Next Article in Journal
Caffeic Acid-Zinc Basic Salt/Chitosan Nanohybrid Possesses Controlled Release Properties and Exhibits In Vivo Anti-Inflammatory Activities
Previous Article in Journal
Oxoglaucine Suppresses Hepatic Fibrosis by Inhibiting TGFβ-Induced Smad2 Phosphorylation and ROS Generation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

New Complexes of Antimony(III) with Tridentate O,E,O-Ligands (E = O, S, Se, Te, NH, NMe) Derived from N-Methyldiethanolamine

Institut für Anorganische Chemie, TU Bergakademie Freiberg, Leipziger Str. 29, 09599 Freiberg, Germany
*
Author to whom correspondence should be addressed.
Molecules 2023, 28(13), 4959; https://doi.org/10.3390/molecules28134959
Submission received: 23 May 2023 / Revised: 16 June 2023 / Accepted: 20 June 2023 / Published: 24 June 2023

Abstract

:
We synthesized a series of new antimony(III) compounds by reaction of Sb(OEt)3 with organic ligands of the type E(CH2-CH2-OH)2, with E = NH, NMe, O, S, Se, and Te. The synthesized compounds have the general composition [E(CH2-CH2-O)2]Sb(OEt). For comparison, the compound (O-CH2-CH2-S)Sb(OEt) was prepared. All compounds are characterized using NMR, IR, and Raman spectroscopy. The molecular structures of the products reveal the formation of chelate complexes, wherein the ligand molecules coordinate as tridentate O,E,O-ligands to the antimony atom. Dimer formation in the solid state allows the antimony atoms to reach pentacoordination. Quantum chemical calculations including topological analysis of electron density reveal that there are polar shared bonds between antimony and the oxygen atoms bound to antimony. The interactions between the donor atom E and the Sb atom and the interactions in the dimers can be characterized as Van der Waals interactions. The reactivity of [MeN(CH2-CH2-O)2]Sb(OEt) was investigated as an example. For this purpose, the compound reacted with a range of organic compounds such as carboxylic acids and carboxylic anhydrides and small molecules like CO2 and NH3. This study establishes a new and easy accessible class of antimony(III) compounds, provides new insights into the chemistry of antimony compounds and opens up new opportunities for further research in this field.

1. Introduction

New main group metal complexes for catalysis have been the subject of increasing research recently due to their good availability and low cost. For example, early main group metals can catalyse many reactions like polymerisation, hydrofunctionalisation and cross-coupling reactions [1]. Heavier pnictogens offers new possibilities of catalytic applications because there are multiple stable oxidation states (usually at least 0, +3 and +5). Additionally, Arduengo and Stewart reported low valent pnictogen complexes of phosphorous to bismuth [2,3]. In the group of pnictogens, the tendency for hybridization lowers with higher atomic number. So the lone pair of pnictogen (III) compounds gains a more s-orbital character and therefore, the lower oxidation states become preferable and the pnictogen compounds become more oxidizing with an increasing atomic number [4]. Furthermore, the heavier pnictogens show a Lewis acidity based on the σ* orbitals. This orbital can be considered as isolobal to the vacant p-orbital in boranes for antimony compounds [5]. The Lewis acidity of antimony compounds can be increased by the introduction of electron withdrawing ligands [6]. So, Lewis acidic antimony compounds are used as fluorine sensors [7,8,9] and catalysts [10,11,12,13]. The redox chemistry between antimony (III) and (V) complexes opens up further possibilities in catalysis [14].
Besides their applications as sensor materials or catalysts, antimony compounds have been used for the treatment of the parasite disease leishmaniasis for more than half a century [15]. Recently, antimony complexes have been investigated for different applications: antimony(V) complexes were tested for their cytotoxic activity [16], for their antibacterial activities and cytotoxicity against macrophages [17]. Different sugars were combined with antimony(V) complexes to modulate the oral bioavailability of antimony containing drugs [18]. The biological potential of antimony(III) dithiocarbamate complexes has been recently reflected in a review article [19]. Antimony(III) and (V) phthalocyanine derivatives absorb near infrared light, which makes these compound attractive as potential photosensitizers for photodynamic cancer therapy and as an infrared filter for plasma display and silicon photodiodes [20].
Diethanolamine is a widely used compound bearing two electron withdrawing hydroxy moieties and a nitrogen donor atom. Complexes of transition metals with diethanolamine and related ligand molecules have been investigated for decades [21,22,23,24]. Also, complexes of diethanolamine with tin have been reported [25] and silatranes were studied extensively, e.g., [26,27,28,29]. Nonetheless, only a few antimony compounds of this type have been reported so far [30,31,32,33].
Herein, we present new antimony complexes with bi- and tridentate ligands of the diethanolamine class with different chalcogens and nitrogen as donor atoms.

2. Results and Discussion

To investigate new antimony complexes with bi- or tridentate ligands, we synthesized the compounds shown in Figure 1. The ligand system was chosen to be as flexible as possible. The ligand system is similar to pincer ligands, but has two (formally) covalent bonds to the central atom (Sb). Compounds 1 to 6 contain an eight-membered ring with a donor atom E in the ligands backbone, while compound 7 is for comparison without this additional donor atom. The donor atoms include all stable chalcogens and nitrogen. So, the influence of the nature of the donor to the antimony atom can be investigated.

2.1. Syntheses

Except for the ligands of 3 and 4, all ligands are commercially available. The selenium and tellurium ligands were synthesized using a literature protocol [34].
Many organometallic compounds are synthesized starting with the element chlorides. Since antimony(III)-chloride is a strongly hygroscopic and soft solid (trivial name: butter of antimony) [35], it is transformed into other starting materials as antimony(III)-acetat [Sb(OAc)3], tris(dimethylamino)antimony [Sb(NMe2)3] or triethoxyantimony [Sb(OEt)3]. The latter two are colourless liquids that can be effortlessly handled with the standard Schlenk technique. For the synthesis of new complexes with three Sb-O bonds, Sb(OEt)3 is an excellent starting material. Ethanol is formed as joint product in the reactions of this starting material with the ligands. The reactions were carried out in cyclohexane, which forms an azeotrope with ethanol boiling at 68 °C. Therefore, the reaction progress was monitored by the temperature at still head. If the temperature rises over 70 °C to 80 °C (boiling point of cyclohexane), the product has been formed in a quantitative yield. After the removal of the solvents under reduced pressure, the crude products were purified by recrystallization from chloroform to remove excess ligands. A similar synthesis procedure was used by Gupta et al. [30].
For temperature sensitive ligands (Se and Te derivatives), another synthesis procedure has been developed: The ligand and Sb(OEt)3 were dissolved together in absolute ethanol and the products crystallized from this solution after several hours. Since the complexes are slightly soluble in ethanol, the yield of this synthesis protocol is lower, but the purity is higher.

2.2. NMR Data

The NMR spectra of the ligands show the signals as expected, but the spectra of the products 17 are more complicated. In the 1H NMR spectra of 2, for example (see Figure S3 in the Supplementary Materials), there are four broad signals for the protons of the tridentate ligand. This indicates that the ring system is not flexible and the equatorial and axial proton positions become non-equivalent. This is a first hint to the coordination of the inner donor atom to antimony. Additionally, the proximity to the quadrupole Sb atom broadens the signals. In the 13C NMR spectrum, the signal of the inner-cyclic O-CH2 at 58.0 ppm becomes less intense and broad because of the proximity to the Sb atom (see Figure S4 in the Supplementary Materials).

2.3. Vibrational Spectroscopy

Since there are only a few vibration frequencies for common covalent trivalent antimony compounds reported in the literature [36,37,38], the vibrational frequencies are calculated using quantum mechanical methods. As an example, the recorded and the calculated IR and Raman spectra of 1 are shown in Figure S5 in the Supplementary Materials. A comparison of the infrared spectra of the free ligand N-methyldiethanolamine and the antimony complex 6 are shown in Figure S6 in the Supplementary Materials. It is obvious that the ν(O-H) above 3000 cm−1 in the ligand cannot be found in the complex. Except for this region, the IR spectra of the ligand N-methyldiethanolamine and the complex 6 look quite similar, since most of the intense vibrations are caused by the light atoms. For a better understanding of the vibrations, including the antimony atom, Raman spectra were used.
According to the quantum chemical calculations, all complexes show characteristic bands at about 600 cm−1, which can be identified as ν(Sb-OEt). There are two bands from 450 to 500 cm−1, which are the symmetric and asymmetric inner-cyclic ν(Sb-O). The band around 1075 cm−1 arises from the inner-cyclic ν(C-O). A ν(Sb-E) can be found in the calculated spectra around 150 to 70 cm−1.
The calculations predicted a ν(Sb…E) with a decreasing wave number and an increasing atomic number of E. These bands can be found in the Raman spectra with the same trend. The presence of the Sb-OEt moiety is also confirmed by the Raman spectra, since there are bands around 600 cm−1 (see Table 1).

