p-Dioxane is non-polar, hence its rotational constants cannot be determined by microwave rotational coherence spectroscopy (RCS). We perform high-resolution gas-phase rotational spectroscopy of para-dioxane-h8 and -d8 using femtosecond time-resolved Raman RCS in a gas cell at T = 293 K and in a pulsed supersonic jet at T130 K. The inertial tensor of p-dioxane-h8 is strongly asymmetric, leading to a large number of asymmetry transients in its RCS spectrum. In contrast, the d8-isotopomer is a near-oblate symmetric top that exhibits a much more regular RCS spectrum with few asymmetry transients. Fitting the fs Raman RCS transients of p-dioxane-h8 to an asymmetric-top model yields the ground-state rotational constants A0 = 5084.4(5) MHz, B0 = 4684(1) MHz, C0 = 2744.7(8) MHz, and (A0 + B0)/2 = 4884.5(7) MHz (±1σ). The analogous values for p-dioxane-d8 are A0 = 4083(2) MHz, B0 = 3925(4) MHz, C0 = 2347.1(6) MHz, and (A0 + B0)/2 = 4002.4(6) MHz. We determine the molecular structure with a semi-experimental approach involving the highly correlated coupled-cluster singles, doubles and iterated triples method and the cc-pCVXZ basis set series from double- to quadruple-zeta (X = D, T, Q). Combining the calculated vibrationally averaged rotational constants A0calc(X),B0calc(X),C0calc(X) for increasing basis-set size X with non-linear extrapolation to the experimental constants A0exp,B0exp,C0exp allows to determine the equilibrium ground state structure of p-dioxane. For instance, the equilibrium C–C and C–O bond lengths are re(CC) = 1.5135(3) Å and re(CO) = 1.4168(4) Å, and the four axial C–H bond lengths are 0.008 Å longer than the four equatorial ones. The latter is ascribed to the trans-effect (anomeric effect), i.e., the partial delocalization of the electron lone-pairs on the O atoms that are oriented trans, relative to the axial CH bonds.

p-Dioxane is a common solvent in many industrial and laboratory applications due to its low polarity and high boiling point. Its six-membered ring can assume three locally stable conformations denoted chair, boat, and twist-boat, which are analogous to those of cyclohexane.1–5 Since p-dioxane has no dipole moment, its rotational constants are not accessible by microwave spectroscopic techniques. Rotational coherence spectroscopy (RCS) is a time-domain spectroscopic method,6–8 which in its Raman variant can deliver accurate rotational constants of non-polar molecules.9–19 

Brown et al. measured the ground state rotational constants of jet-cooled p-dioxane-h8 using high-resolution infrared (IR) diode-laser absorption spectroscopy and assigned the rotational spectrum based on an asymmetric-top model employing Watson’s A-reduced Hamiltonian representation.20 Given the 10 K temperature in their jet expansion, only low-lying rotational levels were populated, which precluded fitting the centrifugal distortion constants.20 Their rotational constants20A0 = 5083.2(3) MHz, B0 = 4684.4(3) MHz, C0 = 2743.6(1) MHz, and (A0 + B0)/2 = 4883.8(3) MHz correspond to a rigid asymmetric-top model with centrifugal distortion constants set to zero.

In this work, we study the gas-phase rotational motion of p-dioxane−h8 and of its fully deuterated d8-isotopomer, using time-resolved femtosecond (fs) Raman RCS, detected by degenerate four-wave mixing (DFWM) both in a supersonic jet and in a room-temperature gas cell. We analyzed the RCS transients in terms of a non-rigid asymmetric top model taking the centrifugal distortion constants from highly correlated coupled-cluster singles, doubles and iterated triples [CCSD(T)] calculations. Since we measure at two temperatures (130 K and 293 K, see below), which are much higher than in Ref. 20, it was mandatory to include the effects of centrifugal distortion in the fit. Thus, while the p-dioxane-h8 rotational constants measured by IR laser spectroscopy20 are expected to differ slightly from those determined here, they constitute a valuable data set for cross-checking and validating our fs Raman RCS experiment and asymmetric-top RCS fitting program. We also combined the rotational constants with results of quantum chemical calculations at the CCSD(T) level and used a semi-experimental procedure based on basis-set extrapolation to obtain the equilibrium structure of p-dioxane, providing accurate bond lengths and angles.21–27 

In a gas-phase electron-diffraction (GED) study at room temperature, Davis and Hassel found that p-dioxane exists exclusively in the chair form.1 They reported the 298 K vibrationally averaged bond lengths ra(C–C), ra(C–O), and ra(C–H) (averaged over the equatorial and axial C–H bonds) and the vibrationally averaged C–C–O and C–O–C bond angles.1 In a later reinvestigation of the p-dioxane structure using GED, Fargher et al. determined all ten structure parameters of the chair form, including the two C–C–H and H–C–H bond angles and the “flap” angle of the six-ring (i.e., the tilt angle of the C–O–C plane relative to that of the four carbon atoms).5 They confirmed the previous finding1 that the GED pattern is explained by the chair form alone.5 Based on the B3LYP/cc-pVTZ calculated ΔE and ΔG298K0 values, they estimated the ratio of the chair relative to the next higher twist-boat conformer as 10000:1 at 298 K. Caminati et al. performed a microwave study on the p-dioxane water complex produced in a supersonic jet expansion.28 In their analysis, they employed the p-dioxane rα structure of Davis and Hassel1 and determined the structure parameters of the complex, but did not report new information on the p-dioxane moiety.28 

The chair chair inversion process in dioxane was investigated by two 1H NMR spectroscopic studies, in which the barrier for the ring-inversion was determined as ΔG=9.4 kcal/mol at 176 K and as 9.7 kcal/mol at 143 K.2,3 An X-ray diffraction investigation over the temperature range 133–285 K showed that the chair form is also the dominant conformation in the solid state.4 

Pickett and Strauss investigated the potential energy surface and the chair chair inversion process in dioxane and in related oxanes, using a potential derived from vibrational and geometrical data.29 They constructed a detailed conformational energy map which they used to estimate the chair chair transition state energy as 11 kcal/mol.29 Chapman and Hester performed a conformational analysis of p-dioxane at the ab initio HF/6-31G*, B3LYP/6-31*, and MP2/6-31* levels, and characterized ten stationary points as minima or transition structures.30 Based on a vibrational analysis, they proposed a detailed inversion mechanism chair [1,4 half-chair] 1,4 twist-boat [1,4 half-chair] chair, with a 1,4-half-chair barrier energy of 12.8 kcal/mol at the MP2/6-31* level.30 

We have measured the Raman RCS transients of p-dioxane-h8 and -d8 in order to obtain accurate ground-state rotational constants. The structure of p-dioxane is analogous to those of the isoelectronic molecules cyclohexane (C6H12)31–34 and piperidine (C5H11N).23,35 While cyclohexane is an oblate symmetric-top molecule that is characterized by an asymmetry parameter κ(2B0A0C0)/(A0C0)=+1.000,31–34 piperidine is a near-oblate asymmetric-top molecule with κ=+0.91 for its parent isotopomer.23 Replacing both methylene groups of cyclohexane by O atoms increases the asymmetry of p-dioxane, rendering it a strongly asymmetric top with κ=+0.656, see below. The increase of asymmetry gives rise to asymmetry transients in the Raman RCS spectrum of p-dioxane, which are not observable for cyclohexane,33,34 and which give detailed information on the rotational constants of p-dioxane. Interestingly, p-dioxane-d8 has a more symmetric mass distribution than the h8-isotopomer, making p-dioxane-d8 a near-oblate asymmetric top. Systematically changing κ from +0.66 to +1.0 for these structurally related molecules allows to study the influence of the inertial tensor on the Raman RCS spectrum. A further point is that the intensity of rotational Raman transitions and thus of the Raman RCS spectrum depends on the anisotropy of the molecular polarizability tensor α^, which is dominantly an electronic property. While the eightfold H/D substitution of p-dioxane-h8 strongly modifies the inertial tensor of p-dioxane, it has no effect on the polarizability tensor, leading to a rotation of α^ with respect to the inertial frame.

The detailed simulation of the asymmetric-top Raman RCS transients of p-dioxane-h8 and -d8 allows us to fit the experimental transients and thereby determine two sets of rotational constants {A0, B0, C0}. Given that p-dioxane has ten independent structure parameters, six rotational constants do not suffice to determine the equilibrium molecular structure. However, combining the experimental rotational constants with CCSD(T) calculations based on a systematic series of basis sets (the Dunning cc-pCVXZ series with X = D, T, and Q) allows us to determine the semi-experimental re structure21–27 of p-dioxane using a non-linear basis-set extrapolation.

The experimental setup used to record fs-DFWM-RCS transients has been described previously.33,34 In short, a 1 kHz repetition rate amplified Ti:sapphire laser system delivers pulses of 75 fs which are split into three beams of equal energy (10–100 μJ/pulse per beam), which act as the Raman pump and dump pulses, followed by the time-delayed probe pulse. The three beams are parallelized and focused by a f = 1000 mm achromatic lens into the probe volume in a folded BOXCARS arrangement.36 Measurements were performed in a gas cell at T = 293 K27,37 and in a pulsed supersonic jet, employing Ar as a carrier gas.34,38

For the supersonic-jet measurements, p-dioxane-h8 was heated to 80 °C, equivalent to a partial pressure of p = 500 mbar;39 the vapor was entrained in Ar carrier gas at a backing pressure of 1000 mbar. The gas mixture was expanded through a pulsed 0.6 mm diameter nozzle (333 Hz repetition rate) into a 0.1 mbar vacuum that is maintained by a Roots blower/rotary-vane vacuum pump combination. The RCS signal was generated in the overlap volume of the pump, dump, and probe laser beams within the core of the jet expansion 2 mm downstream of the nozzle orifice. The pump, dump, and probe pulses were blocked, while the coherently generated RCS signal beam was recollimated, spatially filtered, and detected by a thermoelectrically cooled GaAs photomultiplier. The transients were obtained by scanning the delay time between the pump/dump and the probe pulses in steps of 20 fs. Due to the large quantity (300500 ml) of p-dioxane needed per experiment, jet measurements of p-dioxane-d8 were not feasible.

