research papers
Protonated oxalyl chloride and the ClCO+ cation
aDepartment Chemie, Ludwig-Maximilians Universität, Butenandtstrasse 5-13 (Haus D), D-81377 München, Germany
*Correspondence e-mail: sebastian.steiner@cup.uni-muenchen.de
The reactions of oxalyl chloride were investigated in the binary superacidic systems HF/SbF5 and DF/SbF5. O-Monoprotonated oxalyl chloride was isolated and represents the first example of a protonated acyl chloride. Diprotonated oxalyl chloride is only stable in solution. Salts of the ClCO+ cation were synthesized from the reactions of oxalyl chloride or COClF with SbF5 in 1,1,1,2-tetrafluoroethane (R-134a, CF3CFH2). The colourless salts were characterized by low-temperature vibrational spectroscopy, NMR spectroscopy and single-crystal X-ray diffraction. (1,2-Dichloro-2-oxoethylidene)oxidanium hexafluoridoantimonate(V), [C2O(OH)Cl2][SbF6], crystallizes in the monoclinic P21 and carbonyl chloride hexadecafluoridotriarsenate(V) [ClCO][Sb3F16], in the trigonal P31, with two and three formula units per respectively. Monoprotonated oxalyl chloride and the ClCO+ cation both display very short C—Cl bonds with a strong double-bond character.
1. Introduction
Oxalyl chloride was first prepared by Fauconnier in 1892 by the reaction of diethyl oxalate and phosphorus pentachloride (Fauconnier, 1892). Nowadays, it is commercially produced by the photochlorination of ethylene carbonate (Pfoertner & Oppenländer, 2012). Due to its high reactivity, oxalyl chloride is one of the most versatile organic reagents in chemical syntheses. Among others, it is used in chlorinations, oxidations, reductions, dehydrations, decarboxylations or formylation reactions (Masaki & Fukui, 1977; Omura & Swern, 1978; Shiri & Kazemi, 2017; Denton et al., 2012; Wasserman & Tremper, 1977; Mendelson & Hayden, 1996). The best-known applications are its use in Friedel–Crafts reactions (Alexandrou, 1969; Ketcha & Gribble, 1985) or the Swern oxidation, i.e. the oxidation of primary and secondary to and (Omura & Swern, 1978).
The reactivity of oxalyl chloride towards strong Lewis acids has been thoroughly investigated. Thus, the chlorocarbonyl cation (ClCO+) is observed in the reaction of oxalyl chloride with SbF5 (Prakash et al., 1991). The ClCO+ cation is a representative of the compound class of linear triatomic molecules, such as OCS, ONP or ONS+, for which numerous theoretical calculations have been performed (Peterson et al., 1991; Pak & Woods, 1997). It can be synthesized by reacting SbF5 with either oxalyl chloride, phosgene or carbonyl chloride fluoride. Furthermore, the reaction of carbon monoxide with chlorine in SO2ClF/SbF5 leads to the formation of the ClCO+ cation (Prakash et al., 1991; Bernhardt et al., 1999; Christe et al., 1999). The latter was investigated by Olah in 1991 using NMR spectroscopy (Prakash et al., 1991), while Aubke characterized the species by Raman and IR spectroscopy for the first time in 1999 (Bernhardt et al., 1999). However, it has not yet been possible to elucidate the of the cation due to its high reactivity. This prompted us to isolate the ClCO+ cation and to perform single-crystal X-ray to structurally characterize the cation and to compare its bond lengths with isoelectronic molecules such as OCS, which shows a short C—S bond with a strong C=S double-bond character (Pak & Woods, 1997).
Furthermore, in previous studies by our group, mono- and diprotonated species of oxalic acid, pyruvic acid and parabanic acid were isolated and characterized, whereas the dications constitute vicinal superelectrophiles (Schickinger et al., 2018; Virmani et al., 2022; Beck et al., 2020). In addition, we recently investigated the reactivity of haloacetyl fluorides in superacidic media and observed the protonation of the carbonyl bond, as well as HF addition to the latter, as in the case of dichloroacetyl fluoride (Steiner et al., 2022, 2024). This prompted us to perform investigations on the reactivity of oxalyl chloride in the superacidic system HF/SbF5.
2. Results and discussion
2.1. Syntheses and properties of [C2O(OX)Cl2][SbF6] (X = H or D), [C2(OH)2Cl2][SbnF5n+1]2, [ClCO][Sb3F15Cl] and [ClCO][Sb3F16]
Oxalyl chloride was reacted in the binary superacidic systems HF/SbF5 and DF/SbF5. According to reaction (1) in Scheme 1, [C2O(OH)Cl2][SbF6] (1) and [C2O(OD)Cl2][SbF6] (2) were obtained as O-monoprotonated species of oxalyl chloride in quantitative yields as colourless solids.
To obtain the diprotonated species, oxalyl chloride was reacted in the binary superacidic system HF/SbF5 with an excess of the strong SbF5. However, even with a tenfold excess of SbF5, the isolation of diprotonated oxalyl chloride (3) as a solid was not possible. Instead of the desired species, monoprotonated oxalyl chloride (1) was obtained. To investigate the reaction of oxalyl chloride in anhydrous hydrogen fluoride (aHF) at −60 °C with an excess of SbF5, NMR spectroscopy was performed employing eight equivalents of SbF5. Accordingly, the measured 1H, 19F and 13C NMR spectra indicate the presence of 3 in the solution, as presented in reaction (2) in Scheme 2. Salts 1 and 2 show thermal decomposition at −45 °C.
