Authors' Self-Archive Manuscript: Institution'S Repository (Intelcentru/Icmpp, Iasi
Authors' Self-Archive Manuscript: Institution'S Repository (Intelcentru/Icmpp, Iasi
Title:
“Synthesis, structural characterization and quantum chemical studies on novel
silicon-containing benzoic acid derivatives: 4‒((trimethylsilyl)methoxy)and
4‒(3‒(trimethylsilyl)propoxy)benzoic acids”
by:
Mirela-Fernanda Zaltariov*, Corneliu Cojocaru, Sergiu Shova, Liviu Sacarescu, Maria Cazacu
Department of Inorganic Polymers, “Petru Poni” Institute of Macromolecular Chemistry, Romanian Academy,
Aleea Grigore Ghica Voda 41 A, 700487 Iasi, Romania.
Corresponding author: Mirela-Fernanda Zaltariov
Email: [email protected]
1
Abstract
The present paper is concerned with the synthesis and molecular structure investigation of
two new benzoic acid derivatives having trimethylsilyl tails, 4‒((trimethylsilyl)methoxy) and
4‒(3‒(trimethylsilyl)propoxy)benzoic acids. The structures of title compounds have been
confirmed by X-ray crystallography, Fourier-transform infrared spectroscopy (FTIR) and nuclear
magnetic resonance (1H and 13C NMR). The theoretical studies of molecules were conducted by
using the quantum chemical methods, such as Density Functional Theory (DFT
B3LYP/6‒31+G**), Hartree‒Fock (HF/6‒31+G**) and semiempirical computations (PM3, PM6
and PM7). The optimized molecular geometries have been found to be in good agreement with
experimental structures resulted from the X‒ray diffraction. The maximum electronic absorption
bands observed at 272‒287 nm (UV‒vis spectra) have been assigned to * transitions, which
were in reasonable agreement with the time dependent density functional theory (TD‒DFT)
calculations. The computed vibrational frequencies by DFT method were assigned and compared
with the experimental FTIR spectra. The mapped electrostatic potentials revealed the reactive
sites, which corroborated the observation of the dimer supramolecular structures formed in the
crystals by hydrogen‒bonding. The energies of frontier molecular orbitals (HOMO and LUMO),
energy gap, dipole moment and molecular descriptors for the title compounds were calculated
and discussed.
Keywords: Benzoic acid derivatives; Organosilicon compounds; Computational chemistry;
Vibrational spectra; Electronic absorption spectra; HOMO‒LUMO.
1. Introduction
Benzoic acid derivatives are versatile compounds used as chemicals, pharmaceuticals,
agrochemicals and consumer products.1 Benzoic acid inhibits bacterial development and its
substituted compounds are very important for the development of new materials in food and
pharmaceutical industries.2 In the last years, considerable efforts have been devoted to
theoretically and experimentally study of the benzoic acid derivatives.1‒14 Thus, a number of
these compounds was investigated: p‒hydroxybenzoic, m‒anisic, vanillic, and syringic acids;1
4‒butyl benzoic acid;2 methyl‒ and methoxybenzoic acids;3 m‒trifluoromethylbenzoic acid;4
4‒(2,5‒di‒2‒thienyl‒1H‒pyrrol‒1‒yl)benzoic acid;5 p‒(p‒hydroxyphenoxy)benzoic acid;6 toluic
acid;7 2‒(4‒hydroxyphenylazo)benzoic acid;8 2‒amino‒5‒bromo‒benzoic acid methyl ester;9
4‒(trimethylammonium)benzoic acid chloride;10 2,3,4‒tri‒fluoro‒benzoic acid dimer;11
2‒[(2‒hydroxyphenyl)carbonyloxy]benzoic acid;12 Cu(II) complex with
13
4‒{(Z)‒[(2‒hydroxybenzoyl)hydrazono]methyl}benzoic acid and 4‒[[(substituted
phenyl)imino]methyl]benzoic acids.14 All of the above mentioned molecules were mainly
modeled by density functional theory (DFT), and the simulation outcomes were compared with
the available experimental results (e.g. X‒ray crystallography; FTIR, Raman, UV‒vis and NMR
spectra). The current work aims to report on the synthesis pathway, molecular structure
characterization and theoretical quantum chemical investigations of two novel benzoic acid
derivatives having trimethylsilyl (TMS) tails, i.e. 4‒((trimethylsilyl)methoxy) and
4‒(3‒(trimethylsilyl)propoxy)benzoic acids as potential ligands for the metal ions. The presence
2
of TMS groups creates premises for obtaining metal complexes with increased solubility in
non‒polar media, such as supercritical carbon dioxide (scCO2), a non‒toxic, non‒flammable,
eco-friendly and relatively cheap solvent, where they can act as homogeneous catalyst for
different reactions. The number of metal complexes that are soluble in this environment is very
limited.15 In addition, due to the higher hydrophobicity of the trimethylsilyl group, these
compounds have a high potential for application as bioactive compounds in enzymatic
reactions.16
Previously published articles described the experimental investigations regarding the synthesis
and characterization of benzoic acid derivatives containing organosilicon groups.17‒19 However,
to the best of our knowledge no report has been focused, until now, on both theoretical and
experimental investigations of the molecular structures of silicon‒containing benzoic
derivatives.20 Molecular modeling is an extremely useful tool to complement experimental
techniques providing a potential structural context for understanding the experimental results.21
Note that, some organosilicon compounds have been tested as pharmaceuticals.20
2. Experimental
2.1. Materials
4‒Hydroxybenzoic acid (Aldrich), (Chloromethyl)trimethylsilane 98% (Aldrich),
(3‒chlororpropyl)trimethylsilane 97% (Aldrich), anhydrous K2CO3 (Aldrich), Na2SO4 (Aldrich),
dimethylformamide (Aldrich), chloroform (Aldrich) were used as received.
2.2. Methods
Fourier transform infrared (FT‒IR) measurements were carried out using a Bruker Vertex 70
FT-IR spectrometer. Spectra were recorded in the transmission mode in the range 400‒4000 cm-1
at room temperature with a resolution of 2 cm‒1 and accumulation of 32 scans.
The NMR spectra were recorded on a Bruker Avance DRX 400 MHz Spectrometer equipped
with a 5 mm QNP direct detection probe and Z‒gradients. Spectra were recorded in CDCl3, at
room temperature. The chemical shifts are reported as δ values (ppm). The assignments of all the
signals in the 1D NMR spectra were done using 2D NMR experiments like H,H‒COSY,
H,C‒HMQC and H,C‒HMBC.
UV‒vis absorption spectra measurements were carried out in CHCl3 and DMSO solutions on
a Specord 200 spectrophotometer.