Quantum-chemical calculations and optimizations of the equilibrium (re) geometries of p-dioxane chair-h8 and twist-chair-h8 were performed with the coupled-cluster approach using single and double excitations augmented by a perturbational estimate of connected triple excitations, CCSD(T). All electrons were correlated. The structures were optimized using the correlation-consistent polarized core-valence double-, triple-, and quadruple-zeta basis sets cc-pCVDZ, cc-pCVTZ, and cc-pCVQZ.40 It has been shown that the inclusion of high-order electron correlation and of core-valence correlation is important to achieve highly accurate structures and rotational constants.41–45 

The anharmonic vibrational frequencies, vibrationally averaged rotational constants Av, Bv, and Cv, and the molecular polarizability tensor were calculated for both conformers as well as for p-dioxane-d8 (chair) at the CCSD(T)/cc-pCVDZ level, using analytical second-derivative techniques.46,47 All calculations were carried out with the CFOUR program.48 

The CCSD(T)/cc-pCVQZ calculated structure of the chair isomer is shown in Fig. 1. The Z-matrices and the CCSD(T)/cc-pCVXZ optimized structure parameters of the chair conformer are listed in Tables S2–S5 of the supplementary material, and those of the twist-chair conformer in Tables S6–S8 of the supplementary material. The equilibrium rotational constants Ae, Be, Ce,, v = 0 vibrationally averaged rotational constants A0, B0, C0, and the elements of the electronic polarizability tensor α^ of the chair and twist-boat conformers of p-dioxane-h8 and -d8 are reported in Table I.

FIG. 1.

Equilibrium structure of p-dioxane and definition of nine of the ten structure parameters. The inertial {a, b, c} axes are indicated by dashed lines ( and indicate the down- and up-pointing axes, respectively). Also indicated are the C2 and sigmah symmetry operations of the C2h point group.

FIG. 1.

Equilibrium structure of p-dioxane and definition of nine of the ten structure parameters. The inertial {a, b, c} axes are indicated by dashed lines ( and indicate the down- and up-pointing axes, respectively). Also indicated are the C2 and sigmah symmetry operations of the C2h point group.

Close modal
TABLE I.

CCSD(T) calculated equilibrium, v = 0 vibrationally averaged rotational constants (in MHz), A-reduced centrifugal distortion constants (in kHz), and molecular polarizability tensor components (in atomic units) of p-dioxane(-h8, and -d8) of the chair conformation and of the twist-boat conformation of p-dioxane-h8.

Chair-h8Chair-d8Twist-boat-h8
cc-pCVDZcc-pCVTZcc-pCVQZcc-pCVDZcc-pCVTZcc-pCVQZcc-pCVDZcc-pCVTZ
 Rotational constants (MHz) 
Ae 5094.23 5137.48 5142.90 4036.70 4094.11 4114.13 5151.27 5195.65 
Be 4635.52 4698.16 4723.65 3920.90 3960.52 3965.51 4484.10 4556.92 
Ce 2737.84 2763.36 2771.66 2339.63 2362.56 2368.98 2790.76 2826.38 
A0 5035.77 5079.02a 5084.44a 4000.61 4085.02a 4078.04a 5086.92 5131.30a 
B0 4593.20 4655.84a 4681.33a 3880.14 3919.76a 3924.75a 4444.94 4517.76a 
C0 2708.88 2734.40a 2742.70a 2316.46 2339.39a 2345.81a 2757.32 2792.94a 
κb 0.620 0.639 0.656 0.857 0.811 0.823 0.449 0.475 
 Centrifugal distortion constants (kHz) 
ΔJ 1.247 4   0.6773   1.1766  
ΔK 1.146 5   0.4065   0.9345  
ΔJK −2.188 6   −0.9402   −1.8157  
δJ 0.010 65   −0.0364   −0.0606  
δK 1.391 9   0.0591   0.6596  
 Polarizability tensor elements (bohr3
αaa 54.8   54.7   42.5  
αbb 42.8   42.9   53.8  
αcc 43.4   43.3   43.2  
αab 0.0   0.0   0.0  
αac 0.0   0.0   0.0  
αbc 1.10   1.10   0.0  
α(0)(2) −4.45   −4.49   −4.04  
α(1)(2) −1.10   −1.08   0.0  
α(2)(2) −5.95   −5.90   −5.65  
α(0)(2)/α(2)(2) 0.75   0.76   0.72  
Chair-h8Chair-d8Twist-boat-h8
cc-pCVDZcc-pCVTZcc-pCVQZcc-pCVDZcc-pCVTZcc-pCVQZcc-pCVDZcc-pCVTZ
 Rotational constants (MHz) 
Ae 5094.23 5137.48 5142.90 4036.70 4094.11 4114.13 5151.27 5195.65 
Be 4635.52 4698.16 4723.65 3920.90 3960.52 3965.51 4484.10 4556.92 
Ce 2737.84 2763.36 2771.66 2339.63 2362.56 2368.98 2790.76 2826.38 
A0 5035.77 5079.02a 5084.44a 4000.61 4085.02a 4078.04a 5086.92 5131.30a 
B0 4593.20 4655.84a 4681.33a 3880.14 3919.76a 3924.75a 4444.94 4517.76a 
C0 2708.88 2734.40a 2742.70a 2316.46 2339.39a 2345.81a 2757.32 2792.94a 
κb 0.620 0.639 0.656 0.857 0.811 0.823 0.449 0.475 
 Centrifugal distortion constants (kHz) 
ΔJ 1.247 4   0.6773   1.1766  
ΔK 1.146 5   0.4065   0.9345  
ΔJK −2.188 6   −0.9402   −1.8157  
δJ 0.010 65   −0.0364   −0.0606  
δK 1.391 9   0.0591   0.6596  
 Polarizability tensor elements (bohr3
αaa 54.8   54.7   42.5  
αbb 42.8   42.9   53.8  
αcc 43.4   43.3   43.2  
αab 0.0   0.0   0.0  
αac 0.0   0.0   0.0  
αbc 1.10   1.10   0.0  
α(0)(2) −4.45   −4.49   −4.04  
α(1)(2) −1.10   −1.08   0.0  
α(2)(2) −5.95   −5.90   −5.65  
α(0)(2)/α(2)(2) 0.75   0.76   0.72  
a

Calculated using the cc-pCVDZ ΔA, ΔB, and ΔC values where, e.g., ΔB=BeB0.

b

κ=2BACAC is a measure of the asymmetry of a molecule where a value of +1 corresponds to an oblate symmetric-top and a value of −1 corresponds to a prolate symmetric-top.

The RCS signal I(t) is proportional to the square modulus of its time-dependent third-order susceptibility χ(3)t and can be written as49–51 

I(t)=G(τ)|χ(3)(tτ)|2dτ.
(1)

The experimental apparatus function Gt is determined by the triple convolution of the fs Raman pump, dump, and probe pulses and was measured prior to every RCS experiment using the time-domain Kerr-effect signal of Ar gas in the gas cell or in the supersonic jet at time zero. Gt can be modeled by a Gaussian of 140 fs full width at half maximum (FWHM); the center of Gt was also used to establish time-zero of the RCS experiment. χ(3)(t) associated with non-resonant rotational Raman excitation of asymmetric-top rotational states is given by

χ(3)(t)C+Γ<ΓcΓ,Γsin[(EΓEΓ)t/],
(2)

where ΓJτ and ΓJτ denote the asymmetric-top rotational levels and EΓ and EΓ are their respective rotational energies. In Eq. (2), the modulation amplitude cΓ,Γ is a product of (1) the ground- and excited state population differences eEΓkbTeEΓkbT, (2) the fractional vibrational population of the vibrational level v, pv, (3) the nuclear-spin statistical weight of the rotational level gns,Γ, and (4) the rotational Raman transition probabilities bΓ,Γ. For asymmetric-top molecules, the latter have to be calculated numerically, using the wave functions described in Eq. (4). For linear and symmetric-top molecules, the rotational Raman transition probabilities can be calculated analytically and are known as Placzek-Teller factors (coefficients).52,53

The empirical parameter C accounts for slight distortions of the RCS signal which may arise from stray light that temporally coincides with the probe pulse. In practice, this is only important for very weak RCS signals.

For the calculation of the rotational energy levels Jτ, we choose Watson’s A-reduced Hamiltonian including the first-order (quartic) Watson centrifugal distortion constants ΔJ, ΔK, ΔJK, δJ, and δK,54,55

H^rot=AJ^z+12(B+C)(J^2J^z2)+14(BC)(J^+2+J^2)ΔJJ^4ΔJKJ^z2ΔKJ^z412[δJJ^2+δkJ^z2,J^+2+J^2].
(3)

The CCSD(T) calculated equilibrium rotational constants Ae, Be, Ce, the v = 0 vibrationally averaged constants A0, B0, C0, and the centrifugal distortion constants are listed in Table I.

