Journal of Molecular Spectroscopy 244 (2007) 160–169 www.elsevier.com/locate/jms Rotational spectrum and structure of asymmetric dinitrogen trioxide, N2O3 J. Demaison a a,* , M. Herman a, J. Liévin a, L. Margulès b, H. Møllendal c Laboratoire de Chimie quantique et Photophysique, CP160/09, Université libre de Bruxelles (U.L.B.), ave. F.D. Roosevelt, 50, B-1050 Brussels, Belgium b Laboratoire de Physique des Lasers, Atomes, et Molécules, UMR CNRS 8523, Université de Lille I, F-59655 Villeneuve d’Ascq Cédex, France c Department of Chemistry, University of Oslo, P.O. Box 1033, Blindern, NO-0315 Oslo, Norway Received 11 April 2007; in revised form 7 June 2007 Available online 19 June 2007 Abstract The rotational spectra of the ground vibrational state and the m9 = 1 torsional state have been reinvestigated and accurate spectroscopic constants have been determined. The torsional frequency, m9 = 70(15) cm1, has been determined by relative intensity measurements. The assignment of the infrared spectrum has been slightly revised and an accurate harmonic force field has been calculated. The equilibrium structure has been determined using different, complementary methods: experimental, semi-experimental and ab initio, leading to r(NN) = 1.870(2) Å, in particular. 2007 Elsevier Inc. All rights reserved. Keywords: Microwave; Ab initio; Structure; Force field; N2O3 1. Introduction Due to chemical equilibrium, NO2 and N2O4 always coexist. Adding NO in the gas phase leads to the formation of N2O3. It may exist in two isomeric forms, sym and asym but only the asymmetric species has been detected in gas phase. This heterodimer is more stable than a van der Waals complex but its binding energy is much lower than that of a typical covalent bond. For this reason, the study of the chemical bonding of N2O3 is quite interesting. Particularly, as in N2O4 [1], the NAN bond length is substantially longer than in N2H4 but has not yet been determined accurately. The microwave spectrum of N2O3 was first measured by Kuczkowski [2]. He also obtained the rotational constants for the 15N isotopologues and determined an approximate effective value (r0) for the NAN bond length. These * Corresponding author. Permanent address: Laboratoire de Physique des Lasers, Atomes, et Molécules, Université de Lille I, F-59655 Villeneuve d’Ascq Cédex, France. Fax: +33 3 20 33 70 20. E-mail address: jean.demaison@univ-lille1.fr (J. Demaison). 0022-2852/$ - see front matter 2007 Elsevier Inc. All rights reserved. doi:10.1016/j.jms.2007.06.003 measurements were extended by Brittain et al. [3] who measured the microwave spectra of six isotopologues and derived a complete substitution (rs) structure. The spectra of the four lowest vibrational states were analyzed and approximate vibrational frequencies were obtained from relative intensity measurements. The electric dipole moment was determined for the ground vibrational state. The nuclear quadrupole coupling constants of the nitrogen atoms were also measured by conventional Stark spectroscopy [4] and using a pulsed-beam Fourier transform microwave spectrometer [5,6]. The infrared spectrum of N2O3 was studied repeatedly, starting with the work of D’Or and Tarte [7]. The work prior to 1991 is summarized in Ref. [8]. The more recent investigations can be reviewed as follows. Some controversy remained for a long time in the literature concerning the bands near 260 and 160 cm1 either assigned to the NAN stretching or NO rocking modes. The analysis of the Raman spectrum in a NO matrix and a force field calculation definitely indicated that the band at 266 cm1 is the NAN stretching mode [9]. Another problem in the literature concerned m8, the NO2 out-of-plane wag, assigned J. Demaison et al. / Journal of Molecular Spectroscopy 244 (2007) 160–169 from gas phase infrared spectra to the band at 337 cm1 [10]. The Raman investigation of another band, at 627 cm1, gave unambiguous evidence for a fundamental excitation, therefore attributed to m8. However, this modified assignment disagrees with ab initio predictions [1]. Concerning m1 (NO stretching), the fundamental and first three overtone bands were rotationally analyzed from spectra recorded under high resolution with a Fourier transform spectrometer [11]. The accuracy of the ground state parameters was improved in the latter reference by fitting rotational transitions together with ground state combination differences (GSCDs). More recently, m1 was recorded with a Fourier transform spectrometer under jet-cooled experimental conditions which considerably simplifies the rotational structure [12]. The m3 band at 1304 cm1 was also analyzed using a slit-jet tunable-diode-laser spectrometer [13]. A valence-bond study of the origin of the long, weak NAN bond was recently published [14].The properties of N2O3 (and other nitrogen oxides) have been calculated by the density functional method (DFT) [1]. A good agreement was found with the experimental results except for the torsion, m9 and for m5 and m8 identified around 410 and 630 cm1, respectively. It was suggested to reverse the assignment, namely the band situated at 630 cm1 should be associated with the NO2 in-plane rock and that at 410 cm1 with the NO2 out-of-plane wag, in agreement with the earlier infrared assignment. Despite this body of work, there are still some problems to solve. First, the structure is an empirical rs structure which might be rather inaccurate. Then, the torsional frequency was determined from relative intensity measurements in the microwave range. Its accuracy seems to be rather poor. Furthermore, its value does not agree with the ab initio predictions [1]. Finally, there remains some controversy concerning the assignment of the vibrational spectrum. These various problems will be addressed in the present paper. The paper is organized as follows. Section 2 describes the techniques used for the measurement of the rotational spectrum. Section 3 describes the computational methods used in this study. Section 4 is dedicated to the determination of accurate rotational constants and centrifugal distortion constants for the ground and torsional states. The torsional frequency is also obtained using two different methods (relative intensity measurements and vibrational dependence of the inertial defect). In Section 5, the structure is determined by combining high level ab initio calculations and the available experimental rotational constants. Section 6 discusses the harmonic force field and the assignment of the vibrational spectrum, and Section 7 focuses on the NN bond length. A brief conclusion is provided in Section 8. 