Download as pdf or txt
Download as pdf or txt
You are on page 1of 8

Article

pubs.acs.org/JPCA

Theoretical Characterization of Dimethyl Carbonate at Low


Temperatures
R. Boussessi,† S. Guizani,† and M. L. Senent*
Departamento de Química y Física Teóricas, I. Estructura de la Materia, IEM-CSIC, Serrano 121, Madrid 28006, Spain

N. Jaïdane
Laboratoire de Spectroscopie Atomique, Moléculaire et Applications-LSAMA LR01ES09, Faculté des sciences de Tunis, Université de
Tunis El Manar, 2092, Tunis, Tunisie

ABSTRACT: Highly correlated ab initio methods (CCSD(T)


and RCCSD(T)-F12) are employed for the spectroscopic char-
acterization of the gas phase of dimethyl carbonate (DMC) at
low temperatures. DMC, a relevant molecule for atmospheric
and astrochemical studies, shows only two conformers, cis−cis
and trans−cis, respectively, of C2v and Cs symmetries. cis−cis-
DMC represents the most stable form. Using RCCSD(T)-F12
theory, the two sets of equilibrium rotational constants have
been computed to be Ae = 10 493.15 MHz, Be = 2399.22 MHz,
and Ce = 2001.78 MHz (cis−cis) and to be Ae = 6585.16 MHz,
Be = 3009.04 MHz, and Ce = 2120.36 MHz (trans−cis).
Centrifugal distortions constants and anharmonic frequencies for all of the vibrational modes are provided. Fermi displacements
are predicted. The minimum energy pathway for the cis−cis → trans−cis interconversion process is restricted by a barrier of
∼3500 cm−1. DMC displays internal rotation of two methyl groups. If the nonrigidity is considered, the molecule can be classified
in the G36 (cis−cis) and the G18 (trans−cis) symmetry groups. For cis−cis-DMC, both internal tops are equivalent, and the
torsional motions are restricted by V3 potential energy barriers of 384.7 cm−1. trans−cis-DMC shows two different V3 barriers of
631.53 and 382.6 cm−1. The far-infrared spectra linked to the torsional motion of both conformers are analyzed independently
using a variational procedure and a two-dimensional flexible model. In cis−cis-DMC, the ground vibrational state splits into nine
components: one nondegenerate, 0.000 cm−1 (A1), four quadruply degenerate, 0.012 cm−1 (G), and four doubly degenerate
0.024 cm−1 (E1 and E3). The methyl torsional fundamentals are obtained to lie at 140.274 cm−1 (ν15) and 132.564 cm−1 (ν30).

■ INTRODUCTION
At room temperature, dimethyl carbonate (DMC, CH3O−CO−
detectability.5,6 Not necessarily are the most abundant isomeric
forms in astrophysical sources the more stable ones. For example,
OCH3, dimethyl ester) is a liquid (melting point = 2−4 °C and among the C2H4O2 compounds, methyl formate is more abun-
boiling point = 90 °C) usually employed as a methylating agent dant in the interstellar medium than acetic acid. For C3H6O3
and as a solvent. It is considered to be a green reagent, and it has molecules, nothing can be said because only dihydroxyacetone
been exempted from classification as a volatile organic has been astrophysically discovered.7,8 The facts that only lactic
compound (VOC).1 Because DMC shows very high oxygen acid and DMC are commercialy available and some species show
content, it has been suggested as potential oxygenated fuel very low dipole moment, can explain why these species have not
additive.2 This application has motivated studies of its environ- been detected.
mental impact when it is released into the atmosphere where it Recently, Lovas et al.4 have determined the binding energies of
develops a rich green atmospheric chemistry.2 13 C3H6O3 isomers. In this ranking of stabilities, DMC is placed
The structure DMC has attracted special attention because the in the fourth position with a relative energy of ∼4500 cm−1 with
molecule presents various conformers that intertransform respect to the most stable one, the lactic acid.4 Furthermore,
through internal rotation.3 It has also attracted interest because DMC shows at least two conformers cis−cis and trans−cis, and a
DMC shares the empirical formula of C3H6O3 with a large list of third near-trans−near-trans stable structure has been postulated.3
isomers, which have been the object of previous microwave Recently, the microwave spectrum of the cis−cis form was
spectroscopy studies4 motivated by the astrophysical search in measured by Lovas et al.4 over the frequency ranges of 8.4−25.3
the space where isomerism is a frequent fact.5 Their relative
abundances in extraterrestrial sources have been discussed Received: February 24, 2015
in terms of relative stability and on the basis of molecular Revised: March 31, 2015
properties (i.e., the collisional parameters) that can be keys of the Published: March 31, 2015