[ClCO][Sb3F15Cl] (4) was synthesized by reacting oxalyl chloride with three equivalents of SbF5 in 1,1,1,2-tetrafluoroethane (R-134a, CF3CFH2) at −78 °C. 4 was obtained as a colourless solid according to reaction (3) in Scheme 3. To avoid mixed occupancies of the fluorine positions of the anion with chlorine, carbonyl chloride fluoride was reacted under the same conditions to form [ClCO][Sb3F16] (5), as presented in reaction (4) in Scheme 3 (Prakash et al., 1991; Bernhardt et al., 1999; Christe et al., 1999).
2.2. Vibrational spectroscopy
The low-temperature Raman (Ra) and infrared (IR) spectra of [C2O(OH)Cl2][SbF6] (1), [C2O(OD)Cl2][SbF6] (2) and C2O2Cl2 are illustrated in Fig. 1. The complete vibrational frequencies of 1, 2, 4 and 5, as well as of oxalyl chloride and carbonyl chloride fluoride, are provided in the supporting information (see Figs. S1–S3 and Tables S1–S5) (Bernhardt et al., 1999; Davis et al., 1993; Nielsen et al., 1952).
For the [C2O(OH)Cl2]+ cation with Cs symmetry, 15 fundamental vibrational modes are expected, all of which are Raman and IR active. νs(O—H) is superposed by condensed water in the IR spectra due to the measuring method. Furthermore, the O—H stretching vibration shows low intensity in the Raman spectra due to the poor polarizability of the O—H group, which does not apply to the O—D group. The O—D stretching vibration of the D-isotopomeric species 2 is observed at 2157 cm−1 in the Raman spectrum and at 2388 cm−1 in the IR spectrum. The C=O stretching vibration of the protonated COCl moiety is detected at 1607 cm−1 (Ra) (1 and 2), as well as at 1605 (1) and 1593 cm−1 (2) (IR), and is significantly red-shifted (by approximately 165 cm−1) compared to the starting material. The vibrations are red-shifted by around 100–150 cm−1 in comparison to the corresponding vibrations of protonated acyl fluorides (Steiner et al., 2022, 2024; Bayer et al., 2022). The νs(C=O) of the adjacent unprotonated carbonyl group is not affected by the protonation. The C—Cl stretching vibration of the protonated COCl moiety appears at 817 cm−1 (2) in the Raman spectrum and at 833 cm−1 (1), as well as at 818 cm−1 (2), in the IR spectrum, i.e. blue-shifted by approximately 70 cm−1. The νs(C—Cl) of the neighboring COCl moiety is not affected by protonation and occurs at 623 cm−1 (2) in the Raman spectrum. Additionally, the C—C stretching vibration is detected at 1118 (Ra) (1) and 1117 cm−1 (1 and 2) (IR), and is thus blue-shifted by approximately 20 cm−1 in comparison to oxalyl chloride.
More vibrations are observed for the [SbF6]− anion than expected for ideal octahedral symmetry due to interionic interactions in the solid leading to a symmetry distortion (Weidlein et al., 1988).
2.3. of [C2O(OH)Cl2][SbF6] (1)
The hexafluoridoantimonate of monoprotonated oxalyl chloride (1) crystallizes in the monoclinic P21 with two formula units per The is illustrated in Fig. 2. Crystal data and structure details are provided in Table 1 and in the supporting information (Tables S6 and S7).
Due to the protonation, the C1—O1 bond length [1.225 (8) Å] is significantly elongated compared to the starting material [1.180 (2) Å; Danielson et al., 1995] and is longer than a formal C=O bond (1.18 Å; Allen et al., 1987). The C2—O2 bond length [1.184 (8) Å] of the adjacent unprotonated C=O bond is not affected by the protonation [1.180 (2) Å; Danielson et al., 1995]. Furthermore, the C1—Cl1 bond length [1.647 (7) Å] is significantly shortened compared to the neutral compound [1.747 (3) Å; Danielson et al., 1995] and is in the range between a formal C—Cl single bond (1.76 Å; Allen et al., 1987) and a C=Cl double bond (1.56 Å; Holleman et al., 1987). The same applies for the C2—Cl2 bond length [1.693 (7) Å] [cf. 1.747 (3) Å for oxalyl chloride; Danielson et al., 1995]. The elongation of the C=O bond and the shortening of the C—Cl bonds are consistent with the observed shifts of the νs(C=O) and νs(C—Cl) in the vibrational spectra. The C1—C2 bond length [1.550 (8) Å] is not affected by protonation [cf. 1.545 (8) Å for oxalyl chloride; Danielson et al., 1995]. This is consistent with the results observed for the mono- and diprotonated species of oxalic acid. In both cases, the protonation at the carbonyl groups does not affect the bond lengths of the C—C backbone (Schickinger et al., 2018).
The Sb—F bond lengths are in the range between 1.846 (4) and 1.941 (3) Å, and correspond with values reported in the literature (Minkwitz et al., 1999a,b; Minkwitz & Schneider, 1999). Due to interionic interactions, the anion displays distorted octahedral symmetry. The Sb1—F6 bond [1.941 (3) Å] is significantly longer than the other Sb—F bonds with it being involved in hydrogen bonding.
In the 1, the ions are arranged into chains along the a and b axes by the strong O1(—H1)⋯F6 hydrogen bond [2.421 (7) Å] and the C⋯F interactions C1⋯F2i [2.617 (7) Å; see Fig. 4 for symmetry codes] and C1⋯F4ii [2.565 (7) Å] (see Figs. S4–S5) (Jeffrey, 1997; Bondi, 1964). The chains are linked to each other by the Cl⋯F interaction Cl1⋯F3iii [2.883 (5) Å] to form layers. All interatomic C⋯F and Cl⋯F contacts are below the sum of the van der Waals radii (3.17 and 3.22 Å; Bondi, 1964). Interatomic distances are listed in Table S7.
of2.4. of [ClCO][Sb3F16] (5)
The crystal structures of 4 and 5 were both obtained by recrystallizing the salts from R-134a (CF3CFH2) at −40 °C. As 4 crystallizes as an and shows mixed occupancies of all 16 crystallographic fluorine positions with chlorine, the structural parameters show high standard deviations. Therefore, the of 5 is used for the discussion of all experimental parameters.
