Journal of Molecular Structure 524 (2000) 69–85 www.elsevier.nl/locate/molstruc Microwave spectrum, conformational equilibrium, 14N quadrupole coupling constants, dipole moment, vibrational frequencies and quantum chemical calculations for acrylamide K.-M. Marstokk, H. Møllendal, S. Samdal* Department of Chemistry, University of Oslo, P.O. Box 1033 Blindern, N-0315 Oslo, Norway Dedicated to Professor Peter Klaeboe on the occasion of his 70th birthday. Received 5 May 1999; received in revised form 3 September 1999; accepted 3 September 1999 Abstract The microwave spectrum of acrylamide (CH2yCHCONH2) has been investigated in the 20.0–60.5 GHz spectral region at room temperature. Two conformers were assigned. The most stable rotamer is planar with the carbonyl and vinyl groups in the syn conformation. Five vibrationally excited states of the C–C torsional motion and six other vibrationally excited states have been assigned for this conformer, as have three deuterated species (amide group). The frequencies of most of these excited states have been determined by relative intensity measurements. The dipole moment of this rotamer is (in units of Debye) ma 0:269 3; mb 3:42 2; mc 0:12 24 and mtot 3:44 5: The 14N quadrupole coupling constants were found to be xaa 1:87 26 and xbb 2:73 16 MHz; respectively. The second, less stable skew form is non-planar with the CyC–CyO chain of atoms approximately 155⬚ from syn. The ground vibrational state and two excited states were assigned. The syn form was found to be 6.5(6) kJ mol ⫺1 more stable than the skew conformer by relative intensity measurements. Quantum chemical calculations at HF, MP2 and DFT levels of theory using the rather large basis sets 6-311⫹G ⴱⴱ, 6311⫹⫹G ⴱⴱ and cc-pVTZ have been made to assist the experimental work. 䉷 2000 Elsevier Science B.V. All rights reserved. Keywords: Acrylamide; Microwave spectrum; Conformational equilibrium; Dipole moment; Quantum chemical calculations 1. Introduction This paper is a continuation of our general interest in the molecular structure and conformation of amides in the gaseous state where we have used gas electron diffraction [1–8] and microwave (MW) spectroscopy [9,10] as well as quantum chemical computations [8– 12] in our studies. The molecules we have investigated * Corresponding author. Tel.: ⫹47-22-855-458; fax: ⫹47-22-855441. E-mail address: svein.samdal@kjemi.uio.no (S. Samdal). so far are primary amides where we have focused mostly on the molecular structure and the conformation about the Csp3 –Csp2 bond. In this work we extend our studies to the simplest conjugated primary amide, acrylamide. This compound has several important commercial applications, in particular the production of commonly used plastic [13]. It is an interesting difunctional monomer containing a reactive electron-deficient double bond and an amide group, and it undergoes reactions typical of those two functionalities. The electron withdrawing carboxamide group activates the double bond, which 0022-2860/00/$ - see front matter 䉷 2000 Elsevier Science B.V. All rights reserved. PII: S0022-286 0(99)00362-2 70 K.-M. Marstokk et al. / Journal of Molecular Structure 524 (2000) 69–85 Fig. 1. The numbering of the atoms and the two conformations for acrylamide. consequently reacts readily with nucleophilic reagents by addition. It exhibits both weak acidic and basic properties, and the site of protonation has been studied by ab initio calculations [14]. Acrylamide is highly toxic to humans and animals [15,16]. The question about the planarity of the amide group has been asked, and from the IR spectra in the gaseous state [17] it has been concluded that acrylamide has a planar amide group. In this work no indication of a second conformation was seen. It was concluded that the less stable form must be at least 600 cm ⫺1 (7.2 kJ mol ⫺1) less stable [17]. Some MM2 [18] and ab initio calculations [14,18] have been carried out. They found the syn conformation as shown in Fig. 1 to be the most stable form, in agreement with the evidence from the IR spectra [17]. The ab initio calculations [14] were not conclusive concerning the planarity of the amide group. It is our purpose to get more experimental and theoretical information on the molecular structure and conformational properties of the important acrylamide molecule in the free state. Finally, it should be mentioned that the number of known molecular structures and conformations of the gaseous amides is relatively small owing to their low volatility and tendency to decompose upon heating. There is a considerable interest in the properties of free amides because they are the simplest models for the biologically important peptides. 2. Experimental The sample (m.p. 82–85⬚C specified to be at least 99% pure) utilized in this work was purchased from Fluka AG and used as received. No impurities were seen in the MW spectrum, which was studied using the Oslo spectrometer [19]. The 20–39 and the 53– 60.5 GHz spectral range was investigated thoroughly. Further, selected regions of the 39–60.5 GHz spectral range were also studied. The MW absorption X-band brass cell was held at room temperature in the experiments. The pressure was about 2 Pa when the spectra were recorded, and stored electronically using the computer programs written by Waal [20]. The accuracy of the frequency measurements is presumed to be better than ^0.10 MHz. Radio frequency-microwavefrequency double resonance (RFMWDR) experiments were carried out as described in Ref. [21] using the equipment mentioned in Ref. [22]. The deuterated species (amide group) were produced by seasoning the wave-guide with heavy water and then introducing the parent species. 