X‒Ray crystallographic measurements for 1 and 2 were carried out with an
Oxford‒Diffraction XCALIBUR E CCD diffractometer equipped with graphite‒monochromated
Mo‒Kα radiation. Single crystals were positioned at 40 mm from the detector and 848, and 208
frames were measured each for 25, and 40 s over 1° scan width for 1, and 2, respectively. The
unit cell determination and data integration were carried out using the CrysAlis package of
Oxford Diffraction.22 The structures were solved by direct methods using Olex223 and refined by
full-matrix least-squares on F² with SHELXL-97.24 Atomic displacements for non-hydrogen
atoms were refined using an anisotropic model. All H atoms were introduced in idealized
positions (dCH = 0.96 Å) using the riding model with their isotropic displacement parameters
fixed at 120% of their riding atom. The molecular plots were obtained using the Olex2 program.
The crystallographic data and refinement details are quoted in Table 1, while bond lengths and
angles are summarized in Table 2.
CCDC‒1438222 (1), CCDC‒1438223 (2) contain the supplementary crystallographic data
for this contribution. These data can be obtained free of charge via
www.ccdc.cam.ac.uk/conts/retrieving.html (or from the Cambridge Crystallographic Data Centre,
12 Union Road, Cambridge CB2 1EZ, UK; fax: (+44) 1223-336-033; or [email protected]).
3
2.3. Synthesis
2.3.1. Synthesis of 1
In a 100 mL round bottom flask equipped with magnetic stirring and reflux condenser
protected with CaCl2 tube, 4-hydroxybenzoic (0.98 g, 7 mmol) and
(trimethylsilyl)chlorometylsilane (0. 86 g, 1 mL, 7 mmol) were disolved together in DMF (15
mL). Anhydrous K2CO3 (1.45 g, 10.5 mmol) was added to the resulting solution and the mixture
was stirred under argon athmosphere at 110 oC for 8 hours. The resulting precipitate was filtered
off and the filtrate was poured into water (60 mL) and extracted with chloroform (3 x 15 mL).
The combined organic phases were dried (Na2SO4) and evaporated under reduced pressure. The
resulted beige solid product was dried in vacuo. X‒ray diffraction‒quality single crystals were
obtained by recrystallization from methanol, washed with methanol and diethylether and dried in
air at room temperature. Yield: 1.2 g, 75%. Anal. Calcd for C11H16O3Si (Mr 224.33): C, 58.89, H,
7.19. Found: C, 58.67, H, 7.23.
IRmax (KBr), cm‒1: 3551w, 3474w, 3416w, 2957w, 2901w, 2882w, 2822w, 1688s,
1601vs, 1578s, 1506s, 1421m, 1292s, 1250s, 1215m, 1157s, 1130vw, 1109w, 1007m, 858vs,
847vs, 775m, 758w, 706w, 636m, 619m, 604w, 546w, 513w, 505w, 480vw, 451vw, 378vw.
1
H NMR (CDCl3, 400.13 MHz, δ, ppm): 9.87 (s, 1H, ‒COOH), 7.82 (d, J=8.7 Hz, 2H,
Ar‒H), 7.05 (d, J=8.7 Hz, 2H, Ar‒H), 3.66 (s, 2H,‒CH2‒), 0.16 (s, 9H, ‒Si‒CH3).
13
C NMR (CDCl3, 100.16 MHz, δ, ppm): 190.88 (‒COOH), 166.72 (Ar‒C), 131.91
(Ar‒C), 129.63 (Ar‒C), 114.53 (Ar‒C), 61.78 (‒CH2‒), ‒3.20 (‒Si‒CH3).
UV‒vis (CHCl3), λmax (ε, M‒1cm‒1): 272 (21816); UV‒vis (DMSO), λmax (ε, M‒1cm‒1): 275
(20823).
2.3.2. Synthesis of 2
In a 100 mL round bottom flask equipped with magnetic stirrer and reflux condenser
protected with CaCl2 tube, 4‒hydroxybenzoic (0.83 g, 6 mmol) and
(trimethylsilyl)chloropropylsilane (0. 88 g, 1 mL, 6 mmol) were disolved together in DMF (15
mL). Anhydrous K2CO3 (1.45 g, 10.5 mmol) was added to the resulting solution and the mixture
was stirred under argon athmosphere at 110 oC for 8 hours. The resulting precipitate was filtered
off and the filtrate was poured into water (60 mL) and extracted with chloroform (3 x 15 mL).
The combined organic phases were dried (Na2SO4) and evaporated under reduced pressure. The
solid product was dried in vacuo. X‒ray diffraction‒quality single crystals were obtained by
recrystallization from chloroform/methanol (1:2, v:v), washed with methanol and diethylether
and dried in air at room temperature. Yield: 1.3 g, 86 %. Anal. Calcd for C11H16O3Si (Mr 252.38):
C, 61.87, H, 7.99. Found: C, 61.67, H, 8.13.
IRmax (KBr), cm‒1: 3366m, 3289m, 3155w, 3119w, 3074m, 3040w, 2951vs, 2895s,
2878s, 2826s, 2904s, 2735s, 2573w, 2515vw, 2428vw, 2012w, 1936vw, 1906w, 1693vs, 1601vs,
1578vs, 1510vs, 1470s, 1427s, 1393s, 1313s, 1250vs, 1215vs, 1184s, 1159vs, 1109s, 1063s,
1047s, 1009s, 895s, 854vs, 835vs, 787s, 758s, 694s, 650s, 617s, 606s, 515s, 453w, 436w, 422w,
382w.
1
H NMR (CDCl3, 400.13 MHz, δ, ppm): 9.86 (s, 1H, ‒COOH), 7.82 (d, J=8.7 Hz, 2H,
Ar‒H), 6.97 (d, J=8.7 Hz, 2H, Ar‒H), 3.98 (t, J=6.8 Hz, ‒CH2‒CH2‒CH2‒), 1.84‒1.77 (m, 2H,
‒CH2‒CH2‒CH2‒), 0.63‒0.58 (m, 2H, ‒CH2‒CH2‒CH2‒), 0.02 (s, 9H, ‒Si‒CH3).
4
13
C NMR (CDCl3, 100.16 MHz, δ, ppm): 190.85 (‒COOH), 164.24 (Ar‒C), 132.00
(Ar‒C), 129.68 (Ar‒C), 114.72 (Ar‒C), 70.99 (‒CH2‒CH2‒CH2‒), 23.66 (‒CH2‒CH2‒CH2‒),
12.50 (‒CH2‒CH2‒CH2‒), ‒1.79 (‒Si‒CH3).
UV‒vis (CHCl3), λmax (ε, M‒1cm‒1): 285 (28235); UV‒vis (DMSO), λmax (ε, M‒1cm‒1): 287
(26176).