[ClCO][Sb3F16] (5) crystallizes in the trigonal P31 with three formula units per The is depicted in Fig. 3. Crystal data and structure details are provided in Table 1 and in the supporting information (Tables S6 and S8).
The C1—O1 bond length [1.105 (10) Å] is significantly shortened compared to ClFCO [1.173 (2) Å; Oberhammer, 1980]. It is in the range between a formal C=O double bond (1.19 Å) and a C≡O triple bond (1.07 Å) (Allen et al., 1987). It corresponds with the C=O bond length (1.1562 Å) of the isoelectronic molecule OCS (Pak & Woods, 1997). Furthermore, the C1—Cl1 bond length [1.571 (3) Å] is significantly shortened compared to ClFCO [1.725 (2) Å; Oberhammer, 1980]. It is in the range of a formal C=Cl double-bond length (1.56 Å) and thus displays a significant double-bond character (Allen et al., 1987). It corresponds with the C—Cl bond length [1.67 (2) Å] of ClCN (Beach & Turkevich, 1939). The Cl1—C1—O1 angle [176.1 (8)°] indicates the essentially linear structure of the cation.
The Sb—F bond lengths of the terminal F atoms are in the range between 1.831 (8) and 1.857 (7) Å. The bridging Sb—F bonds are longer than the terminal ones, with bond lengths up to 2.088 (8) Å. The bond angles Sb1—F6—Sb2 [146.0 (4)°] and Sb2—F11—Sb3 [156.6 (4)°] are also in good agreement with the literature (Faggiani et al., 1986; Gerken et al., 2002).
In the 5, the ions form a helical structure along the c axis via the C⋯F interaction C1⋯F5ii (2.75 Å) and the Cl⋯F interaction Cl1⋯F1 [2.56 (1) Å] (see Fig. 4, and Figs. S6–S7 in the supporting information). The cation displays a tetracoordinated C1 atom and further forms the interionic contacts C1⋯F4i (2.968 Å), C1⋯F14iii (3.062 Å), C1⋯F15iv (2.92 Å) and O1⋯F16v [2.79 (2) Å] (see Fig. 4). All interatomic C⋯F, Cl⋯F and O⋯F contacts are below the sum of the van der Waals radii (3.17, 3.22 and 2.99 Å; Bondi, 1964). Selected interatomic distances are listed in Table S8.
of2.5. NMR spectroscopy
To trace the reactivity of oxalyl chloride in aHF and the binary superacidic system HF/SbF5, the 1H, 19F and 13C NMR spectra were measured at −60 °C with acetone-d6 as the The measured NMR spectra and the complete NMR data of oxalyl chloride, 1 and 3 are listed in the supporting information (Figs. S8–S18).
The 13C NMR spectrum of oxalyl chloride dissolved in aHF at −60 °C shows a singlet at 161.1 ppm for both COCl moieties. Furthermore, no chlorine–fluorine exchange or HF addition to the carbonyl bond is observed under these conditions.
By first dissolving equimolar amounts of SbF5 compared to oxalyl chloride in HF and then adding the acyl chloride, monoprotonated oxalyl chloride (1) is formed. The 1H NMR spectrum shows a singlet at 10.03 ppm for the protonated carbonyl group. The 13C NMR spectrum displays two singlets located at 161.7 and 183.4 ppm. These are assigned to the carbonyl groups, whereas the NMR resonance of the protonated COCl moiety is significantly shifted downfield. The 19F NMR spectrum indicates chlorine–fluorine exchange initiated by the protonation of oxalyl chloride. Thus, the NMR resonance at 17.74 ppm indicates the formation of oxalyl fluoride. The 1H, 19F and 13C NMR spectra of (COF)2 at −60 °C in aHF are illustrated in Figs. S10–S11 (see supporting information). Furthermore, in the 19F NMR spectrum, the resonance at −124.78 ppm is assigned to the [SbF6]− anion (Dean & Gillespie, 1969).
Since the diprotonated oxalyl chloride (3) could not be isolated as a solid, the reaction of oxalyl chloride in HF/SbF5 at −60 °C was investigated using NMR spectroscopy, whereas an eightfold amount of SbF5 was applied. Accordingly, the NMR spectra indicate the presence of 3 in the solution. The 1H NMR spectrum shows a singlet at 9.61 ppm for the protonated COCl+ moieties, whereas the 13C NMR spectrum displays a singlet at 183.1 ppm. Compared to the neutral compound, the 13C NMR resonance is significantly shifted downfield. Furthermore, in the 19F NMR spectrum, multiple resonances located in the range between −117.46 and −143.05 ppm are assigned to the [SbnF5n+1]− polyanions (Dean & Gillespie, 1969). Thus, diprotonated oxalyl chloride is stable in solution at −60 °C. However, after removal of the excess HF at −78 °C, it decomposes with the formation of 1. Furthermore, as in the NMR spectra of 1, a chlorine–fluorine exchange is observed, as the signal at 17.46 ppm indicates the formation of oxalyl fluoride. Additional signals in the 19F NMR spectra of 3 at −21.78 and −59.90 ppm are assigned to COF2 and CF3OH (Christe et al., 2007). The 1H, 19F and 13C NMR spectra of COF2 in aHF at −60 °C are depicted in the supporting information (Fig. S12). COF2 is likely formed due to the decomposition of oxalyl fluoride under superacidic conditions.