© 2015 American Chemical Society 4057 DOI: 10.1021/acs.jpca.5b01836


J. Phys. Chem. A 2015, 119, 4057−4064
The Journal of Physical Chemistry A Article

and 227−350 GHz. The analysis and assignments were Thus, anharmonic quadratic, cubic, and quartic force field was
accomplished with the Xiam and Erham programs developed obtained with Möller−Plesset theory (MP2) in connection with
for systems with two identical methyl tops.9,10 The dipole the cc-pVTZ correlation-consistent basis set (denoted in this
moment has been evaluated to be 0.293(3) debye.4 Article by VTZ).26
Measurements of the Raman spectrum of DMC started very The torsional parameters and the torsional energy levels were
early.11,12 In 1966, Collingwood et al. recorded the infrared determined using a variational procedure for solving a vibrational
spectrum between 4000 and 400 cm−1.13 Later, in 1974, Katon Hamiltonian of reduced dimensionality. For this purpose, we
and Cohen reported studies in liquid and solid phases.14 In 1999, employed our code ENEDIM.27 All of the kinetic and potential
Bohets and van der Veken have analyzed and assigned both parameters of the two-dimensional Hamiltonian were computed
infrared and Raman spectra in vapor, and in amorphous and using a set of MP2 geometries. Single point CCSD(T) calcula-
crystalline solid phases between 4000 and 50 cm−1, respectively, tions were performed on the MP2 geometries to obtain refined
on the basis of MP2/6-31G(d,p) and DFT/6-31G(d,p) ab initio potential parameters. The augmented correlation-consistent
calculations.3 Other previous theoretical studies are also basis set aug-cc-pVTZ (denoted in this Article by AVTZ) was
available.15,16 The harmonic frequencies for all of the vibrational used.28
modes were determined at different levels of theory by Sun et al.15 Equilibrium Geometries. After an exhaustive search of
In this Article, we provide accurate theoretical molecular prop- equilibrium structures using MP2/AVTZ, we concluded that
erties that we expect can help future experimental assignments DMC shows only two planar conformers of C2v and Cs symmetry,
of spectra motivated by the search of new molecules in respectively (see Figures 1 and 2). In this Article, we refer to
astrophysical sources. For this purpose, we use state-of-the-art
ab initio methods to focus on the spectroscopic characterization
of DMC at low temperatures. Under these conditions, the lower
torsional energy levels can be populated. DMC displays a
complex far-infrared spectrum because four internal rotations,
the C−OCH3 torsions (denoted by α1 and α2 in this Article) and
the two methyl group torsions (denoted by θ1 and θ2 in this
Article), are responsible for the nonrigid properties. The α1 and
α2 torsions cause the conformer intertransformation, whereas θ1
and θ2 transform equivalent minima. For each conformer, we
analyze the far-infrared spectra (FIR and Raman) using a two-
dimensional variational model for which independent coor-
dinates are θ1 and θ2. The study is applied to the cis−cis and
trans−cis conformers independently, on the basis that the energy
barrier between both structures (∼3500 cm−1) is sufficiently high
to justify this theoretical treatment. Spectroscopic parameters
obtained using second-order perturbation theory are also
provided for all of the vibrational modes, and they are used to Figure 1. CCSD(T) pathway corresponding to the intertransformation
help the variational calculations. of the two conformers of dimethyl carbonate.

■ RESULTS AND DISCUSSION


Computational Details. The structural parameters of the
two DMC conformers and the equilibrium rotational constants
have been computed using the explicitly correlated coupled-
cluster method CCSD(T)-F12b,17,18 implemented in MOLPRO
(2012).19 The default options were selected. The atomic orbitals
were described by the cc-pVTZ-F12 basis set of Peterson et al.20
(denoted in this Article by VTZ-F12) in connection with the
corresponding basis sets21 for the density fitting and the resolu-
tions of the identity. The rotational constants were enhanced
adding a core−valence correlation correction determined with Figure 2. Equilibrium geometry of cis−cis dimethyl carbonate.
CCSD(T) (coupled-cluster theory with singles and doubles sub- Definition of the torsional coordinates.
stitutions, augmented by a perturbative treatment of triple
excitations)22 and the cc-pCVTZ basis set.23,24 them as cis−cis-DMC and trans−cis-DMC. cis−cis-DMC is the
The rotational parameters (rotational constants at the most stable conformer. If the methyl group internal rotational
vibrational levels as well as the centrifugal distortion constants) rotation is considered, the corresponding molecular symmetry
and the harmonic and anharmonic frequencies for all of the groups (MSG) are the G36 and G18. The structural parameters are
vibrational modes were calculated using vibrational second-order shown in Table 1, as well as the rotational constants. The search
perturbation theory (VPT2) as implemented in Gaussian 09.25 for more minimum energy geometries (i.e., near-trans−near-trans3)
To limit computational expense, which is significant given the has been unproductive. We found a structure with imaginary
molecular size (DMC contains six heavy atoms and six hydrogen frequencies.
atoms), different levels of theory were combined. For the first- Figure 1 shows the minimum energy pathway for the cis−cis →
order properties, such as the rotational constants or the harmonic trans−cis conversion process, which is restricted by a barrier
fundamentals, very accurate methods were employed. However, of 3501 cm−1 when CCSD(T) theory is employed. At the
anharmonic effects were computed at lower levels of theory. CCSD(T)-F12 level of theory, the trans−cis-DMC lies 1091.5 cm−1
4058 DOI: 10.1021/acs.jpca.5b01836
J. Phys. Chem. A 2015, 119, 4057−4064
The Journal of Physical Chemistry A Article