2. Experimental details 14 The rotational spectra of the main isotopologue N216O3 were measured by mixing NO and NO2 at about 161 Table 1 Summary of the fitted transitions for the ground vibrational state of N2O3 N#a 142 5 16 5 73 110 a b c d e J 00max 20 2 4 4 11 33 K 00max 10 2 4 1 9 30 (kHz)b Spectral range (GHz) r c GSCDs 20–24 13–40 6–16 40–80 78–255 10400 200 200 50 200 50 Refs. [12] [1] [2] [4] This workd This worke Number of transitions. Mean accuracy. Infrared ground state combination differences. Oslo spectrometer. Lille spectrometer. equal pressure in a cell cooled between 40 and 60 C. The total pressure was about 8 Pa. The microwave spectrum was studied between 40 and 80 GHz with the Oslo Stark spectrometer, which is described in Ref. [15]. The lines are rather weak and some of them are rather broad and skew as they are presumably perturbed by nuclear quadrupole coupling. The accuracy of the measurements is therefore estimated to be no better than 200 kHz. The millimeter wave measurements were performed in Lille with a source-modulated spectrometer using phasestabilized backward wave oscillators working in the frequency range 78–255 GHz [16]. The accuracy of the measurements is about 50 kHz for most lines. A summary of the measurements for the ground state is given in Table 1. One hundred and sixty-nine new lines were measured for the torsional state m9 = 1 with 2 6 J00 6 34 and 0 6 K 00a 6 28. A complete list of frequencies is given as supplementary material (Table S1 for the ground state and Table S2 for the excited torsional state). 3. Methods of computation Most correlated-level ab initio electronic structure computations of the present study have been carried out at two levels: second-order Møller–Plesset perturbation theory (MP2) [17] and coupled cluster (CC) theory with single and double excitation [18] augmented by a perturbational estimate of the effects of connected triple excitations [CCSD(T)] [19]. The Kohn–Sham density functional theory [20] was also used in this study. Both the B3LYP (Becke 3-parameter [21] Lee–Yang–Parr [22]) and B3PW91 (Becke 3-parameter [21] Perdew–Wang-1991 [23]) exchange-correlation functionals were considered. We used correlation-consistent polarized n-tuple zeta basis sets cc-pVnZ [24] with n 2 {D, T, Q}, that are abbreviated as VnZ in the text. To account for the electronegative character of the atoms, the augmented VnZ (aug-cc-pVnZ, AVnZ in short) basis sets [25] were also employed. The core–core and core–valence correlation effects on the computed equilibrium geometries [26], were estimated 162 J. Demaison et al. / Journal of Molecular Spectroscopy 244 (2007) 160–169 thanks to the correlation-consistent polarized weighted core–valence quadruple zeta (cc-pwCVQZ) [27,28] basis sets. For first-row atoms, it is sufficient to use the MP2 method to estimate this correction [29]. The frozen core approximation (hereafter denoted as fc), i.e., keeping the 1s orbitals of N, and O doubly occupied during correlated-level calculations, was used extensively. Some geometry optimizations were also carried out by correlating all electrons (hereafter denoted as ae). The CCSD(T) calculations were performed with MOLPRO [30–33] electronic structure program packages, while most other calculations utilized the GAUSSIAN03 (g03) program [34]. Most calculations were performed on the HP-XC4000 cluster of the ULB/VUB computing center. a negative inertial defect is expected for a planar species [37]. In the present case, however, the inertial defect is positive because of the dominant contribution of the low-frequency in-plane mode m7 = 160 cm1, as further discussed later and presented in Table 3. The planarity of the molecule is confirmed by the so-called planarity defect [35] Ds ¼ scccc s2 Cs1 ¼ 0:0644ð56Þ kHz AþB ð1Þ which presents the expected trends for a planar molecule [38]. 4.2. Torsional state m9 = 1 The assignment of the new data was not difficult because a reliable starting set of constants was available from the literature [13]. A Watson’s Hamiltonian using the A-reduction in the Ir representation was used [35]. The earlier microwave measurements as well as the available infrared GSCDs were included in the procedure. Table 1 summarizes the fitted data. The main difficulty was to find appropriate weights because the measurements are of very different origin and, furthermore, some lines are broadened by unresolved quadrupole hyperfine structure. To solve this problem, a robust regression, the iteratively reweighted least squares (IRLS) method [36], was used. The principle of this method is to estimate the weights from the residuals of a previous iteration. The resulting rotational parameters are listed and compared to the literature values [13] in Table 2. The improvement of the accuracy is significant. One can notice that the value of the inertial defect is small and positive, D0 = 0.135 uÅ2. Actually, when an out-of-plane vibration has a frequency which is much lower than any of the in-plane vibrations, as is the case for N2O3, The assignment of the new data was made starting from the approximate rotational constants of Brittain et al. [3] arising from the observation of eight rotational