3. Conclusions
Oxalyl chloride was reacted in the binary superacidic systems HF/SbF5 and DF/SbF5 to form the O-monoprotonated and its D-isotopomeric species as hexafluoridoantimonates. Both represent the first examples of protonated acyl chlorides. When the SbF5 is applied in eightfold excess, diprotonated oxalyl chloride is formed, which is only stable in solution. By the reaction of oxalyl chloride or carbonyl chloride fluoride in the aprotic solvent 1,1,1,2-tetrafluoroethane (R-134a, CF3CFH2) with a threefold excess of SbF5, salts of the chlorocarbonyl cation were isolated. The colourless salts were characterized by low-temperature vibrational spectroscopy and low-temperature NMR spectroscopy. The crystal structures of [C2O(OH)Cl2][SbF6] and [ClCO][Sb3F16] were determined by single-crystal X-ray Monoprotonated oxalyl chloride and the chlorocarbonyl cation both display very short C—Cl bonds with strong double-bond character.
Supporting information
https://doi.org/10.1107/S2053229624010714/yd3050sup1.cif
contains datablocks I, 5, global. DOI:Structure factors: contains datablock 1. DOI: https://doi.org/10.1107/S2053229624010714/yd30501sup2.hkl
Structure factors: contains datablock 5. DOI: https://doi.org/10.1107/S2053229624010714/yd30505sup3.hkl
Supporting information file. DOI: https://doi.org/10.1107/S2053229624010714/yd3050sup4.pdf
(C2HCl2O2)[SbF6] | F(000) = 336 |
Mr = 363.68 | Dx = 2.741 Mg m−3 |
Monoclinic, P21 | Mo Kα radiation, λ = 0.71073 Å |
a = 6.1616 (8) Å | Cell parameters from 3114 reflections |
b = 10.8379 (10) Å | θ = 3.1–29.8° |
c = 6.8805 (8) Å | µ = 3.80 mm−1 |
β = 106.472 (13)° | T = 106 K |
V = 440.61 (9) Å3 | Plate, colourless |
Z = 2 | 0.31 × 0.17 × 0.12 mm |
Rigaku Xcalibur Sapphire3 diffractometer | 2930 independent reflections |
Radiation source: Enhance (Mo) X-ray Source | 2692 reflections with I > 2σ(I) |
Graphite monochromator | Rint = 0.042 |
Detector resolution: 15.9809 pixels mm-1 | θmax = 32.5°, θmin = 3.1° |
ω scans | h = −9→9 |
Absorption correction: multi-scan (CrysAlis PRO; Rigaku OD, 2020) | k = −15→15 |
Tmin = 0.737, Tmax = 1.000 | l = −10→10 |
8919 measured reflections |
Refinement on F2 | Secondary atom site location: difference Fourier map |
Least-squares matrix: full | Hydrogen site location: difference Fourier map |
R[F2 > 2σ(F2)] = 0.030 | Only H-atom coordinates refined |
wR(F2) = 0.057 | w = 1/[σ2(Fo2) + (0.0195P)2] where P = (Fo2 + 2Fc2)/3 |
S = 1.04 | (Δ/σ)max < 0.001 |
2930 reflections | Δρmax = 1.18 e Å−3 |
122 parameters | Δρmin = −0.65 e Å−3 |
2 restraints | Absolute structure: Refined as an inversion twin |
Primary atom site location: structure-invariant direct methods | Absolute structure parameter: 0.52 (3) |
Geometry. All esds (except the esd in the dihedral angle between two l.s. planes) are estimated using the full covariance matrix. The cell esds are taken into account individually in the estimation of esds in distances, angles and torsion angles; correlations between esds in cell parameters are only used when they are defined by crystal symmetry. An approximate (isotropic) treatment of cell esds is used for estimating esds involving l.s. planes. |