4. Computational details
The silicon‒containing benzoic acid derivatives 1 and 2 were modeled by computational
chemistry approach using the density functional theory (DFT), ab-initio (Hartree-Fock, HF) and
semiempirical (PM3, PM6 and PM7) methods. To this end, the quantum chemical calculations
(DFT, HF and semiempirical) were performed in order to study the molecular structure
characteristics and electronic properties of the titled compounds. The software package
5
Gaussian0925 was employed for DFT and HF calculations of the molecules, using the split-
valence basis set 6-31+G** with added polarization and diffuse functions. The density functional
method was implemented at the level of B3LYP functional (i.e. B3LYP/6-31+G**). The
computations by PM3, PM6 and PM7 methods were performed by means of MOPAC semi
empirical quantum chemistry program.26‒31 The molecular modeling results were analyzed using
graphical-interface computational chemistry software, such as GaussView 5,32 Gabedit33 and
HyperChem.34
Geometry optimization is an important part of computational chemistry, and it deals with
searching by energy minimization of the optimized molecular structures.35 In this work, the
geometry optimization of molecules 1 and 2 was performed by minimizing the energies with
respect to all geometrical parameters without imposing any symmetry constrains. The input
structures for optimization were taken from the CIF-files obtained from X-ray single-crystal
measurements. Further, the geometry optimization was performed on single molecules, in
vacuum. Frequency calculations at the optimized geometries were done to confirm the optimum
conformations (i.e. no imaginary frequencies) and to analyze the vibration modes. The electronic
absorption spectra of the investigated compounds were calculated by the time dependent density
functional theory (TD-DFT).
6
Some structural disparities have been observed in terms of torsion angles (dihedrals). The
information about dihedrals (torsion angles) for 1 and 2 is given extensively in Tables B3 and B6
(ESI), respectively. In Table 2 only the selected torsion angles are summarized.
Such differences could be explained by taking into account the conformational constraints
that occur when the molecules are assembled in crystalline structures. However, in order to
identify the overall conformational similitude between observed molecular structures and the
modeled ones, the overlap plots and the calculation of root-mean-square-deviation (RMSD) are
commonly used as goodness-of-fit estimators.38‒40 The RMSD of the atomic positions is a
similarity measure for the quantitative comparison of one structure with another, and it is given
by:40
RMSD r
1 N i
2
rk rk
j
(1)
N k 1
Where N denotes the number of atoms in the molecule; k is an index over these atoms;
i j
rk , rk – coordinates of atom k in conformations i and j. The minimal value of RMSD is
desirable and it is obtained by the optimal superposition of the two structures. In this work, the
RMSD values have been calculated to compare the observed molecular structures (from X‒ray
crystallography) with their optimized geometries given by quantum chemistry models. The
calculated values of RSMD are summarized in Table 3. In addition, the overlap plots are given in
Fig. 6 and ESI (Figs. A2-A3), showing the superposition of the X‒ray structures of the
compounds and their optimized counterparts. Generally, the predicted geometries are quite
similar with the experimental molecular structures (Fig. 6 and Figs. A2‒A3/ESI). The
semiempirical method PM3 has yield the lowest RSMD values of 0.2553 Å and 0.2029 Å
(Table 3), for the compounds 1 and 2, respectively, showing the best similitude with the
observed structures. This may be attributed to the fact that PM3 method has better predicted the
dihedral angles. Likewise, the methods HF, DFT and PM7 were appropriate for the optimization
of molecular structures of the benzoic-acid derivatives, providing the low values of RMSD, i.e.
RMSD<0.43 Å (Table 3). The semiempirical method PM6 indicated the greatest values of
RMSD, suggesting some conformational discrepancy between modeled structures and observed
ones. For example, in case of compound 2, the RMSD value given by PM6 method is equal to
1.4959 Å, which is significantly greater than RMSD values provided by the other methods. This
structural disparity in terms of atomic positions can be also seen in Fig. 6 (d). In summary, the
optimized molecular geometries computed by DFT, HF, PM3 and PM7 methods were in good
agreement with the observed molecular structures obtained from X‒ray crystallography. In turn,
the optimized geometries calculated by PM6 method fairly matched with the observed structures.
7
the theoretical results listed in Tables 4 and 5, the total energies of molecules computed by HF
and DFT methods are lower with about one order of magnitude than the total energies calculated
by semi empirical methods (PM3, PM6 and PM7). This can be explained by the fact that semi
empirical methods use a simpler Hamiltonian along with empirical data to approximate the
Schrödinger equation, and thus treating only the valence electrons. In turn, the core electrons are
considered frozen. In contrast, HF and DFT methods treat more precisely all type of electrons,
i.e. core and valence electrons. Therefore, the number of double occupied molecular orbitals
taken into account by HF and DFT methods is always greater than that considered by semi
empirical models (Tables 4, 5).
The heats of formation computed by semiempirical methods for the titled compounds 1 and
2 are summarized in Table B7 (ESI). According to PM3 method, the heats of formation for 1 and
2 were equal to ‒144.37 kcal/mol and ‒156.14 kcal/mol, respectively. Other semiempirical
techniques (i.e. PM6 and PM7) converged to similar results. The theoretical calculations revealed
significant values of the dipole moment for the investigated compounds (Tables 4 and 5). For the
molecule 1, the dipole moment ranged from 4.131 to 6.093 Debyes, whereas for the molecule 2,
the calculated dipole moment varied in the extent 3.645‒5.674 Debyes, depending on the
quantum chemistry method applied. Thus, the dipole moment values for compound 2 are lower
than for the compound 1, due to additional content of hydrophobic (‒CH2‒) groups. The
calculated polarizability volumes of titled compounds 1 and 2 were found to be equal to 23.28 Å3
and 26.95 Å3, respectively. The molecular size (dmax) is given in Tables 4 and 5 as the measure of
the maximum atomic distance in the molecules. For instance, according to DFT method, the
molecular dimensions for the compounds 1 and 2 are equal to 11.967 Å and 13.946 Å,
respectively. The molecular surface area (Sm) and molecular volume (Vm) have been calculated
by numerical integration grid techniques implemented in HyperChem program. The calculated
values of these parameters are also reported in Tables 4 and 5.
The energy spectra of molecular orbitals (MOs) in ground state (S0) were computed for
both molecules by means of quantum chemistry methods. Figure 7 shows as an example the
energy spectra of MOs calculated for the compounds 1 and 2 at the level of PM3 model. As
shown in Fig.7, each molecular orbital has an orbital index (j) and a corresponding energy
value (Ej – eigenvalue). Here, the blue lines indicate the eigenvalues of occupied molecular
orbitals (OMO), and the red ones denote the energy levels of unoccupied (virtual) orbitals
(UMO). Such spectra give useful details about the energy values of the main molecular orbitals.
The information about highest occupied (HOMO) and lowest unoccupied (LUMO) molecular
orbitals are important, since they are mostly responsible for the chemical reactivity and
spectroscopic properties of the molecule.36,41
Likewise, the energy difference between frontier molecular orbitals
(HOMO-LUMO), known as energy gap (E), is a measure of the intramolecular charge transfer
and is frequently used in chemical and biochemical activity studies.42
The energy values of frontier molecular orbitals (HOMO, LUMO, HOMO‒1, LUMO+1)
and the corresponding energy gaps (E, E1) computed by different quantum chemistry methods
are summarized in Tables 4 and 5. Thus, the semiempirical and HF methods suggested high
energy gap (E) values for the titled compounds, i.e. from 8.8 to 10.2 eV. The DFT method
provided smaller energy gap values of about 5.1 eV. Such disparity can be attributed to the fact
that DFT method (B3LYP/6‒31+G**) uses the exchange‒correlation functionals43 and therefore
it treats in explicit manner the electronic correlation effects comparing with HF and
semiempirical methods.