Table 1. CCSD(T)-F12/VTZ-F12 Total Electronic Energies (E, in au), Relative Energies (Er, in cm−1), Structural Coordinates
(Distances in Å, Angles in deg), Rotational Constants (in MHz), and MP2/AVTZ Dipole Moment (in debye) of the Two
Conformers of DMC
cis−cis trans−cis cis−cis trans−cis
C2v Cs C2v Cs
O2C1 1.3325 1.3335 E −343.219689 −343.214716
O3C1 1.3325 1.3434 Er 0.0 1091.5
C4O2 1.4315 1.4316 θ1 0.0 0.0
C5O3 1.4315 1.4329 θ2 0.0 0.0
O6C1 1.2043 1.1982 C4O2C1O3 180.0 0.0
H7C4 1.088 1.0854 C5O3C1O2 180.0 180.0
H8C4, H9C4 1.0850 1.0879
H10C5 1.0850 1.0853 calcd exp4 calcd
H11C5, H12C5 1.0850 1.0876 μ 0.3498 3.9212
O3C1O2 108.2 112.2 Ae 10 493.15 6585.16
O6C1O2 125.9 122.8 Be 2399.22 3009.04
O6C1O3 125.9 125.1 Ce 2001.78 2120.36
C4O2C1 113.4 118.6 A0 10 432.29 10 411.707 (5) 6544.72
C5O3C1 113.4 113.3 B0 2379.08 2371.5263 (9) 2981.28
H7C4O2 105.4 105.1 C0 1986.12 1980.2513 (9) 2102.67
H8C4O2, H9C4O2 110.4 110.9
H10C5O3 105.4 105.5
H11C5O3 110.4 110.2
H12C5O3
H7C4O2C1 180.0 180.0
H8C4O2H7−H9C4O2H7 119.6 119.0
H10C5O3C1 180.0 180.0
H11C5O3H10−H12C5O3H10 119.6 119.0

over the cis−cis-form (see Table 1). As was expected, the process with the experimental parameters of ref 4 (A0 = 10 411.707(5)
carries out a significant change of the dipole moment (from MHz, B0 = 2371.5263(9) MHz, and C0 = 1980.2513(9) MHz).
0.3498 to 3.9212 D). However, with the exceptions of three They are shown in Table 1.
angles, O3C1O2, C4O2C1, and O6C1O2, the process occurs with- Full-Dimensional Anharmonic Analysis. The anharmonic
out significant changes in the structural parameters. The two force field has been computed at the MP2/VTZ level of theory.
methyl groups display a slightly distorted C3v structure. In Table 2, VPT2 harmonic and anharmonic frequencies are com-
Table 1 displays the equilibrium and the ground vibrational pared to the experimental infrared and Raman transitions recently
state rotational constants. The last are compared to previous measured by Bohets and van der Veken.3 The experimental band
experimental data derived from microwave spectroscopy.4 cis− centers, which are emphasized in bold, correspond to the Raman
cis-DMC is a near prolate top (κ ≈ −0.91), and trans−cis-DMC is transitions.3 The calculated values have been tentatively assigned to
a clear asymmetric rotor (κ ≈ −0.60). The equilibrium param- local modes (s = stretching, b = bending, and t = torsions) and have
eters of the cis−cis and trans−cis forms were computed to be Ae = been classified following the representation of the C2v and Cs groups.
10 493.15 MHz, Be = 2399.22 MHz, and Ce = 2001.78 MHz and Especially tricky is the correlation of normal torsional modes
to be Ae = 6585.16 MHz, Be = 3009.04 MHz, and Ce = 2120.36 and local modes. For the low frequency modes, this correlation is
MHz, respectively. The ground vibrational state rotational con- almost impracticable because the α coordinates (C−OCH3
stants were obtained using the equation: torsions) are strongly coupled with the θ coordinates (methyl
B0 = Be + ΔBe core + ΔBvib torsions) in the normal coordinates.
The experimental sample3 contains a mixture of the two DMC
Here, ΔBecore regards the core−valence-electron correlation conformers. The main observed bands were assigned to the most
effect on the equilibrium parameters. It was computed at the stable one.3 Some of the unassigned transitions observed at
CCSD(T)-cc-pCVTZ level of theory as the difference between (1105, 1047, 633, and 578 cm−1) were correlated to the second
Be(CV) (calculated correlating both core and valence electrons) conformer. On the basis of the agreement between our anharmonic
and Be(V) (calculated correlating just the valence electrons). values and the experimental data for the previously assigned bands,
ΔBvib represents the vibrational contribution to the rotational we propose new assignments of the unclassified trans−cis transi-
constants derived from the VPT2 αri vibration−rotation interac- tions to the fundamentals band centers ν13, ν14, ν16, and ν17.
tion parameters determined using the MP2 cubic force field (see Focusing on the torsional two-dimensional analysis (see
below). The resulting corrections are below), one of the most beneficial uses of the VPT2 theory is the
form ΔAvib ΔAecore ΔBvib ΔBecore ΔCvib ΔCecore prediction of Fermi interactions. Calculated bands for which
cis−cis −94.61 33.95 −25.46 5.32 −20.47 4.81 the expected Fermi displacements are larger than 5 cm−1 are
trans−cis −58.35 17.91 −34.48 6.72 −22.74 5.05 emphasized in bold. This occurs with ν2(a1) and the ν18(b2)
The ground vibrational state rotational constants of the most methyl stretching fundamentals of the cis−cis form. In the case of
stable cis−cis form has been estimated to be A0 = 10 432.29 MHz, the trans−cis form, three fundamental ν3(a′), ν4(a′), and ν18(a′)
B0 = 2379.08 MHz, and C0 = 1986.12 MHz in good agreement are also displaced.
4059 DOI: 10.1021/acs.jpca.5b01836
J. Phys. Chem. A 2015, 119, 4057−4064
The Journal of Physical Chemistry A Article