3. Results and discussion 3.1. Quantum chemical calculations The quantum chemical computations have been made with the Gaussian94 program package [23] using the IBM RS6000 cluster in Oslo. Calculations were performed at three different levels of theory; Hartree–Fock (HF), Møller–Plesset pertubation (MP2) [24] with all electrons included, as well as density functional theory (DFT) employing the B3LYP method [25]. The basis sets utilized were 6311⫹G ⴱⴱ, 6-311⫹⫹G ⴱⴱ and cc-pVTZ [26]. The Table 1 Molecular structures of the most stable syn conformation of acrylamide from quantum chemical calculations a HF/6-311⫹G ⴱⴱ MP2 full/6-311⫹⫹G ⴱⴱ HF/cc-pVTZ DFT/B3LYP/cc-pVTZ MP2 full/cc-pVTZ 121.97 136.94 149.65 100.58 100.78 133.07 108.68 108.45 108.34 122.12 137.63 149.57 100.73 100.95 133.95 108.74 108.53 108.42 119.41 135.16 149.48 98.78 99.05 131.42 107.48 107.27 107.32 121.78 136.50 149.43 100.30 100.50 132.62 108.44 108.21 108.10 121.79 135.79 148.44 99.91 100.19 132.82 107.70 107.69 107.54 122.15 123.35 114.51 122.63 118.41 118.96 121.00 118.16 120.83 120.00 121.69 118.31 122.50 123.39 114.06 119.41 115.98 116.74 120.51 118.61 120.89 119.72 121.43 118.85 122.34 123.03 114.63 122.34 118.49 119.17 120.86 118.15 120.99 120.39 121.46 118.15 122.23 123.33 114.43 122.48 118.43 119.09 120.99 118.19 120.82 119.86 121.81 118.33 122.68 123.25 114.07 122.05 118.48 119.46 120.18 118.95 120.88 119.04 121.72 119.25 0.03 180.03 180.03 ⫺0.03 0.04 179.98 0.00 180.00 180.00 0.00 ⫺2.98 176.90 161.03 13.51 ⫺21.01 190.87 0.23 179.98 180.01 0.10 0.01 180.01 180.07 ⫺0.03 0.08 179.98 ⫺0.01 179.99 179.99 ⫺0.01 0.01 180.02 180.19 ⫺0.15 0.20 179.87 ⫺0.01 180.02 179.98 0.01 0.09 180.09 180.54 ⫺0.42 0.60 179.64 ⫺0.01 180.00 179.99 ⫺0.01 0.26 3.71 0.00 3.72 0.25 3.78 0.75 3.86 0.11 3.74 0.00 3.74 0.42 3.58 0.01 3.61 0.09 3.89 0.02 3.89 K.-M. Marstokk et al. / Journal of Molecular Structure 524 (2000) 69–85 Bond length (pm) O1yC2 119.54 C2–N3 135.51 C2–C6 149.60 N3–H4 99.06 N3–H5 99.32 C6yC7 131.85 C6–H10 107.72 C7–H8 107.51 C7–H9 107.55 Bond angles (⬚) O1C2N3 122.31 O1C2C6 122.98 N3C2C6 114.70 C2N3H4 122.47 C2N3H5 118.46 H4N3H5 119.07 C2C6C7 120.85 C2C6H10 118.14 C7C6H10 121.01 C6C7H8 120.47 C6C7H9 121.40 H8C7H9 118.13 Dihedral angles (⬚) O1C2C6C7 0.00 O1C2C6H10 180.00 O1C2N3H4 179.99 O1C2N3H5 0.02 C6C2N3H4 ⫺0.02 C6C2N3H5 180.01 C2C6C7H8 ⫺0.01 C2C6C7H9 179.99 H10C6C7H8 180.00 H10C6C7H9 0.00 Dipole moments (Debye) ma 0.02 mb 3.83 mc 0.00 mtot 3.83 DFT/B3LYP/6-311⫹G ⴱⴱ 71 72 HF/6-311⫹G ⴱⴱ Rotational constants DA DB DC DFT/B3LYP/6-311⫹G ⴱⴱ MP2 full/6-311⫹⫹G ⴱⴱ HF/cc-pVTZ ⫺10.31 ⫺14.36 ⫺10.61 ⫺81.72 ⫺11.71 ⫺10.76 339.78 55.24 52.92 DFT/B3LYP/cc-pVTZ MP2 full/cc-pVTZ b (MHz) 313.98 37.24 41.55 24.89 9.16 4.31 39.39 75.40 39.49 a Energies (hartree) for columns 2–7 are: ⫺245.8946737, ⫺247.3774088, ⫺246.7789693, ⫺245.9178246, ⫺247.3958559, ⫺246.9071573, respectively. No symmetry constraints were used in the calculations. b The observed rotational constants are given in Table 7, Ds represent differences between calculated and observed rotational constants, i.e D calc: ⫺ obs: K.-M. Marstokk et al. / Journal of Molecular Structure 524 (2000) 69–85 Table 1 (continued) Table 2 Molecular structures of the least stable skew conformation of acrylamide from quantum chemical calculations HF/6-311⫹G ⴱⴱ MP2 full/6-311⫹⫹G ⴱⴱ HF/cc-pVTZ DFT/B3LYP/cc-pVTZ MP2 full/cc-pVTZ 122.29 137.78 149.17 100.76 100.98 134.20 108.58 108.55 108.47 119.28 135.87 149.44 98.92 99.21 131.55 107.29 107.38 107.32 121.82 136.85 149.23 100.39 100.61 132.81 108.22 108.27 108.16 121.89 136.45 148.04 100.10 100.35 133.09 107.54 107.69 107.60 122.29 120.74 116.95 121.09 116.74 117.70 125.94 112.97 120.95 122.51 121.11 116.35 122.87 121.32 115.75 118.16 115.01 116.29 124.30 114.41 121.10 121.76 120.86 117.35 122.40 120.44 117.14 120.01 115.96 117.09 125.54 113.21 121.14 122.51 120.94 116.53 122.37 120.72 116.88 120.67 116.45 117.57 126.06 112.87 120.94 122.36 121.22 116.39 122.71 121.15 116.10 119.45 115.76 117.35 125.04 114.04 120.82 121.59 121.01 117.37 154.24 ⫺21.53 162.34 6.63 ⫺19.34 184.95 1.06 182.98 176.53 ⫺1.55 4.98 148.75 ⫺26.27 155.05 11.53 ⫺27.76 188.71 1.41 183.08 176.11 ⫺2.22 3.17 155.54 ⫺20.76 158.55 8.54 ⫺22.92 187.08 0.96 182.68 176.98 ⫺1.30 5.71 155.69 ⫺20.19 161.15 7.53 ⫺20.60 185.88 0.98 182.85 176.56 ⫺1.57 5.79 155.54 ⫺20.77 158.36 9.22 ⫺23.83 172.96? 0.41 182.37 176.49 ⫺1.55 4.75 2.45 3.05 0.86 4.01 2.84 3.04 1.24 4.34 2.52 3.00 0.97 4.04 2.26 2.94 0.83 3.80 2.81 3.02 1.00 4.24 73 122.04 137.15 149.49 100.65 100.87 133.24 108.47 108.51 108.40 K.-M. Marstokk et al. / Journal of Molecular Structure 524 (2000) 69–85 Bond length (pm) O1yC2 119.40 C2–N3 136.16 C2–C6 149.61 N3–H4 99.18 N3–H5 99.46 C6yC7 131.95 C6–H10 107.52 C7–H8 107.62 C7–H9 107.55 Bond angles (⬚) O1C2N3 122.39 O1C2C6 120.51 N3C2C6 117.09 C2N3H4 120.20 C2N3H5 116.05 H4N3H5 117.04 C2C6C7 125.41 C2C6H10 113.31 C7C6H10 121.16 C6C7H8 122.59 C6C7H9 121.86 H8C7H9 116.52 Dihedral angles (⬚) O1C2C6C7 153.78 O1C2C6H10 ⫺22.24 O1C2N3H4 158.72 O1C2N3H5 8.27 C6C2N3H4 ⫺22.80 C6C2N3H5 186.75 C2C6C7H8 1.15 C2C6C7H9 182.89 H10C6C7H8 176.87 H10C6C7H9 ⫺1.38 D Ea 6.34 Dipole moments (Debye) ma 2.66 mb 3.08 mc 1.02 mtot 4.20 DFT/B3LYP/6-311⫹G ⴱⴱ 74 Rotational constants b DA DB DC a b HF/6-311⫹G ⴱⴱ DFT/B3LYP/6-311⫹G ⴱⴱ MP2 full/6-311⫹⫹G ⴱⴱ (MHz) 102.7 56.5 66.8 ⫺81.4 ⫺15.3 9.9 ⫺184.2 12.6 43.8 HF/cc-pVTZ 160.4 71.8 72.7 DFT/B3LYP/cc-pVTZ ⫺18.4 5.0 20.5 MP2 full/cc-pVTZ 56.6 65.4 61.1 Energy difference in kJ/mol relative to the most stable conformation, i.e DE Ecalc: ⫺ Esyn: . The observed rotational constants are given in Table 12, Ds represents differences between calculated and observed rotational constants, i.e. D calc: ⫺ obs: K.