2.4. Molecule Structures

The molecule structures of compounds 14, 6 and 7 were determined by single crystal X-ray diffraction. The compounds 13 crystallize in the triclinic space group P-1 as isostructural compounds. Compound 4 crystallizes in the monoclinic space group I2/a. Compound 6 crystallizes with one molecule chloroform in the asymmetric unit in the triclinic space group P-1, which is not isomorphous with 13. Compound 7 crystallizes in the monoclinic space group P21/n. The basic geometric properties of the compounds 14, and 6 are explained using compound 3 as an example. One ethoxy group is bound to the antimony atom. The ethoxy group is disordered in the crystal structures of 2 and 3, and the details are described in the experimental part. The tridentate chelate ligand binds via the lateral oxygen atoms and the central donor atom (Se in compound 3) to the antimony atom. This leads formally to a tetracoordinate antimony atom (Figure 2). The antimony atom is coordinatively unsaturated and forms dimers in the solid state (Figure 3). The dimer is generated by an inversion centre. One oxygen atom (O2) from a neighbouring molecule, coordinates at the antimony atom. The distance Sb1…O2A is 2.580(2) Å in 3.
The intermolecular interaction between two molecules lead to a quadrilateral arrangement of the atoms O2-Sb1…O2A-Sb1A. The atoms are not arranged as a rectangle, but as a parallelogram with the antimony atoms in the centres of the acute angles and the oxygen atoms in the centres of the obtuse angles. For compound 3, these angles are 66.08(8)° for O2-Sb1-O2A and 113.92(8)° for Sb1-O2-Sb1A. Similar values are found for the other dimers of the compounds 1, 2, 4, and 6. The dimer formation leads formally to pentacoordinate antimony atoms. However, one should consider the presence of a free electron pair at the antimony(III) atoms. This supplements the coordination geometry of the antimony atoms and delivers an explanation for the arrangement of the dimer as a parallelogram. The free electron pairs simply need “space” at the periphery of the dimers. The orientation of the electron pair will be discussed in more detail in the quantum chemical section below.
The bond lengths Sb-O in compound 3 are 1.997(2) and 2.008(2) Å for Sb1-O1 and Sb1-O2, respectively. The bond Sb1-O3 is 1.966(2) Å. Similar values are also found in the other molecule structures. The bond to the ethoxy group is always shorter than the bonds to the oxygen atoms from the chelate ligand. These bond lengths agree well with the sum of covalent radii for Sb and O (2.07 Å).
The coordination of the central donor atom E is interesting and deserves closer examination. For this purpose, the data for the atomic distances Sb…E have been compiled in Table 2. In addition, the sums of the atomic radii and the sums of the Van der Waals radii are listed there. Figure 4 provides a graphical illustration of these distances. As can be seen from Figure 4, the distances found in the molecule structures are shorter than the sum of Van der Waals radii for all compounds. The experimental found distances are longer than the sum of covalent radii. These two facts indicate the presence of a donor–acceptor interaction between antimony and the central donor atoms E.
The distance Sb…N in compound 6 is significantly shorter than one would expect in the series of these chemical compounds. This is probably due to the bonding conditions of the nitrogen atom. In contrast to the other donor atoms with E(CH2—)2 units, the nitrogen atom already has three bonding partners within a N(Me)(CH2—)2 unit and additionally exercises the donor function to the antimony atom (Figure 5). If we compare the bond angles C2-E-C3 in the series of molecular structures (Table 3), we find that the elements of nitrogen and oxygen from the second period of the periodic table with angles around 113° are clearly more obtuse than the angles for the elements of the higher periods (99.43 to 102.4°).
The crystal structure of compound 7 gives information about what happens if we replace the tridentate dianionic chelate ligand by a bidentate dianionic ligand. Molecule 7 consists of an antimony(III) ion bound to an ethoxy group and the dianionic ligand. The bond lengths Sb-O have similar values as in the other molecule structures (2.020(2) Å for Sb1-O1 and 1.991(3) Å for Sb1-O2). The Sb-S distance is 2.456(1) Å, which corresponds well with the sum of covalent radii (2.45 Å). The antimony atom is triple-coordinated (Figure 6). This leaves a lot of space for intermolecular interactions, which are found in the crystal structure between each antimony atom and two oxygen atoms from neighbouring molecules (Figure 7). The distances Sb1…O1A and Sb1…O2B are 2.579(3) and 2.563(3) Å, respectively. These interactions make the antimony atoms pentacoordinate. The ladder-type Sb…O contacts lead to infinite chains of molecules parallel to the crystallographic b-axis.

2.5. Quantum Chemical Analysis

For further investigations of the influence of the donor element E and the dimer formation in the crystal, QTAIM (quantum theory of atoms in molecules [41,42]) analysis was performed using the structures optimized at the B97-3c [43] level of theory as a starting point for the structure optimization with the PBE0 [44,45]/def2-TZVPP [46] level. The wave function calculations were made at PBE0/def2-TZVPP with the NOECP keyword to exclude any pseudo-potential.
In all molecules, the Poincaré–Hopf relationship [41] is accomplished. There are bond critical points (BCPs) between Sb and the donor element E as well as two ring critical points (RCPs) inside the five-membered subrings of the chelate ligand-Sb ring system (see Figure 8). The topological data for 3 are shown as an example in Table 4. The data for all other compounds can be found in the Supplementary Materials. The electron density contour plot as well as a plot of the Laplacian of the electron density are shown in Figure 9.
The Sb-O bond is used for comparison with the Sb…E interactions. All BCPs between Sb and O have high electron densities and G/ρ is larger than 1 while H is negative. The graphical representation of the Laplacian (Figure 9) shows an electron depletion at the Sb atom in the direction of the O atom as well as an accumulation of charge at the O atom in the direction of the Sb atom. According to the criteria proposed by Macchi et al. [41,47], these bonds should be considered as polar shared bonds. The BCP between Sb and Se has a quite low electron density, G/ρ is nearly 1 and H > 0. The compounds 1, 2, 4, 5, and 6 feature values for G/ρ > 1 (see Supplementary Materials) and therefore all these Sb…E interactions can be characterized as closed-shell interactions (e.g., Van der Waals interactions). Furthermore, the graphical representation of the Laplacian of the electron density in Figure 9 shows the lone pairs of the Se atom as areas of valence charge concentration. One of the lone pairs shows into the direction of the Sb atom and on this atom the electron density is depleted in this direction, highlighting the nature of the interaction as Van der Waals or ionic interactions.
In addition to the topological analysis, intrinsic bond orbitals (IBO [48]) were calculated at the same level of theory to illustrate the orbital interactions. The lowest unoccupied IBO, the lone pair of Sb, and the Sb…N interaction of compound 6 are shown in Figure 10 as representative examples.
As the IBOs show, the energetically lowest unoccupied IBO of the molecule is located mostly at the Sb atom. The lone pair of the Sb atom can be described as nearly spherical, meaning a high s-orbital character, as one would expect for a fifth period element [49]. The lone pair of the N atom is directed towards the Sb atom, indicating the Sb…N interaction in this molecule.
To investigate the σ-hole at the antimony atom in 6, an electrostatic potential map was created. As can be seen in Figure 11, there is a positive potential at the antinomy atom opposite the three oxygen substituents, which can be interpreted as weak σ-hole.
The formation of dimers in the solid state is another interesting aspect of these compounds. Further quantum chemical investigations were performed to gain insight into the underlying interactions. The solid-state structure of 3 was used as a starting point and all atom positions were optimized without restrictions. The topological analysis shows five BCPs inside the Sb-O-Sb-O-parallelogram. Two indicate the intramolecular Sb-O bonds, two intermolecular Sb…O interactions and one BCP is in the middle of the parallelogram. The central BCP has a very low electron density and is considered as weak contact. The graphical representation of the electron density and the Laplacian of the electron density can be found in Figure 12 and the numeric data in Table 5. Additionally, a weak O…H interaction is found between one O atom and a nearby CH moiety.
Table 5. Results of topological analysis of the Sb-O-Sb-O-plane in the dimeric structure of 3 (all values in atomic units). Parameters are charge density (ρ), its Laplacian (∇2ρ), kinetic energy density (G), total energy density (H), and electronic potential energy density (V). The atom numbering is the same as in Figure 13, resulting from the quantum chemical calculation file (see Supplementary Materials). Only the atom numbering of one molecule is shown for clarity, where applicable.
Table 5. Results of topological analysis of the Sb-O-Sb-O-plane in the dimeric structure of 3 (all values in atomic units). Parameters are charge density (ρ), its Laplacian (∇2ρ), kinetic energy density (G), total energy density (H), and electronic potential energy density (V). The atom numbering is the same as in Figure 13, resulting from the quantum chemical calculation file (see Supplementary Materials). Only the atom numbering of one molecule is shown for clarity, where applicable.
ρ2ρGG/ρHV
Bond critical points
Sb1-Se20.0190.0870.0180.9720.004−0.014
Sb1-O30.1950.7750.2451.260−0.052−0.297
Sb1-O40.1897590.2391.267−0.050−0.289
Sb1-O170.2160.7670.2581.197−0.067−0.325
Sb1…O280.0450.2760.0601.3460.009−0.051
O4…O280.0190.0890.0201.0890.002−0.018
Ring critical points
Sb1-O3-C5-C8-Se20.0130.0530.0131.0200.000−0.013
Sb1-O4-C14-C11-Se20.0120.0520.0110.9250.001−0.010
Sb1-O4-O280.0180.1040.0221.2250.004−0.019
Sb25-O4-O280.0180.1040.0221.2250.004−0.019
As can be seen, the values for the intramolecular Sb…Se (Sb1-Se2) interaction are similar to the values in the monomer and can be characterized as closed-shell interactions. The intramolecular Sb-O bonds (Sb1-O3, Sb1-O4 and Sb1-O17) should be considered as polar shared bonds like in the monomer (see above). The electron densities in the intermolecular Sb…O interactions (Sb1…O28) are higher than in the Sb…Se (Sb1-Se2) interactions and can be characterized as closed-shell Van der Waals interactions [41,47].
For the dimer, IBO [48] calculations were performed to visualize the Sb lone pairs as well as the Sb…O and Sb…Se interactions. Figure 14c shows, that the lone pairs at antimony are mostly spherical with an orientation to the outside of the Sb-O…Sb-O… plane forcing the geometry to be a parallelogram. There are orbitals mostly located at the O atoms, which are directed towards the Sb atom of the neighbouring molecule, which conforms the intermolecular Sb…O interactions (Figure 14b). Additionally, one of the orbitals located at each of the Se atoms are directed to the Sb atoms, confirming an intramolecular interaction of the atoms Se and Sb (Figure 14a).