8
In our case, the large HOMO‒LUMO gaps (> 5 eV) suggest the high excitation energies for
many excited states, good chemical stability and moderate reactivity for the investigated
compounds. The molecular descriptors such as ionization potential (IP), electron affinity (EA),
electronegativity (), chemical potential (µ), chemical hardness () and electrophilicity index ()
can be computed from the eigenvalues of the frontier molecular orbitals. These estimators are
important parameters for quantum chemistry. Note that, electronegativity () is a chemical
property describing the tendency of a molecule to attract electrons (or electron density) towards
itself. By contrast, the electronic chemical potential () determines the escaping tendency of the
electron density from the molecule. The chemical hardness () measures the resistance of the
molecule to lose electrons. In turn, the electrophilicity index () indicates the electrophilic power
of a molecule, its propensity to be saturated with electrons.44
Thus, by using the energy values of HOMO and LUMO the molecular descriptors have
been calculated and reported in Tables 4 and 5. The relationships employed for the calculation of
the molecular descriptors are given in Ref.37 and in Table B8 (ESI). According to the theoretical
results, the chemical hardness values greater than 2 eV and the low chemical potential values
suggested a good stability and resistance of the molecules to lose electrons. The electronegativity
values higher than 3.7 eV indicated the tendency of the molecules to attract the electron density.
However, the electrophilicity index values ranging from 1.4 eV to 3.1 eV, revealed a moderate
electrophilic power.
The plots of the frontier molecular orbitals for the studied molecules (1 and 2) are shown in
Fig.8. In this figure, the spatial distributions of the molecular orbitals were computed at the level
of DFT theory (B3LYP/6‒31+G**). As one can see from Fig.8 (a, e), the HOMO is a delocalized
-orbital distributed on the aromatic ring and over the ether and carboxylic groups. By contrast,
the HOMO-1 levels (Fig. 8 c, g) are localized on the aromatic ring (‒system). The virtual
orbitals (LUMO and LUMO+1) are smaller in size than the occupied molecular orbitals. The
LUMO lobes are spread on benzene ring and over the ether and carboxylic groups (Fig. 8 b, f).
Hence, LUMO represents a delocalized *‒orbital. The levels of LUMO+1 are localized onto the
aromatic ring (Fig. 8 d, h).
9
dipole moments are illustrated as vectors in Fig. 9 (a, c). The origin of each vector is located near
the oxygen atom from ether group and the vector is oriented toward the silyl group. Thus,
the dipole moments act in the directions of these vectors quantities, suggesting the polar
properties of the investigated compounds 1 and 2.
The molecular electrostatic potential map represents a useful three dimensional diagram
to visualize the charge related properties of a molecule. Such maps have been applied extensively
to predict the reactive sites for electophilic and nucleophilic attack, as well as to study the
biological recognition and hydrogen bonding interactions.46 On the basis of partial atomic
charges, the electrostatic potential surface (ESP) surrounding each molecule has been mapped by
using the grid-points technique implemented in GaussView 5 program. Figure 9 (b, d) illustrates
the three-dimensional (3D) mapped isosurfaces of the electrostatic potentials (computed by DFT
method) for the molecules 1 and 2. The red colors denote negative ESP regions followed by
green and blue colors indicating neutral and positive ESP domains, respectively. For both
molecules 1 and 2, the oxygen atom from carbonyl group had the highest negative ESP region
followed by the oxygen atom from hydroxyl group. These negative ESP areas indicated the sites
for electrophilic attack. The most positive ESP region (site of nucleophilic attack) was localized
over hydrogen atom from carboxylic group (Fig.9 b, d). These theoretical results are in good
agreement with the experimental data from X‒ray crystallography. Thus, the sites of electrophilic
and nucleophilic attack identified by ESP computation corroborate the formation in the crystals
of the dimer supramolecular structures by hydrogen‒bonding.
10
(in CHCl3). Thus, the predicted wavelengths were blue‒shifted (hypsochromic effect) with 20‒21
nm compared with experimental values. For the molecule 2, the observed spectral band position
in absorption was pinpointed at 275 nm (in DMSO) and 272 nm (in CHCl3). For this case
(molecule 2), the TD‒DFT simulations also led to shorter theoretical wavelengths for maximum
absorptions, which were equal to 266 nm (in DMSO) and 264 nm (in CHCl3). If compared with
experimental observations, the results of TD‒DFT suggested a less intense hypsochromic effect
(blue‒shift) of 8‒9 nm for the compound 2. The hypsochromic effects for gas phase were more
significant than those for implicit solvents. The maximum absorption peaks that ranged from 256
nm to 287 nm have been assigned to the →* electronic transitions. The TD‒DFT theory
revealed that these excited states belong to HOMO→LUMO and HOMO-1→LUMO+1
electronic transfer. The excited electronic configuration HOMO→LUMO has a major
contribution (94‒96%), whereas HOMO‒1→LUMO+1 has a minor contribution (3‒4%). The
contributions of the excited electronic states were calculated based on the values of coefficients
of the specific configurations. To this end, the open‒source software GaussSum 3.050,51 has been
employed, which was developed for analyzing the output files from several computational
chemistry packages. Hence, the theoretical results suggested that the electronic system
delocalized over the aromatic ring, ether and carboxylic groups (Fig. 8) was mainly responsible
for the maximum absorption peaks of the compounds 1 and 2.
In summary, both observed and theoretical UV‒Vis spectra showed maximum absorption
peaks located into the near‒ultraviolet region (256‒287 nm) that were assigned to →*
electronic transitions. Thus, the experimental UV‒Vis spectra for 1 and 2 were in reasonable
agreement with the theoretical electronic absorption spectra computed by TD-DFT method.
11
For the experimental FTIR spectra, the main assignments were attributed in accordance
with the literature data53 and were summarized in Table 7, for both compounds 1 and 2. The
experimental FTIR spectra recorded in the 400‒4000 cm‒1 region showed the characteristic bands
of stretching vibrations () of the following groups: C‒H, C‒O‒C, O‒H, C=O, Si‒C and C=C-C
(aromatic ring). For details, see Table 7.
Thus, the stretching vibrations (C‒H) from aliphatic groups have been observed in the
region 2822‒2957 cm‒1. The aliphatic (C‒H) vibrations were predicted by DFT method at
wavenumbers (scaled values) ranging from 2888 to 3008 cm‒1 (see ESI, Tables B11‒B12). The
stretching vibrations (C‒H) for the aromatic ring were determined experimentally at 3074 cm‒1
and 3077 cm‒1, for the compounds 1 and 2, respectively. The theoretical counterparts for (C‒H)
(aromatic ring) were predicted by DFT in the range 3090‒3117 cm‒1.