transitions, only. The resulting rotational constants are given in Table 2. As pointed out in Section 1, the torsional frequency could not be directly measured by infrared spectroscopy. It was indirectly determined to be m9 = 63 cm1 from the origins of the m5 + m9 combination and m5 fundamental bands [9]. The value predicted from microwave intensity measurements, m9 = 124(25) cm1 is in striking disagreement with the IR determination [3]. Furthermore, all ab initio values are around 150 cm1 [1]. We have performed relative intensity measurements, using the method described by Esbitt and Wilson [39]. We used a-type R-branch transitions with J quantum numbers ranging from 6 to 10. The highest Ka-members of these series of lines were selected because presenting smaller nuclear quadrupole splittings. For many other lines, the nuclear quadrupole coupling of the three nitrogen nuclei with the overall rotation indeed leads to distortion from the normal Lorentzian shape, therefore lowering the quality of the results. The selected lines also present the advantage of being modulated at comparatively low Stark voltages, which makes it easier to determine the base line accurately. Table 2 Rotational and centrifugal distortion constants of N2O3 Table 3 Experimental and calculated (B3LYP/AVTZ) inertial defects Dm = Ic Ib Ia of N2O3 (uÅ2) 4. Analysis of the rotational spectra 4.1. Ground state Ground state a A B C Dj Djk Dk dj dk Ujk Ukj a b c MHz MHz MHz kHz kHz kHz kHz kHz Hz Hz b Previous This work Previous This work 12454.02(18) 4226.57(2) 3152.99(2) 3.3(2) 7.0(10) 12454.153(25) 4226.6053(10) 3152.9591(10) 3.46221(43) 5.2221(26) 18.15(24) 0.91219(23) 10.0119(94) 0.0420(13) 0.3832(18) 12329.35 4210.80 3156.70 12329.567(33) 4210.9413(14) 3156.7170(13) 3.55303(51) 4.9890(34) 9.80(29) 0.93417(30) 7.524(11) 0.0739(15) 0.4966(26) c 0.86(14) c c c Ref. [13]. Ref. [3]. Not determined. c c c c c c c m(cm1)a m m9 = 1 1 2 3 4 5 6 7 8 9 0 1832 1652 1305 773 627 241 160 414 63 — a b c Calc. Exp. Harm.b Coriolisc Total 0.007 0.003 0.015 0.040 0.089 0.118 0.177 0.000 0.000 0.007 0.016 0.055 0.075 0.130 0.098 0.043 0.777 0.234 0.923 0.016 0.12 0.08 0.22 0.30 0.31 0.20 1.08 0.11 0.80 0.13 Ref. [8]. Harmonic contribution to the inertial defect. Coriolis contribution to the inertial defect. Refs. 0.12 12 0.15 13 0.17 1.18 3 3 0.91 0.14 This work This work J. Demaison et al. / Journal of Molecular Spectroscopy 244 (2007) 160–169 Another feature of the selected transitions is that they actually consist of two coalescing Ka-lines. The MW cell was packed with dry ice during the measurements and the temperature was therefore assumed to be 195 K. The intensities of lines with identical rotational quantum numbers were compared in the ground and m9 states. The a-component of the electric dipole moment and the line width were assumed to be identical for the two sets of transitions. Four separate measurements were made and yielded values for m9 ranging between 78 and 68 cm1. The average value is 70 cm1, with an accuracy estimated to be ±15 cm1. We also estimated the m9 frequency from the changes in the inertial defect upon vibrational excitation, using the formula derived by Hanyu et al. [40] Dðms ¼ 1Þ Dðm ¼ 0Þ ¼ h 1 2p2 c m9 constants, respectively. For most polyatomic molecules, the accuracy of the r0 structure calculated from the experimental I0’s is rather poor. A more reasonable approximation, closer to the equilibrium structure re, is provided by the so-called rs structure. One way to calculate this structure is to assume that the corrections eg’s are isotopologue-invariant. They can then be eliminated using the Kraitchman’s equations. This specific procedure is valid whenever isotopic differences of inertial moments dominate the accuracy [42]. Another way to achieve rs structures is to determine the quantities eg together with the structural parameters in a fit of the inertial moments [43]. However, the rs structure, though expected to be much better than the r0 one, is sometimes worse [44]. For this reason, more sophisticated assumptions have been devised, by Watson et al. among others, defining the so-called mass-dependent ð1Þ structures, rm [45]. A further refinement, leading to the rm structure assumes that e varies as I according to: qffiffiffiffiffi eg ¼ cg I m ð4Þ g ð2Þ The result is m9 = 64 cm1 in good agreement with the present relative intensity measurements and with the literature IR value. It should actually be remembered that the Hanyu formula underestimates the value of the related vibrational frequency [41], thus further strengthening the present result confirming the lower vibrational frequency value. where cg is a parameter to be determined. When there is no small Cartesian coordinate value and no hydrogen atom in the molecule, and provided the system of normal equations is well conditioned [44,45], it often gives results close to the equilibrium structure. The different experimental structures, r0, rs and rð1Þ m , are given in Table 5. Unfortunately, the condition number, j = 682 (for rs), is rather large indicating that the system is not well conditioned. However, at each step of improveð1Þ ment of the model ðr0 ! rs ! rm Þ, the standard deviation 5. Structure 5.1. Empirical structures The experimental rotational constants used in this work are given in Table 4. In order to avoid a possible bias due to measurements of different origins, all rotational constants are taken from Ref. [3]. Furthermore, it was checked that, in the present case, the use of the rotational constants of Table 2 does not modify the results. We will first briefly review the methods used. The experimental data permit to determine the ground state moment of inertia I0, which are different from the equilibrium ones, Ie, with I 0g ¼ I eg ðmi ; d 2ij Þ þ eg ðF kl ; F klm Þ 163 Table 5 Empirical structures of N2O3 from ground state rotational constants (distances in Å and