2.6. Reactivity of 6

Antimony occurs in two stable oxidation states in nature, +III and +V. In the presented complexes, the Sb atom should be in oxidation state +III. Therefore, there is a lone pair located at the Sb atom, and it can react as a Lewis base. Additionally, the electronegative atoms at the ligand causes σ-holes [50], and the complexes can react as Lewis acid, e.g., [51]. So, with the antimony(III) as Lewis acid and the donor atom E as a Lewis base within the same molecule, interesting reactivity could be expected. To investigate the reactivity exemplarily, compound 6 is used in further experiments, since it is soluble in most solvents like chloroform, acetone, ethanol, THF, and acetonitrile.
As nucleophiles, n-propylamine and ammonia gas were used. A nucleophilic attack to the antimony atom was expected. However, with these amines, no new species could be observed in the NMR spectra.
For redox reactions, elementary selenium and tellurium were added to the NMR sample of 6. It could be expected that these chalcogens oxidize the antimony (III) to antimony (V). However, no reaction was observed after one week at room temperature.
The reactions with electrophiles were studied with CO2 and acetic anhydride. These electrophiles can be attacked by the lone pair of the nitrogen atom or the lone pair of the antimony atom. With CO2, no reaction occurs. With acetic anhydride, new signals in the 13C NMR spectrum were found at 166.5, 170.7, 171.0, and a broad one at 178.3 ppm. These signals indicate the presence of acetic anhydride (166 ppm), ethyl acetate (171 ppm), and acetic acid (178 ppm). The signal at 170 ppm might indicate the formation of [MeN(CH2-CH2-O)2]Sb(OAc) with an acetyl instead of an ethoxy group. Sb(OAc)3 has a 13C NMR signal at 181 ppm and can therefore be excluded (see Supplementary Materials, Figures S1 and S2). Therefore, the reaction formulated in Scheme 1 can be assumed to occur in this experiment.
The reaction of 6 with acetic acid instead of acetic anhydride also leads to signals at 170 and 171 ppm in the 13C NMR spectrum, but with much lower intensity. It could be assumed that ethyl acetate is formed and a Sb-O-Sb dimer occurs as a joint product (see Scheme 2), but is not found in the NMR spectra because of the low intensity and very similar signals of the N-methyldiethanolamino ligand system.
The formation of ethyl formate is observed (signal at 161 ppm in the 13C NMR spectrum) in the reaction of 6 with formic acid. A white solid is formed in this reaction. This white solid is slightly soluble in chloroform, but more soluble in methanol. It is assumed that this solid is [MeN(CH2-CH2-O)2]Sb(O2CH) with a formyl moiety instead of an ethoxy moiety with a signal at 165 ppm in the 13C NMR spectrum (see Figure S15 in the Supplementary Materials).

3. Materials and Methods

3.1. General Considerations

All reactions were carried out under argon using the standard Schlenk technique [52,53]. Used chemicals and purification methods can be found in the Supplementary Materials. NMR spectra were recorded in CDCl3 or C6D6 with TMS as internal standard either on a BRUKER DPX 400 spectrometer at 400.13, 100.61 and 79.49 MHz for 1H, 13C and 29Si NMR spectra or on a BRUKER AVANCE III 500 MHz spectrometer at 500.13, 125.76 and 99.36 MHz for 1H, 13C and 29Si NMR spectra, respectively. Raman spectra were measured with Bruker FT-Raman spectrometer RFS 100/S. The device works with an air-cooled Nd:YAG-Laser with a wavelength of 1064 nm and a nitrogen-cooled detector. The samples were filled in glass tubes and sealed with PTFE-paste. The intensity is listed as follows: vs. (very strong) 80–100% of the highest signal, s (strong), m (medium), w (weak), vw (very weak) next 20%, respectively. Infrared spectra were obtained using a Thermo Fisher Nicolet (Waltham, MA, USA) 380 FT-IR-Spectrometer with ATR attachment. The intensities were listed as those of the Raman measurements. Boiling points of liquid samples were measured as described in [54]. Melting points were measured using a Polytherm A hot stage microscope from Wagner and Munz with an attached 52II thermometer from Fluke.

3.2. Syntheses and Characterization

3.2.1. Synthesis of Triethoxyantimony

First, 60.83 g (267 mmol) of freshly distilled SbCl3 (Merck, Darmstadt, Germany) was dissolved in 48.37 g (1.05 mol) abs. ethanol (VWR, stored over molecular sieves 3 Å). To the ice cooled solution, 95.27 g (941 mmol) triethylamine (Chemsolut, distilled from Na/benzophenone) was added. Afterward, 40 mL hexane was added, and a white solid was formed. The suspension was filtrated, and all volatiles were removed in vacuum from the filtrate. The residue is distilled to yield 55.45 g (216 mmol, 81%) of a colourless liquid. B.p.: 68 °C/0.171 mbar; 1H NMR (CDCl3, 400 MHz) δ [ppm] = 1.27 (t, 3JH,H = 7.0 Hz, 9 H, CH3), 4.05 (q, 3JH,H = 7.0 Hz, 6 H, CH2); 13C NMR (CDCl3, 100 MHz) δ [ppm] = 19.8 (CH3), 58.7 (CH2); IR ν [cm−1] = 2963.2 (w), 2918.9 (w), 2862.0 (w), 1477.3 (vw), 1441.6 (vw), 1378.9 (w), 1358.7 (vw), 1154.2 (vw), 1091.6 (m), 1033.7 (vs), 887.1 (m), 630.6 (m), 603.6 (w); Raman ν [cm−1] = 2966.8 (m), 2922.4 (vs), 2866.5 (s), 2746.9 (vw), 2704.5 (vw), 2604.2 (vw), 1476.1 (vw), 1451.0 (vw), 1383.5 (vw), 1360.4 (vw), 1283.2 (vw), 1098.1 (vw), 1063.4 (vw), 891.7 (vw), 805.0 (vw), 596.7 (vw), 324.8 (vw), 114.6 (vw).

3.2.2. Synthesis of 2-[(2-Hydroxyethyl)selanyl]ethan-1-ol/selenium-diglycol

First, 2.051 g (54.2 mmol) sodium borohydride and 2.306 g (57.6 mmol) sodium hydroxide were dissolved in about 30 mL deoxygenated water. 1.345 g (17.0 mmol) selenium was added. The suspension was stirred overnight while the selenium was dissolved to form a white suspension. Then, 3.56 g (44.2 mmol) 2-chloroethanol in 60 mL of THF was added, and the solution was stirred overnight. Two phases were formed. The organic phase was washed with saturated NaCl solution. The aqueous phase was extracted three times with 40 mL chloroform each time. The combined organic phases were dried over Na2SO4. All volatiles were removed in vacuum and a pale yellow liquid was yielded (1.05 g; 6.2 mmol; 37%).1H NMR (D2O, 400 MHz) δ [ppm] = 2.67 (t, 3JH,H = 6.7 Hz, 4 H, Se-CH2), 3.69 (t, 3JH,H = 6.7 Hz, 4 H, O-CH2), 4.69 (s, 2 H, OH); 13C NMR (D2O, 100 MHz) δ [ppm] = 25.6 (Se-CH2), 61.3 (O-CH2); 77Se NMR(D2O, 76 MHz) δ [ppm] = 83.2 ppm; IR ν [cm−1] = 3282.4 (w), 2967.1 (w), 2923.7 (w), 2868.7 (w), 1405.0 (w), 1272.9 (w), 1188.0 (vw), 1150.4 (vw), 1062.6 (m), 1039.5 (vs), 998.0 (s), 929.6 (w), 879.4 (w), 802.3 (vw), 632.6 (m), 615.2 (m), 600.7 (m); Raman ν [cm−1] = 2966.8 (w), 2928.2 (vs), 2876.1 (m), 2718.0 (vw), 1464.5 (vw), 1424.0 (vw), 1281.3 (vw), 1202.2 (vw), 1046.0 (vw), 1007.4 (vw), 934.2 (vw), 882.1 (vw), 671.9 (w), 565.8 (w), 457.8 (vw), 347.9 (vw), 268.8 (vw), 205.2 (vw), 124.2 (vw).