Refinement. Refinement of F2 against ALL reflections. The weighted R-factor wR and goodness of fit S are based on F2, conventional R-factors R are based on F, with F set to zero for negative F2. The threshold expression of F2 > 2sigma(F2) is used only for calculating R-factors(gt) etc. and is not relevant to the choice of reflections for refinement. R-factors based on F2 are statistically about twice as large as those based on F, and R- factors based on ALL data will be even larger. Refined as a 2-component inversion twin. |
x | y | z | Uiso*/Ueq | ||
Sb1 | 0.53137 (5) | 0.40457 (3) | 0.12537 (5) | 0.01795 (9) | |
Cl1 | 0.1619 (3) | 0.69867 (15) | 0.2994 (3) | 0.0292 (4) | |
Cl2 | −0.1153 (3) | 0.56726 (18) | 0.7590 (3) | 0.0328 (4) | |
F6 | 0.4922 (6) | 0.4491 (3) | 0.3867 (5) | 0.0286 (8) | |
F4 | 0.6527 (7) | 0.2572 (3) | 0.2457 (6) | 0.0384 (10) | |
O1 | 0.2117 (8) | 0.5185 (4) | 0.5476 (7) | 0.0213 (9) | |
F3 | 0.5832 (7) | 0.3631 (3) | −0.1181 (6) | 0.0334 (9) | |
F5 | 0.2429 (7) | 0.3389 (4) | 0.0574 (6) | 0.0446 (11) | |
F1 | 0.4165 (8) | 0.5572 (4) | 0.0299 (6) | 0.0433 (11) | |
F2 | 0.8179 (7) | 0.4748 (4) | 0.2204 (6) | 0.0431 (11) | |
C1 | 0.1102 (12) | 0.6133 (6) | 0.4793 (9) | 0.0198 (13) | |
O2 | −0.1927 (10) | 0.7435 (5) | 0.4961 (9) | 0.0382 (14) | |
C2 | −0.0817 (11) | 0.6568 (6) | 0.5680 (10) | 0.0237 (14) | |
H1 | 0.313 (9) | 0.504 (7) | 0.490 (10) | 0.036* |
U11 | U22 | U33 | U12 | U13 | U23 | |
Sb1 | 0.01936 (15) | 0.01674 (14) | 0.01797 (15) | 0.0007 (2) | 0.00563 (11) | −0.0003 (2) |
Cl1 | 0.0418 (10) | 0.0241 (8) | 0.0255 (8) | −0.0010 (7) | 0.0155 (7) | 0.0036 (6) |
Cl2 | 0.0355 (10) | 0.0360 (10) | 0.0331 (10) | −0.0003 (7) | 0.0200 (8) | 0.0037 (7) |
F6 | 0.032 (2) | 0.035 (2) | 0.0212 (18) | 0.0070 (15) | 0.0107 (15) | −0.0026 (14) |
F4 | 0.060 (3) | 0.024 (2) | 0.027 (2) | 0.0170 (19) | 0.0054 (19) | 0.0026 (15) |
O1 | 0.024 (2) | 0.018 (2) | 0.024 (2) | 0.0005 (17) | 0.0105 (18) | −0.0030 (17) |
F3 | 0.044 (2) | 0.036 (2) | 0.0244 (19) | 0.0045 (16) | 0.0163 (17) | −0.0041 (14) |
F5 | 0.028 (2) | 0.066 (3) | 0.036 (2) | −0.0170 (19) | 0.0026 (18) | 0.006 (2) |
F1 | 0.069 (3) | 0.027 (2) | 0.035 (2) | 0.018 (2) | 0.017 (2) | 0.0101 (17) |
F2 | 0.033 (2) | 0.058 (3) | 0.042 (2) | −0.020 (2) | 0.0147 (19) | −0.013 (2) |
C1 | 0.023 (3) | 0.022 (3) | 0.013 (3) | −0.003 (3) | 0.002 (2) | −0.002 (2) |
O2 | 0.039 (3) | 0.038 (3) | 0.040 (3) | 0.017 (3) | 0.015 (2) | 0.011 (3) |
C2 | 0.018 (3) | 0.025 (3) | 0.027 (3) | 0.002 (2) | 0.004 (3) | −0.003 (3) |
Sb1—F1 | 1.846 (4) | Cl1—C1 | 1.647 (7) |
Sb1—F5 | 1.848 (4) | Cl2—C2 | 1.693 (7) |
Sb1—F3 | 1.848 (4) | O1—C1 | 1.225 (8) |
Sb1—F4 | 1.856 (3) | C1—C2 | 1.550 (8) |
Sb1—F2 | 1.863 (4) | O2—C2 | 1.184 (8) |
Sb1—F6 | 1.941 (3) | ||
F1—Sb1—F5 | 90.6 (2) | F5—Sb1—F6 | 88.10 (17) |
F1—Sb1—F3 | 92.44 (17) | F3—Sb1—F6 | 177.29 (18) |
F5—Sb1—F3 | 94.5 (2) | F4—Sb1—F6 | 86.60 (16) |
F1—Sb1—F4 | 174.63 (18) | F2—Sb1—F6 | 85.84 (17) |
F5—Sb1—F4 | 91.05 (19) | O1—C1—C2 | 117.9 (6) |
F3—Sb1—F4 | 92.51 (17) | O1—C1—Cl1 | 125.0 (6) |
F1—Sb1—F2 | 89.3 (2) | C2—C1—Cl1 | 117.2 (6) |
F5—Sb1—F2 | 173.9 (2) | O2—C2—C1 | 119.5 (7) |
F3—Sb1—F2 | 91.58 (18) | O2—C2—Cl2 | 126.5 (6) |
F4—Sb1—F2 | 88.5 (2) | C1—C2—Cl2 | 114.0 (5) |
F1—Sb1—F6 | 88.36 (16) | ||
O1—C1—C2—O2 | −175.1 (7) | O1—C1—C2—Cl2 | 2.6 (7) |
Cl1—C1—C2—O2 | 4.9 (8) | Cl1—C1—C2—Cl2 | −177.4 (4) |
(CClO)[Sb3F16] | Dx = 3.519 Mg m−3 |
Mr = 732.71 | Mo Kα radiation, λ = 0.71073 Å |
Trigonal, P31 | Cell parameters from 2010 reflections |
a = 8.0824 (3) Å | θ = 2.9–31.5° |
c = 18.3341 (8) Å | µ = 6.19 mm−1 |