-M. Marstokk et al. / Journal of Molecular Structure 524 (2000) 69–85 Table 2 (continued) Table 3 Molecular structures of the transition state between the syn and skew forms for acrylamide from quantum chemical calculations HF/6-311⫹G ⴱⴱ a MP2 full/6-311⫹⫹G ⴱⴱ HF/cc-pVTZ DFT/B3LYP/cc-pVTZ MP2 full/cc-pVTZ 121.76 136.65 150.69 100.64 100.91 132.89 108.73 108.51 108.41 122.11 137.19 150.14 100.70 100.98 133.92 108.75 108.59 108.46 119.14 135.25 150.28 98.86 99.17 131.22 107.55 107.45 107.33 121.54 136.30 150.47 100.35 100.64 132.45 108.49 108.27 108.17 121.69 135.84 149.08 100.00 100.31 132.83 107.78 107.69 107.59 122.78 122.53 114.65 121.63 118.64 119.18 123.47 115.47 121.00 121.69 121.36 116.95 122.10 122.81 114.02 119.45 115.15 118.01 122.09 116.62 121.45 121.14 121.25 117.60 122.84 122.57 114.55 121.15 118.24 119.06 123.40 115.40 121.18 121.77 121.27 116.95 122.87 122.53 114.57 121.48 118.46 119.19 123.55 115.55 120.86 121.67 121.42 116.90 123.17 122.91 113.89 120.76 118.26 119.36 122.28 116.90 120.80 120.98 121.23 117.78 83.02 ⫺94.37 186.17 ⫺2.27 8.27 179.73 1.79 182.73 179.04 ⫺0.02 18.9 86.00 ⫺91.61 195.78 ⫺10.61 18.91 172.52 1.45 182.17 178.96 ⫺0.22 13.9 85.67 ⫺93.10 189.58 ⫺4.78 11.78 177.42 0.56 181.47 179.27 0.16 20.1 83.06 ⫺94.58 187.49 ⫺3.26 9.59 178.83 1.69 182.63 179.22 0.15 20.1 83.12 ⫺95.10 189.85 ⫺4.73 11.67 177.10 1.10 181.95 179.25 0.10 19.7 K.-M. Marstokk et al. / Journal of Molecular Structure 524 (2000) 69–85 Bond length (pm) O1yC2 119.27 C2–N3 135.51 C2–C6 150.44 N3–H4 99.13 N3–H5 99.42 C6yC7 131.62 C6–H10 107.78 C7–H8 107.69 C7–H9 107.55 Bond angles (⬚) O1C2N3 122.81 O1C2C6 122.56 N3C2C6 114.59 C2N3H4 121.37 C2N3H5 118.42 H4N3H5 119.13 C2C6C7 123.41 C2C6H10 115.30 C7C6H10 121.27 C6C7H8 121.82 C6C7H9 121.22 H8C7H9 116.96 Dihedral angles (⬚) O1C2C6C7 86.38 O1C2C6H10 ⫺91.96 O1C2N3H4 188.05 O1C2N3H5 ⫺3.94 C6C2N3H4 10.20 C6C2N3H5 178.21 C2C6C7H8 0.80 C2C6C7H9 181.69 H10C6C7H8 179.04 H10C6C7H9 ⫺0.07 DE a 19.9 DFT/B3LYP/6-311⫹G ⴱⴱ Energy difference in kJ/mol relative to the most stable conformation, i.e DE Ecalc: ⫺ Esyn: : 75 76 Table 4 Molecular structures of the anti transition state for acrylamide from quantum chemical calculations Bond length (pm) O1yC2 119.56 C2–N3 135.41 C2–C6 149.72 N3–H4 98.89 N3–H5 99.30 C6yC7 131.90 C6–H10 107.50 C7–H8 107.62 C7–H9 107.54 Bond angles (⬚) O1C2N3 122.00 O1C2C6 119.75 N3C2C6 118.26 C2N3H4 123.38 C2N3H5 117.73 H4N3H5 118.90 C2C6C7 127.19 C2C6H10 112.16 C7C6H10 120.65 C6C7H8 123.58 C6C7H9 120.64 H8C7H9 115.78 DE a 8.35 a DFT/B3LYP/6-311⫹G ⴱⴱ MP2 full/6-311⫹⫹G ⴱⴱ HF/cc-pVTZ DFT/B3LYP/cc-pVTZ MP2 full/cc-pVTZ 122.18 136.64 149.52 100.48 100.48 133.20 108.46 108.54 108.39 122.53 136.63 149.31 100.37 100.37 134.12 108.58 108.57 108.45 119.46 135.05 149.55 98.61 99.02 131.51 107.27 107.38 107.31 121.98 136.22 149.27 100.18 100.46 132.75 108.22 108.30 108.16 122.07 135.58 148.12 99.79 100.14 133.02 107.54 107.70 107.60 121.89 120.15 117.96 123.38 117.83 118.80 127.52 111.94 120.55 123.40 120.95 115.65 7.21 122.23 120.31 117.46 123.21 117.64 119.15 127.04 112.65 121.30 123.22 120.51 116.28 6.89 122.05 119.71 118.24 123.21 117.75 119.03 127.19 112.13 120.68 123.49 120.72 115.80 7.71 122.00 120.15 117.86 123.23 117.78 118.99 127.62 111.84 120.54 123.30 121.04 115.66 7.25 122.34 120.36 117.30 122.92 117.69 119.39 126.98 112.79 120.23 122.80 120.75 116.44 6.83 Energy difference in kJ/mol relative to the most stable conformation, i.e DE Ecalc: ⫺ Esyn: . K.-M. Marstokk et al. / Journal of Molecular Structure 524 (2000) 69–85 HF/6-311⫹G ⴱⴱ K.-M. Marstokk et al. / Journal of Molecular Structure 524 (2000) 69–85 77 Table 5 Calculated frequencies (Cs symmetry only for B3LYP/cc-pVTZ, the other calculations have no symmetry. Parenthesized values are IR intensities in KM/mol) for the syn and skew conformation of acrylamide No. 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 SymCs A 00 A 00 A0 A0 A 00 A0 A 00 A0 A 00 A 00 A 00 A0 A0 A0 A0 A0 A0 A0 A0 A0 A0 A0 A0 A0 Assignment Tors v NH d CyC–C d skel t NH2 d OCN g CH rCH2 g skel g CH g CH rCH2 rNH2 d CCH d CCH scisCH2 scisNH2 strCyC strCyO strCH strCH strCH strNH strNH Syn Skew B3LYP/cc-pVTZ MP2/6-311⫹⫹G 89.6(107) 137.0(77) 276.9(7) 466.2(5) 469.2(9) 616.8(10) 616.9(3) 811.5(6) 821.2(19) 1008.6((2) 1018.8(44) 1037.9(5) 1113.2(3) 1290.0(108) 1356.3(42) 1443.8(80) 1617.3(113) 1689.6(42) 1764.3(243) 3140.0(9) 3151.1(11) 3236.1(2) 3593.3(45) 3727.6(38) 79.5(3) 424.8(86) 278.6(6) 469.9(75) 478.5(74) 591.0(27) 621.8(28) 783.0(21) 831.5(8) 909.4(27) 1001.5(34) 1049.3(5) 1123.8(9) 1292.7(78) 1360.8(71) 1452.6(86) 1635.7(97) 1688.7(22) 1784.5(280) 3189.8(3) 3209.8(10) 3298.1(1) 3619.3(50) 3755.5(43) structural parameters from the different calculations are given in Tables 1–4, and the numbering of the atoms is shown in Fig. 1. All computations predict the syn rotamer to be the most stable form and most of the calculations predict a planar molecular structure. Only the MP2/6311⫹⫹G ⴱⴱ and the MP2/6-31⫹G ⴱ [14] calculations predict a pyramidal amide conformation for the syn form. The non-planarity of the amide moiety found in these computations is in agreement with our previous calculations [12] for other amides. Use of double or triple zeta basis sets with polarization and in particular diffuse functions, gives a non-planar conformation for the amide group when MP2 is used. This