Diego M.
Gil
a,
María E.
Tuttolomondo
a,
Sebastian
Blomeyer
b,
Christian G.
Reuter
b,
Norbert W.
Mitzel
b and
Aída Ben
Altabef‡
*a
aINQUINOA (CONICET-UNT), Instituto de Química Física, Facultad de Bioquímica, Química y Farmacia, Universidad Nacional de Tucumán, San Lorenzo 456, T4000CAN. S. M. de Tucumán. R., Argentina. E-mail: altabef@fbqf.unt.edu.ar; Fax: +54 381 4248169; Tel: +54 381 4311044
bInorganic and Structural Chemistry, Centre for Molecular Materials CM2, Faculty of Chemistry, Bielefeld University, Universitätsstraße 25, 33615 Bielefeld, Germany
First published on 23rd November 2015
The molecular structure and conformational properties of 2,2,2-trichloroethyl chloroformate, ClC(O)OCH2CCl3 were determined experimentally using gas-phase electron diffraction (GED) and theoretically based on quantum-chemical calculations at the MP2 and DFT levels of theory. Further experimental measurements such as UV-visible, IR and Raman spectroscopy were complemented with the corresponding theoretical studies. All experimental results and calculations confirm the presence of two conformers namely anti–gauche (C1 symmetry) and anti–anti (Cs symmetry). The conformational preference was rationalised by NBO and AIM analyses. Molecular properties such as ionisation potential, electronegativity, chemical potential, chemical hardness and softness were deduced from HOMO–LUMO analyses. The TD-DFT approach was applied to assign the electronic transitions observed in the UV-visible spectrum. A detailed interpretation of the infrared and Raman spectra of the title compound are reported. Using calculated frequencies as a guide, IR and Raman spectra also provide evidence for the presence of both C1 and Cs conformers.
The compound ClC(O)OCH2CCl3 is commercially available, but its molecular structure has not yet been studied. In this contribution we report on the gas-phase structure of the title compound determined by gas-phase electron diffraction (GED) and on infrared and Raman spectra recorded in the liquid phase. We compare these experimental data to results of quantum-chemical calculations. Further quantum-chemical exploration of structure and bonding properties concern the internal rotation about the O–CH2 bond indicating that the most stable conformation is anti–anti (Cs symmetry), natural bond orbital (NBO) and AIM analyses to rationalise these results as well as exploration of the HOMO and LUMO frontier molecular orbitals that results in information on ionisation potential (IP), electron affinity (EA), electronegativity (χ), electrophilicity index (ω), hardness (η) and chemical potential (μ). These findings are compared to those previously reported for different acetates6–9 and chloroformates.10–12
Raman intensities were predicted by the procedure outlined below. The Raman activities (Si) were calculated by GAUSSIAN 03 and converted into relative Raman intensity (Ii) using the following relation from the basic theory of Raman scattering:28
(1) |
Natural bond orbital (NBO) calculations were performed at the B3LYP/6-311++G(d,p) level using the NBO 3.1 program29 as implemented in GAUSSIAN 03. These analyses were performed in order to understand various second order interactions between the filled orbitals of one subsystem and vacant orbitals of another subsystem to consequently obtain a measure of intramolecular delocalisation and hyperconjugation. In addition, analyses of the reactivity of the compound was done within Bader's Quantum Theory of Atoms in Molecules (QTAIM) by using the AIM2000 code.30,31
Molecular properties such as ionisation potential, electronegativity, chemical potential, chemical hardness and softness have been deduced from HOMO–LUMO-analysis employing B3LYP/6-311++G(d,p) and PBEPBE/6-311G(3df,2pd) methods. The electronic absorption spectra for the optimised structures were calculated using time dependent DFT (TD-DFT) at PBEPBE/6-311G(3df,2pd) level for the free molecules as well as within the polarisable continuum model (PCM) with methanol as solvent.