angles in degrees) NN NO NOcis NOtrans NNOcis NNOtrans ONN ð3Þ Here g = a,b,c refers to one of the principal inertial axes, mi is the mass of atom i, dij the interatomic distance between atoms i and j, and Fkl and Fklm quadratic and cubic force ð1Þ r0 rs rm 1.889(9) 1.139(4) 1.188(4) 1.217(8) 111.3(6) 116.3(4) 105.2(2) 1.871(4) 1.144(5) 1.192(4) 1.218(4) 112.4(3) 116.6(4) 105.0(3) 1.868(4) 1.137(5) 1.191(4) 1.213(4) 112.2(2) 116.8(4) 105.1(3) Table 4 Ground statea and equilibrium rotational constants (MHz) for N2O3 ONNO2 O15NNO2 ON15NO2 ONN18OcO ONNO18Ot 18 ONNO2 O15N15NO2 a b c A0 B0 C0 D0b Ae Be Ce Dec 12453.59 12294.22 12454.74 11630.73 12150.24 12443.91 12294.86 4226.65 4186.48 4213.18 4202.02 4059.46 4015.72 4172.67 3152.91 3120.34 3145.47 3083.96 3040.34 3033.40 3112.71 0.139 0.139 0.140 0.151 0.136 0.142 0.138 12531.19 12367.91 12531.41 11698.84 12231.90 12521.20 12367.74 4274.15 4233.71 4259.99 4249.65 4102.96 4060.42 4219.20 3186.91 3153.82 3179.02 3117.03 3072.33 3065.88 3145.76 0.009 0.011 0.010 0.013 0.003 0.013 0.011 Ref. [3]. Ground state inertial defect (uÅ2). Equilibrium inertial defect (uÅ2). 164 J. Demaison et al. / Journal of Molecular Spectroscopy 244 (2007) 160–169 of the fit decreases, indicating that its quality improves. It is ð1Þ worth noting that the rs and rm structures are compatible. These structures will be compared and discussed at the end of Section 5.3. 5.2. Ab initio structure The ab initio structure was first calculated with the AVTZ basis set using different level of theory: HF, MP2, CCSD and CCSD(T). The B3LYP and B3PW91 methods were also used. The results are given in Table 6. The large dispersion of the results is not surprising because the value of the T1 diagnostic at 0.022 is larger than the cut-off value of 0.020. It thus indicates that the single-reference CCSD(T) method is no more suitable for properly describing electron correlation effects [46]. At the HF level, the bond lengths are too small, while they are too large at the MP2 level. They are intermediate at the CCSD level. The CCSD(T) values are close to, but smaller than the MP2 values. The CCSD(T) structure is probably the closest one to the equilibrium structure. However, the AVTZ basis set is not yet converged as can be seen in Table 6 where the VTZ, VQZ and AVQZ results are also listed. Adding the core correlation correction (0.002 Å for the NO bond lengths and 0.004 Å for the NN bond) and the effect of basis set enlargement AVTZ fi AVQZ (calculated at the MP2 level) gives the structure listed in Table 7. It should be close to the equilibrium one although it is likely that the NAN bond length is too short. It is also interesting to note that the ab initio NOc and NOt bond lengths are almost identical whereas the experimental values are significantly different as can be seen in Table 5. This might be explained by the inability of the standard AVTZ basis set to describe long range interactions. To check this point, we have used the doubly and triply augmented VTZ basis sets [47]. However, their use do not result into significant asymmetry for the two NO bond lengths. The B3LYP and B3PW91 structures are almost identical, except for the NN bond length. They will be discussed at the end of Section 5.3. The ab initio NAN bond length might be too short because of the basis set superposition error (BSSE), which is a consequence of the incompleteness of the basis set. It causes the molecular interactions to be artifactually too attractive. Thus, the calculated intermolecular distance NAN will be too small when the complex geometry is optimized. The conventional way to correct for BSSE is to use the counterpoise (CP) method [48,49] as implemented in g03. A calculation at the MP2/VQZ level of theory shows that the NAN bond length is increased by 0.010 Å when the counterpoise correction is taken into account. Although this increase is significant, the NAN bond still remains much too short. It is, however, possible to explain the long NAN bond length by performing a multiconfigurational calculation at the complete active space multiconfiguration SCF (CASSCF) level [50,51] using the AVTZ basis set. It appears that a configuration mixing occurs involving a double excitation from the NAN in-plane molecular Table 7 Equilibrium structures of N2O3 (distances in Å and angles in degrees) ð1Þ NN NO NOc NOt NNOc NNOt ONN ONOe rm B3LYPa Ab initiob c rSE e 1.868(2) 1.137(2) 1.191(2) 1.213(2) 112.2(2) 116.8(3) 105.1(2) 131.0 1.869 1.131 1.199 1.199 112.1 116.6 106.3 131.3 1.825 1.140 1.201 1.201 112.1 116.2 105.8 131.7 1.870(2) 1.139(2) 1.199(2) 1.206(2) 111.5(2) 117.4(2) 104.9(1) 131.1 ON + NO2d 1.1408 1.1946 1.1946 133.51 a AVTZ basis set. See Table 6: CCSD(T)/AVTZ + MP2/AVQZ MP2/AVTZ + MP2(ae)/wCVQZMP2/wCVQZ. c Semi-experimental equilibrium structure, see text. d Structure of the constituting species, Ref. [70] for NO and Ref. [71] for NO2. e Value derived from ONO = 360 NNOc NNOt. b Table 6 Ab initio structures of N2O3 (distances in Å and angles in degrees) Basis Method NAN NAO NAOc NAOt NNOc NNOt ONO ONN AVTZ HF MP2 B3LYP B3PW91 CCSD CCSD(T) Range 1.6018 1.8865 1.8693 1.8381 1.7331 1.8389 0.2847 1.1165 1.1529 1.1308 1.1307 1.1406 1.1458 0.0364 1.1671 1.2084 1.1992 1.1946 1.1951 1.2060 0.0413 1.1692 1.2084 1.1989 1.1944 1.1966 1.2063 0.0392 116.69 108.54 112.11 112.31 114.07 111.88 8.15 112.45 119.42 116.62 116.20 114.33 116.44 6.97 130.87 132.04 131.27 131.49 131.60 131.68 1.18 109.46 103.40 106.25 106.37 107.18 105.56 6.06 VTZ VQZ AVQZ wCVQZ wCVQZ reb MP2 MP2 MP2 MP2 MP2 (ae) 1.8785 1.8729 1.8770 1.8729 1.8687 1.8251 1.1542 1.1497 1.1493 1.1494 1.1472 1.1400 1.2075 1.2047 1.2054 1.2045 1.2023 1.2007 1.2069 1.2046 1.2054 1.2044 1.2023 1.2012 108.59 108.74 108.72 108.73 108.77 112.10 119.13 119.18 119.30 119.19 119.10 116.23 132.29 132.08 131.98 132.08 132.13 131.67 103.39 103.55 103.59 103.55 103.61 105.80 a b T1 = 0.0223. CCSD(T)/AVTZ + MP2/AVQZ MP2/AVTZ + MP2(ae)/wCVQZMP2/wCVQZ. J. Demaison et al. / Journal of Molecular Spectroscopy 244 (2007) 160–169 orbital to the corresponding antibonding orbital with respective mixing coefficients of 0.976 and 0.215. Going from SCF to CASSCF corresponds to a lengthening of the NAN bond from 1.603 Å to 1.767 Å. Correlating the whole set of valence electrons using the multiconfiguration-reference configuration interaction (MRCI) method [52,53] enhances the antibonding character and gives a bond length of 1.801 Å. The MRCI calculation confirms that the NAN bond length calculated at the CCSD(T) level is too short. 