Rotational conformers and nuclear spin isomers of carbonyl diisothiocyanate

Eva Gougoula *a, Jonathan Pfeiffer b, Melanie Schnell *ac and Frank Tambornino *b
aDeutsches-Elektronen Synchrotron DESY, Notkestr. 85, 22607 Hamburg, Germany. E-mail: eva.gougoula@desy.de; melanie.schnell@desy.de
bFachbereich Chemie, Philipps-Universität Marburg, Hans-Meerwein-Straße 4, 35043 Marburg, Germany. E-mail: tamborni@chemie.uni-marburg.de
cInstitut für Physikalische Chemie, Christian-Albrechts-Universität zu Kiel, 24118 Kiel, Germany

Received 30th May 2024 , Accepted 16th September 2024

First published on 18th September 2024


Abstract

Nuclear spin isomers of molecules play a pivotal role in our understanding of quantum mechanics and can have significant implications for various fields. In this work, we report the isolation and characterization of stable nuclear spin isomers as well as conformational isomers of a reactive compound, namely carbonyl diisothiocyanate. It can exist as three rotational conformers, two of which, the synsyn and synanti, were observed in a pulsed supersonic jet by chirped pulse Fourier transform microwave spectroscopy in the 2–12 GHz frequency region. The rotational spectra of two distinct nuclear spin isomers of synsyn-carbonyl diisothiocyanate, ortho and para, were recorded and analyzed. Experimental molecular rotational parameters for the identified rotational and nuclear spin isomers were determined, including rotational constants, centrifugal distortion constants, and nuclear quadrupole coupling constants. The two nuclear spin isomers are distinguished by unique hyperfine splitting signatures in their rotational spectra as an outcome of their different nuclear spin states. The relative abundances of the two observed conformers in the gas phase were estimated from the intensity of their rotational transitions. Following detection of singly substituted rare isotopologues of the synsyn conformer, a partial substitution (rs) structure was determined.


Introduction

Molecules with equivalent nuclei with non-zero spin can exist as nuclear spin isomers which are characterized by different sets of rotational quantum numbers that they are allowed to populate.1 Nuclear spin isomers have a pivotal role in the manifestation of quantum mechanics and our understanding of fundamental science as well as across various disciplines. To name a few, the unique properties of nuclear spin isomers may find potential applications in quantum computing,2,3 medical imaging,4,5 and astrophysics where the measured ratios between different nuclear spin isomers can provide an insight into astrophysical processes and stellar environments.6–8 The first observation of nuclear spin isomers was reported in 19299 with the separation and interconversion of ortho- and para-H2, where the orientation of the spin of each hydrogen can be parallel or anti-parallel, respectively. Separation and conversion of nuclear spin isomers was later observed for a number of other molecules, e.g. water,10,11 methanol,12 ethylene,13–15 and fluoromethane,16,17 to name a few. However, in the gas phase, the interconversion between isomers is generally considered improbable, and thus spectroscopic studies in the gas phase can be an ideal way to study isolated nuclear spin isomers.

Advances in modern synthetic chemistry play a key role in engineering molecules with unique symmetry properties that can serve as prototypes to study the properties of isolated nuclear spin isomers. Carbonyl diisothiocyanate (CDIT, C[double bond, length as m-dash]O(NCS)2), a reactive compound that consists of a carbonyl (C[double bond, length as m-dash]O) and two isothiocyanate groups (–N[double bond, length as m-dash]C[double bond, length as m-dash]S), was first reported in 1902,18 and reliable synthetic routes to its formation were established in 198119,20 and reinvestigated in 2021.21 In principle, the compound can exist as three different rotational conformers (Fig. 1), synsyn, synanti, and antianti (with respect to the carbonyl bond relative to the heterocumulene). Additionally, it consists exclusively of non-hydrogen atoms, all of which have a nuclear spin (I) of 0 or 1, and thus are bosons. In the synsyn and antianti forms, due to the overall symmetry of the molecules, the atoms of the –NCS groups are equivalent upon exchange and Bose–Einstein statistics are expected to influence their rotational spectra.


image file: d4cp02226b-f1.tif
Fig. 1 Possible conformers of carbonyl dipseudohalides (synsyn, synanti, and antianti) illustrated by the example of carbonyl diisothiocyanate. syn: carbonyl group and isothiocyanate group on the same side, anti: carbonyl group and isothiocyanate group on different sides.

Information on the structure and relative conformational abundances of carbonyl diisocyanate (C[double bond, length as m-dash]O(NCO)2)22,23 and carbonyl diazide (C[double bond, length as m-dash]O(N3)2)24 is available through gas-phase electron diffraction (GED) and infrared (IR) spectroscopy. The two are closely related to CDIT, both in terms of the available conformations they can adopt, but also by possessing isoelectronic functional groups attached to the carbonyl, and displaying high reactivity. Both the synsyn and synanti forms have been characterized for C[double bond, length as m-dash]O(NCO)2 and C[double bond, length as m-dash]O(N3)2, with the synsyn form being the dominant species in both cases, an observation that is also supported by quantum chemical calculations. Condensed phase studies on CDIT with X-ray diffraction (XRD) only identified the synsyn conformer in the crystal, which is also the most stable form. An important step towards characterizing and rationalizing the properties of CDIT is to obtain spectroscopic insight into the synanti conformation, as well as to experimentally determine its conformational preferences.

In this work, we report the rotational spectrum of CDIT in the gas phase using chirped pulse Fourier transform microwave (CP-FTMW) spectroscopy. The synsyn and synanti conformations are detected, providing access to the molecular properties of the elusive conformer. Their relative abundances fall within the broader expectations based on similar systems and on their computed energies. Interesting effects are observed in the spectrum of the synsyn conformer consistent with the presence of two distinct nuclear spin isomers that follow nuclear spin statistics for bosons. The structure of synsyn-CDIT was partially determined through isotopic substitution of the atoms in the –NCS groups.