Table 2. Harmonic and Anharmonic Fundamental Frequencies (ω and ν, in cm−1) of DMC Calculated with MP2/VTZ
cis−cis trans−cis
assign. ω νa expb assign. ω ν expc
ν1 a1 CH3 s 3220 3082 3037 a′ CH3 s 3219 3081
3036
ν2 CH3 s 3095 3031 2968 CH3 s 3213 3075
2970
ν3 CO s 1812 1784 1774 CH3 s 3097 3016
1774
ν4 CH3 b 1524 1495 CH3 s 3094 3039
ν5 CH3 b 1476 1442 1435 CO s 1838 1808
ν6 CH3 b 1242 1214 1210 CH3 b 1525 1491
1215
ν7 CH3−O s 1159 1126 1130 CH3 b 1521 1480
1130
ν8 CO2 s 936 917 921 CH3 b 1506 1484
921
ν9 CO2 b 521 514 516 CH3 b 1478 1443
517
ν10 COC b 246 241 244 CO2 s 1308 1268
250
ν11 b1 CH3 s 3189 3050 3006 CH3 b 1227 1199
ν12 CH3 b 1509 1468 1456 CH3 b 1211 1183
ν13 CH3 b 1198 1174 1167 CH3−O s 1137 1106 1105
ν14 CO3 b 812 800 798 CH3−O s 1068 1039 1047
ν15 CH3 t 175 169 126 CO2 s 881 861
ν16 CO t 132 128 118 OCO2 b 647 637 633
ν17 b2 CH3 s 3220 3082 3037 CO2 b 585 576 578
3036
ν18 CH3 s 3094 2963 2968 COC b 357 359
2970
ν19 CH3 b 1522 1494 COC b 253 245
ν20 CH3 b 1502 1464 1456 a″ CH3 b 3194 3054
ν21 CO2 s 1334 1293 1293 CH3 s 3190 3052
ν22 CH3 b 1218 1192 CH3 b 1515 1486
ν23 CH3−O s 1020 990 990 CH3 b 1506 1475
ν24 OCO2 b 703 694 693 CH3 b 1197 1172
ν25 C−O−C b 359 356 373 CH3 b 1192 1167
ν26 a2 CH3 s 3189 3050 3006 CO3 s 799 786
ν27 CH3 b 1508 1467 1456 CH3 t 227 218
ν28 CH3 b 1195 1171 CH3 t 185 169
ν29 OC t 212 207 OC t 168 167
ν30 CH3 t 158 154 OC t 138 131
a
Δν = Fermi displacements (emphasized in bold if Δν > 5 cm−1). bInfrared frequencies of ref 3; Raman transitions emphasized in bold. cInfrared
frequencies of ref 3; new assignments [this work].

Strong Fermi interactions among the torsional modes and modes and the remaining vibrations. Under these conditions,
the remaining vibrations are not predicted (see Table 3). This it is possible to define the following torsional Hamiltonian for
validates the two-dimensional model employed below. Few J = 0:29,30
possible displacements are obtained; that is, the combination 2 2
⎛ ∂ ⎞ ⎛ ∂ ⎞
levels ν15 + ν29 and ν15 + ν30 interact with the COC bending Ĥ (θ1 , θ2) = −∑ ∑ ⎜ ⎟Bθθi j (θi , θj)⎜⎜ ⎟⎟
fundamental (ν25). This effect displaces ∼10 cm−1 the torsional i = 1 j = 1 ⎝ ∂θi ⎠ ⎝ ∂θj ⎠
combination bands, in the cis−cis conformer.
In the case of the trans−cis-conformer, the overtone 2ν28 + V eff (θ1 , θ2) (1)
(methyl torsion) is strongly coupled with the COC bending
fundamental (ν18). The corresponding torsional band can be where the independent coordinates θ1 and θ2 are identified with
displaced an amount of 13 cm−1 to the lower frequencies. the two CH3 torsions (see Figure 1). These two coordinates are
defined as a function of three dihedral angles:31
Table 3 displays rotational constants in the torsional
fundamentals and centrifugal distortion constants. θ1 = (H 7C4 O2 C1 + H8C4 O2 C1 + H 9C4 O2 C1)/3 − π
Variational Torsional Analysis. The methyl torsional
energy levels were determined variationally assuming very small, θ2 = (H10C5O3C1 + H11C5O3C1 + H12C5O3C1)/3 − π
almost negligible, interactions between the two methyl torsional (2)

4060 DOI: 10.1021/acs.jpca.5b01836


J. Phys. Chem. A 2015, 119, 4057−4064
The Journal of Physical Chemistry A Article

Table 3. Rotational Spectroscopic Parameters of DMC (in MHz) Calculated with MP2/VTZ
cis−cis trans−cis
κ −0.905150 κ −0.596560
Ae 10 379.8 Ae 6525.35
Be 2387.37 Be 3000.69
Ce 1989.45 Ce 2110.04
A0 10 285.19 A0 6467.00
B0 2361.91 B0 2966.21
C0 1968.98 C0 2087.30
ν15 = 169 cm−1 ν27 = 218 cm−1
A (ν15) 10 272.81 A (ν27 = 1) 6470.63
B (ν15) 2357.30 B (ν27 = 1) 2955.35
C (ν15) 1966.22 C (ν27 = 1) 2084.43
ν30 = 154 cm−1 ν28 = 169 cm−1
A (ν30) 10 271.97 A (ν28 = 1) 6441.19
B (ν30) 2358.38 B (ν28 = 1) 2965.22
C (ν30) 1967.06 C (ν28 = 1) 2087.01
ΔJ 0.1876 × 10−3 ΔJ 0.4510 × 10−3
ΔK 0.4359 × 10−2 ΔK 0.5585 × 10−2
ΔJK 0.1330 × 10−2 ΔJK −0.5872 × 10−3
δJ 0.3109 × 10−4 δJ 0.1526 × 10−3
δK 0.5422 × 10−3 δK 0.6118 × 10−3
Predicted Fermi Displacements (in cm−1)
ν25 (COC bending) 357.5 → 356.0 2ν28 335.4 → 321.9
ν15 + ν30 318.2 → 311.1 2ν29 (CO torsion) 331.4 → 332.0
ν15 + ν29 (CO torsion) 372.7 → 381.3 ν29 + ν28 332.4 → 334.1
ν18 (COC bending) 347.5 → 358.6