The asymmetric-top energy levels are determined by diagonalization of the asymmetric-top Hamiltonian in a symmetric-top basis.54,56 Asymmetric-top eigenfunctions ΓM can be expressed as linear combinations of symmetric-top eigenfunctions JKM,

ΓM=JτM=K=JJa(Γ,K)JKM
(4)

where τ=K1K+1 is the difference of the corresponding prolate and oblate symmetric-top state K quantum numbers. The subscripts −1 and +1 refer to κ=(2BAC)/(AC), where κ=1 corresponds to the prolate symmetric-top and κ=+1.0 to the oblate symmetric-top limit.54,56K=J,J1,,J denotes the projection of the total angular momentum J on the major molecule-fixed principal axis, which for p-dioxane is the c-axis, see also Fig. 1. For p-dioxane-h8, the CCSD(T)/cc-pCVQZ calculations predict κ=0.656, corresponding to a strongly asymmetric top, see Fig. S1 of the supplementary material. In contrast, p-dioxane-d8 is much closer to the oblate-top limit with a calculated κ=+0.823, as shown in Fig. S1 and given in Table I.

The rotational energy levels of asymmetric-top molecules lie between the prolate and oblate limits, as shown in Fig. S1 of the supplementary material. In contrast to symmetric-tops, asymmetric-top states are not degenerate for J-states with the same absolute K value. In fact K is not a good quantum number for asymmetric-tops, instead states are specified by J and τ=K1K+1.54,56 However, for molecules that are only slightly asymmetric, K+1 and K−1 are still approximately good quantum numbers. The rotational Raman selection rules that are discussed below are those for an oblate asymmetric top with KK+1.

The rotational Raman selection rules between levels ΓM and ΓM that contribute to the RCS signal [see Eq. (2)] arise from the rotational Raman transition moment ΓM|α^|ΓM, where α^ is the 3 × 3 molecular polarizability tensor. It is defined in the molecule-fixed axis system whose origin is at the molecular center of mass and whose axes are the principal axes {a, b, c} of the inertial tensor.57,58

The molecular polarizability tensor is an electronic property, so the principal axes of α^ need not coincide with those of the inertial tensor, which is dominantly a nuclear-framework property. Given the C2h symmetry of p-dioxane-h8 and p-dioxane-d8, one axis of α^ coincides with the b axis of the inertial tensor and the other two axes of α^ lie in the σh symmetry plane that contains the a and c inertial axes, see Fig. 1. Since the two axes of α^ that lie in this plane may be tilted with respect to a and c, an off-diagonal element arises, which is calculated to be αbc=1.10 bohr3. The CCSD(T)/cc-pCVDZ calculated elements of α^ are given in part 3 of Table I. With our choice of the molecule-fixed coordinate system for p-dioxane-h8, we obtain αzz=αcc, αyy=αaa, and αxx=αbb.

A detailed derivation of Raman selection rules for symmetric-top molecules is given in Ref. 59. The analogous procedure for finding non-zero values of the rotational Raman transition moment can be applied to asymmetric-top states, as these can be written as a superposition of symmetric-top states, see Eq. (4). Rotational Raman transition moment integrals are best dealt with by decomposing α^ into irreducible spherical polarizability tensor components, αk(j), and then using angular momentum coupling rules.59αk(j) can be regarded as angular momentum states J=j,K=k where k runs from −j to +j. For p-dioxane only the αk(2) tensor components contribute to the RCS signal. Each element in αk(2) can be constructed as a linear combination of the molecular Cartesian polarizability tensor components as shown in Table II. The calculated irreducible spherical tensor components α(0)(2), α(1)(2), and α(2)(2) are given in part III of Table I. The dependence of the allowed changes in K quantum number on the molecular polarizability tensor components allows deeper insight into the simulated RCS transient:60 For example, if one is interested only in coherences with ΔJ=1,2 and ΔK=0, the molecular polarizability is set to values such that αxx=αyyαzz and αoffdiagonal=0.60 

TABLE II.

The relationship of the irreducible spherical polarizability components αk(j) to their molecular polarizability Cartesian counterparts αρσ, from Ref. 59. For p-dioxane, αzz=αcc, αyy=αaa, and αxx=αbb.

ΔJΔKαk(j)αρσ
 α2(2) 12[(αxxαyy)+i(αxy+αyx)] 
±α1(2) 12[(αxz+αzx)+i(αyz+αzy)] 
±α0(2) 16(2αzz(αxx+αyy)) 
−1 α1(2) 12[(αxz+αzx)i(αzy+αyz)] 
 −2 α2(2) 12[(αxxαyy)i(αxy+αyx)] 
ΔJΔKαk(j)αρσ
 α2(2) 12[(αxxαyy)+i(αxy+αyx)] 
±α1(2) 12[(αxz+αzx)+i(αyz+αzy)] 
±α0(2) 16(2αzz(αxx+αyy)) 
−1 α1(2) 12[(αxz+αzx)i(αzy+αyz)] 
 −2 α2(2) 12[(αxxαyy)i(αxy+αyx)] 

Raman RCS transients are characterized by recurring features which are a manifestation of regularities in the structure of the rotational energy levels. Regular frequency spacings occur for (rigid) symmetric-top molecules with rotational energies given by E(J, K) = BJ(J + 1) + (CB)K2 for an oblate symmetric top.8 Felker and co-workers have shown that even for asymmetric molecules, significant subsets of the rotational levels can retain sufficiently regular frequency spacings as to yield rotational recurrences.6,8,61 These subsets produce transients that are characterized by their recurrence time. Transients are named either by their main contributing coherences (J-type and K-type) or by the rotational constants about which they convey information (A-type and C-type). The relevant transient types for p-dioxane are listed in Table S1 of the supplementary material; some corresponding rotational transitions are indicated in Fig. S1 of the supplementary material.8,54 Note that a rotational transition specified by a specific ΔJ and ΔK is not bijectively associated with a single transient type; for instance, rotational transitions with ΔJ=2,ΔK1=0 contribute to both J-type and C-type transients.

The relative intensity of transient types with different |ΔK| values is largely determined by the magnitude of the irreducible spherical polarizability tensor components αk(2) that promote the contributing rotational transitions. The RCS signal is proportional to the fourth power of αk(2); thus, the intensity of J-type transients is proportional to |α(0)(2)|4 whereas the intensity of K-type, A-type, and C-type transients is proportional to |α(2)(2)|4. For both p-dioxane-h8 and p-dioxane-d8, J-, A-, C-, and K-type transients are expected.

The asymmetric-top RCS transients are calculated via Eqs. (1) and (2) using an IDL program (RSI, Inc.); the simulated transient Icalc(t) is then fitted to the experimental RCS transient Iexp(t) using a Levenberg-Marquardt nonlinear least-squares fit.25–27,34,60 The starting values of A0, B0, C0 in the fit were taken from the CCSD(T)/cc-pCVQZ calculations, see Table I. The A-reduced centrifugal distortion constants and the αaa, αcc, and αbc components of the molecular polarizability tensor were fixed to their CCSD(T)/cc-pCVDZ calculated values (see Table I) and were not fitted. The polarizability tensor component αbb was fitted to the experimental transient to balance the relative intensity of ΔK=0 and ΔK=2 transitions.

The number of rotational levels to be considered in the simulation is determined by the cumulative rotational level population at the experimental temperature T and the bandwidth of the fs laser pulse, which limits the highest Raman rotational transition frequencies that can be driven and that contribute to the RCS signal. We considered a cumulative rotational population of 99% at T = 293 K, which corresponds to rotational levels with J = 0–150. Assuming the oblate symmetric-top limit for p-dioxane and the largest possible quantum-number change ΔJ=2/ΔK=2, one obtains ΔνJ,K=4BJ+4(AB)K=4AJ, where for the second step we set K = J since we are only interested in the highest transition frequency. The highest rotational level that can be excited within the BW = 5.6 THz bandwidth of our fs laser system is Jmax=BW/4A=5600GHz/45.1GHz270. This is considerably smaller than the J = 150 upper limit considered, so the laser bandwidth is sufficiently large to drive all possible Raman rotational transitions.

In the gas-cell measurements at T = 293 K, a number of low-lying thermally excited vibrational levels contribute to the RCS signal, see Table III. The v-dependent rotational constants Av, Bv, and Cv can be expressed in terms of the vibration-rotation coupling constants αvA, αvB, and αvC, e.g., for Av,62 

Av=Aei=13N6αiAvi+1/2,
(5)

where Ae is the rotational constant associated with the rigid equilibrium structure, αiA,B,C are the vibration-rotation interaction constants associated with each of the 36 vibrational modes of p-dioxane, and vi=0,1,2, is the quantum number of the ith vibrational mode. The vibrational-averaging corrections for the vibrational zero-point level v = 0 are denoted as ΔA,B,C, e.g., for ΔA,

ΔA=A0Ae=12i3N6αiA.
(6)

The vibrational averaging corrections will be discussed in more detail in Sec. VI.

TABLE III.

CCSD(T)/cc-pCVDZ calculated vibrational frequencies of p-dioxane-h8 and p-dioxane-d8 (in cm−1) up to ν12, relative vibrational populations and relative signal intensities (in %) at 298 K and 130 K, and calculated rotation-vibration constants αvA,B,C (in MHz).