The symmetric stretching of O‒H bond from carboxyl group was observed at 3416 cm‒1
and 3366 cm‒1 for the compounds 1 and 2, respectively. The calculated wavenumber for s(O‒H)
was pinpointed at 3635 cm‒1 in both cases. The O‒H bending vibrations were observed within the
range 1421‒1427 cm‒1 and predicted in the region 1313‒1406 cm‒1. The strong signals identified
experimentally at 1688 cm‒1 (for 1) and 1693 cm‒1 (for 2) were assigned to symmetrical C=O
stretching bands. The theoretical vibration s(C=O) was found at 1718 cm‒1 (for 1 and 2).
In both molecules, the stretching vibrations of C=C‒C from aromatic ring appeared in the
range 1506‒1605 cm‒1 with the intensities varying from medium to strong. The DFT
computations provided the modes for (C=C‒C / aromatic ring) in the region 1489‒1595 cm‒1.
Regarding the ether group, the C‒O‒C asymmetric stretching oscillation was assigned to
the frequency observed at 1292 cm‒1 (for 1) and 1313 cm‒1 (for 2). The vibration a(C‒O‒C) was
identified theoretically at 1227 cm‒1 (only for the compound 1). According to the experimental
data, the signals positioning at 1157 cm‒1 and 1159 cm‒1 were attributed to the C‒O symmetric
stretching vibrations. The DFT calculations provided the modes for s(C=O) at 994‒1063 cm‒1.
According to the experimental FTIR spectra, the stretching oscillation Si‒C was
recognized by the bands at 1250 cm‒1 along with the strong peaks in the region 835‒858 cm‒1.
The predicted modes for (Si‒C) appeared at the ranges 1169‒1238 cm‒1 and 836‒861 cm‒1. The
medium bands observed at wavelength 775 cm‒1 (for 1) and 786 cm‒1 (for 2) were assigned to the
out‒of‒plane bending vibrations of C‒H from the aromatic ring.
In addition, the theoretical IR spectra were computed at the level of HF/6‒31+G**, PM3,
PM6 and PM7 methods and were shown for comparison in Fig. A5 (ESI). In order to estimate
the goodness‒of‒fit between experimental and theoretical vibration frequencies, the linear
correlation coefficients (r2) were calculated and reported in Table B13 (ESI). According to these
data, the greatest values of correlation coefficients were provided by B3LYP/6‒31+G** method,
and were equal to 0.9972 and 0.9971, for the compounds 1 and 2, respectively (Table B13).
Thus, the DFT yielded a better correlation between experimental and calculated vibration
frequencies than other computational chemistry methods applied in this work.
The parity plots between experimental wavenumbers and theoretical (scaled) counterparts
computed by B3LYP/6‒31+G** method are illustrated in Fig. 12 for the investigated compounds
1 (Fig.12a) and 2 (Fig.12b). As shown in Fig.12, the linear correlations emerge between
observed and predicted vibrational frequencies (r2>0.997). The linear regression lines almost
overlap with bisectors for the lower range 500‒1700 cm‒1. For the higher region 2500‒3500
cm‒1, a minor disparity between strait lines and bisectors can be distinguished. Overall, most of
data were scattered nearby the bisectors revealing a good agreement between observed and
12
predicted wavelengths. Hence, both experimental and theoretical vibrational spectra provided
adequate relationships between infrared bands and the molecular structures.
5. Conclusions
Two silicon‒containing organic compounds, namely 4‒((trimethylsilyl)methoxy) and
4‒(3‒(trimethylsilyl)propoxy)benzoic acids (1 and 2) were synthesized for the first time. The
compounds were characterized by X‒ray single‒crystal diffraction, 1H NMR, 13C NMR and
FTIR spectroscopy techniques. Structural data were complemented with quantum chemical
computations using DFT/B3LYP/6‒31+G**, HF/6‒31+G** as well as semiempirical (PM3,
PM6 and PM7) methods. Theoretical optimized geometries for the investigated molecules were
in good agreement with our reported X‒ray structures.
The calculated partial atomic charges and dipole moments (4‒6 Debye) suggested the
polar properties of the investigated compounds. The mapped electrostatic potential surfaces
revealed the sites for electrophilic and nucleophilic attack localized near the carboxylic group, on
oxygen and hydrogen atoms, respectively. These sites identified theoretically corroborated the
observation of the dimer supramolecular structures formed in the crystals by hydrogen‒bonding.
The calculated HOMO‒LUMO gaps (E) greater than 5 eV suggested good chemical stability
and moderate reactivity for the investigated compounds.
Theoretical electronic absorption spectra were developed using TD‒DFT method and
compared with experimental UV‒Vis spectra. A reasonable agreement between experimental and
theoretical electronic spectra was concluded. Both observed and theoretical UV‒Vis spectra
indicated the maximum absorption peaks pinpointed at 256‒287 nm (near‒ultraviolet region) that
were assigned to →* electronic transitions. The theory (TD‒DFT) suggested that the
electronic system delocalized over the aromatic ring, ether and carboxylic groups was mainly
responsible for the most intense transition bands for the investigated compounds 1 and 2.
The theoretical vibrational band assignments were performed at the level of DFT
B3LYP/6‒31+G** model in order to compare the experimental (FTIR) and calculated vibrational
frequencies. The infrared spectra showed the characteristic bands of stretching vibrations of the
following groups: C‒H, C‒O‒C, O‒H, C=O, Si‒C and C=C‒C (aromatic ring). The results
revealed a good agreement between observed and predicted infrared wavenumbers. Both
experimental and theoretical vibrational spectra provided adequate relationships between infrared
bands and the molecular structures of the studied compounds.
Acknowledgement
This research was founded by the Romanian Ministry of National Education under Grant
53/02.09.2013, Cod: PN-II-ID-PCE-2012-4-0261. The author Corneliu Cojocaru is grateful for
the financial support from the H2020 ERA Chairs Project no 667387: SupraChem Lab
Laboratory of Supramolecular Chemistry for Adaptive Delivery Systems, ERA Chair initiative.
13
References
(1) Swislocka, R.; Regulska, E.; Samsonowicz, M.; Lewandowski, W. Experimental and
theoretical study on benzoic acid derivatives. J. Mol. Struct. 2013, 1044, 181‒187.
(2) Karabacak, M.; Cinar, Z.; Kurt, M.; Sudha, S.; Sundaraganesan, N. FT-IR, FT-Raman, NMR
and UV–vis spectra, vibrational assignments and DFT calculations of 4-butyl benzoic acid.
Spectrochim. Acta A 2012, 85, 179‒189.
(3) Verevkin, S. P.; Zaitsau, D.H.; Emelyanenko, V.N.; Stepurko, E.N.; Zherikova, K.V. Benzoic
acid derivatives: Evaluation of thermochemical properties with complementary experimental and
computational methods. Thermochim. Acta 2015, doi:10.1016/j.tca.2015.03.026.
(4) Balachandran, V.; Karpagam, V.; Santhi, G.; Revathi, B.; Ilango, G.; Kavimani, M.