Bθiθj and Veff(θ1,θ2) represent the kinetic energy parameters and 3N − 6


ωk
the effective potential energy surface. This last can be defined as EZPVE(θ1 , θ2) = ∑
the sum of three terms: k=3
2 (4)
To obtain the effective potential energy surfaces, we combine
V eff (θ1 , θ2) = V (θ1 , θ2) + V ′(θ1 , θ2) + V ZPVE(θ1 , θ2)
different levels of ab initio calculations. The geometries are
(3) optimized using MP2/VTZ; the energies are obtained in
where V(θ1, θ2), V′(θ1, θ2), and V ZPVE
(θ1, θ2) are the potential CCSD(T)/AVTZ single point calculations, and zero point
energy surface, the pseudopotential, and the zero point vibrational energy correction is determined with MP2/VTZ. For
vibrational energy correction, respectively.29,30,32 both conformers, the effective surfaces are determined by fitting
One aspect that is important to discuss is the validity of the the effective energies to the double Fourier series transforming
two-dimensional model for DMC because a treatment in four as the totally symmetric representation of the G36 and G18
dimensions appears to be more precise given the strong coupling molecular symmetry groups. Thus, for cis−cis-DMC:
between α and θ coordinates. All of the out-of-plane bending
coordinates involving the O−CH3 groups contribute to the V eff (θ1 , θ2) = 415.378 − 192.83(cos 3θ1 + cos 3θ2)
torsional normal coordinates. In addition, in trans−cis-DMC, the + 1.178 cos 3θ1 cos 3θ2 − 14.582(cos 6θ1 + cos 6θ2)
2ν29 overtone interacts with the COC bending fundamental (see
Table 3). − 0.720(cos 6θ1 cos 3θ2 + cos 3θ1 cos 3θ2)
Although a 4D-model (or a 5D model) could provide more − 0.292 cos 6θ1 cos 6θ2 − 2.166 sin 3θ1 sin 3θ2
accurate frequencies, it is extremely expensive from the com-
putational point of view. Furthermore, the assignments of and for trans−cis-DMC:
previous experimental microwave studies have been performed
using effective Hamiltonian for two identical tops. To our V eff (θ1 , θ2) = 541.588 − 323.084 cos 3θ1
knowledge, effective Hamiltonians depending on many vibra-
− 198.679 cos 3θ2 + 6.958 cos 3θ1 cos 3θ2
tional variables are not contemplated for future rotational
assignments. To avoid large computational effort and to consider − 12.012 cos 6θ1 − 16.712 cos 6θ2
these interactions partially, we have employed a flexible two-
dimensional model where the α coordinates and the remaining + 0.431 cos 6θ1 cos 3θ2 + 0.361 cos 3θ1 cos 3θ2
internal coordinates are allowed to be relaxed during the CH3 + 1.151 cos 6θ1 cos 6θ2 + 1.207 sin 3θ1 sin 3θ2
internal rotation. The potential energy surface is calculated from
the energies of 7 and 10 conformations for selected values of From this effective potential energy surfaces, the energy
the H7C4O2C1 and H10C5O3C1 dihedral angles. A set of 3N − 8 barriers and potential parameters shown in Table 4 can be
internal coordinates (N = number of atoms) are allowed to be derived. This table also shows the kinetic energy parameters
relaxed. For all of the structures, the zero point vibrational energy calculated with the code ENEDIM27 from the optimized geom-
is corrected within the harmonic approximation: etries of the conformers. The effective rotation barriers, V3, of the
4061 DOI: 10.1021/acs.jpca.5b01836
J. Phys. Chem. A 2015, 119, 4057−4064
The Journal of Physical Chemistry A Article

Table 4. Effective Potential and Kinetic Parameters (in cm−1) Table 5. Low Torsional Energy Levels (in cm−1) of cis−cis-
of DMC Calculated with CCSD(T)/AVTZ DMC and trans−cis-DMC Calculated at the CCSD(T)/AVTZ
Level of Theory
cis−cis trans−cis
calcd exp.4 calcd calcd calcd
V3 (θ1) = E(60°,0°) − E(0°,0°) 384.7 398.13 631.5 ν30 ν27
ν15 sym. E Eeff ν28 sym. E Eeff
V3 (θ2) = E(0°,60°) − E(0°,0°) 384.7 398.13 382.6
E(60°,60°) − E(60°,0°) − E(0°,60°) 4.72 7.82 00 A1 0.000 0.000 00 A1 0.000 0.000
B11 5.5883 5.4290 G 0.015 0.012 E1 0.001 0.001
B22 5.5883 5.4956 E1 0.030 0.024 E2 0.013 0.011
B12 −0.2801 −0.1153 E3 0.030 0.024 E3 0.013 0.011
E4 0.013 0.011
10 A3 127.648 132.831 10 A2 132.046 136.105
two equivalent methyl torsions of cis−cis-DMC have been G 127.340 132.564 E1 132.046 136.106
evaluated to be 384.7 cm−1. They have been found to be 631.53 E2 127.057 132.316 E2 131.512 135.615
and 382.6 cm−1 for trans−cis-DMC where the two tops are not E3 127.057 132.316 E3 131.513 135.616
equivalent. One-dimensional cuts of the surfaces are shown in E4 131.513 135.616
Figure 3. The cis-barriers are of the same order of magnitude as 01 A2 135.489 140.528 01 A2 166.779 170.358
G 135.198 140.274 E1 166.738 170.321
E1 134.882 139.999 E2 166.790 170.367
E4 134.882 140.000 E3 166.749 170.330
E4 166.749 170.330
20 A1 230.377 238.919 20 A1 233.558 239.865
G 231.000 239.622 E1 233.558 239.866
E1 237.364 245.570 E2 240.376 246.526
E3 237.358 245.565 E3 240.377 246.527
E4 240.377 246.527
11 A4 231.793 240.554 11 A1 299.420 307.677
G 238.401 246.716 E1 299.393 307.658
E2 239.258 247.627 E2 298.908 307.216
E4 239.262 247.631 E3 298.881 307.195
E4 298.881 307.195
02 A1 264.623 274.279 02 A1 314.896 320.804
G 264.245 273.935 E1 315.727 321.595
E1 263.881 273.599 E2 306.439 312.636
E3 263.874 273.593 E3 306.440 312.637
E4 306.440 312.637
ZPVE 142.507 148.771 158.863 163.256