5.3. Semi-experimental equilibrium structure A more reliable way to estimate the equilibrium structure is to use a set of experimental ground state rotational constants corrected by ab initio computed vibration-rotation interaction constants (a) [54]. The resulting rotational constants can be used to determine the corresponding semiexperimental equilibrium structures, rSE e . Theoretically, one can even initially correct the experimental rotational constants for the magnetic effect [55]. We have estimated the size of this correction by calculating the unknown values of the related g constants using the Gaussian03 program package, at the B3LYP/AVTZ level of theory. The results are: gaa = 0.298; gbb = 0.088; and gcc = 0.026. The corrected values of the rotational constants are given by the relation [56] Bacorr ¼ Baexp 1 þ Mmp gaa ð5Þ in which gaa is expressed in units of the nuclear magneton, m is the electron mass, Mp the proton mass and a = a,b,c. This correction turns out to be negligible, the largest deviation being 2 MHz for A. To calculate the semi-equilibrium structure in N2O3, one needs to calculate the vibration–rotation interaction constants. This requires computing the force field up to cubic terms. As the proper description of electron correlation is important for N2O3, we tried different levels of theory (MP2, B3LYP, B3PW91 and CCSD(T) with the basis sets VTZ and AVTZ). The anharmonic normal coordinate force fields were determined for all isotopologues whose ground state rotational constants are known. The molecular geometry was first calculated. Then, the associated harmonic force field was evaluated analytically in Cartesian coordinates. The cubic (/ijk) and semidiagonal quartic (/ijkk) normal coordinates force constants were determined with the use of a finite difference procedure involving displacements [57] along reduced normal coordinates (step size Dq = 0.03) and the calculation of analytic second derivatives at these displaced geometries. The evaluation of anharmonic spectroscopic constants was based on second-order rovibrational perturbation theory [58]. The resulting force field is expected to rely on the quality of the structure and of the harmonic force field. We 165 Table 8 Experimental and calculated quartic centrifugal distortion constants of N2O3 (kHz) Dj Djk Dk dj dk a Exp. B3LYP/AVTZ B3LYP/AVTZa 3.4621(4) 5.222(3) 18.2(2) 0.9122(2) 10.012(9) 2.450 8.038 11.756 0.640 8.785 3.304 5.212 15.452 0.858 8.881 After scaling, see text. checked that the closest structure to re is from B3LYP/ AVTZ. It is indeed relatively close to the experimental ð1Þ rm structure, see Table 5, as will be later confirmed. Furthermore, as already discussed by Stirling et al. [1], B3LYP/AVTZ reproduces well the experimental fundamental vibrational frequencies. In addition, the quartic centrifugal distortion constants and the inertial defects calculated with this harmonic force field are also in good agreement with the experimental values, see Tables 3 and 8. Also, the B3LYP/AVTZ force field is the one providing the smallest semi-experimental equilibrium inertial defect (which should ideally be zero), see Table 3. The same conclusion was achieved for HONO which also presents both a large T1 diagnostic and a very long NAO bond [59]. The semi-experimental structure is given in Table 7, in which it is compared to other structures. It is rather close ð1Þ to the empirical rm and B3LYP/AVTZ structures. With the exception of the NAN bond length, it is also in good agreement with the CCSD(T) structure. The same comparison occurred in HONO in which the CCSD(T) method was found accurate except for the long NAO bond length [59]. The main discrepancy between the present semi-experimental and B3LYP/AVTZ structures lies with the NOc and the NOt bond lengths. Experimentally, it was found that r(NOt) > r(NOc) while no significant difference results from the ab initio calculations. However, whereas the asymmetry is quite large for the rs structure, Drs = rs(NOt) rs (NOc) = 0.026(2) Å, it is much reduced for the semi-experimental structure, Dre = 0.0069(14) Å. Nevertheless, as further discussed in the next section, this difference seems to be genuine if one assumes that the standard deviation of the fit is a reliable indicator of the uncertainty. 6. Harmonic force field The experimental harmonic force field of N2O3 has already been determined by Nour et al. [9]. However, as noted in the introduction, the assignment of the NO2 out-of-plane wag is in disagreement with ab initio results. Another problem arises from the assumption by Nour et al. that the stretching force constants for the NOc and NOt bonds are identical. Although they are expected to be close, they should differ in the same respect as the two bond lengths, a longer bond being associated with a smaller force constant. 