Results and discussion

The three possible conformers of CDIT (Fig. 1), synsyn, synanti, and antianti, arise through rotation of the isothiocyanate groups –N[double bond, length as m-dash]C[double bond, length as m-dash]S around the C2–N6 or C2–N3 bonds (see Fig. 1 for atom labeling), respectively. The calculated molecular parameters of the conformers of CDIT (Table 1) show that the synsyn and synanti conformers are near-prolate asymmetric tops. The potential energy surface connecting the three possible conformers is calculated at the PBEh-3c level of theory. The two conformers are separated by a high energy barrier (∼16 kJ mol−1), with synanti-CDIT lying approximately 4.5 kJ mol−1 higher than synsyn-CDIT, the global minimum. Antianti-CDIT is a near-oblate asymmetric top, and it is calculated to be approximately 9.5 kJ mol−1 higher in energy than synsyn-CDIT. All three conformers have a sizeable electric dipole moment component along the b-axis of inertia. In synsyn- and antianti-CDIT, due to symmetry, the C2 axis is aligned with the b-axis of inertia and the C[double bond, length as m-dash]O bond, which results in three equivalent pairs of atoms between the two isothiocyanate –N[double bond, length as m-dash]C[double bond, length as m-dash]S groups (Fig. 2).
Table 1 Calculated rotational parameters, including nuclear quadrupole coupling constants of the two 14N nuclei and relative energies of the three possible conformers of CDIT obtained at the B3LYP-D3(BJ)/aug-cc-pVTZ level
synsyn synanti antianti
a Ray's asymmetry parameter image file: d4cp02226b-t1.tif. b Relative energies calculated with respect to synsyn-CDIT at the B3LYP-D3(BJ)/aug-cc-pVTZ level. c Energy barriers between synsyn-CDIT and synanti-CDIT, and synanti-CDIT and antianti-CDIT obtained at the PBEh-3c level and reported with respect to the energetically lower of the two minima it separates, and shown with respect to the relative energies of the conformers at the same level of theory.
A e (MHz) 10[thin space (1/6-em)]335.6 2901.6 1425.6
B e (MHz) 443.7 612.0 1165.6
C e (MHz) 425.4 505.4 641.3
χ aa(N3) (MHz) 1.763 1.618 1.861
χ bb(N3)–χcc(N3) (MHz) 0.936 –1.030 0.823
χ aa(N6) (MHz) 1.763 0.245 1.861
χ bb(N6)–χcc(N6) (MHz) 0.936 0.742 0.823
|μa|, |μb|, |μc| (D) 0, 1.2, 0 0.4, 2.2, 0 0, 2.6, 0
κ –0.99 –0.91 0.34
ΔErelb (cm−1, kJ mol−1) 0 377.4, 4.5 791.9, 9.5
ΔEbarrierc (cm−1, kJ mol−1) 1305, 16 1190, 14



image file: d4cp02226b-f2.tif
Fig. 2 Optimized (re) geometries and relative energies of the three possible conformers of CDIT calculated at the B3LYP-D3(BJ)/aug-cc-pVTZ level.

Spectrum of synsyn-CDIT

The jet-cooled rotational spectrum of CDIT was collected over the 2–12 GHz frequency range, in segments of 2–8 and 8–12 GHz. Upon initial inspection of the 8–12 GHz region, the most intense feature is a pattern consistent with a b-type, Q-branch transition. The lines exhibit hyperfine splitting as expected for a molecule with one or two N nuclei with nuclear spin I = 1. Watson's S-reduction,25 as implemented in Western's PGOPHER,26 was used to perform the initial assignment of the spectrum, without consideration of the hyperfine splitting. The 11,0 ← 10,1 transition was identified at approximately 10[thin space (1/6-em)]615 MHz which gave an excellent initial value of the A0 rotational constant. A preliminary fit including b-type, Q-, P-, and R-branch transitions was performed and yielded the initial set of the A0, B0, and C0 rotational constants which were assigned to synsyn-CDIT, based on agreement (see Table S1, ESI) between the calculated and experimentally determined rotational constants.

Upon inclusion of the nuclear quadrupole coupling terms χaa(14N) and χbb(14N)–χcc(14N) in the fit, the experimental hyperfine splitting appears to vary. For some transitions, it appears to be consistent with one N nucleus (I = 1), while for others with two N (I1 = 1 and I2 = 1) nuclei. In order to rationalize which types of transitions appear to have splitting equivalent to one or two quadrupolar nuclei, two separate fits, Fit 1 and Fit 2, grouping the different types of transitions were performed with PGOPHER,26 and the experimentally determined spectroscopic parameters are summarized in Table S2 (ESI). The types of transitions, Q-, P-, and R-branch, with odd or even image file: d4cp02226b-t2.tif that were included in each fit are indicated in the last rows of Table S2 (ESI). The hyperfine splitting of some representative rotational transitions is showcased in the expanded part of Fig. 3. The two fits yield similar rotational constants which are within the standard deviation of each other, thus excluding the possibility that the variation in the hyperfine splitting is associated with a large amplitude motion within the molecule. This is further supported by the value of DJK, which is in the range of just a couple of kHz. Regarding the nuclear quadrupole coupling constants of the two fits, these are of similar magnitude, however, their signs are reversed as an outcome of only considering one coupling nucleus in Fit 1. These observations suggest that there are two nuclear spin isomers of synsyn-CDIT, ortho and para, associated with nuclear spin states of I = 0, 2 and I = 1, respectively, and nuclear spin statistics dictate the allowed and forbidden transitions. For the ortho isomer, energy levels with even values of image file: d4cp02226b-t3.tif have a non-zero statistical weight, and those with odd values a zero statistical weight, while for the para isomer the opposite is observed. A more detailed explanation on the spin states and the assignment of transitions to the ortho and para isomers will follow in the coming paragraphs.


image file: d4cp02226b-f3.tif
Fig. 3 Representative rotational transitions image file: d4cp02226b-t18.tif of ortho- and para-(synsyn-CDIT) carrying hyperfine splitting due to 14N nuclei. The splitting in the spectrum of the para isomer is consistent with one 14N nucleus, while that for the ortho isomer is analogous to two 14N nuclei. Some I′′, F′′ ← I′, F′ hyperfine transitions are annotated, respectively.

Following the work of Grubbs et al.,27 it is possible to perform a global fit that includes transitions of both ortho- and para-(synsyn-CDIT) using Pickett's SPFIT/SPCAT.28 Watson's S-reduced Hamiltonian25 as implemented in SPFIT was used in the form:

ĤTOT = ĤROT + ĤCD + ĤQ(N) + ĤSS(N)
where ĤROT, ĤCD, ĤQ, and ĤSS are the rotational, centrifugal distortion, nuclear quadrupole coupling, and nuclear spin–nuclear spin coupling terms, respectively. The high degree of similarity between the rotational and centrifugal distortion constants of the two nuclear spin isomers (see Table S2, ESI) allows for simultaneous fitting of the ĤROT and ĤCD terms to both isomers, yielding average values for these constants. The ĤQ term was fitted according to the coupling scheme IN3 + IN6 = Itot and J + Itot = F, which implies coupling between two equivalent quadrupolar nuclei. The results of this fit are summarized in Table 2 and are consistent with those of Table S2 (ESI). The nuclear spin–nuclear spin coupling term is inseparable from the nuclear quadrupole coupling term, and therefore determination of ĤQ is a linear combination of the two.29 As mentioned, there are two nuclear spin wavefunctions that describe the ortho isomer, however, the lines that belong to the I = 0 state exhibit significant overlap and blending with the I = 2 lines. In the current fit, only a handful of well-resolved and well-isolated lines of the I = 0 state were identified. The detailed operator codes that were used for this fit and a complete line list are given in the ESI.