Fig. 1 Torsional potential around O(3)–C(4) bond of ClC(O)OCH2CCl3 calculated at B3LYP, mPW1PW91 and MP2 levels of theory using 6-311++G(d,p) and 6-311G(d,p) basis sets. |
The difference of free B3LYP/6-311++G(d,p) energies, is 1.5 kJ mol−1 indicating the Cs conformation to be more stable than C1. These results agree well with those reported for different compounds of similar structure.6–9 Using the Boltzmann distribution and this difference in free energy, the conformational composition was estimated to be 0.52:0.48 (C1:Cs) at room temperature, taking into account a multiplicity of 2 for the C1 conformer.
The quantum-chemically optimised structural parameters for both conformers of ClC(O)OCH2CCl3 are listed in Table 1. The experimental parameters obtained by gas-phase electron diffraction (GED) are also presented in Table 1 for comparison; they are described in Section 4.3. According to Table 1, structural parameters calculated at the MP2/6-311++G(d,p) level reproduce most experimental gas-phase parameters within experimental error. Some differences between calculated and experimental parameters are observed for the CO, C–O and C–Cl bond lengths. The theoretical description of molecules containing chlorine atoms requires the use of highly polarised basis functions. As was found for related compounds containing sulfonate groups,32–34 the inclusion of extra polarisation functions (beyond a single d function) is necessary to predict the bond lengths in these types of molecules accurately. Table S2 (ESI†) lists the structural parameters obtained using the PBEPBE and MP2 methods with 6-311G(3df,2pd) basis set. These calculations produce a structure close to the experimental one observed in related molecules.8,12,35,36 However, a better match of bond lengths and angles experimentally determined from GED measurements was achieved by the MP2/6-311G(3df,2pd) estimates.
Parametera | C 1 conformer (re) | C s conformer (re) | GED (re)b | ||||
---|---|---|---|---|---|---|---|
B3LYP | mPW1PW91 | MP2 | B3LYP | mPW1PW91 | MP2 | ||
a Bond lengths in Å, angles in degrees. See Fig. 2 for numbering of atoms. b Values for conformer of Cs symmetry, standard deviations given as 3σLS, superscript numbers indicate the regularisation coefficient, subscript letters state if parameters were refined in groups with fixed differences in between. The O(2)C(1)–O(3) angle was not refined explicitly, but results from O(2)C(1)–Cl(8), Cl(8)–C(1)–O(3) and the assumed planarity of the Cl(8)–C(1)–O(2)–O(3) moiety. | |||||||
C(4)–H (mean) | 1.089 | 1.089 | 1.090 | 1.090 | 1.090 | 1.092 | 1.090(15)5.0 |
C(1)–Cl(8) | 1.764 | 1.744 | 1.738 | 1.764 | 1.745 | 1.738 | 1.744(2)a |
C(1)O(2) | 1.187 | 1.184 | 1.195 | 1.189 | 1.186 | 1.197 | 1.184(8)5.0 |
C(1)–O(3) | 1.339 | 1.332 | 1.346 | 1.334 | 1.327 | 1.338 | 1.327(9)5.0 |
O(3)–C(4) | 1.437 | 1.423 | 1.428 | 1.439 | 1.426 | 1.432 | 1.422(10)5.0 |
C(4)–C(5) | 1.531 | 1.524 | 1.525 | 1.527 | 1.519 | 1.520 | 1.535(10)5.0 |
C(5)–Cl | 1.795 | 1.777 | 1.771 | 1.795 | 1.777 | 1.771 | 1.772(2)a |
O(2)C(1)–Cl(8) | 123.5 | 123.6 | 124.2 | 123.9 | 123.9 | 124.5 | 124.9(17)0.5 |
O(2)C(1)–O(3) | 128.1 | 127.8 | 127.7 | 127.3 | 127.0 | 126.7 | 126.2(9) |
Cl(8)–C(1)–O(3) | 108.4 | 108.5 | 108.0 | 108.9 | 109.0 | 108.7 | 108.9(14)0.5 |
C(1)–O(3)–C(4) | 116.9 | 116.7 | 115.7 | 114.8 | 114.4 | 112.9 | 110.6(23)0.5 |
O(3)–C(4)–C(5) | 110.1 | 110.1 | 109.3 | 108.1 | 108.1 | 107.1 | 105.1(10)0.5 |