3.2.3. Synthesis of 2-[(2-Hydroxyethyl)tellanyl]ethan-1-ol/tellurium-diglycol

First, 2.36 g (62.4 mmol) sodium borohydride and 2.62 g (65.5 mmol) sodium hydroxide were dissolved in about 30 mL deoxygenated water and 5 mL THF. Additionally, 2.275 g (17.8 mmol) of tellurium was added. The suspension is stirred at about 40 °C for three days while the tellurium is dissolved to form a white suspension. Then, 3.337 g (41.4 mmol) 2-chloroethanol in 10 mL of THF was added, and the solution was stirred overnight. Two phases were formed. The organic phase was washed with saturated NaCl solution. The aqueous phase was extracted three times with 40 mL chloroform each time. The combined organic phases were dried over Na2SO4. All volatiles were removed in vacuum and a yellow liquid was yielded (1.49 g; 6.8 mmol; 38%).
1H NMR (D2O, 400 MHz) δ [ppm] = 2.79 (t, 3JH,H = 8 Hz, 4 H, Te-CH2), 3.81 (t, 3JH,H = 8 Hz, 4 H, O-CH2), 4.68 (s, 2 H, OH); 13C NMR (D2O, 100 MHz) δ [ppm] = 6.0 (Te-CH2), 63.6 (O-CH2); 125Te NMR(Acetone-d6, 126) δ [ppm] = 127.7 (Product), 158.0 (ditelluride) ppm; IR ν [cm−1] = 2954.6 (w), 2930.4 (w), 2918.9 (w), 2875.5 (w), 2845.6 (m), 2822.5 (m), 2685.5 (w), 1471.5 (w), 1456.1 (w), 1444.5 (w), 1397.2 (m), 1375.1 (w), 1363.5 (w), 1273.8 (m), 1263.2 (m), 1192.8 (w), 1155.2 (w), 1091.6 (w), 1067.5 (s), 1054.9 (s), 1038.5 (vs), 986.5 (s), 960.4 (s), 918.0 (s), 897.7 (s), 804.2 (w), 786.9 (m), 756.0 (m), 666.3; (w); Raman ν [cm−1] = 2970.6 (w), 2928.2 (vs), 2874.2 (m), 2746.9 (vw), 2714.1 (vw), 1466.4 (vw), 1408.6 (vw), 1354.6 (vw), 1265.9 (vw), 1155.9 (w), 1088.4 (vw), 1040.2 (vw), 999.7 (vw), 922.6 (vw), 880.2 (vw), 625.6 (m), 513.8 (w), 434.7 (w), 324.8 (w), 191.7 (vw), 155.1 (vw), 124.2 (vw).

3.2.4. Synthesis of 2-Ethoxy-1,3,6,2-trioxastibocane (1)

First, 2.13 g (20.1 mmol) diethyleneglycole was mixed with 5.13 g triethoxyantimony (20.0 mmol) in 20 mL cyclohexane. The mixture was heated until the head temperature rose to 80 °C (b.p. of cyclohexane). All volatiles were removed in vacuum and the white solid residue was recrystallized from chloroform. Yield: 4.99 g (18.4 mmol; 92%) of white crystals, suitable for single crystal X-ray diffraction.
1H NMR (CDCl3, 400 MHz) δ [ppm] = 1.27 (t, 3JH,H = 8.0 Hz, 3 H, CH3), 3.35–4.43 (m, 10 H, all other H) ; 13C NMR (CDCl3, 100 MHz) δ [ppm] = 19.9 (CH3), 58.3 (O-CH2), 61.9 (CH2 of EtO), 71.6 (CH2-O-Sb); IR ν [cm−1] = 2955.5 (vw), 2901.5 (w), 2848.5 (w), 1466.7 (vw), 1440.6 (vw), 1373.1 (w), 1359.6 (w), 1272.9 (vw), 1250.7 (vw), 1241.0 (w), 1141.7 (vw), 1119.5 (w), 1072.3 (s), 1054.0 (s), 1035.3 (vs), 1009.6 (vs), 915.1 (m), 893.9 (m), 870.7 (vs), 811.9 (m); Raman ν [cm−1] = 2989.9 (vw), 2957.1 (m), 2926.3 (vs), 2914.7 (vs), 2870.4 (s), 2854.9 (s), 2739.2 (vw), 2704.5 (vw), 2679.4 (vw), 2584.9 (vw), 2565.7 (vw), 1456.8 (vw), 1449.1 (vw), 1379.6 (vw), 1360.4 (vw), 1343.0 (vw), 1275.5 (vw), 1254.3 (vw), 1142.4 (vw), 1123.2 (vw), 1092.3 (vw), 1074.9 (vw), 1059.5 (vw), 1049.9 (vw), 1011.3 (vw), 916.8 (vw), 899.4 (vw), 876.3 (vw), 814.6 (vw), 575.5 (vw), 546.5 (vw), 509.9 (w), 469.4 (w), 432.8 (vw), 344.0 (vw), 309.3 (vw), 278.5 (vw), 241.8 (vw), 199.4 (vw), 187.8 (vw), 158.9 (vw), 116.5 (vw); m.p. = 125 °C.

3.2.5. Synthesis of 2-Ethoxy-1,3,6,2-dioxathiastibocane (2)

First, 2.33 g (19.1 mmol) 2.2′-thiodiglycol was mixed with 4.899 g triethoxyantimony (19.1 mmol) in 20 mL cyclohexane. The mixture was heated until the head temperature rose to 80 °C (b.p. of cyclohexane). All volatiles were removed in vacuum and the white slid residue was recrystallized from chloroform. Yield: 5.11 g (17.8 mmol; 93%) of white crystals, suitable for single crystal X-ray diffraction.
1H NMR (CDCl3, 500 MHz) δ [ppm] = 1.27 (t, 3JH,H = 8.0 Hz, 3 H, CH3), 2.29–2.95 (m, 4 H, S-CH3), 3.97–4.51 (m, 6 H, O-CH2 (both)); 13C NMR (CDCl3, 125 MHz) δ [ppm] = 19.9 (CH3), 37.5 (O-CH2), 58.0 (CH2-O-Sb), 62.0 (CH2 of EtO); IR ν [cm−1] = 3319.1 (vw), 2953.6 (vw), 2912.1 (vw), 2844.6 (w), 2701.9 (vw), 1458.0 (w), 1445.5 (w), 1405.9 (w), 1399.2 (w), 1369.3 (w), 1362.5 (w), 1283.5 (w), 1224.6 (w), 1194.7 (w), 1178.4 (w), 1064.6 (s), 1033.7 (vs), 998.0 (vs), 940.2 (m), 890.0 (m), 837.0 (m), 807.1 (m), 729.0 (w), 694.3 (m), 649.9 (m), 612.3 (w); Raman ν [cm−1] = 2957.1 (w), 2914.7 (vs), 2853.0 (m), 2799.0 (vw), 2737.3 (vw), 2704.5 (vw), 1449.1 (vw), 1404.7 (vw), 1368.1 (vw), 1285.1 (vw), 1231.1 (vw), 1223.4 (vw), 1196.4 (vw), 1181.0 (vw), 1096.2 (vw), 1084.6 (vw), 1042.2 (vw), 999.7 (vw), 949.6 (vw), 895.6 (vw), 841.6 (vw), 810.7 (vw), 695.0 (vw), 652.6 (vw), 590.9 (vw), 550.4 (vw), 521.5 (vw), 509.9 (vw), 484.8 (vw), 459.8 (w), 384.5 (vw), 346.0 (vw), 313.2 (vw), 299.7 (vw), 255.3 (vw), 234.1 (vw), 214.8 (vw), 199.4 (vw), 164.7 (vw), 120.3 (vw); m.p. = 132 °C.