Table 2 Final fits of the rotational spectra of synsyn and synanti-CDIT performed with SPFIT/SPCAT using Watson's S-reduction
synsyn-CDIT IN3 + IN6 = Itot synanti-CDIT
a A 0, B0, C0: experimentally determined rotational constants simultaneously determined for the ortho and para nuclear spin isomers of the synsyn-CDIT conformer, and for the synanti-CDIT conformer. b Numbers in parentheses are one standard deviation in units of the last significant figures. c D J , DJK, DK, d1: quartic centrifugal distortion constants simultaneously determined for the ortho and para nuclear spin isomers of the synsyn-CDIT conformer, and for the synanti-CDIT conformer, respectively. d χ aa and χbbχcc: nuclear quadrupole coupling constants. e Magnitude of the calculated dipole moment components. f a-/b-/c-Type transitions. Y and N correspond to “yes” and “no”, respectively, indicating whether the type of transition is detected. g Ray's asymmetry parameter image file: d4cp02226b-t4.tif. h Number of assigned rotational transitions, including hyperfine transitions. i Root mean square deviation of the fit.
A 0 (MHz) 11[thin space (1/6-em)]046.3581(24)b 2938.6797(15)
B 0 (MHz) 449.80178(41) 626.51962(71)
C 0 (MHz) 432.27083(56) 516.14194(50)
D J (kHz) 0.2008(52)
D JK (kHz) 1.281(38) −3.099(39)
D K (kHz) −20.74(16)
d 1 (kHz) −0.0779(19)
χ aa(N3)d (MHz) 1.9269(56) 1.745(11)
χ bb(N3)–χcc(N3) (MHz) −0.742(10) −0.634(14)
χ aa(N6) (MHz) 1.9269(56) 0
χ bb(N6)–χcc(N6) (MHz) −0.742(10) 0.881(17)
|μa|, |μb|, |μc|e (D) 0, 1.2, 0 0.4, 2.2, 0
a-/b-/c-Typef N/Y/N N/Y/N
κ −0.99 −0.91
N 177 161
σ RMS (kHz) 13.2 16.3


The assignment of transitions to the ortho and para nuclear spin isomers of synsyn-CDIT was based on arguments around nuclear spin statistics, appropriate for the exchange of multiple pairs of bosons. The C2v symmetry of this conformer implies that rotation about the b-axis, which coincides with the C2 axis, interchanges equivalent bosons. It holds that the total wavefunction:

 
ψtot = ψelec × ψvib × ψrot × ψns(1)
where ψelec, ψvib, ψrot, and ψns are the electronic, vibrational, rotational, and nuclear spin wavefunctions, respectively, is symmetric and of positive parity. Thus, upon C2 rotation, i.e., exchange of identical bosons, we have:
 
ψ(2, 1) = +ψ(1, 2)(2)
Considering that our experiment probes the vibronic ground state, the electronic and vibrational wavefunctions for synsyn-CDIT both have a positive parity. Therefore, the total parity of ψtot is determined by the product of ψrot × ψns such that:
 
image file: d4cp02226b-t5.tif(3)
where the symbols in parentheses indicate the parity. Upon a C2 rotation about the symmetry axis, the rotational wavefunction is multiplied by image file: d4cp02226b-t6.tif. The parity of ψrot depends on the sum of image file: d4cp02226b-t7.tif specific to each J′′ ← J′ rotational transition, such that:
 
image file: d4cp02226b-t8.tif(4)
For even image file: d4cp02226b-t9.tif the parity of ψrot is positive and for odd image file: d4cp02226b-t10.tif the parity of ψrot is negative. From eqn (3) and (4) it follows that ψns must be of positive parity (ortho-isomers) for even image file: d4cp02226b-t11.tif and of negative parity (para-isomers) for odd image file: d4cp02226b-t12.tif. The ratio between symmetric and antisymmetric nuclear spin wavefunctions is given by:
 
image file: d4cp02226b-t13.tif(5)
There are two symmetric and one antisymmetric spin wavefunctions, corresponding to two ortho-isomers (Itot = 0, 2), associated with even image file: d4cp02226b-t14.tif, and one para-isomer (Itot = 1), associated with odd image file: d4cp02226b-t15.tif.

Sufficient signal to noise ratio for several rotational transitions of the parent synsyn-CDIT allowed for detection of the singly substituted 34S, 13C, and 15N isotopologues in their natural isotopic abundances, 4.4%, 1.1%, and 0.4%, respectively. Detection of these isotopologues further confirms the assignment of the fitted rotational constants to synsyn-CDIT. The C2 axis of symmetry renders the atoms of the two isothiocyanate groups equivalent to each other which results in doubling of the signal intensity of the respective singly substituted isotopologue. It should be noted that it is possible to perform a global fit for each rare isotopologue, including all the available transitions. Single isotopic substitution reduces the C2v symmetry to Cs, meaning that the N, C, and S atoms are no longer equivalent. This holds as further evidence that the observed spectral behavior of the parent synsyn-CDIT isotopologue is an outcome of the presence of two distinct nuclear spin isomers.

The fitted spectroscopic parameters of the singly substituted 34S, 13C, and 15N isotopologues of synsyn-CDIT are collected in Table 3. To generate these parameters, a pseudo-C2v geometry is assumed. Attempts to treat the symmetry as Cs with non-equivalent N atoms prevents the fit from converging and generates unrealistically large nuclear quadrupole coupling constants in the fits for the 34S and 13C isotopologues. The high intensity of the spectrum of the 34S isotopologue allowed for the assignment of 138 transitions, and for the determination of the quartic centrifugal distortion constant DJK and nuclear quadrupole coupling constants χaa(14N) and χbb(14N)–χcc(14N). Unlike for the parent synsyn-CDIT isotopologue, the coupling scheme J + IN3 = F1 and IN6 + F1 = F was used to fit the nuclear quadrupole coupling constants of the 34S and 13C isotopologues. The DJK distortion constant in the 13C isotopologue fit was kept fixed to the value that was determined for the 34S isotopologue. Finally, due to the small number of available transitions of the 15N isotopologue, the DJK and χaa(14N) and χbb(14N)–χcc(14N) constants were both kept fixed to the values determined for the 34S isotopologue. It should be noted that only one quadrupolar nucleus is present for the 15N isotopologue due to the nuclear spin of 15N (I = 0). The experimentally determined A0, B0, and C0 rotational constants will be used in Section 4.5 to determine the atomic coordinates of the atoms in the isothiocyanate groups in synsyn-CDIT.

Table 3 Experimentally determined spectroscopic parameters of the singly substituted 34S, 13C, and 15N isotopologues of synsyn-CDIT in their natural isotopic abundances
34S 13C 15N
a Numbers in parentheses are one standard deviation in units of the last significant figures. b Two equivalent nitrogen atoms assumed in the fits of the 34S and 13C isotopologues. Only one quadrupolar nucleus is considered for the 15N isotopologue. c Number of rotational transitions included in the fit. d Root mean square deviation of the fit.
A 0 (MHz) 11[thin space (1/6-em)]042.9488(27)a 11[thin space (1/6-em)]029.0588(35) 11[thin space (1/6-em)]002.9140(90)
B 0 (MHz) 438.30798(69) 447.6461(18) 449.3334(98)
C 0 (MHz) 421.6377(14) 430.2687(18) 431.7588(90)
D JK (kHz) 1.86(24) [1.86] [1.86]
χ aa(N)b (MHz) 1.9870(80) 1.962(13) [1.9870]
χ bb(N)–χcc(N) (MHz) −0.751(12) −0.692(18) [−0.751]
N 138 66 6
σ RMS (kHz) 14.4 14.9 10.8