Cl–C(5)–Cl | 109.8 | 109.9 | 110.4 | 109.7 | 109.9 | 110.3 | 109.2(9)0.5 |
Cl8–C1–O3–C4 | 178.7 | 178.4 | 176.2 | 179.9 | 179.9 | 179.9 | 179.1(24)0.5 |
C1–O3–C4–C5 | 121.3 | 119.2 | 112.9 | 179.9 | 179.9 | 179.9 | Fixed to 180.0 |
Thermodynamic parameters of both conformations of ClC(O)OCH2CCl3 were also computed. These calculations were performed in order to get reliable data from which relations among energy, structure and reactivity characteristics of the molecule can be obtained. Table S4 (ESI†) shows values of some calculated thermodynamic parameters (such as thermal energy, heat capacity, entropy, zero-point vibrational energies (ZPVEs), rotational constants and rotational temperatures) of both conformers of ClC(O)OCH2CCl3 (B3LYP/6-311++G(d,p)). All values agree well with literature data.6–9 All thermodynamic parameters are smaller in magnitude for the Cs than for the C1 conformer. Dipole moments of C1 and Cs conformers are 1.75 and 1.11 Debye, respectively. Fig. S1 (ESI†) shows a representation of the two conformers with the corresponding directions of dipole moments.
Natural bond orbital (NBO) analyses is a useful tool for understanding delocalisation of electron density from occupied Lewis-type (donor) NBOs into unoccupied non-Lewis type (acceptor) NBOs within a molecule. Table 2 shows the most relevant hyperconjugation interactions for both conformers of ClC(O)OCH2CCl3 (NBO: B3LYP/6-311++G(d,p)). According to the NBO analyses, the stabilising character of the hyperconjugation interactions is more pronounced in the Cs than in the C1 conformer. These results suggest a higher stability of Cs over C1. Similar results were obtained for the related compound CH3C(O)OCH2CCl3.6 The entries of Table 2 demonstrate the hyperconjugative effect LP(O(3)) → σ*(C(4)–C(5)) to be stronger in the anti–gauche than in the anti–anti conformer, indicating that this interaction is important for stabilising the C1 conformer. This interaction helps to rationalise the large dipole moment of the C1 compared with the Cs conformer. In both conformers the lone pair LP(Cl(8)) participates in LP(Cl(8)) → σ*(C(1)–O(2)) and LP(Cl(8)) → σ*(C(1)–O(3)) type interactions. Also in both, the LP(Cl(8)) → σ*(C(1)–O(2)) delocalisation strongly stabilises the molecule. Another strongly stabilising interaction is LP(O(3)) → σ*(C(1)–O(2)) with 229 and 233 kJ mol−1 for conformers C1 and Cs, respectively. Anomeric interactions promoted by electron donation from in-plane (σ) oxygen lone pairs directly affect the bond length of the carbonyl group, mainly in the Cs conformer. A longer bond is expected when the interaction between lone pairs and the antibonding orbital of the carbonyl group increases. As shown in Tables 1 and 2, the CO bond of the Cs conformer is longer than in the C1 one, a fact attributed to the strong LP(O(3)) → σ*(C(1)–O(2)) interaction. The relation between the electron occupation of the σ*(C(4)–O(3)), σ*(C(4)–C(5)) as well as σ*(C(1)–O(2)) and the bond lengths C–C, C–O and CO was investigated in both conformers (results see Table S5, ESI†). The C(4)–O(3) bond in the anti–anti conformer is longer than that of the anti–gauche conformer, which is in agreement with the high occupation of the σ*(C(4)–O(3)) orbital in the Cs conformer. The strength of the LP(O(3)) → σ*(C(4)–C(5)) interaction in the C1 conformer (see Table 2) causes a lengthening of the C–C bond and a shortening of the C–O bond. The latter is attributed to the lower occupation of the σ*(C(4)–O(3)) orbital.