T = 298 KT = 130 K
Normal modeWave numberPopulationSignalPopulationSignalαvAαvBαvC
p-dioxane-h8 
v = 0 0.0 37.7 82.7 98.7 99.4    
ν7 266.3 10.4 6.3 5.4 0.4 1.56 5.08 3.39 
ν8 279.6 9.8 5.6 3.9 0.2 9.61 −0.55 −0.74 
ν9 433.5 4.6 1.3 0.8 0.0 6.61 35.20 3.14 
ν10 445.7 4.4 1.1 0.6 0.0 7.47 −38.36 12.33 
ν11 487.5 3.6 0.7 0.4 0.0 −8.96 2.34 −9.38 
ν12 615.8 1.9 0.2 0.1 0.0 −0.05 3.02 2.63 
2ν7 532.6 2.9 0.5 0.3 0.0    
2ν8 559.1 2.5 0.4 0.2 0.0    
ν7+ν8 545.9 2.7 0.4 0.2 0.0    
p-dioxane-d8 
v=0 0.0 26.3 71.0 78.9 98.2    
ν7 196.1 10.2 10.7 9.0 1.3 3.59 1.09 2.00 
ν8 243.9 8.1 6.7 5.3 0.4 −0.24 5.64 0.36 
ν9 350.3 4.8 2.4 16. 0.0 9.46 5.71 3.79 
ν10 407.1 3.7 1.4 0.9 0.0 2.07 −2.89 14.29 
ν11 421.4 3.4 1.2 0.7 0.0 −10.70 −0.38 −13.79 
ν12 483.3 2.6 0.7 0.4 0.0 2.70 −0.83 1.54 
2ν7 392.2 4.0 1.6 1.0 0.0    
2ν8 487.8 2.5 0.6 0.4 0.0    
ν7+ν8 404.0 3.1 1.0 0.6 0.0    
T = 298 KT = 130 K
Normal modeWave numberPopulationSignalPopulationSignalαvAαvBαvC
p-dioxane-h8 
v = 0 0.0 37.7 82.7 98.7 99.4    
ν7 266.3 10.4 6.3 5.4 0.4 1.56 5.08 3.39 
ν8 279.6 9.8 5.6 3.9 0.2 9.61 −0.55 −0.74 
ν9 433.5 4.6 1.3 0.8 0.0 6.61 35.20 3.14 
ν10 445.7 4.4 1.1 0.6 0.0 7.47 −38.36 12.33 
ν11 487.5 3.6 0.7 0.4 0.0 −8.96 2.34 −9.38 
ν12 615.8 1.9 0.2 0.1 0.0 −0.05 3.02 2.63 
2ν7 532.6 2.9 0.5 0.3 0.0    
2ν8 559.1 2.5 0.4 0.2 0.0    
ν7+ν8 545.9 2.7 0.4 0.2 0.0    
p-dioxane-d8 
v=0 0.0 26.3 71.0 78.9 98.2    
ν7 196.1 10.2 10.7 9.0 1.3 3.59 1.09 2.00 
ν8 243.9 8.1 6.7 5.3 0.4 −0.24 5.64 0.36 
ν9 350.3 4.8 2.4 16. 0.0 9.46 5.71 3.79 
ν10 407.1 3.7 1.4 0.9 0.0 2.07 −2.89 14.29 
ν11 421.4 3.4 1.2 0.7 0.0 −10.70 −0.38 −13.79 
ν12 483.3 2.6 0.7 0.4 0.0 2.70 −0.83 1.54 
2ν7 392.2 4.0 1.6 1.0 0.0    
2ν8 487.8 2.5 0.6 0.4 0.0    
ν7+ν8 404.0 3.1 1.0 0.6 0.0    

The thermally populated vibrational levels of p-dioxane-h8 and -d8 are ν7, ν8, ν9, ν10, ν11, ν12, 2ν7, 2ν8, and ν7+ν8. They account cumulatively for >99% of the total RCS signal for p-dioxane-h8 and for >97% of the RCS signal for p-dioxane-d8, as can be seen from Table III. The CCSD(T)/cc-pCVDZ calculated rotational-vibrational coupling constants for ν7 to ν12 are also given in Table III. The corresponding Av, Bv, Cv constants were calculated via Eq. (6), and the fractional contributions of these levels to χ(3) and to the RCS signal were included in the simulated transients and in the fits. The fit parameters in the simulations of the room-temperature gas-cell transients were A0, B0, and C0, αbb=αxx (see Table II), and the empirical signal distortion parameter C, see Eq. (2).

In the supersonic-jet RCS transient, only J-type transients could be observed, in contrast to the room-temperature gas cell transient. The supersonic-jet transient was fitted with the parameters A0, B0, and C0, the rotational temperature Trot, and the parameter C; the polarizability tensor component αbb was fixed to its CCSD(T)/cc-pCVDZ calculated value. We initially included the above-mentioned vibrationally excited levels in the fit, but found that these do not contribute significantly to the RCS signal at the supersonic-jet temperature. We thus assumed the vibrational temperature Tvib to be equal to the rotational temperature Trot, TvibTrot=130±10 K. This is in agreement with the calculated vibrational populations in Table III.

1. p-Dioxane-h8

An overview of the gas-cell transient for delay time Δt=0650 ps is shown in Fig. 2(a). The transient length is relatively short because of the small anisotropy of the molecular polarizability tensor of p-dioxane, see also Table I. As noted above the RCS signal is proportional to the fourth power of the irreducible spherical polarizability components α(0)(2) and α(2)(2). For p-difluorobenzene,60 we calculated that α(0)(2)=19.0 bohr,3 whereas for p-dioxane-h8α(0)(2) is only 4.4 bohr3. This translates to a decrease of RCS signal intensity by a factor of 250 at equal gas-phase density (pressure). Compared to symmetric-top molecules, the RCS transient decays much faster with increasing Δt because of the relatively large asymmetry of p-dioxane-h8 (κ=0.658). Despite the asymmetry, the RCS transient of p-dioxane-h8 appears rather simple, being dominated by J-type features. Only a close-up view, which is shown in Fig. 3, reveals the fine structure that substantiates the asymmetry of the molecule. Comparison to the gas-cell Raman RCS transient of p-dioxane-d8 in Fig. 2(b) shows that the larger κ=+0.890 of the perdeuterated compound increases the observable time duration of the RCS transient by a factor of 1.5.

FIG. 2.

Raman RCS transients of p-dioxane-h8 and p-dioxane-d8, measured in a gas cell at T = 293 K and p = 30 mbar.

FIG. 2.

Raman RCS transients of p-dioxane-h8 and p-dioxane-d8, measured in a gas cell at T = 293 K and p = 30 mbar.

Close modal
FIG. 3.

Detailed assignment of the T = 293 K gas-cell Raman RCS transient of p-dioxane-h8 for time delays t = 0-400 ps: (a) Magnified and assigned experimental fs Raman RCS transient. (b)–(f) Sub-transients that simulated by restricting the coherences to Raman-allowed ΔJ,ΔK transitions.

FIG. 3.

Detailed assignment of the T = 293 K gas-cell Raman RCS transient of p-dioxane-h8 for time delays t = 0-400 ps: (a) Magnified and assigned experimental fs Raman RCS transient. (b)–(f) Sub-transients that simulated by restricting the coherences to Raman-allowed ΔJ,ΔK transitions.

Close modal

Figure 3(a) compares the experimental transient of p-dioxane-h8 to the simulated sub-transients shown in panels (b)-(f), which correspond to coherences for specific ΔJ and ΔK values. This procedure60 emphasizes the signal contributions from different coherences and allows us to assign even small signals to specific transient types. Since asymmetric-top states are eigenfunctions of the total angular momentum J, see Eq. (4), one can restrict ΔJ to a subset of transitions as outlined in Sec. IV C: The coherences for ΔK transitions are selected by setting specific values for the polarizability tensor elements. Combining ΔJ and ΔK restrictions allows us to simulate five sub-transients with ΔJ = 0/ΔK = 2, ΔJ = 1/ΔK = 0, ΔJ = 1/ΔK = 2, ΔJ = 2/ΔK = 0, and ΔJ = 2/ΔK = 2. Note, however, that changes of the K−1 or K1 quantum numbers cannot be selected. In addition to the intense J-type progression, which occurs at multiples of Δt=51 ps, we can assign C-type (Δt=91 ps), A-type (Δt=49 ps), and K-type (Δt=118 ps) coherences.

The summation of the different ΔJ/ΔK contributions to χ(3) in Eq. (2) gives rise to interferences between different coherences, which strongly affects the relative intensities of the sub-transients in I(t). Thus the total RCS signal in Fig. 3(a) is not obtained by adding the individual sub-transients shown in panels (b)-(f), but by first coherently adding all ΔJ/ΔK sub-transient amplitudes to obtain χ(3)(t)—see Eq. (2)—and then modulus-squaring the result to obtain the RCS signal, see Eq. (1). Note also that collisional dephasing effects become important over the 400 ps delay time of the gas-cell RCS transient in Fig. 3(a); these effects were not included for the sub-transient analysis.

2. Twist-boat conformer

We also searched for signatures of the twist-boat-conformer of p-dioxane-h8. In Fig. 4, we compare the experimental room-temperature gas-cell Raman RCS transient to that which we calculate for the cc-pCVTZ optimized twist-boat conformer. Even the most intense predicted signal of the twist-boat predicted at Δt=193 ps is not observed. This confirms the result of Hedberg et al. who were unable to observe the twist-boat conformer.5 

FIG. 4.

Experimental room-temperature RCS transient of para-dioxane, compared to the calculated RCS transient of the twist-boat conformer. Panel (b) shows the experimental and calculated RCS transients for t = 50–450 ps. The most intense simulated coherences of the twist-boat conformer are magnified horizontally and compared to the experiment in panel (a).

FIG. 4.

Experimental room-temperature RCS transient of para-dioxane, compared to the calculated RCS transient of the twist-boat conformer. Panel (b) shows the experimental and calculated RCS transients for t = 50–450 ps. The most intense simulated coherences of the twist-boat conformer are magnified horizontally and compared to the experiment in panel (a).