Spectrum of synanti-CDIT

Following the assignment of synsyn-CDIT and its singly substituted 34S, 13C, and 15N isotopologues, a large number of lines with hyperfine splitting remained unassigned. A second pattern, with significantly lower intensity than the one identified for synsyn-CDIT, consistent with Q-branch, b-type transitions exhibiting extensive hyperfine splitting was found. Using the calculated rotational constants for synanti-CDIT as a guide, the pattern was fitted to a set of rotational parameters and allowed for identification of more lines. The final spectroscopic parameters were generated with Pickett's SPFIT/SPCAT28 and are summarized in Table 2. The N atoms in synanti-CDIT are not equivalent, and individual χaa(14N) and χbb(14N)–χcc(14N) constants can be determined. The nuclear quadrupole coupling constants were fitted according to the J + IN3 = F1 and IN6 + F1 = F coupling scheme. A representative example of fitted transitions of synanti-CDIT showcasing the hyperfine splitting is given in Fig. 4. The lower intensity of the rotational lines of synanti-CDIT did not allow for detection of singly substituted isotopologues to determine an experimental structure. However, the agreement between experimental and calculated rotational constants, the detected b-type transitions and the significantly lower intensity of this species supports the assignment to synanti-CDIT.
image file: d4cp02226b-f4.tif
Fig. 4 Portion of the experimental spectrum of synanti-CDIT (upper black trace) showcasing the resolution of the hyperfine structure. The simulation (lower purple trace) based on fitted rotational parameters reproduces the hyperfine splitting at a rotational temperature of 0.5 K.

Relative abundances

It is immediately apparent that synsyn-CDIT is the most abundant species in this experiment. This observation is in line with previous condensed phase studies on CDIT21 and other related compounds,24,30–33 and further supported by DFT calculations identifying this conformer as the global minimum, followed by synanti-CDIT approximately 4.5 kJ mol−1 higher in energy. The relative abundances of synsyn and synanti-CDIT can be estimated from the relative intensities of their rotational transitions. According to the procedure described by Quesada-Moreno et al.,34 a set of three or four rotational transitions of the two species with the same image file: d4cp02226b-t16.tif quantum numbers are selected, and their intensities are normalized by the square of the respective electric dipole moment. To get a reliable insight into the relative abundances, it is key that the selected transitions fall within a relatively narrow frequency range, e.g. up to 2 GHz. However, even though synsyn and synanti-CDIT are conformers of the same compound, the experimentally determined A0 rotational constants differ by an order of magnitude. In our rotational spectra there are no assigned transitions that fulfill these requirements and therefore this method would give an unreliable result.

An alternative procedure to estimate the relative abundances of the two species is to use the simulation function of PGOPHER.26 The spectra of synsyn- and synanti-CDIT were simulated simultaneously at a rotational temperature of 0.5 K, using the calculated |μb| dipole moment of 1.2 and 2.2 D, respectively. The rotational temperature of 0.5 K was selected as it can reproduce the experimental intensities of rotational transitions of each species individually. It should be noted that the rotational partition function of each molecule at the given temperature is by default taken into account when the intensities are simulated with PGOPHER. Finally, the simulated abundance of synsyn-CDIT was set to 1 while the abundance of synanti-CDIT was adjusted until the relative intensities of the two simulations could reproduce those of the experimental spectrum. Following this procedure, it is estimated that synsyn-CDIT is approximately 8 to 10 times more abundant than synanti-CDIT under the given experimental conditions.

The estimated 8[thin space (1/6-em)]:[thin space (1/6-em)]1 to 10[thin space (1/6-em)]:[thin space (1/6-em)]1 abundance ratio between synsyn- and synanti-CDIT is consistent with the computed relative energies of the two at the B3LYP-D3(BJ)/aug-cc-pVTZ level. The two conformers are separated by an energetic barrier of 16 kJ mol−1 with the local synanti minimum lying 4.5 kJ mol−1 higher than the global synsyn-CDIT minimum. The antianti conformation is not identified in the spectrum collected under our experimental conditions. Despite the fact that antianti-CDIT possesses the largest |μb| electric dipole moment out of the three conformers, this conformation is energetically disfavored by lying 14 kJ mol−1 above the global minimum. The relative populations of the three conformers of CDIT can be calculated according to their Boltzmann distributions at the temperature of the sample during the experiment (75 °C/348 K), prior to collisional cooling due to supersonic expansion, and by considering their computed energies, relative to the global minimum. Using the expression image file: d4cp02226b-t17.tif where N/N0 is the abundance ratio of two conformations, ΔE is the computed relative energy, k is the Boltzmann constant, and T is the temperature of the sample, we calculate that 80% of CDIT is in the synsyn conformer, while the synanti and antianti conformers take up approximately 17% and 3%, respectively. Considering the double degeneracy of the synanti-CDIT, the relative abundance ratio of the two lowest energy conformers are approximately 80[thin space (1/6-em)]:[thin space (1/6-em)]34. The calculated abundances are broadly consistent with the estimated abundances from the rotational spectrum.

Previous studies have explored the conformational equilibrium of compounds structurally related to CDIT, both in the condensed and gas phase. A summary of their determined relative abundances as well as their barriers to interconversion are summarized in Table 4. Compounds with the general formula X(C[double bond, length as m-dash]O)NCO (X = F, Cl, Br) consist of a central carbonyl group, which is substituted by a halogen atom and a pseudohalide group. In relation to CDIT, those can also exist as different rotational conformers: syn and anti (see Fig. 1). Both IR and Raman spectroscopy for fluoro,31–33,35 chloro,32,36 and bromocarbonyl37 isocyanate show that their conformers are in an equilibrium state at room temperature. For X = F the syn conformer is with 75(12)%33 preferred, according to gas phase electron diffraction (GED), while for X = Cl, and X = Br the anti-conformer is the dominant form. The syn to anti ratio for chlorocarbonyl isocyanate calculated from GED is 25(8)[thin space (1/6-em)]:[thin space (1/6-em)]75(8).36

Table 4 An overview of the (syn–)syn to (syn–)anti ratio of CDIT and other related compounds, determined through different methods, alongside computed energy barriers
Compound (syn–)syn[thin space (1/6-em)]:[thin space (1/6-em)](syn–)anti ratio ΔE (kJ mol−1) Barrier (kJ mol−1) Method
a Ratio, computed energy difference, and barrier to interconversion between synsyn- and synanti-CO(N3)2 from ref. 24. b Ratio between synsyn- to synanti-CO(NCO)2 and energy difference and barrier from ref. 23 and 40. c synsyn- to synanti-CO(NCS)2 ratio determined in this work, computed energy difference and barrier to interconversion from this work and ref. 41. d Ratio from ref. 31, 33, 36 and 42, and computed energy and barrier from ref. 33, 36 and 42. e From ref. 36 and 42. f From ref. 37 the ratio is not explicitly determined, however, the anti-conformer is favoured. Decreased electronegativity of the halogen atom [χ(F) > χ(Cl) > χ(Br)] favors the adoption of an anti-orientation of the –NCO group. g From ref. 39.
CO(N3)2a 88[thin space (1/6-em)]:[thin space (1/6-em)]12 (∼7[thin space (1/6-em)]:[thin space (1/6-em)]1) 5–6.7 40.6 Matrix IR, GED
CO(NCO)2b 66[thin space (1/6-em)]:[thin space (1/6-em)]34 (∼2[thin space (1/6-em)]:[thin space (1/6-em)]1) 3.8(8) 12.6 GED
CDIT CO(NCS)2c 10[thin space (1/6-em)]:[thin space (1/6-em)]1 2.9–4.5 8–16 CP-FTMW
F(C[double bond, length as m-dash]O)NCOd 77[thin space (1/6-em)]:[thin space (1/6-em)]23 2.1–5.0 GED
Cl(C[double bond, length as m-dash]O)NCOe 25(8)[thin space (1/6-em)]:[thin space (1/6-em)]75(8) 1.7–6.7 GED
Br(C[double bond, length as m-dash]O)NCOf Mainly anti IR and Raman
Cl(C[double bond, length as m-dash]O)NCSg 84(6)[thin space (1/6-em)]:[thin space (1/6-em)]16(6) 1.3 5.9 GED
Cl(C[double bond, length as m-dash]O)SCNg Mainly syn 7.1 38.9 IR