is probably due to an artefact of the basis set and the MP2 method when applied to amides. Some bonds lengths (Tables 1–4) are underestimated at the HF level of theory as compared with the bond lengths predicted from MP2 and DFT. ⴱⴱ B3LYP/cc-pVTZ MP2/6-311⫹⫹G ⴱⴱ 99.9 353.8 277.6 432.0 525.1 561.3 592.9 819.0 824.1 986.4 1040.3 1043.9 1116.6 1311.6 1359.9 1455.7 1617.6 1684.8 1758.6 3137.9 3178.8 3219.3 3584.2 3749.7 108.6 410.0 279.3 490.9 529.1 558.9 611.2 798.1 938.8 937.4 1022.6 1048.6 1129.9 1304.8 1370.1 1464.8 1628.5 1668.7 1774.6 3189.0 3232.3 3287.3 3616.3 3711.1 DFT and MP2 give similar bond length. All bond angles are remarkably similar at the three levels of theory even with different basis sets. This is also the case for the dihedral angles except for the nonplanarity of the amide group which is predicted only in the MP2 calculation, as mentioned above. Comparison of the calculated rotational constants of the syn form with the experimental ones as shown in Tables 1 and 7, indicates that the DFT/B3LYP structure is better than both the HF and MP2 structures, because the experimental and theoretical rotational constants are in better agreement. This is a useful result because the rotational constants and the dipole moment obtained from theory often are used to predict the frequencies of the strongest transitions in the MW spectrum. Good estimates will of course make the assignments easier. All calculations predict the existence of a second skew conformation whose theoretical structures are given in Table 2. The dihedral angle O1C2C6C7 is 78 K.-M. Marstokk et al. / Journal of Molecular Structure 524 (2000) 69–85 Table 6 Experimental and theoretical findings for the lowest vibrational frequencies of acrylamide IR MW Ab initio (syn) Proposed assignment Obs. (cm ⫺1) Assign. Obs. (cm ⫺1) Assign. Calc. (cm ⫺1) Assign. 262.8 n 2(1 ← 0)wag NH torsion wag NH d C C–C 89.6 137.0 276.9 torsion wag NH d CyC–C 344.7 352.7 378.5 n 2(2 ← 1)wag NH fund. out-of-pl n 2(3 ← 2)wag NH ⬇ 90 ⬇ 235 ^ 40 ⬇ 307 ^ 40 ⬇ 398 ^ 60 n1 ⫹ n2 ⬇ 420 ^ 60 n1 ⫹ n3 366.4 466.2 469.2 616.8 616.9 n1 ⫹ n3 456.4 t (CN) 614.9 Comp. band about 155⬚ from syn (0⬚) in all computations except for one (148.75⬚; MP2/6-311⫹⫹G ⴱⴱ). All calculations predict a pyramidal amide group for this rotamer, which is computed to be from 3 to 6 kJ mol ⫺1 less stable than the syn conformation. These numbers are smaller than the estimate of at least 600 cm ⫺1 (7.2 kJ mol ⫺1) from the IR spectra [17], and our value of 6.5(6) kJ mol ⫺1 from our relative intensity measurements. We succeeded in assigning the skew form (see below). It is shown in the same table that the best agreement between the calculated and experimental rotation constants again is found for the DFT/ B3LYP calculations, which are therefore assumed to produce the most accurate structure. The structures of the two transitional states are given in Tables 3 and 4. The transition state between the syn and the skew conformations occurs for a dihedral angle where the vinyl group is approximately perpendicular to the amide group. The barrier height is computed to be about 20 kJ mol ⫺1 relative to syn from all calculations except for the MP2/6311⫹⫹G ⴱⴱ computations, which yield a barrier height of 14 kJ mol ⫺1. The anti barrier is predicted to be between 1.2 and 3.7 kJ mol ⫺1 in the different calculations. 3.2. Vibrational frequencies The calculated frequencies for the two conformations are given in Table 5. Tentative assignments based on the displacement of the atoms for each d OyC–N g CH n 1 ⬇ 90 n 2262.8 n 3 ⬇ 307 ^ 40 n 2(2 ← 1) 344.7 n 1 ⫹ n 2 ⬇ 353 n 2(3 ← 2) 378.5 n1 ⫹ n3 n4 n5 n6 n7 CC torsion wag NH sk bend wag NH wag NH/n 1 ⫹ n 3 sk bend CN torsion sk bend CH out-of-pl fundamental mode for the syn conformations are also given in this table. For the skew conformation the calculated frequencies are quite similar at both levels of theory. It should be noted (see above) that these two calculations both predict a non-planar amide group for the skew form. However, the corresponding computations of the frequencies for the syn form yield a widely different result. The N–H wagging frequency changes radically from 137.0 to 424.8 cm ⫺1 in these two calculations. It is remarked above that the B3LYP/cc-pVTZ procedure predicts a planar conformation while the MP2/6311⫹⫹G ⴱⴱ method predicts a pyramidal amide group for the syn rotamer. There are not many experimentally reported observed frequencies for acrylamide, but some are available [14,17,27]. In Table 6 the experimental findings for the lowest frequencies are summarized. The crucial point about the amide planarity is the IR assignment [17] of the very weak band NH wag n2 3 ! 2 378:5 cm⫺1 which can also be the n1 ⫹ n3 combination band which is expected to fall in this region (Table 6). The assignment of the NH wag n2 3 ! 2 378:5 cm⫺1 is likely to be correct, and then the IR investigation shows that the amide unit is planar. Vibrational frequencies can be obtained from MW spectroscopy by means of relative intensity measurements which have been carried out as described in Ref. [28]. The large uncertainties for this method are due to the difficulties of estimating the position of the base line. K.