V = 1037.22 (9) Å3 | T = 102 K |
Z = 3 | Plate, colourless |
F(000) = 984 | 0.19 × 0.14 × 0.11 mm |
Rigaku Xcalibur Sapphire3 diffractometer | 3334 independent reflections |
Radiation source: Enhance (Mo) X-ray Source | 3029 reflections with I > 2σ(I) |
Graphite monochromator | Rint = 0.044 |
Detector resolution: 15.9809 pixels mm-1 | θmax = 32.0°, θmin = 2.2° |
ω scans | h = −9→12 |
Absorption correction: multi-scan (CrysAlis PRO; Rigaku OD, 2020) | k = −12→9 |
Tmin = 0.882, Tmax = 1.000 | l = −23→27 |
6310 measured reflections |
Refinement on F2 | Primary atom site location: structure-invariant direct methods |
Least-squares matrix: full | Secondary atom site location: difference Fourier map |
R[F2 > 2σ(F2)] = 0.038 | w = 1/[σ2(Fo2) + (0.0262P)2] where P = (Fo2 + 2Fc2)/3 |
wR(F2) = 0.077 | (Δ/σ)max < 0.001 |
S = 1.03 | Δρmax = 1.36 e Å−3 |
3334 reflections | Δρmin = −1.09 e Å−3 |
199 parameters | Absolute structure: Twinning involves inversion, so Flack parameter cannot be determined |
0 restraints |
Geometry. All esds (except the esd in the dihedral angle between two l.s. planes) are estimated using the full covariance matrix. The cell esds are taken into account individually in the estimation of esds in distances, angles and torsion angles; correlations between esds in cell parameters are only used when they are defined by crystal symmetry. An approximate (isotropic) treatment of cell esds is used for estimating esds involving l.s. planes. |
Refinement. Refined as a 4-component inversion twin |
x | y | z | Uiso*/Ueq | ||
Sb2 | 0.63376 (11) | −0.03511 (10) | 0.6093 (6) | 0.01747 (16) | |
Sb3 | 0.61862 (11) | −0.00397 (11) | 0.3929 (6) | 0.01723 (15) | |
Sb1 | 0.36234 (12) | 0.24398 (12) | 0.5809 (6) | 0.01933 (17) | |
Cl1 | 0.0670 (5) | 0.5221 (5) | 0.4818 (6) | 0.0259 (6) | |
F6 | 0.4921 (10) | 0.1038 (10) | 0.6276 (7) | 0.0217 (15) | |
F11 | 0.5841 (11) | −0.0061 (11) | 0.5059 (7) | 0.0259 (16) | |
F14 | 0.3543 (10) | −0.1218 (11) | 0.3896 (7) | 0.0253 (16) | |
F12 | 0.8757 (10) | 0.1112 (12) | 0.4172 (7) | 0.0307 (18) | |
F3 | 0.1582 (11) | 0.0903 (10) | 0.6393 (7) | 0.0289 (17) | |
F7 | 0.8440 (10) | 0.2034 (10) | 0.5984 (8) | 0.0304 (18) | |
F1 | 0.2479 (12) | 0.3677 (12) | 0.5370 (7) | 0.035 (2) | |
F15 | 0.6529 (11) | 0.0015 (11) | 0.2933 (8) | 0.0297 (19) | |
F13 | 0.6347 (12) | 0.2325 (11) | 0.3976 (7) | 0.0287 (18) | |
F2 | 0.2737 (11) | 0.0514 (11) | 0.5132 (7) | 0.0297 (17) | |
F9 | 0.6597 (12) | −0.0508 (12) | 0.7083 (8) | 0.034 (2) | |
F16 | 0.5937 (11) | −0.2420 (10) | 0.4091 (7) | 0.0288 (17) | |
F4 | 0.5920 (11) | 0.3809 (11) | 0.5326 (7) | 0.0313 (18) | |
F8 | 0.3967 (11) | −0.2487 (11) | 0.6050 (7) | 0.0314 (18) | |
F5 | 0.4618 (10) | 0.4138 (11) | 0.6581 (7) | 0.0267 (16) | |
F10 | 0.7572 (12) | −0.1617 (12) | 0.5794 (7) | 0.0349 (19) | |
O1 | −0.0886 (17) | 0.7424 (16) | 0.4599 (8) | 0.043 (3) | |
C1 | −0.029417 | 0.646600 | 0.467900 | 0.031 (3) |
U11 | U22 | U33 | U12 | U13 | U23 | |
Sb2 | 0.0197 (4) | 0.0198 (4) | 0.0159 (4) | 0.0121 (3) | −0.0007 (3) | 0.0003 (3) |
Sb3 | 0.0173 (4) | 0.0188 (4) | 0.0146 (4) | 0.0083 (3) | 0.0003 (3) | −0.0002 (3) |
Sb1 | 0.0240 (4) | 0.0221 (4) | 0.0182 (4) | 0.0162 (4) | −0.0027 (3) | −0.0026 (3) |
Cl1 | 0.0318 (17) | 0.0268 (16) | 0.0233 (15) | 0.0177 (14) | −0.0020 (13) | 0.0022 (13) |
F6 | 0.026 (4) | 0.033 (4) | 0.015 (3) | 0.022 (3) | 0.000 (3) | 0.000 (3) |