Fluoro- and chlorocarbonyl isothiocyanate (X(C[double bond, length as m-dash]O)NCS (X = F, Cl))38 as well as their isomers, fluoro- and chlorocarbonyl thiocyanate (X(C[double bond, length as m-dash]O)SCN (X = F, Cl)),39 also show an equilibrium state in the gas and liquid phase. In the solid state, solely the thermodynamically more stable syn conformers of F(C[double bond, length as m-dash]O)NCS, F(C[double bond, length as m-dash]O)SCN, Cl(C[double bond, length as m-dash]O)SCN are present, as it was deduced from single-crystal XRD and solid-state Raman spectroscopy. The structure of Cl(C[double bond, length as m-dash]O)NCS was determined by GED and the conformer syn to anti ratio is 84(6)[thin space (1/6-em)]:[thin space (1/6-em)]16(6),39 being the only determined ratio. As already discussed, CDIT and its related dipseudohalides, carbonyl diisocyanate (CO(NCO)2) and carbonyl diazide (CO(N3)2), can in principle exist as three different conformers: synsyn, synanti, and antianti (see Fig. 1).22 Both CO(NCO)2 and CO(N3)2 show an equilibrium in the gas phase between their synsyn and synanti conformers with a ratio of 62[thin space (1/6-em)]:[thin space (1/6-em)]3840 and 88[thin space (1/6-em)]:[thin space (1/6-em)]1224 (values estimated from IR spectroscopy), respectively. For carbonyl diisocyanate the ratio was refined from GED and is 66(3)[thin space (1/6-em)]:[thin space (1/6-em)]34(3).23 These values fit well with the computed energy difference of 4.3 kJ mol−1 and 6.9 kJ mol−1 (B3LYP/6-311+G(3d,f) level of theory) between the synsyn- and synanti-conformers of CO(NCO)2 and CO(N3)2, respectively.23,24,40 For both compounds, the antianti conformer is not found spectroscopically, probably as it lies much higher in energy, as a result of steric repulsion between the pseudohalide groups both being anti oriented with respect to the carbonyl group. In the solid state, both crystallize with their thermodynamically stable rotamer, which is the synsyn conformer.23,24 Pfeiffer et al.41 discussed a correlation between the synsyn to synanti conformational ratio and the computed energy difference between conformers. The general trend suggests that a higher synsyn to synanti ratio is consistent with a higher energy difference between the two conformers. From Table 4, we also observe that a higher ratio is associated with a lower energy barrier to interconversion, suggesting a lower conversion rate to synanti, and the findings for CDIT fit well with this observation.

Nuclear quadrupole coupling constants

The nuclear quadrupole coupling constants determined for synsyn- and synanti-CDIT were diagonalized according to a common set of orthogonal x, y, z axes. These are positioned on each individual N nucleus, yielding the diagonalized χxx, χyy, and χzz constants summarized in Table 5. The diagonalization was performed with QDIAG,43 and a visualization of the orientation of the axes is shown in Fig. S1 in the ESI. The off-diagonal |χab| constants were obtained through DFT calculations at the B3LYP-D3(BJ)/aug-cc-pVTZ level and given a 10% uncertainty. The same procedure was performed for other related compounds with a N nucleus within a –NCS group as well as for compounds with a –NH2 group. The calculated values of |χab| are given in the footnotes of Table 5.
Table 5 Diagonalized nuclear quadrupole coupling constants χxx, χyy, and χzz for CDIT and other related compounds
χ xx (MHz) χ yy (MHz) χ zz (MHz)
a This work. Calculated off-diagonal components for synsyn-CDIT |χab(N3/N6)| = 1.30(13) MHz and for synanti-CDIT |χab(N3)| = 2.04(20) MHz and |χab(N6)| = 2.08(21) MHz at the B3LYP-D3(BJ)/aug-cc-pVTZ level. b Calculated off-diagonal component |χab| = 1.33(13) MHz at the B3LYP-D3(BJ)/aug-cc-pVTZ level and from ref. 46. c From ref. 44. d Calculated off-diagonal component |χab| = 2.09(12) MHz at the B3LYP-D3(BJ)/aug-cc-pVTZ level and from ref. 47. e Calculated off-diagonal component |χab| = 0.215(22) MHz at the B3LYP-D3(BJ)/aug-cc-pVTZ level and from ref. 45.
synsyn-CDITa N3/N6 −1.790(81) 2.382(81) −0.5924(60)
synanti-CDITa N3 −2.24(16) 2.79(16) −0.554(12)
N6 −1.87(21) 2.31(21) −0.441(13)
Ethoxycarbonyl isothiocyanateb N −1.685(86) 2.338(86) −0.6532(25)
Ph–NCSc N −1.4635 1.94656 −0.483069
Anilined N 4.20(13) 0 −4.20(6)
Ureae N 2.311(18) 1.778(18) −4.0889(26)


From the values in Table 5, we observe that the local electronic environment of a N nucleus in a –NCS group displays significant differences to that of a –NH2 group, e.g., in phenyl isothiocyanate (Ph-NCS)44 and urea.45 These differences can be partially rationalized by looking at the chemical environment around each N nucleus. In the case of the molecules with one or more –NCS groups, the N atom exhibits sp hybridization in a linear arrangement with the neighboring atoms, while the N atom in –NH2 is sp3 hybridized in a slightly pyramidal configuration. Comparing synsyn- and synanti-CDIT to the structurally related urea, the magnitude of |χzz|, which in a planar or almost planar molecule is approximately equal to |χcc|, provides an indication of the degree of delocalization of the electron pair on the N nucleus. The –NCS group can contribute to resonance forms and, in combination with the electron withdrawing character of the group and the higher electronegativity of the S atom, the lone pair of N exhibits a higher degree of delocalization. The more localized electron pair in –NH2 is reflected by the significantly larger magnitude of |χzz|. The closely related molecules to CDIT, ethoxycarbonyl isothiocyanate46 and Ph–NCS,44 display a similar electronic environment around the N nucleus suggesting that the component attached to the –NCS group has little effect on the local electronic environment of the N nucleus. Synsyn and synanti-CDIT exhibit similar electronic environments around the N nuclei, suggesting that the reactivity of the molecule is not exclusively associated with a certain conformer. However, following the earlier arguments around the relative abundances, the synsyn form is expected to partake in most reactions.