3.2.6. Synthesis of 2-ethoxy-1,3,6,2-dioxaselenastibocane (3)

First, 1.285 g triethoxyantimony (5.0 mmol) was dissolved in 5 mL abs. ethanol. Next, 0.600 g selenium-diglycole (3.5 mmol) was added. After a few minutes, colourless crystals were obtained, suitable for single crystal X-ray diffraction. Yield: 0.963 g (2.9 mmol; 83%) of white crystals.
1H NMR (CDCl3, 500 MHz) δ [ppm] = 1.27 (t, 3JH,H = 8.0 Hz, 3 H, CH3), 2.44–2.96 (m, 4 H, Se-CH2), 4.05 (q, 3JH,H = 8.0 Hz, 2 H, CH2 of EtO) 4.11–4.64 (m, 4 H, O-CH2); 13C NMR (CDCl3, 125 MHz) δ [ppm] = 19.8 (CH3), 32.2 (O-CH2), 58.7 (CH2-O-Sb), 62.3 (CH2 of EtO); 77Se NMR (CDCl3, 76 MHz) δ [ppm] = 50.2; IR ν [cm−1] = 3278.6 (w), 2954.6 (vw), 2912.1 (w), 2841.7 (w), 2702.9 (w), 1699.1 (vw), 1455.1 (w), 1398.2 (m), 1361.6 (m), 1277.7 (m), 1211.1 (w), 1197.6 (w), 1179.3 (w), 1165.8 (w), 1062.6 (s), 1038.5 (vs), 1007.7 (vs), 976.8 (vs), 929.6 (s), 889.1 (m), 816.7 (m), 785.9 (m), 667.3 (m), 612.3 (w); Raman ν [cm−1] = 2980.3 (w), 2968.7 (w), 2922.4 (vs), 2893.5 (w), 2851.1 (m), 2791.3 (vw), 2706.4 (vw), 1449.1 (vw), 1402.8 (vw), 1370.0 (vw), 1289.0 (vw), 1277.4 (vw), 1213.8 (vw), 1196.4 (vw), 1179.1 (vw), 1082.7 (vw), 1049.9 (vw), 1015.2 (vw), 976.6 (vw), 936.1 (vw), 789.5 (vw), 671.9 (vw), 589.0 (vw), 575.5 (w), 500.3 (w), 469.4 (vw), 452.0 (w), 374.9 (vw), 320.9 (vw), 284.3 (vw), 257.3 (vw), 216.8 (vw), 199.4 (vw), 157.0 (vw), 116.5 (vw); m.p. = 109 °C.

3.2.7. Synthesis of 2-Ethoxy-1,3,6,2-dioxatellurastibocane (4)

First, 1.54 g triethoxyantimony (6 mmol) was dissolved in 5 mL abs. ethanol. Next, 1.49 g tellurium-diglycole (6.8 mmol) was added. After a few minutes, colourless crystals were obtained, suitable for single crystal X-ray diffraction. The solid was filtrated and washed 3 times with 2 mL of ethanol. After removing volatiles in vacuum, 0.91 g (2.4 mmol; 40%) of a white powder was yielded. Crystals in the solution decompose in a few months to a black residue and a yellowish solution.
1H NMR (CDCl3, 400 MHz) δ [ppm] = 1.72 (t, 3JH,H = 6.9 Hz, 3 H, CH3), 2.74 (m, 4 H, Te-CH2), 4.06 (q, 3JH,H = 7.0 Hz, 2 H, CH2 of EtO), 4.19–4.82 (m, 4 H, O-CH2); 13C NMR (CDCl3, 100 MHz) δ [ppm] = 16.0 (CH3), 19.9 (O-CH2), 58.4 (CH2-O-Sb), 64.1 (CH2 of EtO); 125Te NMR (CDCl3, 126 MHz) δ [ppm] = 1.3 (br); IR ν [cm−1] = 2954.6 (w), 2930.4 (w), 2918.9 (w), 2875.5 (w), 2845.6 (m), 2822.5 (m), 2685.5 (w), 1471.5 (w), 1456.1 (w), 1444.5 (w), 1397.2 (m), 1375.1 (w), 1363.5 (w), 1273.8 (m), 1263.2 (m), 1192.8 (w), 1155.2 (w), 1091.6 (w), 1067.5 (s), 1054.9 (s), 1038.5 (vs), 986.5 (s), 960.4 (s), 918.0 (s), 897.7 (s), 804.2 (w), 786.9 (m), 756.0 (m), 666.3 (w); Raman ν [cm−1] = 2976.4 (m), 2920.5 (vs), 2853.0 (s), 2835.6 (s), 2702.6 (w), 1454.9 (vw), 1397.0 (vw), 1362.3 (vw), 1267.8 (vw), 1192.6 (vw), 1175.2 (vw), 1078.8 (vw), 631.4 (vw), 598.6 (vw), 535.0 (w), 494.5 (m), 469.4 (s), 444.3 (w), 423.1 (vw), 322.8 (w), 278.5 (vw), 189.8 (m), 143.5 (w); m.p. = 105 °C (decomp.).

3.2.8. Synthesis of 2-Ethoxy-1,3,6,2-dioxazastibocane (5)

First, 2.20 g (20.9 mmol) diethanolamine was mixed with 4.988 g triethoxyantimony (19.4 mmol) in 20 mL cyclohexane. The mixture was heated until the head temperature rose to 80 °C (b.p. of cyclohexane). All volatiles were removed in vacuum and the white solid residue was recrystallized from chloroform. Yield: 4.82 g (17.9 mmol; 92%) of white powder, slightly contaminated with a residue of cyclohexane.
1H NMR (CDCl3, 400 MHz) δ [ppm] = 1.29–1.52 (m, 8 H, CH3 and cyclohexane), 1.89–2.73 (m, 4 H, N-CH3), 3.64–4.15 (m, 4 H, O-CH2), 4.36 (q, 3JH,H = 8.0 Hz, 2 H, O-CH2 of EtO); 13C NMR (CDCl3, 100 MHz) δ [ppm] = 21.3 (CH3), 57.8 (N-CH2), 63.1 (CH2 of EtO), 78.1 (CH2-O); IR ν [cm−1] = 3203.3 (w), 2898.6 (w), 2839.8 (m), 2694.2 (w), 1446.4 (w), 1358.7 (w), 1125.3 (w), 1087.7 (m), 1043.4 (vs), 1013.5 (vs), 916.1 (m), 871.7 (s), 819.6 (w), 667.3 (m), 653.8 (m), 612.3 (m); Raman ν [cm−1] = 2916.6 (vs), 2858.8 (vs), 2733.4 (vw), 2696.8 (vw), 1452.9 (w), 1368.1 (vw), 1350.7 (vw), 1263.9 (vw), 1240.8 (vw), 1150.2 (vw), 1125.1 (vw), 1096.2 (vw), 1063.4 (vw), 1017.1 (vw), 920.7 (vw), 876.3 (vw), 820.4 (vw), 801.1 (vw), 531.1 (s), 450.1 (vw), 419.3 (vw), 388.4 (vw), 293.9 (vw), 205.2 (w), 114.6 (vw); m.p. could not be determined, because of waxy consistency.

3.2.9. Synthesis of 2-Ethoxy-6-methyl-1,3,6,2-dioxazastibocane (6)

First, 2.668 g (22.4 mmol) N-methyldiethanlamine were mixed with 5.801 g triethoxyantimony (22.6 mmol) in 20 mL cyclohexane. The mixture was heated until the head temperature rose to 80 °C (b.p. of cyclohexane). All volatiles were removed in vacuum and the white slid residue was recrystallized from chloroform. Yield: 5.770 g (20.3 mmol; 90%) of white crystals, suitable for single crystal X-ray diffraction.
1H NMR (CDCl3, 400 MHz) δ [ppm] = 1.18–1.29 (m, 3 H, CH3), 2.54 (s, 3 H, N-CH3), 2.56–2.85 (m, 4 H, N-CH2), 3.94–4.32 (m, 6 H, O-CH2(both)); 13C NMR (CDCl3, 100 MHz) δ [ppm] = 20.2 (CH3), 41.7 (CH3-N), 57.5 (CH2 of EtO and N-CH2), 61.7 (CH2-O); IR ν [cm−1] = 2946.8 (vw), 2893.8 (vw), 2859.1 (vw), 2825.3 (vw), 2683.6 (vw), 1460.9 (vw), 1449.3 (w), 1416.5 (w), 1373.1 (w), 1361.6 (w), 1352.9 (vw), 1338.4 (vw), 1287.3 (vw), 1266.1 (w), 1259.4 (w), 1237.2 (vw), 1204.4 (vw), 1161.0 (vw), 1142.7 (vw), 1067.5 (s), 1042.4 (vs), 1030.8 (vs), 991.3 (s), 913.2 (w), 879.4 (vs), 870.7 (s), 757.0 (m), 675.0 (w); Raman ν [cm−1] = 2961.0 (s), 2951.4 (s), 2907.0 (vs), 2862.6 (vs), 2837.6 (vs), 2808.6 (w), 2733.4 (vw), 2712.2 (vw), 2694.9 (vw), 2669.8 (vw), 1470.3 (w), 1454.9 (vw), 1418.2 (vw), 1375.8 (vw), 1350.7 (vw), 1289.0 (vw), 1262.0 (vw), 1238.9 (vw), 1206.1 (vw), 1163.7 (vw), 1144.4 (vw), 1090.4 (vw), 1071.1 (vw), 1049.9 (vw), 1034.4 (vw), 995.9 (vw), 914.9 (vw), 885.9 (vw), 872.5 (vw), 758.7 (vw), 560.0 (vw), 521.5 (w), 500.3 (w), 440.5 (vw), 401.9 (vw), 361.4 (vw), 303.5 (vw), 272.7 (vw), 228.3 (vw), 189.8 (vw), 137.7 (vw); m.p. = 82 °C.