Conformational stability, vibrational (FT-IR and FT-Raman) spectra and computational analysis
of m-trifluoromethyl benzoic acid. Spectrochim. Acta A 2015, 137, 165‒175.
(5) Kurt, M.; Babur Sas, E.; Can, M.; Okur, S.; Icli, S.; Demic, S.; Karabacak, M.; Jayavarthanan,
T.; Sundaraganesan, N. Synthesis and spectroscopic characterization on 4-(2,5-di-2-thienyl-1H-
pyrrol-1-yl)benzoic acid: A DFT approach. Spectrochim. Acta A 2016, 152, 8‒17.
(6) Balachandran, V.; Lalitha, S.; Rajeswari, S.; Rastogi, V. K. Theoretical investigations on the
molecular structure, vibrational spectra, thermodynamics, HOMO–LUMO, NBO analyses and
paramagnetic susceptibility properties of p-(p-hydroxyphenoxy)benzoic acid. Spectrochim. Acta
A 2014, 121, 575‒585.
(7) Babu, P. D. S.; Periandy, S.; Ramalingam, S. Vibrational spectroscopic (FTIR and FTRaman)
investigation using ab initio (HF) and DFT (LSDA and B3LYP) analysis on the structure of
Toluic acid. Spectrochim. Acta A 2011, 78, 1321‒1328.
(8) Cinar, M.; Yildiz, N.; Karabacak, M.; Kurt, M. Determination of structural, spectrometric and
nonlinear optical features of 2-(4-hydroxyphenylazo)benzoic acid by experimental techniques
and quantum chemical calculations. Spectrochim. Acta A 2013, 105, 80‒87.
(10) Szafran, M.; Katrusiak, A.; Dega-Szafran, Z.; Kowalczyk, I. The structure of
4-(trimethylammonium)benzoic acid chloride studied by X-ray diffraction, DFT calculations,
NMR and FTIR spectroscopy. J. Mol. Struct. 2011, 996, 75‒81.
(11) Mukherjee, V.; Singh, N. P.; Yadav, R. A. Experimental and calculation aspects of
vibrational spectra and optimized geometry of 2,3,4-tri-fluoro-benzoic acid dimer. Spectrochim.
Acta A 2009, 74, 1107‒1114.
14
(12) Muthu, S.; IsacPaulraj, E. Spectroscopic and molecular structure (monomeric and dimeric
structure) investigation of 2-[(2-hydroxyphenyl) carbonyloxy] benzoic acid by DFT method: A
combined experimental and theoretical study. J. Mol. Struct. 2013, 1038, 145‒162.
(13) Chen, S. L.; Liu, Z.; Liu, J.; Han, G. C.; Li, Y. H. Synthesis, characterization, crystal
structure and theoretical approach of Cu(II) complex with 4-{(Z)-[(2-
hydroxybenzoyl)hydrazono]methyl} benzoic acid. J. Mol. Struct. 2012, 1014, 110‒118.
(14) Marinkovic, A. D.; Jovanovic, B. Z.; Assaleh, F. H.; Vajs, V. V.; Juranic, M. I. Linear free
energy relationships applied to the reactivity and the 13C NMR chemical shifts in 4-[[(substituted
phenyl)imino]methyl]benzoic acids. J. Mol. Struct. 2012, 1011, 158‒165.
(15) Montilla, F.; Rosa, V.; Prevett, C.; Avilés, T.; Nunes da Ponte, Mealli, D. M. C.
Trimethylsilyl-substituted ligands as solubilizers of metal complexes in supercritical carbon
dioxide. Dalton Trans. 2003, 2170‒2176.
(17) Bott, R.W.; Eaborn, C.; Jones, P.W. Organosilicon compounds. XXXVIII. The preparation
and optical resolution of p-[ethylmethyl(p-methoxyphenyl)silyl]benzoic acid. J. Organomet.
Chem. 1966, 6, 484‒489.
(19) Weinberg, J. M.; Wooley, K. L. The investigation of a transsilylation reaction for the
preparation of silyl esters: reactivity correlated with 29Si NMR resonance frequencies. J.
Organomet. Chem. 1997, 542, 235‒240.
(20) Bains, W.; Tacke, R. Silicon chemistry as a novel source of chemical diversity in drug
design, Curr. Opin. Drug. Discov. Devel. 2003, 6, 526‒543.
(21) Steinberg, X.; Lespay-Rebolledo, C.; Brauchi, S. A structural view of ligand-dependent
activation in thermo TRP channels. Front. Physiol. 2014, 5, Article 171.
(22) CrysAlis RED, Oxford Diffraction Ltd., Version 1.171.36.32, 2003.
15
(25) 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.; Caricato, M.; Li, X.;
Hratchian, H. P.; Izmaylov, A. F.; Bloino, J.; Zheng, G.; Sonnenberg, J. L.; Hada, M.; Ehara, M.;
Toyota, K.; Fukuda, R.; Hasegawa, J.; Ishida, M.; Nakajima, T.; Honda, Y.; Kitao, O.; Nakai, H.;
Vreven, T.; Montgomery, J. A.; Peralta, J. E.; Ogliaro, F.; Bearpark, M.; Heyd, J. J.; Brothers, E.;
Kudin, K. N.; Staroverov, V. N.; Kobayashi, R.; Normand, J.; Raghavachari, K.; Rendell, A.;
Burant, J. C.; Iyengar, S. S.; Tomasi, J.; Cossi, M.; Rega, N.; Millam, J. M.; Klene, M.; Knox, J.
E.; Cross, J. B.; Bakken, V.; Adamo, C.; Jaramillo, J.; Gomperts, R.; Stratmann, R. E.; Yazyev,
O.; Austin, A. J.; Cammi, R.; Pomelli, C.; Ochterski, J. W.; Martin, R. L.; Morokuma, K.;
Zakrzewski, V. G.; Voth, G. A.; Salvador, P.; Dannenberg, J. J.; Dapprich, S.; Daniels, A. D.;
Farkas, Ö.; Foresman, J. B.; Ortiz, J. V.; Cioslowski, J.; Fox, D. J. Gaussian 09. Gaussian. Inc.,
Wallingford CT, 2009. Official Website: http://www.gaussian.com/.
(26) Stewart, J. J. P. MOPAC 2012, Stewart Computational Chemistry, Version 14.139W. 2012,
Colorado Springs, Website: http://openmopac.net.
(27) Stewart, J. J. P. MOPAC: A semiempirical molecular orbital program. J. Comp. Aided Mol.
Des. 1990, 4, 1‒103.
(28) Maia, J. D. C.; Carvalho, G. A. U.; Mangueiram C. P.; Santana, S. R.; Cabral, L. A. F.;
Rocha, G. B. GPU linear algebra libraries and GPGPU programming for accelerating MOPAC
semiempirical quantumchemistry calculations. J. Chem. Theory Comput. 2012, 8, 3072–3081.
(32) Dennington, R.; Keith, T.; Millam, J. GaussView, Version 5, Semichem Inc., Shawnee
Mission, KS. 2009. Website: http://www.gaussian.com/g_prod/gv5.htm.