intense than those of ν27(a″). Because the 2D-potential energy


Figure 3. One-dimensional cuts of the 2D potential energy barriers
surfaces, V3cis−cis (θ = θ1), V3trans−cis (θ = θ1), and V3trans−cis (θ = θ2), of the
surface shows nine minima, each level splits into nine
cis−cis (in red) and trans−cis (in black) conformers. components. For cis−cis-DMC, the sublevels are one non-
degenerate (Ai, i = 1,2,3,4), two two-degenerate (Ei and Ej,
i = 1,2; j = 3,4)), and one four times degenerate (G). The G
for acetone (367.07 cm−1)33 and for methanol (377.94 cm−1).34 components of the two fundamentals, which are expected to
Table 4 shows also the difference of energies E(60°,60°) − be the prominent subcomponents, are calculated to lie at
E(60°,0°) − E(0°,60°) that can be understood as a measure of 140.274 cm−1 (ν15) and 132.564 cm−1 (ν30) with the isotopically
the two methyl group interactions. This difference is very small corrected potential. The (0, 0) ground-state subcomponents are
as compared to other molecules such as acetone33 or dimethyl- found to be 0.000 cm−1 (A1), 0.012 cm−1 (G), and 0.024 cm−1
ether.35,36 The more significant interaction terms of the surfaces (E1 and E3).
cos 3θ1 cos 3θ2 and sin 3θ1 sin 3θ2 present expansion coefficients The trans−cis levels split into one doubly degenerate sublevel
almost negligible if they are compared to the heights of the (Ai; i = 1, 2) and four quadruply degenerate (Ei; i = 1, 2, 3, 4). In
torsional barriers. this case, the two fundamentals are found at 170.358 cm−1 (ν28)
The torsional energy levels of both conformers are shown and 136.105 cm−1 (ν27) and the (0, 0) sublevels at 0.000 cm−1
in Table 5. The first columns are devoted to the calculations (A1), 0.001 cm−1 (E1), and 0.0011 cm−1 (E2, E3, and E4). The
performed using V(θ1,θ2), and the second columns to those energy differences between subcomponents are smaller in the
performed with Veff(θ1,θ2). Both groups of energies are denoted trans−cis-form where one of the barriers is double higher than
by E and Eeff. The energy levels are classified following the cis−cis-DMC.
representations of the G36 and the G18 MSG and the vibrational Unfortunately, there are no experimental data available with
quanta. cis−cis-DMC presents two methyl torsional modes which to compare our results. To our knowledge, the unique
ν15(b1) and ν30(a2), one active in infrared and the second one observation in this spectrum region is from Bohets and Veken,3
inactive. In the case of trans−cis-DMC, both modes are infrared who have observed unassigned Q branches of Type C bands
active. The bands assigned to ν28(a″) are expected to be more at 126 and 118 cm−1. It seems reasonable to assign them to
4062 DOI: 10.1021/acs.jpca.5b01836
J. Phys. Chem. A 2015, 119, 4057−4064
The Journal of Physical Chemistry A Article

torsional out-of-plane modes. What is not possible is to clarify if


they have be assigned to methyl group torsion or to the CO bond
■ REFERENCES
(1) Tundo, P.; Selva, M. The Chemistry of Dimethyl Carbonate. Acc.
torsions. Furthermore, both types of torsions are participating Chem. Res. 2002, 35, 706−716.
in all of the torsional normal modes, and this makes difficult (2) Bilde, M.; Møgelberg, T. E.; Sehested, J.; Nielsen, O. J.; Wallington,
the calculation of the zero point vibrational correction. Large T. J.; Hurley, M. D.; Japar, S. M.; Dill, M.; Orkin, V. L.; Buckley, T. J.;
differences between E and Eeff energies can be derived from these et al. Atmospheric Chemistry of Dimethyl Carbonate: Reaction with
interactions. OH Radicals, UV Spectra of CH3OC(O)OCH2 and CH3OC(O)-
When the behavior of the cis−cis and trans−cis forms is OCH2O2 Radicals, Reactions of CH3OC(O)OCH2O2 with NO and
compared, it can be concluded that band centers are displaced to NO2, and Fate of CH3OC(O)OCH2O Radicals. J. Phys. Chem. A 1997,
101, 3514−3525.
the large frequencies in the second conformer, which shows the
(3) Bohets, H.; Van der Veken, B. J. On the Conformational Behavior
smaller splittings. of Dimethyl Carbonate. Phys. Chem. Chem. Phys. 1999, 1, 1817−1826.