3.2.10. Synthesis of 2-Ethoxy-1,4-oxathiastibpentan (7)

First, 1.583 g (6.2 mmol) triethoxyantimony was dissolved in 5 mL ethanol and 0.520 g (6.6 mmol) 2-mercaptoethanol was added. Within 2 h, white crystals were formed. The solid was filtrated and washed three times with fresh 2 mL ethanol. After removing all volatiles in vacuum, 0.85 g (3.5 mmol; 56%) white crystals, suitable for single crystal X-ray diffraction, were received.
1H NMR (CDCl3, 400 MHz) δ [ppm] = 1.27 (t, 3JH,H = 6.96 Hz, 3 H, CH3), 3.10 (m, 2 H, S-CH2), 4.01 (q, 3JH,H = 6.93 Hz, 2 H, CH2 of EtO); 4.39 (m, 2 H, O-CH2); 13C NMR (CDCl3, 100 MHz) δ [ppm] = 19.7 (CH3), 35.0 (S-CH2), 59.5 (CH2 of EtO), 71.5 (br, CH2-O); IR ν [cm−1] = 2910.4 (w), 2844.6 (m), 1456.1 (w), 1414.6 (w), 1378.0 (w), 1349.0 (w), 1269.0 (m), 1206.3 (w), 1170.6 (w), 1089.6 (w), 1033.7 (vs), 1000.0 (vs), 932.5 (m), 879.4 (s), 841.8 (s), 754.1 (w), 660.2 (s), 609.4 (m); Raman ν [cm−1] = 2949.4 (m), 2914.7 (vs), 2853.0 (m), 2824.1 (vw), 2746.9 (vw), 2704.5 (vw), 2685.2 (vw), 1456.8 (vw), 1449.1 (vw), 1422.1 (vw), 1352.6 (vw), 1271.6 (vw), 1211.9 (vw), 1171.4 (vw), 1094.2 (vw), 1046.0 (vw), 1009.4 (vw), 939.9 (vw), 849.3 (vw), 666.1 (vw), 569.7 (vw), 540.8 (vw), 471.3 (w), 459.8 (w), 419.3 (w), 349.8 (m), 303.5 (vw), 282.3 (m), 228.3 (vw), 195.6 (vw), 164.7 (vw), 131.9 (w), 108.8 (vw); m.p. = 93 °C.

3.3. Reactivity Studies

First, 0.73 g 6 was dissolved in 4.672 g CDCl3. To 0.5 mL of this solution were added in four different NMR tubes:
(1)
0.075 g n-Propylamine;
(2)
0.121 g formic acid and a white residue was formed;
(3)
Tip of a spatula elementar tellurium;
(4)
Tip of a spatula elementar selenium.
Next, 0.52 g 6 was dissolved in 1.467 g acetonitril-d3. Through 0.5 mL of this solution, we introduced into separate NMR tubes:
(5)
CO2 (Linde, 5.3);
(6)
NH3 (Nippon Gases, 4.0).
Then, 0.56 g 6 was dissolved in 1.756 g CDCl3. To 0.5 mL of this solution, we added in separate NMR tubes:
(7)
0.2 mL acetic acid (freshly distilled from Ac2O);
(8)
0.2 mL acetic anhydride (freshly distilled).

3.4. Crystal Structure Analyses

Single crystal X-ray diffraction data of 14, 6, and 7 were collected on a STOE IPDS-II image plate diffractometer equipped with a low-temperature device with Mo-Kα radiation (λ = 0.71073 Å) using ω and φ scans. Crystal data and details of structure refinement are summarized in Table 6 and Table 7. Software for data collection was X-AREA, for cell refinement, X-AREA, and for data reduction, X-RED [55]. Preliminary structure models were derived by direct methods [56] and the structures were refined by full-matrix least-squares calculations based on F2 for all reflections using SHELXL [57]. All hydrogen atoms were included in the models in calculated positions and were refined as constrained to the bonded atoms. The ethoxy groups in the crystal structures of 2 and 3 are disordered and were refined with split atom models for the atoms C5 and C6. Parts A and B of the disordered ethyl groups have site occupation factors of 0.72/0.28 in compound 2 and 0.78/0.22 in compound 3.
The crystal under investigation of compound 1 was twinned with three components (second component with 40.9%, third component with 1.8%). Overlap from the second and third domain causes errors in the intensities of some reflections, leading to spurious peaks and holes of residual electron density.
Compound 6 crystallizes with one molecule chloroform in the asymmetric unit.
The compounds form intermolecular interactions in the crystal lattice via Sb1…O1 and Sb1…O2 interactions. This leads to four membered rings of Sb1…O…Sb1A…O and causes short distances between neighbouring oxygen atoms. This leads to alert level A or B warnings in the CheckCIF routine [58]. These short contacts, O1…O1 and O2…O2, should not be considered as an error, but they are an intrinsic property of these solid state structures.
CCDC 2263270-2263275 contains the supplementary crystallographic data for this paper. These data can be obtained free of charge from The Cambridge Crystallographic Data Centre via www.ccdc.cam.ac.uk/data_request/cif.

3.5. Quantum Chemical Calculations

Initial geometries were optimized using the B97-3c [43] method as implemented in the ORCA software package version 5 [59]. The calculation of Hessian matrices verified the presence of local minima with zero imaginary frequencies. The vibration frequencies (infrared and Raman spectra) were calculated with the same method and scaled according to [60]. The B97-3c method is recommended because of the good cost–benefit-ratio for structure optimization and frequency calculations [61]. Additionally, a benchmark of the accuracy of the IR calculations performed with this method was made by Katysuba et al. showing that this method offers similarly good results as the well-recognized B3LYP functional [60]. The data of the experimental molecular structures (see Section 2.4) were used as a starting point for the geometry optimization. As no experimental molecular structure of 5 was obtained, the data of 6 were adopted and used. The RMSD of the optimized and experimental structures on all non-hydrogen atoms was calculated using the curcuma software package [62] and are below 0.4 Å (see Supplementary Materials).
Since quantum chemical calculations cannot output band widths and band forms for vibrations, a 10 cm−1 line width with a Gaussian line shape is assumed for comparison. Considering this, the fingerprints of experimental and calculated spectra look quite similar.
The atom coordinates of the optimized molecules can be found in the Supplementary Materials. The wave function for QTAIM calculations were performed at the PBE0 [44,45]/def2-TZVPP [46] level of theory with the NOECP option to exclude any pseudopotentials. The topological analysis was performed with the program Multiwfn [63]. IBO calculations were performed at PBE0/def2-TZVPP level of theory with the program IBOview [48].

4. Conclusions

A simple and straightforward method for the synthesis of antimony(III) complexes with tridentate O,E,O-ligands was developed. The complexes were synthesized from triethoxyantimony with the ligand in ethanol in good yields. Nitrogen and chalcogen atoms were used as donor atom E in the ligand system. In the Raman spectra, a ν(Sb-E) can be found as predicted with quantum mechanical methods.
The single crystal structure determination of compounds 14, and 6 confirms the formation of chelate complexes, wherein the ligand molecules coordinate as tridentate O,E,O-ligands to the antimony atom. Furthermore, the formation of dimers in the solid state was observed in all these crystal structures. This allows the antimony atoms to reach pentacoordination. Compound 7 bears a bidentate ligand leading to the molecular composition (O-CH2-CH2-S)Sb(OEt). This triple-coordinated complex compensates for the coordinative unsaturation in the crystalline state by associating with two neighbouring molecules each and thus also achieves the fivefold coordination of antimony.
According to QTAIM calculations, there are polar shared bonds between antimony and the oxygen atoms of the ethoxy groups and between antimony and the lateral oxygen atoms of the chelate ligands. The interactions between the donor atom E and the Sb atom can be characterized as Van der Waals interactions. This could be confirmed by IBO analysis, which shows a displacement of orbitals located at the donor atom E towards the Sb atom. The formation of dimers in the solid state is observed and analysed. There are Sb…O Van der Waals interactions according to QTAIM and IBO analysis. Additionally, the IBO analysis shows that the lone pair of the antimony is located spherically around the Sb atom, which indicates a high s-character of this orbital.
Despite the lone pair at the antimony atom and the Lewis acid properties caused by the σ hole formation at the same atom [64], no reactions with nucleophiles were observed at room temperature. An oxidation of antimony(III) with selenium or tellurium was also not observed. The complexes react with acetic anhydride and more slowly with carboxylic acids, whereby the ethoxy moiety is substituted.
Further investigations on these complexes are part of the research in our group.

Supplementary Materials

The following Supplementary Materials can be downloaded at: https://www.mdpi.com/article/10.3390/molecules28134959/s1, Statement of origin and purification method of used chemicals; Table S1. NMR spectra (1H and 13C{1H}) of CDCl3 solutions of Sb(OAc)3 and 2; Figures S1–S4. IR and Raman spectra of 1 and 6; Figures S5 and S6. Additional figures of the molecular structures; Figures S7–S14. Further data of the topological analysis of compounds 1, 2, 4, 5, 6; Section S7. The atomic coordinates of all quantum chemically optimized structures; Section S8. The RMSD of the experimental and optimized structures; Section S9. The 13C NMR spectrum of the reaction product of 6 with formic acid; Figure S15.

Author Contributions

M.H.: Conceptualization, formal analysis, investigation, data curation, writing—original draft, writing—review and editing, visualization; U.B.: formal analysis, investigation, data curation, writing—original draft, writing—review and editing, visualization. All authors have read and agreed to the published version of the manuscript.