Close modal

3. p-Dioxane-d8

As discussed above, with κ=0.890, p-dioxane-d8 is much closer to the oblate symmetric-top limit than the h8-isotopomer. Therefore the rotational level structure of p-dioxane-d8 has many more close level coincidences than the h8-isotopomer, which in turn increases the time-delay over which we can observe recurrences with a sufficiently high S/N ratio to Δt=0900 ps, as shown in Fig. 2(b). The closeness to the oblate symmetric-top case renders the p-dioxane-d8 transient rather simple, with a dominant J-type progression. At large magnification, three additional C-type recurrences are observed at Δt=320 ps, Δt=430 ps, and Δt=745 ps. In this case, sub-transient decomposition was not necessary for the assignments.

The complete experimental gas-cell RCS transient of p-dioxane-h8 at T = 293 K is shown in the top panel of Fig. 5. Magnified sections of the experimental data are compared to the calculated transients in the 16 following sub-panels. The v = 0 vibrational ground state dominates the RCS signal at room temperature (82.7% of the signal, see Table III). Nevertheless, the transient is strongly influenced by the rotational constants of the thermally populated vibrational states, especially ν7 (6.3%) and ν8 (5.6%).

FIG. 5.

Experimental and simulated femtosecond rotational Raman RCS spectrum of p-dioxane-h8 in a room-temperature gas cell for delay time t = 50–670 ps. The fine structure is well reproduced by the simulation upon including signal contributions from the low-lying vibrational levels ν7, ν8, ν9, ν10, ν11, ν12, 2ν7, 2ν8, and ν7+ν8, see also Table III.

FIG. 5.

Experimental and simulated femtosecond rotational Raman RCS spectrum of p-dioxane-h8 in a room-temperature gas cell for delay time t = 50–670 ps. The fine structure is well reproduced by the simulation upon including signal contributions from the low-lying vibrational levels ν7, ν8, ν9, ν10, ν11, ν12, 2ν7, 2ν8, and ν7+ν8, see also Table III.

Close modal

As an illustration of the contributions from the thermally populated vibrations to χ(3)(t), Fig. 6 shows a section of the RCS signal with a large A-type transient at 542 ps and a small C-type transient at 547 ps. Since the rotational constants of the excited vibrational levels differ from those of the v = 0 level by their respective vibration-rotation coupling constants, see Eqs. (5) and (6) and Table III, they produce coherences that are slightly time-shifted with respect to the v = 0 level, as indicated for the A-type transients by the vertical arrows in Figs. 6(a)–6(g). These shifts reflect the change of the Av rotational constant, as is expected for A-type transients. We construct the total χ(3)(t) in Fig. 6(h) by summing traces (a)-(g) for all thermally populated vibrational fundamentals, including additional overtone and combination levels, as listed in Fig. 6. The relative population of these levels at T = 293 K and respective αvA,B,C values are given in Table III. The simulated I(t) in Fig. 6(i) is obtained from χ(3)(t) via Eq. (1) and is fitted to the observed I(t) as shown in Fig. 6(i). The fit not only involves the rotational level eigenenergies (as is typical for microwave spectroscopy) but also the rotational Raman transition intensities, which are based on the rotational wave functions, see Eq. (4).

FIG. 6.

The calculated χ(3)(t) contributions to Eq. (2) for Δt=538554 ps at T = 293 K from (a) the v = 0 vibrational level and (b)–(g) the v = 1 levels of modes ν7ν12. The vertical arrows mark the positive maximum of the contributions to χ(3). These are shifted relative to the trace (a) due to their slightly different rotational constants Av, Bv, Cv, see Eqs. (6) and (7). Trace (h) shows the summed contributions from (a)-(g) plus the contributions from the overtone levels 2ν7, 2ν8, and the ν1+ν2 combination; see also Table III. Trace (i) gives the modulus-square of χ(3)(t), see Eqs. (1) and (2), which is compared to the experimental signal IDFWM|χ(3)|2.

FIG. 6.

The calculated χ(3)(t) contributions to Eq. (2) for Δt=538554 ps at T = 293 K from (a) the v = 0 vibrational level and (b)–(g) the v = 1 levels of modes ν7ν12. The vertical arrows mark the positive maximum of the contributions to χ(3). These are shifted relative to the trace (a) due to their slightly different rotational constants Av, Bv, Cv, see Eqs. (6) and (7). Trace (h) shows the summed contributions from (a)-(g) plus the contributions from the overtone levels 2ν7, 2ν8, and the ν1+ν2 combination; see also Table III. Trace (i) gives the modulus-square of χ(3)(t), see Eqs. (1) and (2), which is compared to the experimental signal IDFWM|χ(3)|2.

Close modal

We performed analogous fits for all 16 delay-time windows in Fig. 5 and averaged the resulting rotational constants, giving A0 = 5084.4(5) MHz, B0 = 4684(1) MHz, C0 = 2744.7(8) MHz, and (A0 + B0)/2 = 4884.5(7) MHz, with ±1σ uncertainty given in parentheses, see Table IV. As Table S1 of the supplementary material shows, the most intense J-type transients only give information on (A0 + B0)/2; on the other hand, the weak asymmetry transients allow us to fit all three rotational constants even without relying on J-type transients. This demonstrates that the presence of asymmetry transients is much more important than large signal intensities when it comes to the determination of rotational constants.

TABLE IV.

Experimental and CCSD(T) calculated ground state rotational constants of p-dioxane-h8 and p-dioxane-d8 (in MHz, ±1σ uncertainties in parentheses).

p-dioxane-h8p-dioxane-d8
ConstantCCSD(T)/ cc-pCVQZafs-Raman RCS supersonic jetfs-Raman RCS gas-cellIR diode laser jet bCCSD(T)/ cc-pCVQZafs-DFWM gas-cellb
A0 5084.44  5084.4(5) 5083.2(3) 4078.04 4083(2) 
B0 4681.33  4684(1) 4684.4(3) 3924.75 3925(4) 
C0 2742.70  2744.7(8) 2743.6(1) 2345.81 2347.1(6) 
(A0 + B0)/2 4882.885 4887(1) 4884.5(7) 4883.8(3) 4001.395 4002.4(6) 
p-dioxane-h8p-dioxane-d8
ConstantCCSD(T)/ cc-pCVQZafs-Raman RCS supersonic jetfs-Raman RCS gas-cellIR diode laser jet bCCSD(T)/ cc-pCVQZafs-DFWM gas-cellb
A0 5084.44  5084.4(5) 5083.2(3) 4078.04 4083(2) 
B0 4681.33  4684(1) 4684.4(3) 3924.75 3925(4) 
C0 2742.70  2744.7(8) 2743.6(1) 2345.81 2347.1(6) 
(A0 + B0)/2 4882.885 4887(1) 4884.5(7) 4883.8(3) 4001.395 4002.4(6) 
a

Calculated using the cc-pCVDZ ΔA, ΔB, and ΔC values where, e.g., ΔB=BeB0.

b

Reference 20, fit without centrifugal distortion constants, see Eq. (4).

The Raman RCS rotational constants of p-dioxane-h8 agree nicely with the previous values obtained by Brown et al.20 by IR laser spectroscopy in a very cold (Trot = 10 K) supersonic jet. Their rotational constants are included in Table IV, and the comparison confirms the analysis of the fs Raman RCS experiments and our RCS fits. Small differences between the two sets of constants are to be expected, since we had to include centrifugal distortion constants in our non-rigid rotor fits at T = 130 and 293 K, while Brown et al. could employ a rigid-rotor model at their low rotational temperatures.20 

The supersonic-jet signal intensity was low, so only the most intense four J-type recurrences could be observed, as is shown in Fig. 7. Although these recurrences do not yield very accurate rotational constants, we used them to check the room-temperature fitted (A0 + B0)/2. The supersonic-jet fitted (A0 + B0)/2 = 4887(1) MHz agrees with the room-temperature fitted value of (A0 + B0)/2 = 4884.5(7) MHz within the mutual 2σ values. The difference could be due to the poorer statistics of the analysis of the supersonic-jet transient, as well as to small errors of the calculated rotation-vibration coupling constants αvA, αvB, and αvC that were employed for the room-temperature gas-cell fit.

FIG. 7.

Experimental and simulated femtosecond Raman RCS transients of p-dioxane-h8 in an Ar supersonic-jet at Trot = Tvib = 130 K.

FIG. 7.

Experimental and simulated femtosecond Raman RCS transients of p-dioxane-h8 in an Ar supersonic-jet at Trot = Tvib = 130 K.

Close modal

Figure 8 shows the experimental and simulated room-temperature gas cell RCS transient of p-dioxane-d8. The 12 observed J and C-type transients allow for an determination of (A0 + B0)/2 = 4002.4(6) MHz and C0 = 2347.1(6) MHz with a similar accuracy as for p-dioxane-h8, see Table IV. However, the individual rotational constants A0 = 4083(2) MHz and B0 = 3925(4) MHz of p-dioxane-d8 have an almost 10x larger uncertainty than the combined constant (A0 + B0)/2. This again shows that transient complexity is much more important when determining rotational constants than transient length.

FIG. 8.

Experimental and simulated femtosecond rotational Raman RCS spectrum of p-dioxane-d8 in a room-temperature gas cell for delay time t = 60–920 ps. Low-lying vibrational levels ν7, ν8, ν9, ν10, ν11, ν12, 2ν7, 2ν8, and ν7+ν8 were included in the fit, see also Table III.

FIG. 8.

Experimental and simulated femtosecond rotational Raman RCS spectrum of p-dioxane-d8 in a room-temperature gas cell for delay time t = 60–920 ps. Low-lying vibrational levels ν7, ν8, ν9, ν10, ν11, ν12, 2ν7, 2ν8, and ν7+ν8 were included in the fit, see also Table III.