-M. Marstokk et al. / Journal of Molecular Structure 524 (2000) 69–85 79 Table 7 Spectroscopic constants (A-reduction, I r representation [44]) for the ground vibrational state and five C2–C6 torsion vibrational states Av (MHz) Bv (MHz) Cv (MHz) DJ (kHz) DJK (kHz) DK (kHz) dJ (kHz) dK (kHz) FJ (Hz) FJK (Hz) FKJ (Hz) FK (Hz) fJ (Hz) fJK (Hz) fK (Hz) IC ⫺ IA ⫺ IB (10 ⫺20u m 2) No a RMS b a b v0 v1 v2 v3 v4 v5 10732.8296(34) 4218.7012(13) 3030.7434(13) 0.7043(33) 3.370(22) 5.403(8) 0.2417(7) 3.20(4) ⫺0.066(6) ⫺0.222(34) ⫺1.54(13) 0.57(4) 0.00170(34) ⫺0.60(5) 1.98(16) ⫺0.131300(34) 354 0.081 10660.7092(33) 4216.4693(11) 3038.0502(17) 0.696(14) 3.15(7) 5.100(24) 0.2394(26) 3.37(7) ⫺0.208(33) ⫺2.3(4) ⫺3.5(5) 0.38(4) ⫺0.0022(4) ⫺0.74(12) 1.85(17) ⫺0.91431(7) 322 0.063 10594.1266(38) 4214.2765(14) 3044.8733(16) 0.776(7) 3.658(31) 5.055(15) 0.2418(19) 3.41(5) ⫺0.007(7) ⫺0.65(11) ⫺0.57(17) 0.11(10) ⫺0.0107(26) 0.64(13) 0.40(34) ⫺1.64738(6) 317 0.090 10531.951(6) 4212.1010(22) 3051.3677(26) 0.824(9) 3.94(5) 5.119(24) 0.2403(35) 3.52(9) 0.036(11) ⫺1.27(28) 2.3(7) ⫺1.6(5) ⫺0.20(6) 1.19(25) ⫺5.5(15) ⫺2.34419(10) 206 0.112 10473.259(21) 4209.902(15) 3057.626(15) 0.81(11) 3.69(15) 5.9(8) 0.227(8) 3.26(17) ⫺0.066 ⫺0.222 ⫺1.54 0.57 0.0017 ⫺0.60 1.98 ⫺3.01477(35) 42 0.193 10417.592(23) 4207.719(20) 3063.765(19) 0.87(15) 3.87(9) 5.1(5) 0.241(6) 3.44(13) ⫺0.066 ⫺0.222 ⫺1.54 0.57 0.0017 ⫺0.60 1.98 ⫺3.6661(4) 48 0.202 Number of transitions. Root-mean-squares deviation. For the C2–C6 torsional frequency of the syn rotamer relative intensity measurements gave 90 ^ 10 cm⫺1 : This frequency can also be estimated for planar molecules from the inertial defect [29,30]. Oka [30] has shown that the torsional fundamental can be calculated from the formula: 1=2 D0 ⫺33:715=n1 ⫹ 0:0186Icc where D 0 is the inertial defect for the ground state and Icc is the principal moment of inertia about the c-axis. This gives a value of 91 cm ⫺1 in our case. Hanyu, Britt and Boggs [29] have obtained yet another formula for the calculation of the torsional frequency from the difference between the inertial defect of two consecutive torsionally excited states: Dnt⫹1 ⫺ Dnt ⫺67=nt where D n t⫹1 and D n t are the inertia defects of the two states and n t is the torsional frequency. This yielded a frequency of 85 cm ⫺1. The other frequencies of the syn form given in Table 6 were obtained from relative intensity measurements. 3.3. MW spectrum and assignment of the syn form The MW spectrum consists of a relatively strong and dense b-type spectrum. The most stable conformation is predicted from the quantum chemical computations to be the syn form. The largest component of the dipole moment of this rotamer is mb ⬇ 3:7 D (Table 1). Searches were first made for the strong bQ-branch transitions using rotational constants obtained from the B3LYP/6-311⫹G ⴱⴱ calculations as the starting point. These ground state transitions were readily identified close to their predicted frequencies. The b-type R-branch transitions were assigned next. The strongest bQ-branch transitions for the first vibrationally excited states of the torsional motion about the C2–C6 bond were then assigned. Following the progression of these transitions we were able to assign altogether five vibrational excited states of this molecule. The results (A reduction, I r representation) for the ground state and the five vibrationally excited states of the torsion are given in Table 7. The n 2 wag NH was expected to have rotational constants close to the ground state rotational constants mainly because the light H-atoms attached to the 80 K.-M. Marstokk et al. / Journal of Molecular Structure 524 (2000) 69–85 Table 8 Spectroscopic constants a for six vibrational excited states (same units as in Table 7. The seven sextic centrifugal distortions constants are fixed at zero) Av Bv Cv DJ DJK DK dJ dK IC ⫺ IA ⫺ IB No. RMS n 2 d CyC–C n 3 wag NH n1 ⫹ n2 n1 ⫹ n3 2n 1 ⫹ n 2 2n 1 ⫹ n 3 10765.263(13) 4217.684(4) 3025.270(4) 0.725(12) 4.06(9) 5.30(5) 0.243(5) 3.40(12) 0.28333(14) 57 0.144 10724.935(10) 4209.4293(31) 3029.3310(33) 0.784(10) 3.77(16) 5.31(13) 0.225(4) 3.21(8) ⫺0.35208(12) 65 0.139 10684.740(19) 4214.551(15) 3032.921(14) 1.18(12) 3.85(12) 5.1(7) 0.239(6) 3.32(13) ⫺0.5809(4) 38 0.159 10652.745(20) 4206.784(16) 3035.993(16) 0.86(12) 4.05(11) 5.0(7) 0.236(6) 2.92(12) ⫺1.1130(4) 44 0.185 10614.352(14) 4212.680(5) 3040.947(6) 0.814(20) 4.15(11) 4.2(6) 0r.251(7) 3.03(13) ⫺1.38762(23) 36 0.153 10587.472(16) 4205.643(11) 3043.632(10) 0.67(8) 3.41(11) 7.5(6) 0.220(7) 4.75(14) ⫺1.85584(27) 28 0.131 N-atom are involved in this out-of-plane motion. By comparing the Stark effect of the strongest unassigned bQ-branch transitions in the vicinity of the ground state transitions, this vibrationally excited state was identified. The negative inertial defect confirms that this excited state belongs to an out-of-plane motion [32,33]. The n 3 in-plane skeleton bending transition was searched for further away from the ground state transitions, and the positive inertial defect confirms [32,33] that the assigned excited state belongs to an in-plane motion. The two n1 ⫹ n2 and n1 ⫹ n3 vibrational excited states were readily assigned from their estimation of their rotational constants as: Xn1⫹n2 Xn1 ⫹ Xn2 ⫺ X0 where X is the rotational constants A, B or C. The estimated values for the n1 ⫹ n2 rotational constants are 10693.2, 4215.5 and 3032.7 MHz which can be compared with observed values (Table 8) 10684.7, 4214.5 and 3032.9 MHz. The estimated values for the n1 ⫹ n3 rotational constants are 10652.8, 4207.2 and 3036.4 MHz which can be compared with observed values (Table 9) 10652.7, 4206.8 and 3036.0 MHz. The same procedure was used to assign Table 10 Stark coefficients and dipole moment of acrylamide (uncertainties represent one standard deviation) Table 9 Spectroscopic constants (same units as in Table 7. D4 means that the H-atom numbering 4 (see Fig. 1) has been replaced with deuterium. The seven sextic centrifugal distortions constants are fixed to zero) for three isotopic species D4 A B C DJ DJK DK dJ dK IC ⫺ IA ⫺ IB No RMS D5 D4D5 10122.825(25) 10716.108(22) 10116.491(18) 4139.638(9) 4017.565(9) 3947.488(6) 2940.775(6) 2924.594(9) 2842.336(4) 0.72 0.72 0.72 3.10(19) 3.59(15) 3.13(13) 5.43 5.43 5.43 0.243(17) 0.202(15) 0.184(8) 3.1 3.1 3.1 ⫺0.1553(4) ⫺0.1499(5) ⫺0.17731(24) 14 16 17 0.104 0.143 0.082 Transition 93,6 ← 92,7 兩M兩 9 8 7 6 5 83,5 ← 82,6 8 7 6 5 4 53,2 ← 52,3 5 Dipole moment a (Debye) ma 0:269 3 mb 3:42 2 a Dn E ⫺2/10 ⫺5 (MHz V ⫺2 cm 2) Obs. Calc. 3.360 2.680 2.000 1.460 0.987 1.870 1.380 1.020 0.637 0.402 ⫺1.390 3.390 2.661 2.014 1.461 0.986 1.865 1.412 1.009 0.673 0.397 ⫺1.390 mc 0:12 24 1 Debye 3:33564 × 10⫺30 C m: K.