scholar]
  • 68.Zhao Y., Truhlar D.G. Density functionals with broad applicability in chemistry. Acc. Chem. Res. 2008;41:157–167. doi: 10.1021/ar700111a. [DOI] [PubMed] [Google Scholar]
  • 69.Weigend F., Ahlrichs R. Balanced basis sets of split valence, triple zeta valence and quadruple zeta valence quality for H to Rn: Design and assessment of accuracy. Phys Chem Chem Phys. 2005;7:3297–3305. doi: 10.1039/b508541a. [DOI] [PubMed] [Google Scholar]
  • 70.Krishnan R.B., Binkley J.S., Seeger R., J.A Pople Self consistent molecular orbital methods. XX. A basis set for correlated wave functions. J. Chem. Phys. 1980;72:650–654. [Google Scholar]
  • 71.McWeeny R. Proceedings of the Royal Society of London A: Mathematical, Physical and Engineering Sciences. The Royal Society; 1956. The density matrix in self-consistent field theory. II. Applications in the molecular orbital theory of conjugated systems. [Google Scholar]
  • 72.Frisch M.J., Trucks G.W., Schlegel H.B., Scuseria G.E., Robb M.A., Cheeseman J.R., Scalmani G., Barone V., Mennucci B., Petersson G.A., Nakatsuji H. Gaussian, Inc.; Wallingford CT: 2009. Gaussian 09, revision D. 01. [Google Scholar]
  • 73.Dennington R., Keith T., Millam J. Semichem Inc.; Shawnee Mission KS: 2009. In GaussView. [Google Scholar]

Articles from Heliyon are provided here courtesy of Elsevier

RESOURCES