166 J. Demaison et al. / Journal of Molecular Spectroscopy 244 (2007) 160–169 Selecting the value 627 cm1 for the NO2 in-plane rock (m5) and 414 cm1 for the NO2 out-of-plane wag (m8) gives scaling factors k5 = 0.84(2) and k8 = 0.85(4). These show the expected order of magnitude [62], comforting the revised assignment based on the ab initio calculations [10]. The scaled force field reproduces well the experimental harmonic frequencies as well as the quartic centrifugal distortion constants, see Tables 8 and 11. However, it is not accurate enough to point out the difference between the two stretching force constants for the NOc and NOt bonds. It has to be noted that the range of the scaling factors is much larger than usual [62]. It demonstrates the difficulty to calculate an accurate theoretical force field for N2O3. To improve the force field, we determined the diagonal force constants from the experimental data, fixing most of the non-diagonal constants at the scaled ab initio values. The result, given at the bottom of Table 10, does not reveal any significant difference between the cis and trans NO bond lengths. The harmonic force field was first calculated in Cartesian coordinates at the B3LYP/AVTZ level of theory (as in Section 5.3). It was then transformed into a representation based on a set of nonredundant internal coordinates which are defined in Table 9. Then, this theoretical force field was scaled in the usual way [60] using nine different scaling factors. These were refined by fitting them with the program ASYM40 [61] to the observed data described ðthÞ below. Practically, the theoretical force constants F ij are scaled by the factors ki to give the scaled force constants ðscÞ F ij pffiffiffiffiffiffiffiffi ðthÞ ðscÞ F ij ¼ k i k j F ij ð6Þ We used the same experimental data as Nour et al. [9], except that the fundamental frequencies mi were transformed into harmonic frequencies xi using the B3LYP/ AVTZ anharmonic force field of Section 5.3. Furthermore, the five quartic centrifugal distortion constants of Table 2 were also used. The results are given in Table 10. 7. Discussion of the NN bond length Table 9 Symmetry coordinates of N2O3 An interesting result of this work is the determination of NAN bond length. It appears to be quite long re(NAN) = 1.870 Å, as expected. It is much longer than in N2, re(N„N) = 1.098 Å [63] and in N2H2, re(N@N) = 1.247 Å [64]. It is probably more informative to compare it to other single bonds. From this point of view, the prototype of the NAN single bond is found in hydrazine, N2H4. Although it has been extensively studied by microwave spectroscopy [65] and by electron diffraction [66], there is no equilibrium structure available, the highest level of ab initio calculation being MP2/6-311+G(3df, 2p) Symmetry coordinate a0 S1 = Dr(ON1) S2 = Dr(N1N2) S3 = Dr(N2Ot) S4 = Dr(N2Oc) S5 = D\(ON1N2) S6 = 61/2[2D\(N1N2Oc) D\(OcN2Ot) D\((N1N2Ot)] S7 = 21/2[D\(OcN2Ot) D\((N1N2Ot)] a00 S8 = Dc(N1N2OcOt)a S9 = Ds(ON1N2Oc) a Out-of-plane bending. Table 10 Harmonic force constants and scaling factors for N2O3 i kia Theoretical (unscaled) harmonic force fieldb 1 2 3 4 5 6 7 8 9 0.9153(13) 0.7157(67) 0.8639(38) 1.0911(63) 0.837(23) 1.195(17) 0.8662(66) 0.8465(44) 0.2624(86) 16.593 0.721 0.634 0.752 0.328 0.077 0.420 0.357 0.005 Experimental harmonic force constants for N2O3 1 15.263(12) 2 0.583 3 0.564 4 0.752 5 0.287 6 0.080 7 0.374 8 0.30187(82) 9 0.00246 a b 1.029 0.036 0.012 0.180 0.164 0.090 11.075 1.698 0.170 0.348 0.333 10.934 0.232 0.096 0.494 1.337 0.320 0.182 0.796 0.300 0.993 10.73(19) 1.341(11) 0.145 0.354 0.288 10.76(19) 0.222 0.110 0.480 1.460(26) 0.089(14) 0.155 0.647(14) 0.305 0.9134(29) 0.067 0.7345(37) 0.0281 0.0110 0.140 0.152 0.071 0.01762(31) Scaling factor (uncertainty in parentheses). The units correspond to energies in aJ, bond lengths in Å and angles in radians. J. Demaison et al. / Journal of Molecular Spectroscopy 244 (2007) 160–169 Table 11 Harmonic frequencies xi (cm1) of N2O3 i mi(exp)a xi(exp)b xi(calc)c Exp. calc. 1 2 3 4 5 6 7 8 9 1829.5 1610.0 1304.3 773.0 615.5 241.0 205.0 414.0 63.0 1853.2 1690.1 1329.3 784.7 621.6 248.6 212.7 424.4 65.8 1853.4 1689.9 1329.0 784.2 621.7 247.4 216.0 424.4 65.8 0.2 0.2 0.3 0.5 0.1 1.2 3.3 0.0 0.0 a Estimated gas phase values, Refs. [8,9]. b Estimated harmonic frequency: xi ¼ mi ðexpÞ þ ½xi mi ðB3LYP= AVTZÞ. c Calculated with the experimental force constants of Table 10. Table 12 Ab initio CCSD(T) structure of hydrazine, N2H4 (distances in Å and angles in degrees) VTZ NAN NAHoc NAHi NNHo NNHi HoNNHo HoNNHi 1.4448 1.0120 1.0154 106.208 110.789 154.723 89.919 VQZ 1.4382 1.0105 1.0137 106.711 111.189 153.915 89.639 wCVQZ aea fca 1.4340 1.0092 1.0123 106.943 111.374 152.561 90.464 1.4377 1.0106 1.0138 106.716 111.198 153.794 89.726 V5Z reb 1.4361 1.0103 1.0134 106.953 111.400 152.934 90.021 1.4324 1.0089 1.0119 107.180 111.576 151.701 90.759 a ae = all electrons correlated, fc = frozen core approximation. V5Z + wCVQZ(ae) wCVQZ(fc). Derived parameter: \(HNH) = 107.588 degrees. c The hydrogen atoms which take the inner and outer positions are denoted as i and o, respectively. b Table 13 YZ bond lengths (Å) and \(XYZ) bond angles (degrees) in XYZ2 molecules X Y Z YZc YZt XYZc XYZt Refs. ON HO HO HO HN@ HN@ H2NCH@ HOCH@ H2N N N B B C C C C N O O H F H F H H H 1.206 1.192 1.188 1.313 1.086 1.305 1.077 1.077 1.009 111.5 115.6 120.4 122.3 124.2 128.2 121.8 121.9 111.6 117.4 114.1 116.8 119.4 118.8 123.4 120.1 119.6 107.2 This work [72] [73] [74] [75] [76] [77] [72] This work 1.199 1.208 1.194 1.323 1.090 1.320 1.081 1.082 1.012 [67]. For this reason, we have calculated the structure of N2H4 in the present work using the CCSD(T) method and basis sets up to quintuple zeta. The results are reported in Table 12. The T1 diagnostic being only 0.0073 at the CCSD(T)/wCVQZ(ae) level of theory, the derived structure is expected to be reliable. In N2H4, re(NAN) = 1.432 Å is therefore much shorter than in N2O3. Another interesting molecule is N2O2, the C2v dimer of NO for ð1Þ which rm ¼ 2:263ð1Þ Å [68]. The harmonic force field of this dimer has also been determined [69]. It gives for the 167 Table 14 Structure of a few XNO and XNO2 molecules ((distances in Å and angles in degrees) XNO NO HNO FNO ClNO BrNO HONO NCNO ONNO a b XNO2 r(NO) Refs. 1.151 1.209 1.132 1.136 1.133 1.169 1.205 1.152 70 59 59 59 79 [59] NO2 