(33) Allouche, A. R. Gabedit – A graphical user interface for computational chemistry softwares.
J. Comput. Chem. 2011, 32, 174‒182.
(34) HyperChem (TM) Release 8.0, Hypercube, Inc., 1115 NW 4th Street, Gainesville, Florida
32601, USA, Website: http://www.hyper.com/
(35) Bernhard Schlegel, H. Geometry optimization. Comput. Mol. Sci. 2011, 1, 790‒809.
16
(36) Cojocaru, C.; Rotaru, A.; Harabagiu, V.; Sacarescu, L. Molecular structure and electronic
properties of pyridylindolizine derivative containing phenyl and phenacyl groups: Comparison
between semi-empirical calculations and experimental studies. J. Mol. Struc. 2013, 1034,
162‒172.
(37) Cojocaru, C.; Airinei, A.; Fifere, N. Molecular structure and modeling studies of azobenzene
derivatives containing maleimide groups. Springer Plus 2013, 2:586, 1‒19.
(38) Zeyrek, C. T.; Unver, H.; Arpaci, O. T.; Polat, K.; Iskeleli, N. O.; Yildiz, M. Experimental
and theoretical characterization of the 2-(4-bromobenzyl)-5-ethylsulphonyl-1,3-benzoxazole. J.
Mol. Struc. 2015, 1081, 22‒37.
(39) Saglam, E. G.; Ebinc, A.; Zeyrek, C. T.; Unver, H.; Hokelek, T. Structural studies on some
dithiophosphonato complexes of Ni(II), Cd(II), Hg(II) and theoretical studies on a
dithiophosphonato Ni(II) complex using density functional theory. J. Mol. Struc. 2015, 1099,
490‒501.
(41) Chitradevi, A.; Kumar, S. S.; Athimoolam, S.; Bahadur, S. A.; Sridhar, B. Single crystal
XRD, vibrational spectra, quantum chemical and thermal studies on a new semi-organic crystal:
4-Aminium antipyrine chloride. J. Mol. Struct. 2015, 1099, 58‒67.
(42) Mabkhot, Y. N.; Aldawsari, F. D.; Al-Showiman, S. S.; Barakat, A.; Soliman, S. M.;
Choudhary, M. I.; Yousuf, S.; Mubarak, M. S.; Hadda, T. B. Novel enaminone derived from [2,3-
b] thiene: Synthesis, x-ray crystal structure, HOMO, LUMO, NBO analyses and biological
activity. Chem. Central J. 2015, 9:24, 1‒11.
(43) Zhang, G.; Musgrave, C. B. Comparison of DFT methods for molecular orbital eigenvalue
calculations. J. Phys. Chem. A 2007, 111, 1554‒1561.
(44) Parr, R. G.; Szentpaly, L. V.; Liu, S. Electrophilicity Index. J. Am. Chem. Soc. 1999, 121,
1922‒1924.
(45) Sidir, I.; Sidir, Y. G.; Kumalar, M.; Tasal, E. Ab initio Hartree–Fock and density
functional theory investigations on the conformational stability, molecular structure and
vibrational spectra of 7-acetoxy-6-(2,3-dibromopropyl)-4,8-dimethylcoumarin molecule. J.
Mol. Struc. 2010, 964, 134‒151.
(46) Barakat, A.; Al-Majid, A. M.; Soliman, S. M.; Mabkhot, Y. N.; Ali, M.; Ghabbour, H. A.;
Fun, H. K.; Wadood, A. Structural and spectral investigations of the recently synthesized
chalcone (E)-3-mesityl-1-(naphthalen-2-yl) prop-2-en-1-one, a potential chemotherapeutic agent.
Chem. Cent. J. 2015, 9:35, 1‒15.
17
(47) Grimme, S. Calculation of the electronic spectra of large molecules. Rev. Comput. Chem.
2004, 20, 153‒218.
(48) Mahadevan, D.; Periandy, S.; Karabacak, M.; Ramalingam, S.; Puviarasan, N. Spectroscopic
(FT‒IR, FT‒Raman and UV‒vis) investigation and frontier molecular orbitals analysis on 3-
methyl-2-nitrophenol using hybrid computational calculations. Spectrochim. Acta A 2012,
86,139‒151.
(49) Fleming, S.; Mills, A.; Tuttle, T. Predicting the UV‒vis spectra of oxazine dyes. Beilstein J.
Org. Chem. 2011, 7, 432‒441.
(50) O'Boyle, N. M.; Tenderholt A. L.; Langner, K. M. A library for package-independent
computational chemistry algorithms. J. Comput. Chem. 2008, 29, 839‒845.
(51) O'Boyle, N. M. GaussSum open-source program, Version 3.0, 2013, available online at
http://gausssum.sourceforge.net/.
(52) Johnson III, R. D. NIST Computational Chemistry Comparison and Benchmark Database,
NIST Standard Reference Database Number 101, Release 16a, August 2013. CCCBDB III.B.3.a.
(XIII.C.1.); Precomputed vibrational scaling factors. Retrieved from web-page:
http://cccbdb.nist.gov/vibscalejust.asp.
(53) Socrates, G. Infrared and Raman characteristic group frequencies. Tables and Charts, John
Wiley & Sons, LTD, Chichester, 2001, 3rd edition, pp. 50‒246.
18
Table 1. Crystallographic data and structure refinement parameters for the compounds 1 and 2.
Compound 1 2
Empirical formula C11H16O3Si C13H 20O3Si
Formula weight 224.331 g/mol 252.385 g/mol
Temperature 200.00(14) °K 200.05(10) °K
Crystal system triclinic monoclinic
Space group P-1 P21/c
a = 6.6120(3) Å a = 16.508(13) Å
b = 7.6894 (4) Å b = 10.026(2) Å
c = 12.8140(10) Å c = 8.8482(15) Å
Unit cell dimensions
101.928(6)° 90.00°
98.644(6)° 95.89(3)°
100.425(4) ° 90.00°
Volume 614.86(6) Å3 1456.8(12) Å3
Z 2 4
Calculated density 1.212 g/cm3 1.151 g/cm3
Absorption coefficient, μ 0.177 mm-1 0.156 mm-1
F(000) 240.0 544.0
Crystal size 0.15 × 0.1 × 0.1 mm 0.1 × 0.05 × 0.05 mm
Radiation (wavelength) MoKα (λ = 0.71073 Å) MoKα (λ = 0.71073 Å)
2Θ range for data collection 5.54 to 46.5° 4.76 to 46.5°
Final R indexes [all data] R1 = 0.0931, wR2 = 0.2187 R1 = 0.1745, wR2 = 0.2275
Largest diff. peak and hole 0.88 and -0.38 e Å-3 0.34 and -0.25 e Å-3
19
Table 2. Summary of selected structural parameters: bond lengths (Å), angles (o deg) and
dihedrals (o deg), for the investigated molecules 1 (C11H16O3Si) and 2 ( C13H 20O3Si ).