Interaction (donor → acceptor)a | anti–gauche (C1) | anti–anti (Cs) |
---|---|---|
a LP indicates electron lone pair on the specified atom (see Fig. 2 for numbering of atoms). | ||
LP(O(2)) → σ*(C(1)–Cl(8)) | 170 | 175 |
LP(O(2)) → σ*(C(1)–O(3)) | 132 | 131 |
LP(O(3)) → σ*(C(1)–O(2)) | 229 | 233 |
LP(O(3)) → σ*(C(4)–C(5)) | 17 | 5 |
LP(O(3)) → σ*(C(1)–Cl(8)) | 5 | 5 |
LP(O(3)) → σ*(C(4)–H(6)) | 11 | 22 |
LP(O(3)) → σ*(C(4)–H(7)) | 19 | 22 |
LP(Cl(8)) → σ*(C(1)–O(2)) | 116 | 117 |
LP(Cl(8)) → σ*(C(1)–O(3)) | 25 | 25 |
LP(Cl(9)) → σ*(C(4)–C(5)) | 19 | 20 |
LP(Cl(10)) → σ*(C(4)–C(5)) | 16 | 16 |
LP(Cl(11)) → σ*(C(4)–C(5)) | 20 | 20 |
Total | 781 | 790 |
Atomic charges affect molecular polarisabilities, dipole moments, electronic structures and even more molecular properties. Here we use Mulliken charge calculations to investigate them. The charge distributions of C1 and Cs conformers of the title compound were calculated at the B3LYP/6-311++G(d,p) level of theory; they are listed in Table S6 (ESI†). In both conformers the atomic charges for the three different carbon atoms C(1), C(4) and C(5) are found to be +0.76 e, −0.09 e and −0.18 e, respectively. For the oxygen and chlorine atoms there is no significant difference in atomic charges between the two conformers, either. When changing from the anti–anti to anti–gauche, the hydrogen atoms become chemically inequivalent. Hence, their atomic charges differ marginally at +0.22 e and +0.24 e in the C1 conformer, whereas in the Cs conformer they are the same (+0.22 e). The negative charges on O(2) and O(3) makes C(1) positively charged and therefore a preferred site for nucleophilic attack. The negative charges are mainly located on O(2) and O(3). Hence, these atoms are supposed to interact with the positive part of a receptor.
The Quantum Theory of Atoms in Molecules (QTAIM) is used to characterise bonding interactions through a topological analysis of the electron densities.30 In QTAIM the nature of bonding interactions is determined by the charge density ρ and its Laplacian ∇2(ρ) at the bond critical points (BCPs). Atomic charges within molecules are obtained by integrating the charge density over atomic basins defined by charge density topology.30Table 3 lists characteristic QTAIM parameters and Fig. S2 (ESI†) shows the molecular graphs consisting of atomic interactions lines (also called bond paths) and bond critical points for both conformers of ClC(O)OCH2CCl3. The charge densities at the BCP along C(1)–O(3) are relatively high for both conformers and the corresponding Laplacians of the electron densities are negative. This means the charge density is located in the internuclear region and indicates highly covalent character for these bonds. The charge density at the BCP in the Cs (0.312 e a0−3) and the C1 conformer (0.309 e a0−3) vary only insignificantly. The electron density at the CO BCPs in the C1 conformer (0.435 e a0−3) and the Cs conformer (0.434 e a0−3) are also the same and the Laplacian of the charge density is negative, too. These results agree with the NBO data (see Table 2) that give similar values for the interaction LP(O(3)) → σ*(C(1)–O(2)) in both conformers.