FNO2 ClNO2 BrNO2 O2NNO2 a 68 \(ONO) r(NOc) 133.5 135.7 131.9 131.0 134.8 1.195 1.177 1.191 1.196 1.191 HONO2 130.5 ONNO2 131.1 1.208 1.199 r(NOt) Refs. [71] [59] b [78] [80] 1.192 1.206 [72] b ð1Þ rm structure using the rotational constants of Ref. [81]. This work. NAN stretching force constant, fNN = 0.193(14) aJÅ2. This value is smaller than for N2O3, in agreement with the fact that the NN bond length is longer in N2O2. Another important point is the asymmetry of the NO2 group. The \(NNOt) angle at 117.4(2) degrees is much larger than the \(NNOc) angle whose value is only 111.5(2) degrees. Generally, in XYZ2 molecules, when the two \(XYZ) bond angles are significantly different, the longest bond length YZ corresponds to the largest \(XYZ) bond angle, see Table 13. This result confirms the likely asymmetry of the two r(NO) bond lengths, the r(NOt) bond being longer as expected. However, it seems difficult to determine an accurate value of the difference between the two lengths, the most reliable value being probably the semi-experimental one, Dre = 0.0069(14) Å. The NO bond length is given for a few molecules in Table 14. It varies from 1.132 Å in FNO to 1.208 Å in HNO3 (NOc). 8. Conclusions The rotational spectra of the ground vibrational state and of the m9 = 1 torsional state of N2O3 have been reinvestigated in the present work and accurate spectroscopic constants have been determined. The torsional frequency, m9 = 70(15) cm1, has been determined by relative intensity measurements, confirming the lowest of the two values proposed in the literature, as suggested from previous IR results. The assignment of the infrared spectrum has been slightly revised and an accurate harmonic force field has been determined. The equilibrium structure has been determined using different, complementary methods and the specificity of N2O3 as a complex has been confirmed by the determination of the NN bond length, re(NN) = 1.870(2) Å. Acknowledgments This work was sponsored, in Brussels, by the Fonds National de la Recherche Scientifique (FNRS, contracts FRFC and IISN) and the «Action de Recherches Concertées 168 J. Demaison et al. / Journal of Molecular Spectroscopy 244 (2007) 160–169 de la Communauté française de Belgique». It was also performed within the ‘‘LEA HiRes ’’ and the EU project QUASAAR (MRTN-CT-2004-512202). The authors are indebted to ULB for providing an invited international chair to Dr. Demaison (2006). [29] Appendix A. Supplementary data [30] Supplementary data for this article are available on ScienceDirect (www.sciencedirect.com) and as part of the Ohio State University Molecular Spectroscopy Archives (http://msa.lib.ohio-state.edu/jmsa_hp.htm). References [1] A. Stirling, I. Papai, J. Mink, D.R. Salahub, J. Chem. Phys. 100 (1994) 2910–2923. [2] R.L. Kuczkowski, J. Am. Chem. Soc. 87 (1965) 5259–5260. [3] A.H. Brittain, A.P. Cox, R.L. Kuczkowski, Trans. Faraday Soc. 65 (1969) 1963–1974. [4] A.P. Cox, D.J. Finnigan, J.C.S. Faraday Trans. II 69 (1973) 49–54. [5] S.G. Kukolich, J. Am. Chem. Soc. 104 (1982) 6927–6929. [6] A.P. Cox, J. Randell, A.C. Legon, Chem. Phys. Lett. 126 (1986) 481–486. [7] L. D’Or, P. Tarte, Bull. Soc. R. Sci. Liege 22 (1953) 276–284. [8] F. Mélen, M. Herman, J. Phys. Chem. Ref. Data 21 (1992) 831–881. [9] E.M. Nour, L.-H. Chen, J. Laane, J. Phys. Chem. 87 (1983) 1113–1120. [10] C.H. Bibart, G.E. Ewing, J. Chem. Phys. 61 (1974) 1293–1299. [11] L.A. Chewter, I.W.M. Smith, G. Yarwood, Mol. Phys. 63 (1988) 843–864. [12] R. Georges, J. Liévin, M. Herman, A. Perrin, Chem. Phys. Lett. 256 (1996) 675–678. [13] A.M. Andrews, J.L. Domenech, G.T. Fraser, W.J. Lafferty, J. Mol. Spectrosc. 163 (1994) 428–435. [14] R.D. Harcourt, P.P. Wolynec, J. Phys. Chem. A 104 (2000) 2138–2143. [15] (a) G.A. Guirgis, K.M. Marstokk, H. Møllendal, Acta Chem. Scand. 45 (1991) 482–490; (b) H. Møllendal, A. Leonov, A. de Meijere, J. Phys. Chem. A 109 (2005) 6344; (c) H. Møllendal, G.C. Cole, J.-C. Guillemin, J. Phys. Chem. A 110 (2005) 921. [16] F. Willaert, H. Møllendal, E. Alekseev, M. Carvajal, I. Kleiner, J. Demaison, J. Mol. Struct. 795 (2006) 4–8. [17] C. Møller, M.S. Plesset, Phys. Rev. 46 (1934) 618–622. [18] G.D. Purvis III, R.J. Bartlett, J. Chem. Phys. 76 (1982) 1910–1918. [19] K. Raghavachari, G.W. Trucks, J.A. Pople, M. Head-Gordon, Chem. Phys. Lett. 157 (1989) 479–483. [20] W. Kohn, L.J. Sham, Phys. Rev. A 140 (1965) 1133–1138. [21] A.D. Becke, J. Chem. Phys. 98 (1993) 5648–5652. [22] C.T. Lee, W.T. Yang, R.G. Parr, Phys. Rev. B 37 (1988) 785–789. [23] J.P. Perdew, J.A. Chevary, S.H. Vosko, K.A. Jackson, M.R. Pederson, D.J. Singh, C. Fiolhais, Phys. Rev. B 46 (1992) 6671–6687 (and references therein). [24] T.H. Dunning Jr., J. Chem. Phys. 90 (1989) 1007–1023. [25] R.A. Kendall, T.H. Dunning Jr., R.J. Harrison, J. Chem. Phys. 96 (1992) 6796–6806. [26] A.G. Császár, W.D. Allen, J. Chem. Phys. 104 (1996) 2746–2748. [27] K.A. Peterson, T.H. Dunning Jr., J. Chem. Phys. 117 (2002) 10548–10560. [28] Basis sets were obtained from the Extensible Computational Chemistry Environment Basis Set Database, Version 02/02/06, as developed and distributed by the Molecular Science Computing Facility, Environmental and Molecular Sciences Laboratory which is part of [31] [32] [33] [34] [35] [36] [37] [38] [39] [40] [41] [42] [43] [44] [45] [46] [47] [48] [49] [50] [51] [52] [53] the Pacific Northwest Laboratory, P.O. Box 999, Richland, Washington 99352, USA, and funded by the US Department of Energy. The Pacific Northwest Laboratory is a multi-program laboratory operated by Battelle Memorial Institute for the US Department of Energy under Contract DE-AC06-76RLO 1830. Contact David Feller or Karen Schuchardt for further information. L. Margulès, J. Demaison, H.D. Rudolph, J. Mol. Struct. 599 (2001) 23–30. MOLPRO 2000 is a package of ab initio programs written by H.