-M. Marstokk et al. / Journal of Molecular Structure 524 (2000) 69–85 Table 11 14 N nuclear quadrupole splitting (Eq) of acrylamide and diagonal elements of the 14N quadrupole coupling tensor Transition F0 ← F Eq (MHz) Obs. 2←1 0.65 22,1 ← 11,0 22,1 ← 11,0 3←2 ⫺0.07 22,0 ← 11,1 3←2 0.26 31,3 ← 20,2 3←2 0.64 42,3 ← 41,4 4←4 ⫺1.10 52,4 ← 51,5 5←5 ⫺1.13 6←6 ⫺1.15 62,5 ← 61,6 71,6 ← 70,7 7←7 ⫺1.49 81,7 ← 80,8 8←8 ⫺1.38 82,6 ← 81,7 8←8 ⫺0.82 92,7 ← 91,8 9←9 ⫺0.91 14 N nuclear quadrupole coupling constants a (MHz) xaa 1:86 26 xbb 2:73 16 Calc. 0.68 ⫺0.10 0.31 0.70 ⫺1.19 ⫺1.13 ⫺1.09 ⫺1.45 ⫺1.32 ⫺0.81 ⫺0.93 a Uncertainties represent one standard deviation. Some constants (xaa ,xbb )/MHz for other amides are: formamide [45] (1.94(3),1.90(3)); acetamide [46] (1.9368(24),2.0090(22)); 2fluoroacetamide [46] (1.6595(26),2.0413(27)); propionamide [8] (2.2(8),2.3(5)). the 2n1 ⫹ n2 and the 2n1 ⫹ n3 vibrational excited state. We believed that we would be able to calculate very accurately the shift in the rotational constants of the ground state to the respective isotopic species using the B3LYP/6-311⫹G ⴱⴱ molecular structure. This was done for the all the 13C isotopic species, but we were not able to assign any of these isotopic species in their natural abundance owing to insufficient intensities. This 81 was also done for the three isotopomers obtained when the H-atoms in the amide group were replaced with deuterium. All the three isotopic species were assigned in a straightforward manner. 3.4. Dipole moment of the syn form The 11 Stark coefficients (Table 10) of transitions with comparatively small quadrupole splittings were used to determine the dipole moments following the standard procedure [31]. Field strengths were typically in the 500–1000 V cm ⫺1 range. Quadrupole interaction is small and is not taken into account in the least-squares fit. All three principal axis dipole moment components were fitted in order to get information of the planarity of the amide group. 3.5. 14N quadrupole coupling constants of the syn form All transitions were split by a small amount by quadrupole interaction caused by the 14N nucleus. Resolved quadrupole components were only observed in a few cases. The components appearing in Table 11 were used to determine the quadrupole coupling constants of the 14N nucleus following the procedure of Ref. [31]. The spectroscopic constants in Table 7 were used to predict the unperturbed center frequencies. The 14N nuclear quadrupole splittings (Eq in Table 11) were found by subtracting the calculated center frequencies from the frequencies of the resolved components. A rigid-rotor Hamiltonian was used in the least squares fitting procedure in order to Table 12 Spectroscopic constants (same units as in Table 7. The seven sextic centrifugal distortions constants are fixed to zero.) for the skew conformation Av Bv Cv DJ DJK DK dJ dK IC –IA –IB No. RMS a Ground state v 0 10049.549(22) 4287.924(12) 3035.806(12) 1.14(9) 4.20(12) 8.0(6) 0.283(9) 3.23(15) ⫺1.67697(35) 53 0.196 Ex. state v torsion 10002.8(5) 4292.160(24) 3050.385(33) 1.15(16) 3.35(18) 8.0 0.28 3.2 ⫺2.591(4) 14 0.129 a-type R-branch transitions from the MWRFDR experiment. Ex. state a v dCyC–C 9999.4(47) 4287.50(27) 3047.86(30) 1.1 4.2 8.0 0.28 3.2 ⫺2.60(5) 6 0.122 82 K.-M. Marstokk et al. / Journal of Molecular Structure 524 (2000) 69–85 Table 13 Comparison of calculated (calculated rotational constants using nT 0 through nT 5: and observed (observed rotational constants are listed in Table 7) rotational constants (MHz) using B 2:8 nT A (MHz) Calc.⫺obs. B (MHz) Calc.⫺obs. 0 ⫺0.76 ⫺0.004 1 1.57 ⫺0.004 2 0.36 0.000 3 ⫺0.89 ⫺0.011 4 ⫺0.98 ⫺0.013 5 1.19 ⫺0.006 Calculated rotational constants (MHz) An 10774.3(21)⫺159(5)具z 2典n ⫺11.4(14)具z 4典n Bn 4219.86(2)⫺4.13(4)具z 2典n ⫺0.832(11)具z 4典n An 3026.66(17)⫺15.3(4)具z 2典n ⫹ 1.55(11)具z 4典n C (MHz) Calc.⫺obs. 0.059 ⫺0.085 ⫺0.011 0.095 0.140 ⫺0.034 determine the 14N nuclear quadrupole coupling constants x aa and x bb. The results are shown in Table 11, and some constants for other amides are given for comparison. 3.6. Assignment of the skew form This rotamer is predicted to have a sizeable dipole moment along the a-inertial axis. MWRFDR-searches were therefore carried out in order to assign the a-type R-branch transitions of this conformer. The strongest transitions were expected to occur in the 50–60 GHz region, and the searches were carried out here using the DFT/B3LYP rotational constants to predict the spectrum. The rather weak aR-lines were readily found using this method. Confirmation of their assignments were obtained from Stark effect studies, rigid-rotor fit and b-type Q-branch transitions that were subsequently assigned. The resulting spectroscopic constants are listed in Table 12. The dipole moment of this rotamer could not be measured owing to insufficient intensities. The quadrupole coupling constants of the 14N nucleus with the molecular rotation were not determined because of the weakness of the spectrum. Two vibrationally excited states were found for this rotamer; their spectroscopic constants are found in Table 12. Rough relative intensity measurements yielded 91(20) cm ⫺1 for what is assumed to be the C–C torsional vibration, and ca. 208(70) cm ⫺1 for the lowest heavy-atom bending vibration. These two values compare well with the corresponding frequencies in Table 5 (99.9 and 277.6 cm ⫺1, B3LYP). 