anti–gauche | anti–anti | |
---|---|---|
C(1)–O(3) | ||
ρ | 0.309 | 0.312 |
∇2(ρ) | −0.57 | −0.54 |
d(BCP–C(1)) | 0.886 | 0.877 |
d(BCP–O(3)) | 1.646 | 1.645 |
d/Å | 1.339 | 1.334 |
q(C(1)) | −0.364 | −0.277 |
q(O(3)) | 0.134 | 0.125 |
ν/cm−1 | 1143 | 1167 |
C(1)O(2) | ||
ρ | 0.435 | 0.434 |
∇2(ρ) | −0.04 | −0.05 |
d(BCP–C(1)) | 0.767 | 0.768 |
d(BCP–O(2)) | 1.476 | 1.478 |
d/Å | 1.187 | 1.189 |
q(C(1)) | −0.364 | −0.277 |
q(O(2)) | −0.143 | −0.152 |
ν/cm−1 | 1847 | 1840 |
O(3)–C(4) | ||
ρ | 0.239 | 0.239 |
∇2(ρ) | −0.354 | −0.378 |
d(BCP–O(3)) | 1.773 | 1.771 |
d(BCP–C(4)) | 0.945 | 0.949 |
d/Å | 1.437 | 1.439 |
q(O(3)) | 0.134 | 0.125 |
q(C(4)) | −0.421 | −0.577 |
ν/cm−1 | 985 | 1002 |
The energies and topologies of the highest occupied molecular orbital (HOMO) and lowest unoccupied molecular orbital (LUMO) are important parameters for predicting chemical reactivity. The HOMO acts primarily as an electron donor and the LUMO as electron acceptor and their energies correspond to the ionisation potential (IP) and the electron affinity (EA): IP = −EHOMO; EA = −ELUMO. The energy difference between HOMO and LUMO (ΔEHOMO–LUMO) is an important parameter to estimate the molecular chemical stability as well as electrical transport by means of electron conductivity.37 In order to understand biological properties including drug design and the possible eco-toxicological characteristics of drug molecules, several new chemical reactivity descriptors have been proposed. Conceptual DFT based descriptors have helped in many ways to understand the structure of molecules and their reactivity by calculating the chemical potential, global hardness and electrophilicity index. Using HOMO and LUMO energy values for both conformers, the electronegativity (χ), chemical potential (μ), chemical hardness (η), chemical softness (S) and global electrophilicity index (ω) can be calculated by using the following equations: χ = (IP + EA)/2, μ = −(IP + EA)/2, η = (IP − EA)/2, S = 1/(2η), ω = μ2/(2η), where IP and EA are the ionisation potential and electron affinity, respectively.38 The above-mentioned parameters were calculated at B3LYP/6-311++G(d,p) and PBEPBE/6-311G(3df,2pd) levels of theory and are listed in Table 4. Using the B3LYP method, the HOMO–LUMO energy gap is predicted to be 7.14 and 7.16 eV for C1 and Cs conformers, respectively. The energy gaps obtained from the PBEPBE calculations are comparable but slightly lower. The energy gap of both conformers is relatively large indicating that both exhibit high chemical stability and lower reactivity. The global electrophilicity index measures the stabilisation energy when the system acquires an additional electronic charge from the environment. Electrophilicity encompasses both, the ability of an electrophile to acquire additional electronic charge and the resistance of the system to exchange electronic charge with the environment. It contains information about electron transfer (chemical potential) as well as the stability (hardness) and is the better descriptor of global chemical reactivity. The hardness signifies the resistance towards deformation of the electronic cloud of chemical systems under small perturbations encountered during chemical processes. Molecules with large HOMO–LUMO gap are considered hard, those with small HOMO–LUMO gap soft. This concept relates molecular stability to hardness in a way that smaller HOMO–LUMO gaps are associated with more reactive species.