F11 | 0.040 (4) | 0.033 (4) | 0.015 (3) | 0.026 (4) | 0.001 (3) | 0.000 (3) |
F14 | 0.021 (4) | 0.033 (4) | 0.019 (4) | 0.011 (3) | −0.004 (3) | −0.005 (3) |
F12 | 0.026 (4) | 0.038 (5) | 0.028 (4) | 0.015 (4) | −0.005 (3) | −0.008 (3) |
F3 | 0.026 (4) | 0.033 (4) | 0.030 (4) | 0.017 (4) | −0.002 (3) | 0.002 (3) |
F7 | 0.019 (4) | 0.019 (4) | 0.046 (5) | 0.004 (3) | 0.002 (3) | −0.001 (3) |
F1 | 0.049 (5) | 0.043 (5) | 0.028 (4) | 0.034 (5) | −0.011 (4) | −0.002 (4) |
F15 | 0.032 (4) | 0.042 (5) | 0.016 (4) | 0.019 (4) | 0.002 (3) | −0.003 (3) |
F13 | 0.044 (5) | 0.023 (4) | 0.023 (4) | 0.019 (4) | 0.004 (3) | 0.002 (3) |
F2 | 0.036 (4) | 0.032 (4) | 0.027 (4) | 0.021 (4) | −0.012 (3) | −0.009 (3) |
F9 | 0.047 (5) | 0.045 (5) | 0.021 (5) | 0.031 (4) | −0.006 (4) | 0.001 (4) |
F16 | 0.035 (4) | 0.023 (4) | 0.032 (4) | 0.017 (3) | −0.007 (4) | −0.004 (3) |
F4 | 0.036 (4) | 0.029 (4) | 0.031 (4) | 0.019 (4) | 0.007 (4) | 0.003 (4) |
F8 | 0.032 (4) | 0.020 (4) | 0.037 (5) | 0.009 (3) | 0.008 (4) | 0.003 (3) |
F5 | 0.034 (4) | 0.028 (4) | 0.023 (4) | 0.019 (4) | −0.005 (3) | −0.007 (3) |
F10 | 0.042 (5) | 0.043 (5) | 0.034 (5) | 0.032 (4) | −0.005 (4) | −0.002 (4) |
O1 | 0.058 (7) | 0.056 (7) | 0.043 (7) | 0.049 (6) | 0.003 (6) | 0.000 (6) |
C1 | 0.041 (8) | 0.037 (8) | 0.014 (6) | 0.019 (7) | 0.000 (5) | −0.005 (6) |
Sb2—F8 | 1.831 (8) | Sb3—F12 | 1.857 (7) |
Sb2—F10 | 1.834 (7) | Sb3—F11 | 2.088 (8) |
Sb2—F7 | 1.835 (7) | Sb1—F3 | 1.833 (8) |
Sb2—F9 | 1.838 (8) | Sb1—F2 | 1.834 (7) |
Sb2—F11 | 1.978 (8) | Sb1—F4 | 1.844 (8) |
Sb2—F6 | 1.992 (6) | Sb1—F1 | 1.852 (7) |
Sb3—F15 | 1.845 (8) | Sb1—F5 | 1.852 (7) |
Sb3—F13 | 1.852 (7) | Sb1—F6 | 2.075 (7) |
Sb3—F14 | 1.855 (7) | Cl1—C1 | 1.571 (3) |
Sb3—F16 | 1.855 (7) | O1—C1 | 1.105 (10) |
F8—Sb2—F10 | 93.8 (4) | F16—Sb3—F12 | 89.7 (4) |
F8—Sb2—F7 | 165.6 (3) | F15—Sb3—F11 | 179.1 (4) |
F10—Sb2—F7 | 94.4 (4) | F13—Sb3—F11 | 84.4 (3) |
F8—Sb2—F9 | 94.6 (4) | F14—Sb3—F11 | 85.0 (3) |
F10—Sb2—F9 | 98.3 (3) | F16—Sb3—F11 | 83.6 (3) |
F7—Sb2—F9 | 95.9 (4) | F12—Sb3—F11 | 83.0 (3) |
F8—Sb2—F11 | 83.8 (4) | F3—Sb1—F2 | 89.0 (3) |
F10—Sb2—F11 | 88.8 (3) | F3—Sb1—F4 | 169.1 (3) |
F7—Sb2—F11 | 84.6 (4) | F2—Sb1—F4 | 91.9 (4) |
F9—Sb2—F11 | 172.9 (3) | F3—Sb1—F1 | 95.5 (4) |
F8—Sb2—F6 | 85.0 (3) | F2—Sb1—F1 | 94.3 (4) |
F10—Sb2—F6 | 172.2 (3) | F4—Sb1—F1 | 95.3 (4) |
F7—Sb2—F6 | 85.3 (3) | F3—Sb1—F5 | 88.1 (3) |
F9—Sb2—F6 | 89.5 (3) | F2—Sb1—F5 | 170.8 (3) |
F11—Sb2—F6 | 83.5 (3) | F4—Sb1—F5 | 89.3 (4) |
F15—Sb3—F13 | 95.1 (3) | F1—Sb1—F5 | 94.6 (4) |
F15—Sb3—F14 | 95.8 (3) | F3—Sb1—F6 | 85.3 (3) |
F13—Sb3—F14 | 90.0 (3) | F2—Sb1—F6 | 84.6 (3) |
F15—Sb3—F16 | 96.9 (4) | F4—Sb1—F6 | 83.9 (3) |
F13—Sb3—F16 | 168.0 (4) | F1—Sb1—F6 | 178.6 (3) |
F14—Sb3—F16 | 88.5 (3) | F5—Sb1—F6 | 86.5 (3) |
F15—Sb3—F12 | 96.2 (3) | Sb2—F6—Sb1 | 146.0 (4) |
F13—Sb3—F12 | 89.3 (4) | Sb2—F11—Sb3 | 156.6 (4) |
F14—Sb3—F12 | 168.0 (3) | O1—C1—Cl1 | 176.1 (8) |
Footnotes
‡Deceased
Acknowledgements
We are grateful to the Department of Chemistry of the Ludwig Maximilian University, the Deutsche Forschungsgemeinschaft (DFG) and F-Select GmbH for the financial support of this work. Special thanks go to Professor Dr Konstantin Karaghiosoff for help with this work after Professor Dr Andreas J. Kornath passed away. Open access funding enabled and organized by Projekt DEAL.