Funding

The authors thank TU Bergakademie Freiberg (Freiberg, Germany) for financial support. The authors acknowledge computing time on the compute cluster of the Faculty of Mathematics and Computer Science of Technische Universität Bergakademie Freiberg, operated by the computing center (URZ) and funded by the Deutsche Forschungsgemeinschaft (DFG) under DFG grant number 397252409.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are contained within the article or Supplementary Materials.

Acknowledgments

The authors gratefully acknowledge help from Beate Kutzner (NMR) and Regina Moßig (Raman). Erica Brendler (Institute for Analytical Chemistry, TU Bergakademie Freiberg) is acknowledged for help with the NMR measurements. The authors thank Jörg Wagler for help with the data treatment of the three-component twin (“Drilling”) of the crystal structure of 1 and Conrad Hübler for help with the quantum chemical calculations.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Not applicable.

References

  1. Harder, S. (Ed.) Early Main Group Metal Catalysis; Wiley-VCH Verlag GmbH & Co. KgaA: Weinheim, Germany, 2020. [Google Scholar] [CrossRef]
  2. Stewart, C.A.; Harlow, R.L.; Arduengo, A.J. Chemistry and structure of the first 10-Sb-3 species. J. Am. Chem. Soc. 1985, 107, 5543–5544. [Google Scholar] [CrossRef]
  3. Arduengo, A.J.; Stewart, C.A.; Davidson, F.; Dixon, D.A.; Becker, J.Y.; Culley, S.A.; Mizen, M.B. The synthesis, structure, and chemistry of 10-Pn-3 systems: Tricoordinate hypervalent pnictogen compounds. J. Am. Chem. Soc. 1987, 109, 627–647. [Google Scholar] [CrossRef]
  4. Lipshultz, J.M.; Li, G.; Radosevich, A.T. Main Group Redox Catalysis of Organopnictogens: Vertical Periodic Trends and Emerging Opportunities in Group 15. J. Am. Chem. Soc. 2021, 143, 1699–1721. [Google Scholar] [CrossRef] [PubMed]
  5. Marczenko, K.M.; Zurakowski, J.A.; Bamford, K.L.; MacMillan, J.W.M.; Chitnis, S.S. Hydrostibination. Angew. Chem. Int. Ed. 2019, 58, 18096–18101. [Google Scholar] [CrossRef]
  6. Tofan, D.; Gabbaï, F.P. Fluorinated antimony(V) derivatives: Strong Lewis acidic properties and application to the complexation of formaldehyde in aqueous solutions. Chem. Sci. 2016, 7, 6768–6778. [Google Scholar] [CrossRef] [Green Version]
  7. Ke, I.-S.; Myahkostupov, M.; Castellano, F.N.; Gabbaï, F.P. Stibonium Ions for the Fluorescence Turn-On Sensing of F in Drinking Water at Parts per Million Concentrations. J. Am. Chem. Soc. 2012, 134, 15309–15311. [Google Scholar] [CrossRef]
  8. Hirai, M.; Gabbaï, F.P. Lewis acidic stiborafluorenes for the fluorescence turn-on sensing of fluoride in drinking water at ppm concentrations. Chem. Sci. 2014, 5, 1886–1893. [Google Scholar] [CrossRef] [Green Version]
  9. Li, L.; Zhang, Y.; Li, Y.; Duan, Y.; Qian, Y.; Zhang, P.; Guo, Q.; Ding, J. Polymeric Membrane Fluoride-Selective Electrodes Using Lewis Acidic Organo-Antimony(V) Compounds as Ionophores. ACS Sens. 2020, 5, 3465–3473. [Google Scholar] [CrossRef] [PubMed]
  10. Akiba, K.-y.; Ohnari, H.; Ohkata, K. Oxidation of alpha-Hydroxyketones with Triphenylantimony Dibromide and its Catalytic Cycle. Chem. Lett. 1985, 14, 1577–1580. [Google Scholar] [CrossRef] [Green Version]
  11. Huang, Y.; Shen, Y.; Chen, C. Bromodiphenylstibine-Mediated Oxidation of Benzyl Alcohols by Bromine. Synthesis 1985, 1985, 651–652. [Google Scholar] [CrossRef]
  12. Yang, M.; Pati, N.; Bélanger-Chabot, G.; Hirai, M.; Gabbaï, F.P. Influence of the catalyst structure in the cycloaddition of isocyanates to oxiranes promoted by tetraarylstibonium cations. Dalton Trans. 2018, 47, 11843–11850. [Google Scholar] [CrossRef] [PubMed]
  13. Yang, M.; Tofan, D.; Chen, C.-H.; Jack, K.M.; Gabbaï, F.P. Digging the Sigma-Hole of Organoantimony Lewis Acids by Oxidation. Angew. Chem. Int. Ed. 2018, 57, 13868–13872. [Google Scholar] [CrossRef] [PubMed]
  14. Poddel’sky, A.I.; Smolyaninov, I.V.; Shataeva, A.I.; Baranov, E.V.; Fukin, G.K. Binuclear Triphenylantimony(V) Catecholates through N-Donor Linkers: Structural Features and Redox Properties. Molecules 2022, 27, 6484. [Google Scholar] [CrossRef]
  15. Lizarazo-Jaimes, E.; Monte-Neto, R.; Reis, P.; Fernandes, N.; Speziali, N.; Melo, M.; Frézard, F.; Demicheli, C. Improved Antileishmanial Activity of Dppz through Complexation with Antimony(III) and Bismuth(III): Investigation of the Role of the Metal. Molecules 2012, 17, 12622–12635. [Google Scholar] [CrossRef] [Green Version]
  16. de Oliveira, L.G.; Silva, M.M.; de Paula, F.C.S.; Pereira-Maia, E.C.; Donnici, C.L.; de Simone, C.A.; Frézard, F.; da Silva Júnior, E.N.; Demicheli, C. Antimony(V) and Bismuth(V) Complexes of Lapachol: Synthesis, Crystal Structure and Cytotoxic Activity. Molecules 2011, 16, 10314–10323. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  17. Islam, A.; Silva, J.D.; Berbet, F.; da Silva, S.; Rodrigues, B.; Beraldo, H.; Melo, M.; Frézard, F.; Demicheli, C. Novel Triphenylantimony(V) and Triphenylbismuth(V) Complexes with Benzoic Acid Derivatives: Structural Characterization, in Vitro Antileishmanial and Antibacterial Activities and Cytotoxicity against Macrophages. Molecules 2014, 19, 6009–6030. [Google Scholar] [CrossRef] [Green Version]
  18. Ferreira, W.; Islam, A.; Andrade, A.; Fernandes, F.; Frézard, F.; Demicheli, C. Mixed Antimony(V) Complexes with Different Sugars to Modulate the Oral Bioavailability of Pentavalent Antimonial Drugs. Molecules 2014, 19, 5478–5489. [Google Scholar] [CrossRef] [Green Version]
  19. Adeyemi, J.O.; Onwudiwe, D.C. Chemistry and Some Biological Potential of Bismuth and Antimony Dithiocarbamate Complexes. Molecules 2020, 25, 305. [Google Scholar] [CrossRef] [Green Version]
  20. Kubiak, R.; Janczak, J. Synthesis, Structure, and UV/Vis Characterization of Antimony(III) Phthalocyanine: [(SbPc)2(Sb2I8)(SbBr3)]2. Molecules 2022, 27, 1839. [Google Scholar] [CrossRef]
  21. Sankhla, B.S.; Kapoor, R.N. Mono-, Di- and Tri-ethanolamine Derivatives of Samarium. Bull. Chem. Soc. Jpn. 1967, 40, 1381–1383. [Google Scholar] [CrossRef] [Green Version]
  22. Davies, C.W.; Patel, B.N. Complexes of the cupric ion with mono-, di-, and tri-ethanolamine. J. Chem. Soc. A 1968, 1968, 1824–1828. [Google Scholar] [CrossRef]
  23. Gharia, K.S.; Singh, M.; Mathur, S.; Sankhla, B.S. Ethanolamine Derivatives of Lanthanons(III). Synth. React. Inorg. Met.-Org. Chem. 1980, 10, 403–416. [Google Scholar] [CrossRef]
  24. Ghadwal, R.S.; Singh, A. Synthesis and Spectroscopic Characterisation of a New Class of Heterobimetallic Homoleptic Diethanolaminate Complexes of Niobium(V) and Tantalum(V). J. Chem. Res. 2006, 2006, 451–455. [Google Scholar] [CrossRef]
  25. Sharma, K.; Sharma, M.; Singh, A.; Mehrotra, R.C. Synthesis and Characterization of Homo- and Heterobimetllic Tri- and Di-Ethanolaminate Derivatives Containing Dibutyltin(IV) Moiety. Main Group Met. Chem. 2003, 26, 131–140. [Google Scholar] [CrossRef]
  26. Verkade, J.G. Main group atranes: Chemical and structural features. Coord. Chem. Rev. 1994, 137, 233–295. [Google Scholar] [CrossRef]