Parameter | C 1 conformer | C s conformer | ||
---|---|---|---|---|
B3LYPa | PBEPBEb | B3LYP | PBEPBE | |
a B3LYP/6-311++G(d,p). b PBEPBE/6-311G(3df,2pd). | ||||
E HOMO/eV | −8.915 | −7.558 | −8.932 | −7.544 |
E LUMO/eV | −1.778 | −2.445 | −1.776 | −2.454 |
ΔEHOMO–LUMO/eV | 7.137 | 5.113 | 7.156 | 5.090 |
Electronegativity χ/eV | 5.346 | 5.001 | 5.354 | 4.999 |
Chemical potential μ | −5.346 | −5.001 | −5.354 | −4.999 |
Chemical hardness η/eV | 3.569 | 2.556 | 3.578 | 2.545 |
Chemical softness, S/eV | 0.140 | 0.196 | 0.139 | 0.196 |
Global electrophilicity index, ω/eV−1 | 4.004 | 4.892 | 4.006 | 4.909 |
Dipole moment, μ/D | 1.755 | 1.598 | 1.112 | 0.995 |
Molecular orbitals are defined as Eigen-functions of the Fock operator, which exhibits the full symmetry of the nuclear point group. They necessarily from a basis for irreducible representations of full point-group symmetry. The pictorial illustration of the frontier molecular orbitals and their respective positive and negative regions for both conformers of the title compound is shown in Fig. 3. Both HOMO and LUMO frontier molecular orbitals play an important role in the electrical and optical properties, as well as in the UV-visible spectra and chemical reactions. The HOMO of the anti–gauche conformer (Fig. 3a) is primarily composed of p-type orbitals located on oxygen and chlorine atoms. The LUMO is spread over the entire molecule except the chloroformate group. The HOMO of the anti–anti conformer (Fig. 3b) is composed of p-type orbitals located on the CCl3 moiety. The LUMO plot reveals that this molecular orbital is mostly spread over the CH2CCl3 group.
Fig. 4 shows the experimental UV-visible spectrum measured in methanol solution and the calculated ones for both conformations of ClC(O)OCH2CCl3. Over all a good agreement between calculated and experimental spectra is observed. Absorption bands for a HOMO−2 → LUMO transition calculated including a PCM model for solvation in methanol appear at 215 nm for the C1 conformer and at 213 nm for the conformer with Cs symmetry. According to these calculations, we can assign the experimentally observed absorptions at 211 nm and 209 nm to the C1 and Cs conformers, respectively. Other transitions were obtained in quantum-chemical calculations corresponding to the HOMO → LUMO and HOMO−1 → LUMO transitions but they were not observed experimentally, probably due to the smaller oscillator strengths (f) calculated for both conformers.
Fig. 4 UV-visible spectrum measured in methanol as well as the calculated ones for both conformations of ClC(O)OCH2CCl3 (PBEPBE/6-311G(3df,2pd)/PCM(methanol)). |
Nine pseudo conformers with C(1)–O(3)–C(4)–C(5) dihedral angles in the range of 100°–180° in steps of 10° were used and weighted by means of their relative energy via a Boltzmann distribution at the temperature of experiment (328 K). Differences in parameters between the pseudo conformers were kept fixed to their computed values. Thus, sixteen independent geometrical parameters were refined during the least-squares procedure. Refinement of most of these was supported by flexible restraints along the ideas of Bartell et al.39 and the SARACEN40 method as are implemented in the UNEX program.41
Quadratic and cubic force fields were calculated at the MP2/6-311G(d,p) level of theory for both conformers and, using the SHRINK program,42 two sets of amplitudes and anharmonic vibrational corrections were obtained. Each of these was used for a certain range of pseudo conformers (anti–anti: 150° ≤ Φ ≤ 180°, anti–gauche: 100° ≤ Φ ≤ 140°). Scale factors of vibrational amplitudes were refined in groups corresponding to distinct peaks on the radial distribution curve, whereas the ones of the C(1)–O(2), C(1)–O(3), O(3)–C(4) and C(4)–C(5) bond lengths had to be refined individually with weak restraints. Experimental structural data are summarised in Table 1 and the radial distribution curve is shown in Fig. 5. The finally achieved R-factor was RG = 5.0%.