3.7. Energy difference between syn and skew form Relative intensity measurements were carried out as described in Ref. [28] to determine the energy difference. This difference depends on the ratio of the dipole moments. The values taken from the B3LYP calculations (Tables 1 and 2) were used to derive a value of 6.5(6) kJ mol ⫺1, slightly larger than the B3LYP/cc-pVTZ value of 5.8 kJ mol ⫺1. The estimated standard deviation of ^0.6 kJ mol ⫺1 is assumed to take both systematic and random errors into account. 3.8. Heavy atom planarity and potential function of the syn form The smooth variation of the rotational constants for the vibrationally excited torsional fundamental vibration of the syn rotamer indicates that there is no barrier to the planar syn form [32], and that this torsional mode is nearly harmonic. Gwinn et al. [34] and Legon [35] have given a quantitative treatment of this problem. They have shown that it is possible to define a potential function for the torsion as: V A 具z4 典 ⫹ B具z4 典 where z is a dimensionless coordinate. If B is positive, the heavy atom skeleton has a symmetry plane, if B is negative, the equilibrium conformation will be nonplanar. According to this theory the rotational constants can be expanded in a power series of the expectation values 具z2 典 and 具z4 典; where bn is the An, Bn or Cn rotational constants in the nth excited state of the torsion. The empirical parameters b0, b2 and b4 are adjusted to give the best fit to the data. The values 具z2 典n and 具z4 典n depend only on the constant B in the equation given above. The rotational constants of successively excited states of the torsional vibration were least-squares fitted to the equation: Bn b0 ⫹ b2 具z2 典n ⫹ b4 具z4 典n employing a program described in Ref. [36] for a series of B-values. It was found that the value B 2:8 yield the best overall fit (Table 13). This positive B-value indeed shows that the heavy atom skeleton is K.-M. Marstokk et al. / Journal of Molecular Structure 524 (2000) 69–85 83 Fig. 2. Potential function 22:2 具z4 典 ⫹ 2:8具z2 典 describing the torsional vibration around the C2–C6 bond. Eigenstates are indicated. planar. The constant A was then adjusted to reproduce the torsional fundamental frequency of 90 cm ⫺1. This was achieved with A 22:2 cm⫺1 : The potential function derived in this manner is sketched in Fig. 2. 3.9. Planarity of the amide group of the syn form A non-planar conformation of the amide group of the syn form would produce a component of the dipole moment along the principal inertial c-axis mc that would be non-zero. The theoretical calculations indicate a very small mc except for MP2/6-311⫹G ⴱⴱ where mc 0:75 D: The dipole moment is a rather sensitive test of planarity. However, with the transitions we have used, the dipole moment along the principal inertial c-axis is not very accurately determined (mc 0:12(24) D, Table 10) which indicates that the amide group is less non-planar than the MP2/6311⫹G ⴱⴱ calculations predict. When mc is fixed to zero the fit is insignificantly larger. A definite conclusion cannot be drawn from the mc dipole moment component, but it can safely be stated that a large deviation from planarity is ruled out. Another evidence that acrylamide is planar, or very nearly planar, comes from the value of inertial defect which is zero for a completely rigid planar molecule. The inertial defect for syn acrylamide is as large as ⫺0:131300 34 × 10⫺20 u m2 (Table 7). However, an amide group with floppy hydrogen atoms could produce a deviation from the ideal zero value. This has in fact been found for the related compound formamide [37,38] which has an inertial defect of ⫺0:006510 2 × 10⫺20 u m2 [37] and is planar. If the contribution from the out-of-plane hydrogen atoms of the methyl group is arbitrarily assumed to be a pseudo-inertial defect 3:20 × 10⫺20 u m2 ; (10 ⫺20 u m 2) of some other amides have been estimated to be: ⫺0.122 for acetamide [39], ⫺0.078 for 2-fluoroacetamide [11], ⫺0.314 for propionamide [8] and ⫺0.156 for acetidinone [36]. The values for some other related compounds are: 0.0169 for trans acrolein [40], ⫺0.0195 for cis acrolein [41], ⫺0.041 and ⫺0.049 for cis and trans acrylic acid [42]. All of these molecules are assumed to be planar. Acrylamide is also planar in the solid state [43] where it again prefers the syn conformation. The inertial defects of the normal (Table 7) and the deuterated species (Table 9) vary little. This is expected for a planar or nearly planar molecule. We are again led to the conclusion that the syn rotamer has a planar, or very nearly planar amide group. There is no sufficient experimental evidence 84 K.-M. Marstokk et al. / Journal of Molecular Structure 524 (2000) 69–85 to say anything definite about the planarity of the amide group of the skew conformer. at the University of Oslo. Anne Horn is thanked for the artwork. 