Table 3. Root mean square deviation (RMSD) values to compare the goodness‒of‒fit between
observed and predicted structures of the investigated molecules 1 and 2.
20
Table 4. Summary of theoretical results obtained for optimized geometries of the investigated
molecule 1 (C11H16O3Si) using different QM methods.
DFT HF PM3 PM6 PM7
No. of occ. MO 60 60 41 41 41
Et, (kcal / mol) -592411 -589522 -57765 -57958 -58936
D, (Debye) 4.324 4.131 3.968 6.093 4.674
dmax, (Å) 11.967 11.909 11.812 11.598 11.865
Sm, (Å2) 458.61 454.29 464.16 462.98 456.92
3
Vm, (Å ) 735.71 730.11 734.35 734.84 730.44
E LUMO+1 (eV) -0.623 1.697 0.163 0.124 -0.039
E LUMO (eV) -1.396 1.212 -0.324 -0.347 -0.412
E HOMO (eV) -6.528 -8.851 -9.287 -9.117 -9.310
E HOMO-1 (eV) -7.332 -9.587 -10.534 -10.172 -10.285
E, (eV) 5.132 10.063 8.963 8.770 8.898
E1 , (eV) 6.709 11.284 10.697 10.296 10.246
IP (eV) 6.528 8.851 9.287 9.117 9.310
EA (eV) 1.396 -1.212 0.324 0.347 0.412
(eV) 3.962 3.820 4.806 4.732 4.861
µ (eV) -3.962 -3.820 -4.806 -4.732 -4.861
(eV) 2.566 5.032 4.482 4.385 4.449
(eV) 3.059 1.450 2.576 2.553 2.656
Table 5. Summary of theoretical results obtained for optimized geometries of the investigated
molecule 2 ( C13H 20O3Si ) using different QM methods.
DFT HF PM3 PM6 PM7
No. of occ. MO 68 68 47 47 47
Et, (kcal / mol) -641757 -638518 -64663 -64873 -65853
D, (Debye) 4.098 3.937 3.645 5.674 4.329
dmax, (Å) 13.946 13.892 13.869 14.410 13.814
Sm, (Å2) 522.38 515.47 520.50 520.46 511.10
3
Vm, (Å ) 837.37 831.29 833.35 838.23 832.64
E LUMO+1 (eV) -0.647 1.643 0.088 -0.009 -0.088
E LUMO (eV) -1.418 1.311 -0.388 -0.472 -0.459
E HOMO (eV) -6.555 -8.878 -9.406 -9.307 -9.384
E HOMO-1 (eV) -7.358 -9.616 -10.156 -9.984 -10.332
E, (eV) 5.137 10.189 9.018 8.835 8.925
E1 , (eV) 6.711 11.259 10.244 9.975 10.244
IP (eV) 6.555 8.878 9.406 9.307 9.384
EA (eV) 1.418 -1.311 0.388 0.472 0.459
(eV) 3.986 3.784 4.897 4.890 4.922
µ (eV) -3.986 -3.784 -4.897 -4.890 -4.922
(eV) 2.568 5.095 4.509 4.418 4.463
(eV) 3.094 1.405 2.659 2.706 2.714
21
Table 6. Summary of data related to the observed and theoretical electronic absorption spectra
for the investigated compounds 1 and 2 (absorption wavelength, (nm); excitation energy,
E (eV) and oscillator strength, f).
Table 7. Observed vibrational frequencies and the corresponding main assignments for the
investigated compounds 1 and 2.
22
CH3
O
C OH Cl Si CH3
HO n
CH3
110 oC DMF
8h K2CO3
CH3
O
C O Si CH3
HO n CH
3
23
Fig. 2. Molecular structures of the investigated compounds determined by X‒ray structure
analysis (ORTEP‒3 view): (a) 1 (C11H16O3Si), 4‒((trimethylsilyl)methoxy)benzoic acid;
(b) 2 ( C13H 20O3Si ), 4‒(3‒(trimethylsilyl)propoxy)benzoic acid.
24
Fig. 3. Partial packing diagrams for the compounds 1 (a) and 2 (b) showing the dimmers
structures formed via hydrogen-bond bridges (O…H-O). Thermal ellipsoids are drawn at 50%
probability level. H-bonds parameters of 1.: O4−H∙∙∙O3 [O4−H 0.82 Å, H∙∙∙O3 1.83 Å, O4∙∙∙O3
(3 – x, 3 - y, 1 - z) 2.625(6) Å, O4−H∙∙∙O3 161.5°]. H-bonds parameters of 2: O2−H∙∙∙O6 [O2−H
0.82 Å, H∙∙∙O6 1.80 Å, O2∙∙∙O6 (2 – x, 1 - y, 2 - z) 2.613(6) Å, O2−H∙∙∙O6 171.0°].
25
Fig. 4. Optimized molecular geometries of the investigated compounds 1 (a) and 2 (b), computed
at ground state (S0) by DFT method (B3LYP/6‒31+G**); balls and sticks rendering models with
full atomic numbering.
26
Fig. 5. Histograms of interatomic angles distributions (a) and bond lengths distributions (b) for
the investigated compounds 1 and 2.
27
Fig. 6. Superposition of the X‒ray structures (blue) of the investigated compounds and their
optimized counterparts (red) computed by different methods: (a) 1 (HF/6‒31+G**); (b) 1 (PM3);
(c) 2 (B3LYP/6‒31+G**); (d) 2 (PM6).
28
Fig. 7. Energy spectra of molecular orbitals (eigenvalues profiles) computed by PM3 method for
the investigated molecules 1 (a) and 2 (b) in their ground state (S0).
29
Fig. 8. Plots of frontier molecular orbitals for optimized structures of investigated compounds 1
and 2 computed at the level of DFT method (B3LYP/6‒31+G**): (a), (b), (c), (d) molecular
orbitals for 1; (e), (f), (g), (h) molecular orbitals for 2.
30
Fig. 9. Atomic charges distributions along with dipole moments orientations and electrostatic
potential surfaces for the investigated compounds 1 and 2; computations performed at the level of
DFT theory (B3LYP 6‒31+G**): (a) atomic charge distribution for 1; (b) electrostatic potential
map for 1; (c) atomic charge distribution for 2; (d) electrostatic potential map for 2.
31
Fig.10. UV‒Vis spectra of the investigated compounds in DMSO solvent; (a) Experimental
UV‒Vis spectra for the compounds 1 and 2; (b) Theoretical electronic absorption spectrum
(TD‒DFT) for the compound 1; (c) Theoretical electronic absorption spectrum (TD‒DFT) for
the compound 2.
32
Fig.11. IR spectra of the investigated compounds: experimental vs. theoretical
(B3LYP/6‒31+G**); (a) IR spectra for the compound 1; (b) IR spectra for the compound 2.
33
Fig.12. Parity plots between experimental and theoretical (scaled) wavenumbers;
(a) parity plot for the compound 1; (b) parity plot for the compound 2; theoretical vibrational
frequencies calculated by B3LYP/6‒31+G** method.
34