The C(1)–O(3)–C(4) angle, 110.6(23)° (for Cs) in GED, exhibits the most significant difference between the experiment and the different levels of theory, as the latter yield it at least 2° wider. Despite the sterically demanding CCl3 group, this angle is smaller than in the ethyl esters F3CC(O)OCH2CF3 (114.4°),7 H3CC(O)OCH2CH3 (115.7°)43 and in the methyl esters H3CC(O)OCH3 (116.4°)44 as well as F3CC(O)OCH3. Structure refinement of the latter was done with a small-amplitude motion model for the rotation of the CF3 group8 as well as with a dynamic model,45 resulting in a C–O–C angle of 114.2° and 116.3°, respectively. Surprisingly this angle is wider for molecules bearing sterically less demanding substituents at O(3). This effect can be neglected within the error limit for ethyl acetate but is of higher significance for F3CC(O)OCH2CF3 and ClC(O)OCH2CCl3. Additionally, this angle is found at 117.6° in ClC(O)OCF3.46 So, one can conclude that the electronic effect of the chlorine substituent at C(1) on this angle is quite small and, if at all, widens it. Therefore, the narrow C–O–C angle in ClC(O)OCH2CCl3 could be attributed to an attractive interplay between the CCl3 or CF3 groups and the carbonyl group, however, NBO analyses did not provide any indication for this in our case.
The C(1)–O(3) and O(3)–C(4) bond lengths depend on the strength of the electron-withdrawing nature of the substituents at C(1) and – if present – C(5), as one can see for the above-mentioned molecules. The shortest C(1)–O(3) bond can be observed in F3CC(O)OCH37 and ClC(O)OCH2CCl3 with 1.326 Å and 1.327 Å, respectively. The unsubstituted molecules show significantly longer C(O)–O bond lengths of 1.345 Å (H3CC(O)OCH2CH3)43 and 1.360 Å (H3CC(O)OCH3).45 The same applies to the O(3)–C(4) bond, which is determined to 1.421 Å in F3CC(O)OCH3 and ClC(O)OCH2CCl3, but is longer in methyl and ethyl acetate with 1.443 Å and 1.448 Å, respectively. Interestingly these bonds are shorter in the title compound than in F3CC(O)OCH2CF3 (C(O)–O: 1.336 Å, O–C: 1.423 Å), which one could expect to show a similar or even stronger electron-with drawing effect of the substituents.
Additionally it should be noted that in all quantum-chemical calculations the O(3)–C(4)–C(5) angle is 2–3° larger than in the GED structure refinement at 105.1(10)°. In contrast to the above-mentioned C(1)–O(3)–C(4) angle, the O(3)–C(4)–C(5) angle is in good compliance with those determined for other ethyl esters, namely CH3C(O)OCH2CH3 (108.2°)43 and CF3C(O)OCH2CF3 (107.8°).7 Note, that for this angle we observe significant differences between theory and experiment for all ethyl esters investigated to date.
Due to the complex shape and low height of the potential energy profile, it was not possible to find a parameterised function suitable for refining its parameters and, hence, the rotational barrier itself. Therefore the refinement of GED data did not allow providing an experimental ratio of conformers, however, the failure of one- and two-conformer models proves the existence of both, anti–anti and anti–gauche conformers in the gas-phase and justifies the description by a dynamic model. Regarding the potential used to describe the populations of the pseudo conformers we conclude that both of them supply a good description of the gas-phase composition, since they yield the same structural parameters as well as R-factors.