Close modal

By combining the experimentally measured rotational constants with the calculated equilibrium and vibrationally averaged rotational constants and the corresponding equilibrium structure parameters, it is possible to determine a semi-experimental equilibrium structure.21–23,37,38,60,63 In short, we calculated the equilibrium re molecular structure and the equilibrium rotational constants Ae, Be, Ce using the CCSD(T) method and Dunning’s correlation-consistent core-valence polarized basis sets series cc-pCVXZ from double- to quadruple-zeta (X = D, T, Q). With the cc-pCVDZ basis set, we additionally calculated the cubic force field necessary for the vibrational-averaging corrections ΔA,ΔB,ΔC in Eq. (7). With our current computer resources, these could not be calculated with the larger (X = T,Q) basis sets, so we used the CCSD(T)/cc-pCVDZ values in conjunction with all basis sets.

We generalized the previous two-point linear basis-set extrapolation method37,38,60,63 to a three-point non-linear extrapolation, in which a CCSD(T) calculated vibrationally averaged rotational constant (e.g., A0) is plotted vs. a given calculated equilibrium structure parameter of p-dioxane [e.g., the equilibrium bond length re(C–C)]. Thus, for A0 and re(C–C) which are calculated as a function of the basis-set size X, we plot A0=C1+C2exp[C3re(C–C)]. The semi-experimental re(C–C) bond length is obtained by extrapolating the calculated A0/re(C–C) curve to the experimental A0 value and determining the extrapolated re(C–C) value at the intersection. Analogous extrapolations are made for B0/re(C–C) and C0/re(C–C). The procedure is shown exemplarily in Fig. 9 for the rotational constants A0, B0, and C0 of p-dioxane-h8 and the constants C0 and (A0 + B0)/2 of p-dioxane-d8, all of which are plotted vs. re(C–C). The experimental uncertainties of these five rotational constants are so small that they do not contribute significantly to the extrapolated re(C–C) values. However, the extrapolations yield slightly different values of re(C–C); Fig. 9 shows that the spread for re(C–C) is 0.0009 Å and the corresponding 1σ is 0.0003 Å. We report the semi-experimental structure parameters in Table V as the average and ±1σ standard deviation of the five values determined via the different rotational constants, e.g., re(C–C) = 1.5135(3) Å.

FIG. 9.

CCSD(T) calculated equilibrium rotational constants () and vibrational ground state rotational constants () of p-dioxane, plotted vs. the equilibrium re(C–C) bond length. The extrapolated vibrational ground state rotational constants are connected by red lines. The semi-experimental reSE(C–C) bond length is determined by intersecting the extrapolation function with the experimental rotational constant (drawn as horizontal black lines). This procedure is shown for the rotational constants A0, B0, C0, (A0 + B0)/2 of p-dioxane-h8 (left) and p-dioxane-d8 (right).

FIG. 9.

CCSD(T) calculated equilibrium rotational constants () and vibrational ground state rotational constants () of p-dioxane, plotted vs. the equilibrium re(C–C) bond length. The extrapolated vibrational ground state rotational constants are connected by red lines. The semi-experimental reSE(C–C) bond length is determined by intersecting the extrapolation function with the experimental rotational constant (drawn as horizontal black lines). This procedure is shown for the rotational constants A0, B0, C0, (A0 + B0)/2 of p-dioxane-h8 (left) and p-dioxane-d8 (right).

Close modal
TABLE V.

CCSD(T) calculated, and semi-experimental bond lengths (in Å) and bond angles (in degrees) for p-dioxane. ±1σ uncertainties in parentheses. Experimental rα,294K and ra,294K bond lengths and bond angles from Ref. 5 are given in the last two columns.

reCCSD(T)rerarαra
pCVDZpCVTZpCVQZfs-Raman RCSGED, Ref. 5 
Bond length 
r(C–C) 1.5267 1.5154 1.5139 1.5135(3) 1.5221(3) 1.510(2) 1.512(2) 
r(C–O) 1.4254 1.4191 1.4172 1.4168(4) 1.4232(4) 1.417(1) 1.419(1) 
r(C–Heq1.1037 1.0892 1.0886 1.0885(1) 1.0935(1) 1.082(2) 1.094(2) 
r(C–Hax1.1120 1.0971 1.0964 1.0963(1) 1.1016(1) 1.090(2) 1.102(2) 
reCCSD(T)rerarαra
pCVDZpCVTZpCVQZfs-Raman RCSGED, Ref. 5 
Bond length 
r(C–C) 1.5267 1.5154 1.5139 1.5135(3) 1.5221(3) 1.510(2) 1.512(2) 
r(C–O) 1.4254 1.4191 1.4172 1.4168(4) 1.4232(4) 1.417(1) 1.419(1) 
r(C–Heq1.1037 1.0892 1.0886 1.0885(1) 1.0935(1) 1.082(2) 1.094(2) 
r(C–Hax1.1120 1.0971 1.0964 1.0963(1) 1.1016(1) 1.090(2) 1.102(2) 
Bond angle 
C–O–C 109.292 109.618 109.907 110.1(1) 110.5(1) 110.9(5) 110.7(5) 
C–C–O 110.556 110.621 110.556 a 110.6b 111.1(2) 110.9(2) 
C–C–Heq 110.964 110.859 110.898 a 110.9b 106.9(28)  
C–C–Hax 109.142 109.411 109.474 109.481(7) 109.553(7) 104.6(28)  
H–C–H 108.990 109.067 109.118 109.123(5) 108.989(5) 108.0(13)  
Flapc 52.639 52.331 52.305 52.299(6) 51.924(6) 50.6(4) 51.1(4) 
Bond angle 
C–O–C 109.292 109.618 109.907 110.1(1) 110.5(1) 110.9(5) 110.7(5) 
C–C–O 110.556 110.621 110.556 a 110.6b 111.1(2) 110.9(2) 
C–C–Heq 110.964 110.859 110.898 a 110.9b 106.9(28)  
C–C–Hax 109.142 109.411 109.474 109.481(7) 109.553(7) 104.6(28)  
H–C–H 108.990 109.067 109.118 109.123(5) 108.989(5) 108.0(13)  
Flapc 52.639 52.331 52.305 52.299(6) 51.924(6) 50.6(4) 51.1(4) 
a

The C–C–O and C–C–Heq angles do not extrapolate monotonically with increasing basis-set size (see the text).

b

The CCSD(T)/cc-pCVQZ calculated values were used for the estimation of ra,294K.

c

The angle between the COC plane and the plane of the four carbon atoms.

Figure 9 also shows sub-plots involving the A0 and B0 constants of p-dioxane-d8. Since these are 10 times less precise than the other experimental constants, we did not include them in the analysis, but note that the respective extrapolated re(C–C) values are consistent with those of the other sub-plots. A further consistency check was done by comparing to the average rotational constant (A0 + B0)/2 of p-dioxane-h8 (lower left in Fig. 9), which gives re(C–C) = 1.5136 Å, in excellent agreement with the value above.

Previously we have employed linear two-point extrapolations using basis sets with X = 2 and 3.37,38,60,63 Figure 9 now reveals that the extrapolation functions can be concave or convex, depending on the rotational constant. While our previous assumption of linearity is acceptable for some combinations of rotational constants and structure parameters, e.g., for (A0 + B0)/2 vs. re(C–C), it is clearly not valid for other combinations, e.g., A0 or B0 vs. re(C–C) bond length, see Fig. 9.

The analogous plots for the semi-experimental equilibrium bond lengths re(C–O), re(C–Heq), re(C–Hax) and the equilibrium bond angles C–O–C, C–C–Hax, H–C–H, and the flap angle are given in Figs. S2–S8 of the supplementary material; the corresponding semiexperimental structure parameters are collected in the fourth data column of Table V.

While the extrapolations in Fig. 9 and Figs. S2–S8 of the supplementary material are ordered with respect to the basis-set size X, X is not a variable in this procedure, in contrast to usual basis-set extrapolation schemes. Thus, the extrapolation curve through the calculated points should approach the experimental rotational constant monotonically as a function of X and should not cross the experimental line; this is indeed observed in Fig. 9 and Figs. S2–S8 of the supplementary material. In general, the extrapolated semi-experimental structure parameters lie slightly beyond the CCSD(T)/cc-pCVQZ calculated values, indicating that the CCSD(T)/cc-pCVXZ combination is rather close to the basis-set limit. This has also been noted in several other studies.38,43–45

However, the condition of monotonicity is not fulfilled for the (C–C–O) angle (Fig. S9 of the supplementary material) and for the (C–C–Heq) angle (not shown); the extrapolation procedure fails for these two structure parameters. The CCSD/pCVTZ and pCVQZ data columns of Table V indicate the approximate uncertainty of these two angles. Demaison et al. have noted similar problems when predicting semi-experimental bond angles and especially dihedral angles at the MP2/cc-pVTZ level.23 

In Table V, we list the extrapolated semi-experimental bond lengths and angles. The last two columns compare Raman RCS-based values to the gas-phase electron diffraction structure parameters rα and ra that are taken from Ref. 5. These T = 294 K thermally averaged values5 cannot be directly compared to the equilibrium structure parameters, since chemical bonds expand with increasing temperature and vibrational excitation.64,65 Using CFOUR and the CCSD(T)/cc-pCVDZ cubic force field, we computationally converted the semi-experimental re values to the ra(294 K) values, which are listed in the fifth data column of Table V. The uncertainties of the Raman RCS-based ra values contain only the uncertainties of the semi-experimental re parameters, the errors of the computed force field are not known. The calculations predict that all bond lengths and angles increase with increasing temperature, except for the H–C–H bond angle and the flap angle. The semi-experimental Raman RCS ra(C–H) bond lengths and bond angles are in good agreement with the GED values. However, the corresponding ra(C–C) and ra(C–O) bond lengths differ significantly, being 0.012 Å and 0.004 Å longer than the GED values; the flap angle differs by 0.9°. While there might be contributions to these differences that arise from the CCSD(T)/cc-pCVDZ computed cubic force field, we believe that these differences are significant.