■ CONCLUSIONS
The exhaustive search of equilibrium structures of DMC leads to
(4) Lovas, F. J.; Plusquellic, D. F.; Widicus Weaver, S. L.; McGuire, B.
A.; Blake, G. A. Organic Compounds in the C3H6O3 Family: Microwave
Spectrum of cis−cis Dimethyl Carbonate. J. Mol. Spectrosc. 2010, 264,
10−18.
two planar conformers of C2v and Cs symmetry. The corre- (5) Remijan, A. J.; Hollis, J. M.; Lovas, F. J.; Plusquellic, D. F.; Jewell, P.
sponding energy difference has been calculated to be 1091.5 cm−1 R. Interstellar Isomers: The Importance of Bonding Energy Differences.
and the equilibrium rotational constants to be Ae = 10 493.15 Astrophys. J. 2005, 632, 333−339.
MHz, Be = 2399.22 MHz, and Ce = 2001.78 MHz (cis−cis) and to (6) Dumouchel, F.; Faure, A.; Lique, F. The Rotational Excitation of
be Ae = 6585.16 MHz, Be = 3009.04 MHz, and Ce = 2120.36 MHz HCN and HNC by He: Temperature Dependence of the Collisional
(trans−cis). The ground vibrational state rotational constants of Rate Coefficients. Mon. Not. R. Astron. Soc. 2010, 406, 2488−2492.
the most stable cis−cis form have been estimated to be A0 = (7) Widicus Weaver, S. L.; Blake, G. A. 1,3-Dihydroxyacetone in
Sagittarius B2(N-LMH): The First Interstellar Ketose. Astrophys. J. lett
10 432.29 MHz, B0 = 2379.08 MHz, and C0 = 1986.12 MHz.
2005, 624, 33−36.
The MP2/VTZ anharmonic force field shows that the (8) Apponi, A. J.; Halfen, D. T.; Ziurys, L. M.; Hollis, J. M.; Remijan, A.
correlation of normal torsional modes and local modes is J.; Lovas, F. J. Investigating the Limits of Chemical Complexity in
especially tricky given the coupling among all of the out-of-plane Sagittarius B2(N): A Rigorous Attempt to Confirm 1,3-Dihydrox-
vibrations of the molecule. Strong Fermi interactions among the yacetone. Astrophys. J. 2006, 643, 29−32.
torsional modes and the remaining vibrations are not predicted (9) Hartwig, H.; Dreizler, H. The Microwave Spectrum of trans-2,3-
with the exception of the trans−cis overtone 2ν28 (methyl Dimethyloxirane in Torsional Excited States. Z. Naturforsch. 1996, 51a,
torsion), which is strongly coupled with the COC bending 923−932.
fundamental (ν18). (10) Groner, P. Effective Rotational Hamiltonian for Molecules with
The effective rotation barriers, V3, of the two equivalent methyl Two Periodic Large-Amplitude Motions. J. Chem. Phys. 1997, 107,
4483−4498.
torsions of cis−cis-DMC have been evaluated to be 384.7 cm−1 (11) Kohlrausch, K. W. F.; Pongratz, A. Raman-Effekt und
and to be 631.53 and 382.6 cm−1 for trans−cis-DMC where the Konstitutionsprobleme, 4. Mitteil.: Carbonyl-Frequenz und Molekul-
two tops are not equivalent. The two fundamentals of cis−cis- Konstitution. Ber. Dtsch. Chem. Ges. 1933, 66, 1355−1369.
DMC and trans−cis-DMC are found at 140.274 cm−1 (ν15) and (12) Kubo, M.; Morino, Y.; Mizushima, S. Internal Rotation VIII:
132.564 cm−1 (ν30) and at 170.358 cm−1 (ν28) and 136.105 cm−1 Molecular Structure of Carbonic Ester. Sci. Pap. Inst. Phys. Chem. Res.
(ν27), respectively. The energy differences between torsional (Jpn.) 1937, 32, 129−131.
splittings are smaller in the trans−cis-form where one of the (13) Collingwood, B.; Lee, H.; Wilmshurst, J. K. The Structures and
barriers is double that in cis−cis-DMC. Vibrational Spectra of Methyl Chloroformate and Dimethyl Carbonate.


Aust. J. Chem. 1966, 19, 1637−1649.
(14) Katon, J. E.; Cohen, M. D. The Vibrational Spectra and Structure
AUTHOR INFORMATION of Dimethyl Carbonate and its Conformational Behavior. Can. J. Chem.
Corresponding Author 1975, 53, 1378−1386.
(15) Sun, H.; Mumby, S. J.; Maple, J. R.; Hagler, A. T. An ab Initio
*Phone: +34915616800. E-mail: ml.senent@csic.es. CFF93 All-Atom Force Field for Polycarbonates. J. Am. Chem. Soc. 1994,
Present Address 116, 2978−2987.
† (16) Labrenz, D.; Schröer, W. Conformational Analysis of Symmetric
Laboratoire de Spectroscopie Atomique, Moléculaire et
Applications-LSAMA LR01ES09, Faculté des sciences de Carbonic Acid Esters by Quantum Chemical Calculations and Dielectric
Measurements. J. Mol. Struct. 1991, 249, 327−341.
Tunis, Université de Tunis El Manar, 2092, Tunisie. (17) Knizia, G.; Adler, T. B.; Werner, H.-J. Simplified CCSD (T)-F12
Notes Methods: Theory and Benchmarks. J. Chem. Phys. 2009, 130, 054104-
The authors declare no competing financial interest. 1−20.