Mode | Experimental | B3LYP/6-311++G(d,p)c | Approximate description of moded | ||
---|---|---|---|---|---|
IR (liq.)a | Raman (liq.)b | C 1 conf. | C s conf. | ||
a sh, shoulder; s, strong; w, weak; m, medium; v, very. b Relative band heights in parentheses. c IR intensities are shown in parenthesis. d ν, stretching; δ, bending; ρ, rocking; ω, wagging; τω, twisting; τ, torsion. See Fig. 2 for numbering of atoms. | |||||
1 | 3022 w | 3018 (8) | 3158 (3) | 3142 (<1) | ν a CH2 |
2 | 2967 w | 2966 (31) | 3092 (11) | 3081 (6) | ν s CH2 |
3 | 1782 vs | 1781 (7) | 1847 (376) | 1840 (412) | ν CO |
4 | 1445 w (Cs) | 1444 (6) | 1478 (14) | 1482 (13) | δ CH2 |
1440 sh (C1) | 1439 (5) | ||||
5 | 1370 w (Cs) | 1370 (2) | 1402 (17) | 1405 (13) | ω CH2 |
1362 sh (C1) | 1360 sh | ||||
6 | 1276 w (C1) | 1274 (3) | 1305 (38) | 1278 (16) | τω CH2 |
1257 sh (Cs) | 1256 (2) | ||||
7 | 1156 sh (Cs) | 1153 sh | 1143 (728) | 1167 (857) | ν C(1)–O(3) |
1136 s (C1) | 1135 (2) | ||||
8 | 1068 sh (Cs) | 1067 sh | 1077 (154) | 1094 (35) | ν C(4)–C(5) |
1060 w (C1) | 1060 (4) | ||||
9 | 1050 sh (C1) | — | 1072 (18) | 1055 (30) | ρ CH2 |
1031 vw (Cs) | |||||
10 | 983 vw | 984 (10) | 985 (9) | 1002 (3) | ν O(3)–C(4) |
11 | 878 vw (Cs) | 877 (10) | 845 (36) | 883 (32) | ν C(1)–Cl(8) |
839 vw (C1) | 839 (17) | ||||
12 | 802 m | 803 (9) | 777 (173) | 765 (209) | ν a CCl3 |
13 | 721 m | 721 (14) | 695 (122) | 701 (136) | ν a CCl3 |
14 | 680 w | 680 (1) | 682 (15) | 685 (8) | δ out-of-plane CO |
15 | 576 w | 575 (56) | 564 (32) | 568 (41) | ν s CCl3 |
16 | 498 vvw | 494 (24) | 489 (40) | 482 (14) | δ out-of-plane CO + O(2)–C(1)–Cl(8) |
17 | 467 vvw | 468 (26) | 471 (11) | 457 (14) | δ Cl(8)–C(1)–O(3) |
18 | — | 383 (100) | 382 (4) | 377 (5) | δ s CCl3 |
19 | — | 349 (10) | 347 (2) | 347 (2) | δ a CCl3 |
20 | — | 287 (47) | 287 (<1) | 287 (<1) | δ a CCl3 |
21 | — | 260 (10) (C1) | 272 (1) | 255 (1) | δ C(1)–O(3)–C(4) |
247 (20) (Cs) | |||||
22 | — | 231 (6) | 229 (<1) | 207 (2) | δ O(3)–C(4)–C(5) |
23 | — | 206 (28) | 204 (1) | 201 (<1) | ρ CCl3 |
24 | — | 186 (22) | 181 (1) | 175 (<1) | τ C(4)–C(5) |
25 | — | — | 67 (1) | 86 (1) | ρ CCl3 |
26 | — | — | 48 (<1) | 48 (0) | τ CCl3 |
27 | — | — | 32 (1) | 23 (2) | τ C(4)–O(3) |
Fig. 7 IR spectrum calculated for both conformers of ClC(O)OCH2CCl3 with the experimental one in the range of 1300–1000 cm−1. |
Footnotes |
† Electronic supplementary information (ESI) available. See DOI: 10.1039/c5cp05295e |
‡ Member of the Research Career of CONICET, Argentina. |
This journal is © the Owner Societies 2016 |