References
Alexandrou, N. E. (1969). J. Chem. Soc. C, pp. 536–537. Google Scholar
Allen, F. H., Kennard, O., Watson, D. G., Brammer, L., Orpen, A. G. & Taylor, R. (1987). J. Chem. Soc. Perkin Trans. 2, pp. S1–S19. CrossRef Web of Science Google Scholar
Bayer, M. C., Kremser, C., Jessen, C., Nitzer, A. & Kornath, A. J. (2022). Chem. A Eur. J. 28, e202104422. CrossRef Google Scholar
Beach, J. Y. & Turkevich, A. (1939). J. Am. Chem. Soc. 61, 299–303. CrossRef CAS Google Scholar
Beck, S., Raljic, M., Jessen, C. & Kornath, A. J. (2020). Eur. J. Org. Chem. 2020, 4521–4527. CrossRef CAS Google Scholar
Bernhardt, E., Willner, H. & Aubke, F. (1999). Angew. Chem. Int. Ed. 38, 823–825. CrossRef CAS Google Scholar
Bondi, A. (1964). J. Phys. Chem. 68, 441–451. CrossRef CAS Web of Science Google Scholar
Christe, K. O., Hegge, J., Hoge, B. & Haiges, R. (2007). Angew. Chem. Int. Ed. 46, 6155–6158. CrossRef CAS Google Scholar
Christe, K. O., Hoge, B., Boatz, J. A., Prakash, G. K. S., Olah, G. A. & Sheehy, J. A. (1999). Inorg. Chem. 38, 3132–3142. CrossRef CAS Google Scholar
Danielson, D. D., Hedberg, L., Hedberg, K., Hagen, K. & Trtteberg, M. (1995). J. Phys. Chem. 99, 9374–9379. CrossRef CAS Google Scholar
Davis, J. F., Wang, A. & Durig, J. R. (1993). J. Mol. Struct. 293, 27–30. CrossRef CAS Google Scholar
Dean, P. A. W. & Gillespie, R. J. (1969). J. Am. Chem. Soc. 91, 7260–7264. CrossRef CAS Google Scholar
Denton, R. M., An, J., Lindovska, P. & Lewis, W. (2012). Tetrahedron, 68, 2899–2905. CrossRef CAS Google Scholar
Faggiani, R., Kennepohl, D. K., Lock, C. J. L. & Schrobilgen, G. J. (1986). Inorg. Chem. 25, 563–571. CrossRef CAS Google Scholar
Farrugia, L. J. (2012). J. Appl. Cryst. 45, 849–854. Web of Science CrossRef CAS IUCr Journals Google Scholar
Fauconnier, A. (1892). C. R. Acad. Sci. 114, 122–123. Google Scholar
Gerken, M., Dixon, D. A. & Schrobilgen, G. J. (2002). Inorg. Chem. 41, 259–277. Web of Science CrossRef PubMed CAS Google Scholar
Holleman, A. F., Wiberg, E. & Wiberg, N. (2017). In Anorganische Chemie. Berlin, Boston: De Gruyter. Google Scholar
Jeffrey, G. A. (1997). In An Introduction to Hydrogen Bonding: Topics in Physical Chemistry. New York: Oxford University Press Inc. Google Scholar
Ketcha, D. M. & Gribble, G. W. (1985). J. Org. Chem. 50, 5451–5457. CrossRef CAS Google Scholar
Masaki, M. & Fukui, K. (1977). Chem. Lett. 6, 151–152. CrossRef Google Scholar
Mendelson, W. L. & Hayden, S. (1996). Synth. Commun. 26, 603–610. CrossRef CAS Google Scholar
Minkwitz, R., Hartfeld, N. & Hirsch, C. (1999a). Z. Anorg. Allg. Chem. 625, 1479–1485. CrossRef CAS Google Scholar
Minkwitz, R., Hirsch, C. & Berends, T. (1999b). Eur. J. Inorg. Chem. 1999, 2249–2254. CrossRef Google Scholar
Minkwitz, R. & Schneider, S. (1999). Angew. Chem. Int. Ed. 38, 210–212. CrossRef CAS Google Scholar
Nielsen, A. H., Burke, T. G., Woltz, P. J. H. & Jones, E. A. (1952). J. Chem. Phys. 20, 596–604. CrossRef CAS Google Scholar
Oberhammer, H. (1980). J. Chem. Phys. 73, 4310–4313. CrossRef CAS Google Scholar
Omura, K. & Swern, D. (1978). Tetrahedron, 34, 1651–1660. CrossRef CAS Google Scholar
Pak, Y. & Woods, R. C. (1997). J. Chem. Phys. 107, 5094–5102. CrossRef CAS Google Scholar
Peterson, K. A., Mayrhofer, R. C. & Woods, R. C. (1991). J. Chem. Phys. 94, 431–441. CrossRef CAS Google Scholar
Pfoertner, K.-H. & Oppenländer, T. (2012). Photochemistry. Ullmann's Encyclopedia of Industrial Chemistry. Weinheim: Wiley-VCH. Google Scholar
Prakash, G. K. S., Bausch, J. W. & Olah, G. A. (1991). J. Am. Chem. Soc. 113, 3203–3205. CrossRef CAS Google Scholar
Rigaku OD (2020). CrysAlis PRO. Rigaku Oxford Diffraction Ltd, Yarnton, Oxfordshire, England. Google Scholar
Schickinger, M., Saal, T., Zischka, F., Axhausen, J., Stierstorfer, K., Morgenstern, Y. & Kornath, A. J. (2018). ChemistrySelect, 3, 12396–12404. CrossRef CAS Google Scholar
Sheldrick, G. M. (2015a). Acta Cryst. A71, 3–8. Web of Science CrossRef IUCr Journals Google Scholar
Sheldrick, G. M. (2015b). Acta Cryst. C71, 3–8. Web of Science CrossRef IUCr Journals Google Scholar
Shiri, L. & Kazemi, M. (2017). Res. Chem. Intermed. 43, 6007–6041. CrossRef CAS Google Scholar
Spek, A. L. (2020). Acta Cryst. E76, 1–11. Web of Science CrossRef IUCr Journals Google Scholar
Steiner, S., Jessen, C. & Kornath, A. J. (2022). Z. Anorg. Allge Chem. 648, e202200060. CrossRef Google Scholar
Steiner, S., Nitzer, A., Jessen, C. & Kornath, A. J. (2024). Z. Anorg. Allge Chem. 650, e202400013. CrossRef Google Scholar
Virmani, A., Pfeiffer, M., Jessen, C., Morgenstern, Y. & Kornath, A. J. (2022). Z. Anorg. Allge Chem. 648, e202200005. CrossRef Google Scholar
Wasserman, H. H. & Tremper, A. W. (1977). Tetrahedron Lett. 18, 1449–1450. CrossRef Google Scholar
Weidlein, J., Müller, U. & Dehnicke, K. (1988). In Schwingungsspektroskopie. Eine Einführung. Stuttgart: Thieme. Google Scholar
This is an open-access article distributed under the terms of the Creative Commons Attribution (CC-BY) Licence, which permits unrestricted use, distribution, and reproduction in any medium, provided the original authors and source are cited.