(18) Werner, H.-J.; Adler, T. B.; Manby, F. R. General Orbital Invariant
MP2-F12 Theory. J. Chem. Phys. 2007, 126, 164102-1−18.
ACKNOWLEDGMENTS (19) MOLPRO, version 2012.1, a package of ab initio programs; Werner,
This research was supported by the MINECO of Spain grant H.-J.; Knowles, P. J.; Manby, F. R.; Schütz, M.; Celani, P.; Knizia, G.;
FIS2013-40626-P and by the Marie Curie International Research Korona, T.; Lindh, R.; Mitrushenkov, A.; Rauhut, G.; et al.; see http://
www.molpro.net.
Staff Exchange Scheme Fellowship within the seventh European
(20) Peterson, K. A.; Adler, T. B.; Werner, H.-J. Systematically
Community Framework Program under grant no. PIRSES-GA- Convergent Basis Sets for Explicitly Correlated Wavefunctions: The
2012-31754. We acknowledge the COST Actions CM1405 atoms H, He, B−Ne, and Al−Ar. J. Chem. Phys. 2008, 128, 084102-1−
“MOLIM” and CM1401 “Our Astrochemical History”. We also 12.
acknowledge the CTI (CSIC) and CESGA for computing (21) Yousaf, K. E.; Peterson, K. A. Optimized Auxiliary Basis Sets for
facilities. Explicitly Correlated Methods. J. Chem. Phys. 2008, 129, 184108-1−7.

4063 DOI: 10.1021/acs.jpca.5b01836


J. Phys. Chem. A 2015, 119, 4057−4064
The Journal of Physical Chemistry A Article

(22) Raghavachari, K.; Trucks, G. W.; Pople, J. A.; Head-Gordon, M. A


Fifth-Order Perturbation Comparison of Electron Correlation Theories.
Chem. Phys. lett. 1989, 157, 479−483.
(23) Woon, D. E.; Dunning, T. H., Jr. Gaussian Basis Sets for Use in
Correlated Molecular Calculations. V. Core Valence Basis Sets for
Boron Through Neon. J. Chem. Phys. 1995, 103, 4572−4585.
(24) Peterson, K. A.; Dunning, T. H., Jr. Accurate Correlation
Consistent Basis Sets for Molecular Core−Valence Correlation Effects:
The Second Row Atoms Al−Ar, and the First Row Atoms B−Ne
Revisited. J. Chem. Phys. 2002, 117, 10548−10560.
(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.; et al. Gaussian 09, revision A.02; Gaussian, Inc.:
Wallingford, CT, 2009.
(26) Dunning, T. H., Jr. Gaussian Basis Sets for Use in Correlated
Molecular Calculations. I. The Atoms Boron Through Neon and
Hydrogen. J. Chem. Phys. 1989, 90, 1007−1023.
(27) ENEDIM, “A Variational Code for Non-Rigid Molecules”; Senent,
M. L., 2001; see http://tct1.iem.csic.es/PROGRAMAS.htm for more
details.
(28) Kendall, R. A.; Dunning, T. H., Jr.; Harrison, R. J. Electron
Affinities of the Firstrow Atoms Revisited. Systematic Basis Sets and
Wave functions. J. Chem. Phys. 1992, 96, 6796−6806.
(29) Senent, M. L. Determination of the Kinetic Energy Parameters of
Non-Rigid Molecules. Chem. Phys. Lett. 1998, 296, 299−306.
(30) Senent, M. L. Ab Initio Determination of the Roto-Torsional
Energy Levels of trans-1,3-Butadiene. J. Mol. Spectrosc. 1998, 191, 265−
275.
(31) Szalay, V.; Császár, A. G.; Senent, M. L. Symmetry Analysis of
Internal Rotation. J. Chem. Phys. 2002, 117, 6489−6492.
(32) Császár, A. G.; Szalay, V.; Senent, M. L. Ab Initio Torsional
Potential and Transition Frequencies of Acetaldehyde. J. Chem. Phys.
2004, 120, 1203−1207.
(33) Smeyers, Y. G.; Senent, M. L.; Botella, V.; Moule, D. C. An ab
Initio Structural and Spectroscopic Study of AcetoneAn Analysis of
the Far Infrared Torsional Spectra of Acetoneh6 and d6. J. Chem. Phys.
1993, 98, 2754−2767.
(34) Muñoz-Caro, C.; Niño, A.; Senent, M. L. Theoretical Study of the
Effect of Torsional Anharmonicity on the Thermodynamic Properties of
Methanol. Chem. Phys. Lett. 1997, 273, 135−140.
(35) Senent, M. L.; Moule, D. C.; Smeyers, Y. G. An ab Initio and
Spectroscopic Study of Dimethyl Ether: An Analysis of the Infrared and
Raman Spectra. Can. J. Phys. 1995, 73, 425−431.
(36) Villa, M.; Senent, M. L.; Domínguez-Gómez, R.; Á lvarez-Bajo, O.;
Carvajal, M. CCSD(T) Study of Dimethyl-Ether Infrared and Raman
Spectra. J. Phys. Chem. A 2011, 115, 13573−13580.

4064 DOI: 10.1021/acs.jpca.5b01836


J. Phys. Chem. A 2015, 119, 4057−4064

You might also like