Structural analysis

Experimental determination of the rotational constants of the parent synsyn-CDIT (Table 2) and its singly substituted 34S, 13C, and 15N isotopologues (Table 3) allows for Kraitchman48,49 analysis to be performed to obtain the substitution (rs) coordinates of the atoms in the –NCS groups. The analysis was performed with the program KRA from the PROSPE website.43 The resulting rs coordinates, alongside their fractional uncertainties50 to partially account for rotation–vibration effects, are shown in Table 6. The rs method can only provide the magnitude of each coordinate, so the signs were inferred from the calculated DFT geometry. The |c| coordinates were calculated as zero, consistent with planarity of the molecule, which is also supported by the value of the inertia defect51Δ0 = IcIaIb = –0.182 u Å2, where Ix corresponds to the respective moment of inertia for each axis. The magnitude and the sign of Δ0 is consistent with low-lying out-of-plane motions of the –NCS groups at the zero point. The rs coordinates were evaluated with EVAL43 to yield bond lengths and angles relevant to the –NCS groups and are summarized in Table 6.
Table 6 Atomic substitution (rs) coordinates of synsyn-CDIT determined experimentally alongside the calculated (re) coordinates, and structural parameters determined with XRD and CP-FTMW spectroscopy
synsyn-CDIT |a|a (Å) b (Å)
a Only the magnitude of the a-coordinates is provided here. Due to the C2v symmetry, to describe the positions of the atoms in either one of the –NCS groups, the a-coordinates need to all carry a positive or negative sign, respectively.
N r e 1.1426 −0.4410
r s 1.087(12) −0.4273(35)
C r e 2.3366 −0.2639
r s 2.3302(12) −0.2690(56)
S r e 3.8920 −0.1624
r s 3.8687(13) −0.086(17)

XRD r s
C[double bond, length as m-dash]O (Å) 1.181(16)
Ccarbonyl–N (Å) 1.381(5)
N[double bond, length as m-dash]C (Å) 1.207(4) 1.253(12)
C[double bond, length as m-dash]S (Å) 1.526(4) 1.5497(28)
∠(SCN) (°) 173.7(4)–175.7(4) 179.5(8)


The experimentally determined rs coordinates are shown alongside the re calculated coordinates in Table 6 and Fig. 5. The high level of agreement between the two provides strong evidence in support of the geometry for synsyn-CDIT. The values of the rs bond lengths r(C[double bond, length as m-dash]S) and r(N[double bond, length as m-dash]C), and the angle ∠(SCN) are compared to those determined by single-crystal X-ray diffraction (XRD).21 A slight elongation of all determined bond lengths is observed for the gas phase structure while the ∠(SCN) angle shows an essentially linear arrangement in the atoms of the –NCS groups. The more compact structure as well as the slight distortion from linearity in the ∠(SCN) angle that is observed with XRD is likely due to packing effects in the crystal. It is worth noting that only the synsyn conformer crystallizes and therefore there is no structural information for the synanti conformer.21


image file: d4cp02226b-f5.tif
Fig. 5 The experimental rs coordinates (solid spheres) of synsyn-CDIT compared to the calculated B3LYP-D3(BJ)/aug-cc-pVTZ re coordinates (translucent underlying structure).

Conclusions

The rotational spectra of two conformers of the reactive compound carbonyl diisothiocyanate, synsyn and synanti, were recorded with chirped pulse Fourier transform microwave spectroscopy in the 2–12 GHz frequency region. We estimate that the synsyn form is approximately eight to ten times more abundant than the synanti form, which fits well with the observed relative abundances of conformers of other related compounds. The spectrum of synsyn-carbonyl diisothiocyanate was manifested as two distinct nuclear spin isomers, ortho and para, that exhibit characteristics of nuclear spin statistics for bosons, and were identified by the varying hyperfine structure of their rotational transitions. Observation of singly substituted isotopologues allowed for partial structure determination of synsyn-carbonyl diisothiocyanate and determination of the atomic coordinates of the atoms in the –NCS groups. The gas-phase structure of isolated synsyn-carbonyl diisothiocyanate shows a slight elongation of the bond lengths between the atoms of the –NCS groups when compared to the structure determined with XRD using crystals.

Carbonyl diisothiocyanate is a unique example of a molecule composed exclusively of eight bosons (I = 0 and 1), six of which, in the synsyn and antianti forms, appear as three equivalent pairs (–NCS groups). Synsyn-carbonyl diisothiocyanate is present as two nuclear spin isomers, ortho and para, which are distinguished by their nuclear quadrupole coupling signatures, as an outcome of differing symmetries in their nuclear states while their nuclear quadrupole coupling constants χaa and χbbχcc are the same for the two nuclear spin isomers, as seen in Table 3. Observation of ortho and para nuclear spin isomers in molecules that exhibit nuclear spin statistics for bosons have previously been reported for several systems, however, these mainly include small linear molecules52,53 (e.g., D2, N2) and asymmetric tops such as non-covalently bound complexes54–56 (e.g., CO⋯N2, N2⋯HCl, N2⋯OCS) where the C2 axis coincides with the a-axis of inertia. In terms of nuclear spin statistics, these complexes were shown to exhibit characteristics of their respective boson counterparts that exchange upon a C2 rotation about the symmetry axis, and the linkage between their nuclear spin isomers lies within the J and Ka quantum numbers. In the case of synsyn-carbonyl diisothiocyanate, the C2 axis is aligned with the b-axis of inertia, meaning that inclusion of the Kc quantum number is necessary to link the two isomers. Considering the lack of examples of complex molecules in the literature with nuclear spin isomers where the J, Ka, and Kc quantum numbers are required to properly describe them, synsyn-carbonyl diisothiocyanate makes for a noteworthy case.

Data availability

The data supporting this article have been included as part of the ESI. Lists of fitted transition frequencies as well as detailed operator codes used in the fits are summarized in the ESI.

Conflicts of interest

There are no conflicts to declare.

Acknowledgements

J. P. acknowledges Fonds der chemischen Industrie for a PhD scholarship. F. T. acknowledges funding through the German Research Council (DFG) TA 1357/5-1. The authors would like to thank Prof. Dr Daniel Obenchain for his help in constructing the input files for SPFIT/SPCAT. The authors would also like to thank Dr Denis Tikhonov for useful discussions around the NEB calculations.