3.10. Structures References Comparison of the experimental (Table 7 and 12) and the theoretical rotational constants (Table 1 and 2) show the best agreement in the case of the B3LYP computations. This agreement is not believed to be fortuitous but to reflect the fact that the B3LYP structures of the syn and skew forms are indeed accurate. The B3LYP structures are taken to be plausible structures of the two forms. It is expected that any experimental structures that are determined in the future for the syn and skew rotamers will be very close to the B3LYP structures in Tables 1 and 2. [1] [2] [3] [4] [5] [6] [7] [8] [9] [10] 4. Conclusions Gaseous acrylamide exists in the heavy-atom syn and skew conformations. The syn form (Fig. 1) is the most stable conformation being 6.5(6) kJ mol ⫺1 more stable than skew. The syn rotamer has a planar heavy atom skeleton and a planar, or very nearly planar, amide group. Eleven vibrationally excited states, the dipole moment and the quadrupole coupling constants of the 14N nucleus were determined for this rotamer. The torsion around the C–C single bond is close to harmonic near the syn bottom of the potential well. The torsional fundamental frequency is 90(10) cm ⫺1. Less information is available for the skew rotamer which is obtained from the syn form by rotating approximately 155⬚ around the C–C bond. The quantum chemical calculations predict a non-planar amide group for the skew form, but there is insufficient evidence to confirm or reject this prediction. Elaborate quantum chemical calculations have been carried out for the two forms. The best agreement with experimental values is found in the B3LYP/ cc-pVTZ calculations. Acknowledgements We are grateful to The Research Council of Norway (Programme for Supercomputing) for a grant of computer time at the IBM RS6000 cluster [11] [12] [13] [14] [15] [16] [17] [18] [19] [20] [21] [22] [23] [24] [25] [26] [27] [28] S. Samdal, R. Seip, J. Mol. Struct. 52 (1979) 195. S. Samdal, R. Seip, J. Mol. Struct. 62 (1980) 131. M. Hargittai, S. Samdal, R. Seip, J. Mol. Struct. 71 (1981) 147. L. Fernholt, S. Samdal, R. Seip, J. Mol. Struct. 72 (1981) 217. S. Samdal, R. Seip, J. Mol. Struct. 413/414 (1997) 423. S. Gundersen, S. Samdal, R. Seip, D.J. Shorokhov, T.G. Strand, J. Mol. Struct. 445 (1998) 229. S. Gundersen, S. Samdal, R. Seip, D.J. Shorokhov, J. Mol. Struct. 477 (1999) 225. K.-M. Marstokk, H. Møllendal, S. Samdal, J. Mol. Struct. 376 (1996) 11. S. Gundersen, V.P. Novikov, S. Samdal, R. Seip, D.J. Shorokhov, V.A. Sipachev, J. Mol. Struct. 485–486 (1999) 97. K.-M. Marstokk, H. Møllendal, S. Samdal, Acta Chem. Scand. 50 (1996) 845. K.-M. Marstokk, H. Møllendal, J. Mol. Struct. 22 (1974) 287. S. Samdal, J. Mol. Struct. 440 (1998) 165. D.C. MacWilliams, in: R.H. Yocum, E.B. Nyquist (Eds.), Functional Monomers, 1, Marcel Dekker, New York, 1973, pp. 1–197. C.K. Lin, S.Y. Chen, M.H. Lien, J. Phys. Chem. 99 (1995) 1454. D. McCollister, F. Oyen, V.K. Rowe, Toxicol. Appl. Pharmacol. 6 (1964) 172. P.S. Spencer, H.H. Schaumburg, Environ. Health Perspect. (EHP) 11 (1975) 129. R.A. Kydd, A.R.C. Dunham, J. Mol. Struct. 69 (1980) 79. U. Berg, N. Bladh, J. Comput. Chem. (1996) 17396. G.A. Guirgis, K.-M. Marstokk, H. Møllendal, Acta Chem. Scand. 45 (1991) 482. Ø. Waal, personal communication, 1994. F.J. Wordarczyk, E.B. Wilson, J. Mol. Spectrosc. 37 (1971) 445. K.-M. Marstokk, H. Møllendal, Acta Chem. Scand. Ser. A 42 (1988) 374. M.J. Frisch, G.W. Trucks, H.B. Schlegel, P.M.W. Gill, B.G. Johnson, M.A. Robb, J.R. Cheeseman, T. Keith, G.A. Petersson, J. A. Montgomery, K. Raghavachari, M.A. AlLaham, V.G. Zakrzewski, J.V. Ortiz, J.B. Foresman, J. Cioslowski, B.B. Stefanov, A. Nanayakkara, M. Challacombe, C.Y. Peng, P.Y. Ayala, W. Chen, M.W. Wong, J.L. Anders, E.S. Replogle, R. Gomperts, R.L. Martin, D.J. Fox, J.S. Binkley, D.J. Defrees, J. Baker, J.P. Stewart, M. HeadGordon, C. Gonzales, J.A. Pople, Gaussian94 (Revision E.2) Gaussian, Inc., Pittsburgh, PA, 1995. C. Møller, M.S. Plesset, Phys. Rev. 46 (1934) 618. A.D. Becke, J.Chem. Phys. 98 (1993) 5648. T.H. Dunning Jr, J. Chem. Phys. 90 (1987) 1007. J.S. Suh, K.H. Michaelien, J. Raman Spectrosc. 18 (1987) 409. A.S. Esbitt, E.B. Wilson, Rev. Sci. Instrum. 34 (1963) 901. K.-M. Marstokk et al. / Journal of Molecular Structure 524 (2000) 69–85 [29] Y. Hanyu, C.O. Britt, J.E. Boggs, J. Chem. Phys. 45 (1966) 4725. [30] T. Oka, J. Mol. Struct. 352/353 (1995) 225. [31] K.-M. Marstokk, H. Møllendal, Acta Chem. Scand. Ser. A 36 (1982) 517. [32] D.R. Herschbach, V.W. Laurie, J. Chem. Phys. 37 (1962) 1668. [33] D.R. Herschbach, V.W. Laurie, J. Chem. Phys. 40 (1964) 3142. [34] W.D. Gwinn, A.S. Gaylord, International Review of Science, Series Two, 3, Butterworth, London, 1976, p. 205. [35] A.C. Legon, Chem. Rev. 80 (1980) 231. [36] K.-M. Marstokk, H. Møllendal, S. Samdal, E. Uggerud, Acta Chem. Scand. 43 (1989) 351. [37] E. Hirota, R. Sugisaki, C.J. Nielsen, G.O. Sørensen, J. Mol. Spectrosc. 49 (1974) 251. 85 [38] R.D. Brown, P.D. Godfrey, B. Kleibömer, J. Mol. Spectrosc. 124 (1987) 34. [39] T. Koijima, E. Yano, K. Nakagawa, S. Tsunekawa, J. Mol. Spectrosc. 122 (1987) 408. [40] M. Winnewisser, G. Winnewisser, T. Honda, E. Hirota, Z. Naturforsch. 30a (1975) 1001. [41] C.E. Blom, A. Bauder, Chem. Phys. Lett. 88 (1982) 88. [42] K. Bolton, N.L. Owen, J. Sheridan, Nature 218 (1968) 266. [43] I.V. Isakov, Zh. Strukt. Khim. 7 (1966) 898 (in Russian). [44] J.K.G. Watson, in: J.R. Durig (Ed.), Vibrational Spectra and Structure, 6, Elsevier, Amsterdam, 1977, p. 1. [45] J.C. Ribes, S.G. Ables, P.D. Godfrey, R.D. Brown, Aust. J. Phys. 26 (1973) 79. [46] N. Heineking, H. Dreizler, Z. Naturforsch. 48a (1993) 787.