-J. Werner, P.J. Knowles, with contributions from R.D. Amos, A. Bernhardsson, A. Berning, P. Celani, D.L. Cooper, M.J.O. Deegan, A.J. Dobbyn, F. Eckert, C. Hampel, G. Hetzer, T. Korona, R. Lindh, A.W. Lloyd, S.J. McNicholas, F.R. Manby, W. Meyer, M.E. Mura, A. Nicklass, P. Palmieri, R. Pitzer, G. Rauhut, M. Schütz, H. Stoll, A.J. Stone, R. Tarroni, T. Thorsteinsson. MOLPRO 2000 is a package of ab initio programs written by H.-J. Werner, P.J. Knowles, with contributions from R.D. Amos, A. Bernhardsson, A. Berning, P. Celani, D.L. Cooper, M.J.O. Deegan, A.J. Dobbyn, F. Eckert, C. Hampel, G. Hetzer, T. Korona, R. Lindh, A.W. Lloyd, S.J. McNicholas, F.R. Meyer, W. Manby, M.E. Mura, A. Nicklass, P. Palmieri, R. Pitzer, G. Rauhut, M. Schütz, H. Stoll, A.J. Stone, R. Tarroni, T. Thorsteinsson. C. Hampel, K.A. Peterson, H.-J. Werner, Chem. Phys. Lett. 190 (1992) 1–12, and references therein. M.J.O. Deegan, P.J. Knowles, Chem. Phys. Lett. 227 (1994) 321–326. M.J. Frisch, G.W. Trucks, H.B. Schlegel, G.E. Scuseria, M.A. Robb, J.R. Cheeseman, J.A. Montgomery, Jr., T. Vreven, K.N. Kudin, J.C. Burant, J.M. Millam, S.S. Iyengar, J. Tomasi, V. Barone, B. Mennucci, M. Cossi, G. Scalmani, N. Rega, G.A. Petersson, H. Nakatsuji, M. Hada, M. Ehara, K. Toyota, R. Fukuda, J. Hasegawa, M. Ishida, T. Nakajima, Y. Honda, O. Kitao, H. Nakai, M. Klene, X. Li, J.E. Knox, H.P. Hratchian, J.B. Cross, C. Adamo, J. Jaramillo, R. Gomperts, R.E. Stratmann, O. Yazyev, A.J. Austin, R. Cammi, C. Pomelli, J.W. Ochterski, P.Y. Ayala, K. Morokuma, G.A. Voth, P. Salvador, J.J. Dannenberg, V.G. Zakrzewski, S. Dapprich, A.D. Daniels, M.C. Strain, O. Farkas, D.K. Malick, A.D. Rabuck, K. Raghavachari, J.B. Foresman, J.V. Ortiz, Q. Cui, A.G. Baboul, S. Clifford, J. Cioslowski, B.B. Stefanov, G. Liu, A. Liashenko, P. Piskorz, I. Komaromi, R.L. Martin, D.J. Fox, T. Keith, M.A. AlLaham, C.Y. Peng, A. Nanayakkara, M. Challacombe, P.M.W. Gill, B. Johnson, W. Chen, M.W. Wong, C. Gonzalez, J.A. Pople, Gaussian, Inc., Revision D.01, Pittsburgh, PA, 2003. J.K.G. Watson, in: J.R. Durig (Ed.), Vibrational Spectra and Molecular Structure, vol. 6, Elsevier, Amsterdam, 1977, p. 1. L.C. Hamilton, Regression with Graphics, Duxbury, Belmont, CA, 1992 (Chapter 6). T. Oka, J. Mol. Struct. 352–353 (1995) 225–233. J. Demaison, M. Le Guennec, G. Wlodarczak, B.P. Van Eijck, J. Mol. Spectrosc. 159 (1993) 357–362. A.S. Esbitt, E.B. Wilson, Rev. Sci. Instrum. 34 (1963) 901–907. Y. Hanyu, C.O. Britt, J.E. Boggs, J. Chem. Phys. 45 (1966) 4725–4728. T. Oka, J. Mol. Struct. 352/353 (1995) 225–233. C.C. Costain, J. Chem. Phys. 29 (1958) 864–874. H.D. Rudolph, in: M. Hargittai, I. Hargittai (Eds.), Advances in Molecular Structure Research, vol. 1, JAI Press, Greenwich, CT, 1995, p. 63. J. Demaison, H.D. Rudolph, J. Mol. Spectrosc. 215 (2002) 78–84. J.K.G. Watson, A. Roytburg, W. Ulrich, J. Mol. Spectrosc. 196 (1999) 102–109. T.J. Lee, P.R. Taylor, Int. J. Quant. Chem. Symp. 23 (1989) 199–207. D.E. Wong, T.H. Dunning Jr., J. Chem. Phys. 100 (1994) 2975–2988. S.F. Boys, F. Bernardi, Mol. Phys. 19 (1970) 553–566. S. Simon, M. Duran, J. Chem. Phys. 105 (1996) 11024–11031. H.-J. Werner, P.J. Knowles, J. Chem. Phys. 82 (1985) 5053–5063. P.J. Knowles, H.-J. Werner, Chem. Phys. Lett. 115 (1985) 259–267. H.-J. Werner, P.J. Knowles, J. Chem. Phys. 89 (1988) 5803–5814. P.J. Knowles, H.-J. Werner, Chem. Phys. Lett. 145 (1988) 514–522. J. Demaison et al. / Journal of Molecular Spectroscopy 244 (2007) 160–169 [54] I.M. Mills, in: C.W. Mathews (Ed.), Molecular Spectroscopy: Modern Research, vol. 1, Academic Press, New York, 1972, pp. 115–140. [55] W. Gordy, R.L. Cook, Microwave Molecular Spectra, Wiley, New York, 1984 (Chapter VIII). [56] W. Gordy, R.L. Cook, Microwave Molecular Spectra, Wiley, New York, 1984 (Chapter XI). [57] W. Schneider, W. Thiel, Chem. Phys. Lett. 157 (1989) 367–373. [58] I.M. Mills, in: C.W. Mathews (Ed.), Molecular Spectroscopy: Modern Research, vol. 1, Academic Press, New York, 1972, p. 115. [59] J. Demaison, A.G. Császár, A. Dehayem-Kamadjeu, J. Phys. Chem. A 110 (2006) 13609–13617. [60] P. Pulay, G. Fogarasi, G. Pongor, J.E. Boggs, A. Vargha, J. Am. Chem. Soc. 105 (1983) 7037–7047. [61] L. Hedberg, I.M. Mills, J. Mol. Spectrosc. 203 (2000) 82–95. [62] M.W. Wong, Chem. Phys. Lett. 256 (1996) 391–399 . [63] T.A. Roden, T. Helgaker, P. Jørgensen, J. Olsen, J. Chem. Phys. 121 (2004) 5874–5884. [64] J.M.L. Martin, P.R. Taylor, Mol. Phys. 96 (1999) 681–692. [65] I. Gulaczyk, J. Pyka, M. Kreglewski, J. Mol. Spectrosc. 241 (2007) 75–89. [66] K. Kohata, T. Fukuyama, K. Kuchitsu, J. Phys. Chem. 86 (1992) 602–606. [67] A. Chung-Phillips, K.A. Jebber, J. Chem. Phys. 102 (1995) 7080– 7087. 169 [68] A.R.W. McKellar, J.K.G. Watson, B.J. Howard, Mol. Phys. 86 (1995) 273–286. [69] K.-X. Au Yong, J.M. King, A.R.W. McKellar, J.K.G. Watson, J. Mol. Spectrosc. 238 (2006) 127–134. [70] A.H. Saleck, G. Winnewisser, K.M.T. Yamada, Mol. Phys. 76 (1992) 1443–1455. [71] Y. Morino, M. Tanimoto, Can. J. Phys. 62 (1984) 1315–1322. [72] C. Gutle, to be published. [73] J. Demaison, J. Liévin, M. Herman, Int. Rev. Phys. Chem. in press. [74] J. Breidung, J. Demaison, J.-F. D’Eu, L. Margulès, D. Collet, E.B. Mkadmi, A. Perrin, W. Thiel, J. Mol. Spectrosc. 228 (2004) 7–22. [75] L. Margulès, J. Demaison, P.B. Sreeja, J.-C. Guillemin, J. Mol. Spectrosc. 238 (2006) 234–240. [76] C. Puzzarini, A. Gambi, J. Phys. Chem. A 108 (2004) 4138–4145. [77] E. Askeland, H. Møllendal, E. Uggerud, J.C. Guillemin, J.R. Aviles Moreno, J. Demaison, T.R. Huet, J. Phys. Chem. A 110 (2006) 12572–12584. [78] F. Kwabia Tchana, J. Orphal, I. Kleiner, H.D. Rudolph, H. Willner, P. Garcia, O. Bouba, J. Demaison, B. Redlich, . Mol. Phys. 102 (2004) 1509–1521. [79] C. Degli Esposti, F. Tamassia, G. Cazzoli, Z. Kisiel, J. Mol. Spectrosc. 170 (1995) 582–600. [80] Q. Shen, K. Hedberg, J. Phys. Chem. A 102 (1998) 6470–6476. [81] R. Dickinson, G.W. Kirby, J.G. Sweeny, J.K. Tyler, J. Chem. Soc. Faraday Trans. II 74 (1978) 1393–1402.