References

  1. L. D. Landau and E. M. Lifshitz, Quantum mechanics: non-relativistic theory, Elsevier, 2013, vol. 3 Search PubMed.
  2. E. Moreno-Pineda, M. Damjanović, O. Fuhr, W. Wernsdorfer and M. Ruben, Nuclear Spin Isomers: Engineering a Et4N[DyPc2] Spin Qudit, Angew. Chem., Int. Ed., 2017, 56, 9915–9919 CrossRef CAS PubMed.
  3. A. Gaita-Ariño, F. Luis, S. Hill and E. Coronado, Molecular spins for quantum computation, Nat. Chem., 2019, 11, 301–309 CrossRef PubMed.
  4. S. S. Kaushik, Z. I. Cleveland, G. P. Cofer, G. Metz, D. Beaver, J. Nouls, M. Kraft, W. Auffermann, J. Wolber, H. P. McAdams and B. Driehuys, Diffusion-weighted hyperpolarized 129Xe MRI in healthy volunteers and subjects with chronic obstructive pulmonary disease, Magn. Reson. Med., 2011, 65, 1154–1165 CrossRef PubMed.
  5. J. R. Birchall, M. R. H. Chowdhury, P. Nikolaou, Y. A. Chekmenev, A. Shcherbakov, M. J. Barlow, B. M. Goodson and E. Y. Chekmenev, Pilot Quality-Assurance Study of a Third-Generation Batch-Mode Clinical-Scale Automated Xenon-129 Hyperpolarizer, Molecules, 2022, 27, 1327 CrossRef CAS PubMed.
  6. T. Hama, A. Kouchi and N. Watanabe, Statistical ortho-to-para ratio of water desorbed from ice at 10 kelvin, Science, 2016, 351, 65–67 CrossRef CAS PubMed.
  7. H. Kawakita, J. Watanabe, H. Ando, W. Aoki, T. Fuse, S. Honda, H. Izumiura, T. Kajino, E. Kambe, S. Kawanomoto, K. Noguchi, K. Okita, K. Sadakane, B. Sato, M. Takada-Hidai, Y. Takeda, T. Usuda, E. Watanabe and M. Yoshida, The Spin Temperature of NH3 in Comet C/1999S4 (LINEAR), Science, 2001, 294, 1089–1091 CrossRef CAS PubMed.
  8. J. E. Dickens and W. M. Irvine, The Formaldehyde Ortho/Para Ratio as a Probe of Dark Cloud Chemistry and Evolution, Astrophys. J., 1999, 518, 733–739 CrossRef CAS PubMed.
  9. K. F. Bonhoeffer and P. Harteck, Experimente über Para- und Orthowasserstoff, Naturwissenschaften, 1929, 17, 182 CrossRef CAS.
  10. D. A. Horke, Y.-P. Chang, K. Długołęcki and J. Küpper, Separating Para and Ortho Water, Angew. Chem., Int. Ed., 2014, 53, 11965–11968 CrossRef CAS PubMed.
  11. P. Cacciani, J. Cosléou and M. Khelkhal, Nuclear spin conversion in H2O, Phys. Rev. A: At., Mol., Opt. Phys., 2012, 85, 12521 CrossRef.
  12. Z.-D. Sun, M. Ge and Y. Zheng, Separation and conversion dynamics of nuclear-spin isomers of gaseous methanol, Nat. Commun., 2015, 6, 6877 CrossRef CAS PubMed.
  13. P. L. Chapovsky, J. Cosléou, F. Herlemont, M. Khelkhal and J. Legrand, Separation and conversion of nuclear spin isomers of ethylene, Chem. Phys. Lett., 2000, 322, 424–428 CrossRef CAS.
  14. Z.-D. Sun, K. Takagi and F. Matsushima, Separation and Conversion Dynamics of Four Nuclear Spin Isomers of Ethylene, Science, 2005, 310, 1938–1941 CrossRef CAS PubMed.
  15. P. L. Chapovsky, V. V. Zhivonitko and I. V. Koptyug, Conversion of Nuclear Spin Isomers of Ethylene, J. Phys. Chem. A, 2013, 117, 9673–9683 CrossRef CAS PubMed.
  16. B. Nagels, M. Schuurman, P. L. Chapovsky and L. J. F. Hermans, Nuclear spin conversion in molecules: experiments on 13CH3F support a mixing-of-states model, Phys. Rev. A: At., Mol., Opt. Phys., 1996, 54, 2050 CrossRef CAS PubMed.
  17. P. L. Chapovskii, Conversion of nuclear spin modifications of CH3F molecules in the gaseous phase, Sov. J. Exp. Theor. Phys., 1990, 70, 895 Search PubMed.
  18. A. E. Dixon, The action of metallic thiocyanates upon carbonyl chloride, Proc. R. Soc. London, 1902, 18, 235 Search PubMed.
  19. R. Bunnenberg and J. C. Jochims, Carbonyldiisothiocyanat, Chem. Ber., 1981, 114, 2075–2086 CrossRef CAS.
  20. R. Bunnenberg, J. C. Jochims and H. Härle, Zur Darstellung und Chlorierung von Carbonyl-diisothiocyanat, Chem. Ber., 1982, 115, 3587–3596 CrossRef CAS.
  21. J. Pfeiffer, C. Trost, A. Pachkovska and F. Tambornino, A Crystallographic, Spectroscopic, and Computational Investigation of Carbonyl and Oxalyl Diisothiocyanate, Inorg. Chem., 2021, 60, 10722–10728 CrossRef CAS PubMed.
  22. W. J. Balfour, S. G. Fougere, D. Klapstein and W. M. Nau, The infrared and Raman spectra of carbonyl diisocyanate, Spectrochim. Acta, Part A, 1994, 50, 1039–1046 CrossRef.
  23. T. M. Klapötke, B. Krumm, S. Rest, R. Scharf, J. Schwabedissen, H.-G. Stammler and N. W. Mitzel, Carbonyl Diisocyanate CO(NCO)2: Synthesis and Structures in Solid State and Gas Phase, J. Phys. Chem. A, 2016, 120, 4534–4541 CrossRef PubMed.
  24. X. Zeng, M. Gerken, H. Beckers and H. Willner, Synthesis and Characterization of Carbonyl Diazide, OC(N3)2, Inorg. Chem., 2010, 49, 9694–9699 CrossRef CAS PubMed.
  25. J. K. G. Watson, Determination of Centrifugal Distortion Coefficients of Asymmetric-Top Molecules. III. Sextic Coefficients, J. Chem. Phys., 2003, 48, 4517–4524 CrossRef.
  26. C. M. Western, PGOPHER: a program for simulating rotational, vibrational and electronic spectra, J. Quant. Spectrosc. Radiat. Transfer, 2017, 186, 221–242 CrossRef CAS.
  27. G. S. Grubbs II, D. A. Obenchain, H. M. Pickett and S. E. Novick, H2-AgCl: a spectroscopic study of a dihydrogen complex, J. Chem. Phys., 2014, 141, 114306 CrossRef PubMed.
  28. H. M. Pickett, The fitting and prediction of vibration-rotation spectra with spin interactions, J. Mol. Spectrosc., 1991, 148, 371–377 CrossRef CAS.
  29. R. F. Code and N. F. Ramsey, Molecular-Beam Magnetic Resonance Studies of HD and D2, Phys. Rev. A: At., Mol., Opt. Phys., 1971, 4, 1945–1959 CrossRef.
  30. D. W. Ball, Carbonyl diazide, OC(N3)2: calculated thermodynamic properties, Comput. Theor. Chem., 2011, 965, 176–179 CrossRef CAS.