The accurate semi-experimental structure of p-dioxane derived here allows interesting comparisons to the isoelectronic molecules piperidine and cyclohexane, whose equilibrium geometries are also chair-shaped. Their equilibrium semi-experimental structures have been determined by Demaison et al. based on microwave spectroscopic data for piperidine23,35 and deuterated cyclohexane isotopomers31,32 using a mixed estimation method.23 The semi-experimental equilibrium structure of cyclohexane has also been determined by combining Raman RCS of the h12 and −d12 isotopomers with CCSD(T) calculations and basis-set interpolation.33,34

We first compare the semiexperimental C–C, C–N, and C–O bond lengths: The re(C–C) bond distance of cyclohexane determined by Raman RCS is 1.5260 Å,33,34 in excellent agreement with the 1.5258(6) Å value from microwave spectroscopy.23,31,32 The re(C–N) bond distance in piperidine is 1.4589(2) Å,23 and the re(C–O) distance in p-dioxane is 1.4168(4) Å. The bond length decreases by 0.067 Å in piperidine and by 0.109 Å in p-dioxane relative to cyclohexane can be explained by the contraction of the valence orbitals of N and O due to the increased nuclear charge.23 

When comparing the C–C bond distances along this series, one observes that the C2–C3 (and C5–C6) bond lengths in piperidine decrease to 1.5212(2) Å,23 and the C2–C3 (and C5–C6) bonds in p-dioxane decrease to 1.5135(3) Å, corresponding to contractions by 0.005 and 0.013 Å relative to cyclohexane. We interpret this as arising from a slight shrinkage of the sp3-orbitals of the carbon atoms that are involved in the polar C–N and C–O bonds. That the shrinkage is larger in p-dioxane than in piperidine is in line with the larger polarity of the C–O bonds.

When comparing C–H bond lengths, we note that already in cyclohexane the semi-experimental re(C–H)ax = 1.096 Å is 0.004 Å longer than the equatorial C–H bond length re(C–H)eq.23,31–34 A similar difference between the axial and equatorial C–H bond distances is found at the C4 atom of piperidine (the C atom farthest from the nitrogen). However, in piperidine, the axial and equatorial C–H bond lengths at the C2 and C5 atoms differ by 0.012 Å. In p-dioxane, the four axial C–H bond lengths are 0.008 Å longer than the four equatorial C–H bond lengths. In piperidine, these differences in equatorial and axial C–H bond lengths have been rationalized in terms of the trans-effect,23 in which a lone-pair (lp) at a heteroatom X that is chemically bound to a C–H group can delocalize into an antibonding σ* orbital of this C–H bond, if the latter is oriented trans to this lone-pair (i.e., trans H–C–X–lp). In pyranose sugars, this is also referred to as the anomeric (or Edward-Lemieux) effect.66–68 

We measured the femtosecond Raman rotational coherence transients of p-dioxane-h8 and p-dioxane-h8 in a gas-cell at T = 293 K. For p-dioxane-h8, Raman RCS measurements were performed in a supersonic-jet experiment, allowing to cool the sample to TrotTvib130 K. Fitting the experimental RCS transients to an asymmetric-top model allowed to determine the v=0 ground-state rotational constants for p-dioxane-h8 and p-dioxane-d8. The RCS transient of p-dioxane-h8 exhibits a number of A-, C-, and K-type asymmetry coherences that were analyzed using a sub-transient decomposition method,60 allowing to separate coherences with specific rotational quantum number changes (ΔJ=0,1,2 combined with ΔK=0,2).

All of the coherences in the Raman RCS transient could be assigned to the most stable chair conformer of p-dioxane. No signals of the twist-boat conformer were observed by Raman RCS, confirming its very low concentration, in agreement with previous investigations using gas-phase electron diffraction.1,5

Perdeuteration of p-dioxane-h8 changes the inertial tensor from a strongly asymmetric top (κ=0.66) to a near-oblate symmetric top (κ=0.89 for p-dioxane-d8). This increase of symmetry leads to (1) a smaller number of asymmetry transients for the d8-isotopomer as compared to the h8 species, (2) an increase in the delay time over which recurrences can be observed, i.e., 0–900 ps for p-dioxane-d8 as compared to 0–600 ps for p-dioxane-h8, and (3) to a decrease of the information content of the RCS spectrum, which results in 5–10 times larger uncertainties of the A0 and B0 constants of p-dioxane-d8 relative to those of p-dioxane-h8.

Combining the experimentally determined rotational constants with CCSD(T)/cc-pCVXZ (X = D,T,Q) calculations allowed to determine accurate semi-experimental values for eight of the ten equilibrium structure parameters (bond lengths and angles). The exceptions are the C–C–O and C–C–Heq bond angles, which do not change monotonically when increasing the basis set from double- to quadruple-zeta. The semi-experimental equilibrium re(C–C) bond length is 1.5135(3) Å, which is 0.013(1) Å shorter than the value 1.526(1) Å determined for cyclohexane.33,34 The four axial C–H bond lengths are determined to be re(C–H)ax = 1.096 Å, which is 0.008 Å longer than the four equatorial C–H bond lengths. This difference is rationalized in terms of trans-effect, which is also known as the anomeric effect in the context of pyranose sugars.66–68 

We computationally converted the semi-experimental re values to the corresponding ra values at T = 294 K for comparison to the gas-phase electron diffraction results. While most values are in good agreement with those determined by gas-phase electron diffraction,5 the thermally averaged ra(C–C) = 1.512(2) Å measured by GED does not agree well with ra(CC) = 1.522 Å that we compute based on our semi-experimental re = 1.5135(3) Å value and the CCSD(T)/cc-pCVDZ cubic force field. We also note a 0.004 Å discrepancy between the computed/semi-experimental C–O distance ra = 1.4232(4) Å and the GED ra value.5 

See supplementary material for an energy-level diagram of asymmetric top molecules, rotational Raman transient types, Z-matrices, and CCSD(T) calculated structure parameters of the chair and twist-boat forms of p-dioxane, and eight figures with additional semi-experimental structure parameter extrapolations.

This work has been supported by the Schweizerische Nationalfonds (SNSF) through Grant Nos. 200020-144490 and 200020-130376.