  31. D. Klapstein and W. M. Nau, Conformational properties of carbonyl isocyanates—stereoelectronic effects favouring the cisoid conformation, J. Mol. Struct., 1993, 299, 29–41 CrossRef CAS.
  32. D. Klapstein and W. M. Nau, Spectroscopy of acyl and carbonyl isocyanates, Spectrochim. Acta, Part A, 1994, 50, 307–316 CrossRef.
  33. H.-G. Mack, C. O. Della Védova and H. Willner, Structures and conformations of carbonyl isocyanates and carbonyl azides. An experimental and theoretical investigation, J. Mol. Struct., 1993, 291, 197–209 CrossRef CAS.
  34. M. M. Quesada-Moreno, A. Krin and M. Schnell, Analysis of thyme essential oils using gas-phase broadband rotational spectroscopy, Phys. Chem. Chem. Phys., 2019, 21, 26569–26579 RSC.
  35. J. R. Durig, G. A. Guirgis, K. A. Krutules, H. Phan and H. D. Stidham, Raman and infrared spectra, conformational stability, barriers to internal rotation and ab initio calculations of fluorocarbonyl isocyanate, J. Raman Spectrosc., 1994, 25, 221–232 CrossRef CAS.
  36. H.-G. Mack, H. Oberhammer and C. O. Della Védova, How reliable are ab initio calculations? Experimental and theoretical investigation of the structure and conformation of chlorocarbonyl isocyanate, ClC(O)NCO, J. Mol. Struct.: THEOCHEM, 1989, 200, 277–288 CrossRef.
  37. C. O. D. Védova, Preparation and properties of bromocarbonyl isocyanate, BrC(O)NCO, Spectrochim. Acta, Part A, 1992, 48, 1179–1185 CrossRef.
  38. L. A. Ramos, S. E. Ulic, R. M. Romano, M. F. Erben, C. W. Lehmann, E. Bernhardt, H. Beckers, H. Willner and C. O. Della Védova, Vibrational Spectra, Crystal Structures, Constitutional and Rotational Isomerism of FC(O)SCN and FC(O)NCS, Inorg. Chem., 2010, 49, 11142–11157 CrossRef CAS PubMed.
  39. L. A. Ramos, S. E. Ulic, R. M. Romano, M. F. Erben, Y. V. Vishnevskiy, C. G. Reuter, N. W. Mitzel, H. Beckers, H. Willner, X. Zeng, E. Bernhardt, M. Ge, S. Tong and C. O. Della Védova, Spectroscopic Characterization and Constitutional and Rotational Isomerism of ClC(O)SCN and ClC(O)NCS, J. Phys. Chem. A, 2013, 117, 2383–2399 CrossRef CAS PubMed.
  40. Q. Liu, H. Li, Z. Wu, D. Li, H. Beckers, G. Rauhut and X. Zeng, Photolysis of Carbonyl Diisocyanate: Generation of Isocyanatocarbonyl Nitrene and Diazomethanone, Chem. – Asian J., 2016, 11, 2953–2959 CrossRef CAS PubMed.
  41. J. Pfeiffer, J. P. Wagner and F. Tambornino, Photolytic Decarbonylation of Oxalyl Diisothiocyanate in Solid Argon Matrices to synanti Carbonyl Diisothiocyanate and its Isomerization, Eur. J. Inorg. Chem., 2023, e202300290 CrossRef CAS.
  42. M. Tho Nguyen, M. R. Hajnal and L. G. Vanquickenborne, How reliable are ab initio calculations?, J. Mol. Struct.: THEOCHEM, 1991, 231, 185–193 CrossRef.
  43. Z. Kisiel, PROSPE – Programs for ROtational SPEctroscopy, 2024, https://info.ifpan.edu.pl/~kisiel/prospe.htm.
  44. W. Sun, W. G. D. P. Silva and J. van Wijngaarden, Rotational Spectra and Structures of Phenyl Isocyanate and Phenyl Isothiocyanate, J. Phys. Chem. A, 2019, 123, 2351–2360 CrossRef CAS PubMed.
  45. U. Kretschmer, D. Consalvo, A. Knaack, W. Schade, W. Stahl and H. Dreizler, The 14N quadrupole hyperfine structure in the rotational spectrum of laser vaporized urea observed by molecular beam Fourier transform microwave spectroscopy, Mol. Phys., 1996, 87, 1159–1168 CrossRef CAS.
  46. Y. Xu, W. Li, J. Zhang and G. Feng, Conformations and structures of ethoxycarbonyl isothiocyanate revealed by rotational spectroscopy, Chin. J. Chem. Phys., 2022, 35, 875–882 CrossRef CAS.
  47. A. Hatta, M. Suzuki and K. Kozima, Nuclear Quadrupole Effects in the Microwave Spectrum and Dipole Moment of Aniline, Bull. Chem. Soc. Jpn., 1973, 46, 2321–2323 CrossRef CAS.
  48. J. Kraitchman, Determination of Molecular Structure from Microwave Spectroscopic Data, Am. J. Phys., 1953, 21, 17–24 CrossRef CAS.
  49. H. D. Rudolph, Extending Kraitchman's equations, J. Mol. Spectrosc., 1981, 89, 430–439 CrossRef CAS.
  50. C. Costain, Further comments on the accuracy of rs substitution structures, Trans. Am. Crystallogr. Assoc., 1966, 2, 157–164 CAS.
  51. R. K. Bohn, J. A. Montgomery, H. H. Michels and J. A. Fournier, Second moments and rotational spectroscopy, J. Mol. Spectrosc., 2016, 325, 42–49 CrossRef CAS.
  52. F. G. Brickwedde, R. B. Scott and H. S. Taylor, The Difference in Vapor Pressures of Ortho and Para Deuterium, J. Chem. Phys., 1935, 3, 653–660 CrossRef CAS.
  53. S. Fleischer, I. S. Averbukh and Y. Prior, Selective Alignment of Molecular Spin Isomers, Phys. Rev. Lett., 2007, 99, 93002 CrossRef PubMed.
  54. C. Xia, A. R. W. McKellar and Y. Xu, Infrared spectrum of the CO–N2 van der Waals complex: assignments for CO-paraN2 and observation of a bending state for CO-orthoN2, J. Chem. Phys., 2000, 113, 525–533 CrossRef CAS.
  55. R. S. Altman, M. D. Marshall and W. Klemperer, The microwave spectrum and molecular structure of N2–HCl, J. Chem. Phys., 1983, 79, 57–64 CrossRef CAS.
  56. J. P. Connelly, S. P. Duxon, S. K. Kennedy, B. J. Howard and J. S. Muenter, Hyperfine and Tunneling Effects in the Microwave Spectrum of N2–OCS, J. Mol. Spectrosc., 1996, 175, 85–98 CrossRef CAS.

Footnotes

Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d4cp02226b
Single isotopic substitution reduces the symmetry of synsyn-CDIT from C2v to Cs, making the atoms in the –N[double bond, length as m-dash]C[double bond, length as m-dash]S groups non-equivalent. Pseudo-C2v refers to the treatment of the Cs as C2v for the sake of obtaining a fit that can converge with reasonable values of nuclear quadrupole coupling constants.

This journal is © the Owner Societies 2024
Click here to see how this site uses Cookies. View our privacy policy here.