1.
M.
Davis
and
O.
Hassel
,
Acta Chem. Scand.
17
,
1181
(
1963
).
2.
F. A. L.
Anet
and
J.
Sandstrom
,
J. Chem. Soc. D
23
,
1558
(
1971
).
3.
F. R.
Jensen
and
R. A.
Neese
,
J. Am. Chem. Soc.
93
,
6329
(
1971
).
4.
J.
Buschmann
,
E.
Mueller
, and
P.
Luger
,
Acta Crystallogr., Sect. C: Cryst. Struct. Commun.
42
,
873
(
1986
).
5.
M.
Fargher
,
L.
Hedberg
, and
K.
Hedberg
,
J. Mol. Struct.
1071
,
41
(
2014
).
6.
P. M.
Felker
and
A. H.
Zewail
,
J. Chem. Phys.
86
,
2460
(
1987
).
7.
P. M.
Felker
,
J. Phys. Chem.
96
,
7844
(
1992
).
8.
P. M.
Felker
and
A. H.
Zewail
, “
Molecular structures from ultrafast coherence spectroscopy
,” in
Femtosecond Chemistry
, edited by
J.
Manz
and
L.
Wöste
(
VCH
,
Weinheim
,
1995
), Chap. 5, Vol. I.
9.
H.-M.
Frey
,
P.
Beaud
,
T.
Gerber
,
B.
Mischler
,
P. P.
Radi
, and
A. P.
Tzannis
,
Appl. Phys. B
68
,
735
(
1999
).
10.
H. M.
Frey
,
P.
Beaud
,
T.
Gerber
,
B.
Mischler
,
P. P.
Radi
, and
A. P.
Tzannis
,
J. Raman Spectrosc.
31
,
71
(
2000
).
11.
C.
Riehn
,
A.
Weichert
, and
B.
Brutschy
,
J. Phys. Chem. A
105
,
5618
(
2001
).
12.
H. M.
Frey
,
P.
Beaud
,
T.
Lang
, and
M.
Motzkus
, “
High resolution femstosecond CARS spectroscopy
,” in
Femtochemistry and Femtobiology: Ultrafast Dynamics in Molecular Science
, edited by
A.
Douhal
and
J.
Santamaria
(
World Scientific Publishing
,
Singapore
,
2002
), Chap. 1, p.
99
.
13.
H. M.
Frey
,
A.
Müller
, and
S.
Leutwyler
,
J. Raman Spectrosc.
33
,
855
(
2002
).
14.
P.
Hobza
,
C.
Riehn
,
A.
Weichert
, and
B.
Brutschy
,
Chem. Phys.
283
,
331
(
2002
).
15.
V. V.
Matylitsky
,
W.
Jarzeba
,
C.
Riehn
, and
B.
Brutschy
,
J. Raman Spectrosc.
33
,
877
(
2002
).
16.
W.
Jarzȩba
,
V. V.
Matylitsky
,
C.
Riehn
, and
B.
Brutschy
,
Chem. Phys. Lett.
368
,
680
(
2003
).
17.
C.
Riehn
,
V. V.
Matylitsky
,
W.
Jarzeba
,
B.
Brutschy
,
P.
Tarakeshwar
, and
K.
Kim
,
J. Am. Chem. Soc.
125
,
16455
(
2003
).
18.
D. S.
Kummli
,
H. M.
Frey
,
M.
Keller
, and
S.
Leutwyler
,
J. Chem. Phys.
123
,
054308
(
2005
).
19.
C.
Riehn
,
V. V.
Matylitsky
,
M. F.
Gelin
, and
B.
Brutschy
,
Mol. Phys.
103
,
1615
(
2005
).
20.
P.
Brown
,
P.
Davies
,
G. M.
Hansford
, and
N. A.
Martin
,
J. Mol. Spectrosc.
158
,
468
(
1993
).
21.
J.
Demaison
,
Mol. Phys.
105
,
3109
(
2007
).
22.
J.
Demaison
,
H. D.
Rudolph
, and
A. G.
Császár
,
Mol. Phys.
111
,
1539
(
2013
).
23.
J.
Demaison
,
N. C.
Craig
,
P.
Groner
,
P.
Ecija
,
E. J.
Cocinero
,
A.
Lesarri
, and
H. D.
Rudolph
,
J. Phys. Chem. A
119
,
1486
(
2015
).
24.
D. S.
Kummli
,
H. M.
Frey
, and
S.
Leutwyler
,
J. Chem. Phys.
124
,
144307
(
2006
).
25.
D. S.
Kummli
,
H. M.
Frey
, and
S.
Leutwyler
,
Chimia
60
,
212
(
2006
).
26.
D. S.
Kummli
,
H. M.
Frey
, and
S.
Leutwyler
,
J. Phys. Chem. A
111
,
11936
(
2007
).
27.
D. S.
Kummli
,
S.
Lobsiger
,
H. M.
Frey
,
S.
Leutwyler
, and
J. F.
Stanton
,
J. Phys. Chem. A
112
,
9134
(
2008
).
28.
W.
Caminati
,
A.
Dell’Erba
,
S.
Melandri
, and
P. G.
Favaro
,
J. Am. Chem. Soc.
120
,
5555
(
1998
).
29.
H. M.
Pickett
and
H. L.
Strauss
,
J. Am. Chem. Soc.
92
,
7281
(
1970
).
30.
D. M.
Chapman
and
R. E.
Hester
,
J. Phys. Chem. A
101
,
3382
(
1997
).
31.
J.
Dommen
,
T.
Brupbacher
,
G.
Grassi
, and
A.
Bauder
,
J. Am. Chem. Soc.
112
,
953
(
1990
).
32.
E.
Białkowska-Jaworska
,
M.
Jaworski
, and
Z.
Kisiel
,
J. Mol. Struct.
350
,
247
(
1995
).
33.
G.
Brügger
,
H.-M.
Frey
,
P.
Steinegger
,
F.
Balmer
, and
S.
Leutwyler
,
J. Phys. Chem. A
115
,
9567
(
2011
).
34.
G.
Brügger
,
H.-M.
Frey
,
P.
Steinegger
,
P.
Kowalewski
, and
S.
Leutwyler
,
J. Phys. Chem. A
115
,
12380
(
2011
).
35.
H.
Ehrlichmann
,
J.-U.
Grabow
,
H.
Dreizler
,
N.
Heineking
, and
M.
Andolfatto
,
Z. Naturforsch., A
44
,
841
(
1989
).
36.
A. C.
Eckbreth
,
Appl. Phys. Lett.
32
,
421
(
1978
).
37.
D. S.
Kummli
,
H.-M.
Frey
, and
S.
Leutwyler
,
Chem. Phys.
367
,
36
(
2010
).
38.
T. S.
Den
,
H. M.
Frey
, and
S.
Leutwyler
,
J. Chem. Phys.
141
,
194303
(
2014
).
39.
C. G.
Vinson
and
J. J.
Martin
,
J. Chem. Eng. Data
8
,
74
(
1963
).
40.
D. E.
Woon
and
T. H.
Dunning
 Jr.
,
J. Chem. Phys.
103
,
4572
(
1995
).
41.
K. L.
Bak
,
J.
Gauss
,
P.
Jørgensen
,
J.
Olsen
,
T.
Helgaker
, and
J. F.
Stanton
,
J. Chem. Phys.
114
,
6548
(
2001
).
42.
C.
Puzzarini
,
M.
Heckert
, and
J.
Gauss
,
J. Chem. Phys.
128
,
194108
(
2008
).
43.
C.
Puzzarini
,
J. F.
Stanton
, and
J.
Gauss
,
Int. Rev. Phys. Chem.
29
,
273
(
2010
).
44.
C.
Puzzarini
,
Phys. Chem. Chem. Phys.
15
,
6595
(
2013
).
45.
M.
Hochlaf
,
C.
Puzzarini
, and
M.
Senent
,
Mol. Phys.
113
,
1671
(
2015
).
46.
J.
Gauss
and
J. F.
Stanton
,
Chem. Phys. Lett.
276
,
70
(
1997
).
47.
J. F.
Stanton
,
C. L.
Lopreore
, and
J.
Gauss
,
J. Chem. Phys.
108
,
7190
(
1998
).
48.
CFOUR, a quantum chemical program package written by
J. F.
Stanton
,
J.
Gauss
,
M. E.
Harding
,
P. G.
Szalay
, with contributions from
A. A.
Auer
,
R. J.
Bartlett
,
U.
Benedikt
,
C.
Berger
,
D. E.
Bernholdt
,
Y. J.
Bomble
,
L.
Cheng
,
O.
Christiansen
,
M.
Heckert
,
O.
Heun
,
C.
Huber
,
T.-C.
Jagau
,
D.
Jonsson
,
J.
Jusélius
,
K.
Klein
,
W. J.
Lauderdale
,
D. A.
Matthews
,
T.
Metzroth
,
D. P.
O’Neill
,
D. R.
Price
,
E.
Prochnow
,
K.
Ruud
,
F.
Schiffmann
,
W.
Schwalbach
,
S.
Stopkowicz
,
A.
Tajti
,
J.
Vázquez
,
F.
Wang
, and
J. D.
Watts
, and the integral packages MOLECULE (
J.
Almlöf
and
P. R.
Taylor
), PROPS (
P. R.
Taylor
), ABACUS (
T.
Helgaker
,
H. J. A.
Jensen
,
P.
Jørgensen
, and
J.
Olsen
), and ECP routines by
A. V.
Mitin
and
C.
van Wüllen
, for the current version, see http://www.cfour.de.
49.
S.
Mukamel
,
Principles of Nonlinear Spectroscopy
(
Oxford University Press
,
New York
,
1995
).
50.
B.
Grimberg
,
V.
Lozovoy
,
M.
Dantus
, and
S.
Mukamel
,
J. Phys. Chem. A
106
,
697
(
2002
).
51.
E. J.
Brown
,
Q.
Zhang
, and
M.
Dantus
,
J. Chem. Phys.
110
,
5772
(
1999
).
52.
G.
Placzek
and
E.
Teller
,
Z. Phys.
81
,
209
(
1933
).
53.
J. J.
Barrett
and
A.
Weber
,
J. Mol. Spectrosc.
50
,
205
(
1974
).
54.
W.
Gordy
and
R. L.
Cook
,
Microwave Molecular Spectra
, 3rd ed. (
Wiley-Interscience
,
New York
,
1984
).
55.
J. K. G.
Watson
,
Mol. Phys.
15
,
479
(
1968
).
56.
C. H.
Townes
and
A. L.
Schawlow
,
Microwave Spectroscopy
, reprint edition (
Dover Publications
,
New York
,
1975
).
57.
D. A.
Long
,
Raman Spectroscopy
(
McGraw-Hill International Book Company
,
1977
).
58.
H. W.
Kroto
,
Molecular Rotation Spectra
(
Dover Publications, Inc.
,
New York
,
1992
).
59.
D. A.
Long
,
The Raman Effect: A Unified Treatment of the Theory of Raman Scattering by Molecules
(
John Wiley & Sons, Ltd
,
2002
).
60.
T. S.
Den
,
H. M.
Frey
,
P. M.
Felker
, and
S.
Leutwyler
,
J. Chem. Phys.
143
,
144306
(
2015
).
61.
P.
Joireman
,
L. C.
Connell
,
S. M.
Ohline
, and
P. M.
Felker
,
J. Chem. Phys.
96
,
4118
(
1992
).
62.
G.
Herzberg
,
Molecular Spectra and Molecular Structure: III. Electronic Spectra and Electronic Structure of Polyatomic Molecules
(
D. Van Nostrand Company
,
1966
).
63.
P.
Kowalewski
,
H.-M.
Frey
,
D.
Infanger
, and
S.
Leutwyler
,
J. Phys. Chem. A
119
,
11215
(
2015
).
64.
Accurate Molecular Structures: Their Determination and Importance
, edited by
A.
Domenicano
and
I.
Hargittai
(
Oxford University Press
,
1992
).
65.
J. R.
Durig
,
Equilibrium Structural Parameters
(
Elsevier
,
Amsterdam
,
2000
), Vol. 24.
66.
J. T.
Edward
and
I.
Puskas
,
Can. J. Chem.
40
,
711
(
1961
).
67.
R. U.
Lemieux
and
A. R.
Morgan
,
Can. J. Chem.
43
,
2205
(
1965
).
68.
S.
Wolfe
,
A.
Rauk
,
L. M.
Tel
, and
I. G.
Czismadia
,
J. Chem. Soc. B
0
,
136
(
1971
).

Supplementary Material