  27. Puri, J.K.; Singh, R.; Chahal, V.K. Silatranes: A review on their synthesis, structure, reactivity and applications. Chem. Soc. Rev. 2011, 40, 1791–1840. [Google Scholar] [CrossRef]
  28. Srivastav, N.; Mutneja, R.; Singh, N.; Singh, R.; Kaur, V.; Wagler, J.; Kroke, E. Diverse Molecular Architectures of Si and Sn [4.4.3.01,6]Tridecane Cages Derived from a Mannich Base Possessing Semi-Rigid Unsymmetrical Podands. Eur. J. Inorg. Chem. 2016, 2016, 1730–1737. [Google Scholar] [CrossRef]
  29. Sidorkin, V.F.; Belogolova, E.F.; Wang, Y.; Jouikov, V.; Doronina, E.P. Electrochemical Oxidation and Radical Cations of Structurally Non-rigid Hypervalent Silatranes: Theoretical and Experimental Studies. Chem.-Eur. J. 2017, 23, 1910–1919. [Google Scholar] [CrossRef]
  30. Gupta, A.K.S.; Bohra, R.; Mehrotra, R.C. Heterocyclic compounds containing antimony: Part 3-Synthesis and spectral studies of (PriO)Sb[O(CH2)2X(CH2)2O] and [EO(CH2)2X(CH2)2O] Sb[O(CH2)2X(CH2)2O] (E=H, Li, Na or K; and X=O or S). Indian J. Chem. Sect. A Inorg. Phys. Theor. Anal. 1991, 30A, 588–591. [Google Scholar]
  31. Voronkov, M.G.; Pestunovich, V.A.; Albanov, A.I.; Kuznetsova, G.A.; Selbst, É.A.; Baryshok, V.P. 1-Fluoro-2-hydrostibatrane. J. Struct. Chem. 2006, 47, 527–531. [Google Scholar] [CrossRef]
  32. Ladilina, E.Y.; Semenov, V.V.; Fukin, G.K.; Gushchin, A.V.; Dodonov, V.A.; Zhdanovich, I.V.; Finet, J.-P. One-step synthesis of pentavalent triphenylantimony derivatives Ph3Sb(OSiR3)2, Ph3Sb(OCH2CH2)2NH and Ph3Sb(OCH2CH2NMe2)2: X-ray molecular structure of Ph3Sb(OSiMe3)2. J. Organomet. Chem. 2007, 692, 5701–5708. [Google Scholar] [CrossRef]
  33. Moaven, S.; Villanueva, O.H.; Unruh, D.K.; Cozzolino, A.F. The complicating role of pnictogen bond formation in the solution-phase and solid-state structures of the heavier pnictogen atranes. Dalton Trans. 2022, 51, 11335–11339. [Google Scholar] [CrossRef]
  34. Milton, M.D.; Khan, S.; Singh, J.D.; Mishra, V.; Khandelwal, B.L. A facile access to chalcogen and dichalcogen bearing dialkylamines and diols. Tetrahedron Lett. 2005, 46, 755–758. [Google Scholar] [CrossRef]
  35. Herrmann, H.; Bucksch, H. Butter of antimony. In Dictionary Geotechnical Engineering/Wörterbuch GeoTechnik; Springer: Berlin/Heidelberg, Germany, 2014; p. 179. [Google Scholar] [CrossRef]
  36. Goel, R.G.; Maslowsky, E.; Senoff, C.V. Organoantimony compounds. III. Far-infrared and Raman spectroscopic studies on triorganoantimony(V) derivatives, R3SbX2. Inorg. Chem. 1971, 10, 2572–2577. [Google Scholar] [CrossRef]
  37. Kolondra, W.; Schwarz, W.; Weidlein, J. Spektroskopische und röntgenographische Untersuchungen an methylenverbrückten Organoantimonverbindungen. Z. Anorg. Allg. Chem. 1983, 501, 137–145. [Google Scholar] [CrossRef]
  38. Lopes, L.J.S.; Guerra, A.C.O.; Comerlato, N.M.; Turci, C.C.; Ferreira, G.B. Vibrational and Electronic Spectroscopy of Neutral Antimony Coordination Compounds of the 1,3-Dithiole-2-thione-4,5-dithiolate (dmit). J. Phys. Chem. A 2012, 116, 2244–2260. [Google Scholar] [CrossRef]
  39. Pauling, L. Die Natur der Chemischen Bindung; Verl. Chemie: Weinheim, Germany, 1962. [Google Scholar]
  40. Mantina, M.; Chamberlin, A.C.; Valero, R.; Cramer, C.J.; Truhlar, D.G. Consistent van der Waals Radii for the Whole Main Group. J. Phys. Chem. A 2009, 113, 5806–5812. [Google Scholar] [CrossRef] [Green Version]
  41. Bader, R.F.W. Atoms in Molecules; Clarendon Press: Oxford, UK, 1994. [Google Scholar]
  42. Kumar, P.S.V.; Raghavendra, V.; Subramanian, V. Bader’s Theory of Atoms in Molecules (AIM) and its Applications to Chemical Bonding. J. Chem. Sci. 2016, 128, 1527–1536. [Google Scholar] [CrossRef] [Green Version]
  43. Brandenburg, J.G.; Bannwarth, C.; Hansen, A.; Grimme, S. B97-3c: A revised low-cost variant of the B97-D density functional method. J. Chem. Phys. 2018, 148, 064104. [Google Scholar] [CrossRef]
  44. Perdew, J.P.; Ernzerhof, M.; Burke, K. Rationale for mixing exact exchange with density functional approximations. J. Chem. Phys. 1996, 105, 9982–9985. [Google Scholar] [CrossRef]
  45. Adamo, C.; Barone, V. Toward reliable density functional methods without adjustable parameters: The PBE0 model. J. Chem. Phys. 1999, 110, 6158–6170. [Google Scholar] [CrossRef]
  46. Weigend, F.; Ahlrichs, R. Balanced basis sets of split valence, triple zeta valence and quadruple zeta valence quality for H to Rn: Design and assessment of accuracy. Phys. Chem. Chem. Phys. 2005, 7, 3297. [Google Scholar] [CrossRef] [PubMed]
  47. Macchi, P.; Proserpio, D.M.; Sironi, A. Experimental Electron Density in a Transition Metal Dimer: Metal-Metal and Metal-Ligand Bonds. J. Am. Chem. Soc. 1998, 120, 13429–13435. [Google Scholar] [CrossRef]
  48. Knizia, G. Intrinsic Atomic Orbitals: An Unbiased Bridge between Quantum Theory and Chemical Concepts. J. Chem. Theory Comput. 2013, 9, 4834–4843. [Google Scholar] [CrossRef] [Green Version]
  49. Janiak, C.; Meyer, H.-J.; Gudat, D.; Kurz, P. Riedel Moderne Anorganische Chemie; De Gruyter: Berlin, Germany, 2018. [Google Scholar] [CrossRef]
  50. Benz, S.; Poblador-Bahamonde, A.I.; Low-Ders, N.; Matile, S. Catalysis with Pnictogen, Chalcogen, and Halogen Bonds. Angew. Chem. Int. Ed. 2018, 57, 5408–5412. [Google Scholar] [CrossRef] [Green Version]
  51. Qiu, J.; Song, B.; Li, X.; Cozzolino, A.F. Solution and gas phase evidence of anion binding through the secondary bonding interactions of a bidentate bis-antimony(III) anion receptor. Phys. Chem. Chem. Phys. 2018, 20, 46–50. [Google Scholar] [CrossRef]
  52. Herzog, S.; Dehnert, J. Eine rationelle anaerobe Arbeitsmethode. Z. Chem. 1964, 4, 1–11. [Google Scholar] [CrossRef]
  53. Böhme, U. Inertgastechnik; De Gruyter: Berlin, Germany; Boston, MA, USA, 2020. [Google Scholar] [CrossRef]
  54. Herbig, M.; Kroke, E. Low cost apparatus for rapid boiling point determination of small air sensitive samples under inert atmosphere. Thermochim. Acta 2017, 654, 81–84. [Google Scholar] [CrossRef]
  55. STOE & Cie GmbH. X-RED (Version 1.53) and X-AREA, Version 1.55; STOE & Cie GmbH: Darmstadt, Germany, 2009. [Google Scholar]
  56. Sheldrick, G.M. A short history of SHELX. Acta Crystallogr. Sect. A Found. Crystallogr. 2007, 64, 112–122. [Google Scholar] [CrossRef] [Green Version]
  57. Sheldrick, G.M. Crystal structure refinement with SHELXL. Acta Crystallogr. Sect. C Struct. Chem. 2015, 71, 3–8. [Google Scholar] [CrossRef] [Green Version]
  58. Spek, A.L. CheckCIF validation ALERTS: What they mean and how to respond. Acta Crystallogr. Sect. E Crystallogr. Commun. 2020, 76, 1–11. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  59. Neese, F. Software update: The ORCA program system—Version 5.0. WIREs Comput. Mol. Sci. 2022, 12, e1606. [Google Scholar] [CrossRef]
  60. Katsyuba, S.A.; Zvereva, E.E.; Grimme, S. Fast Quantum Chemical Simulations of Infrared Spectra of Organic Compounds with the B97-3c Composite Method. J. Phys. Chem. A 2019, 123, 3802–3808. [Google Scholar] [CrossRef] [PubMed]
  61. Bursch, M.; Mewes, J.-M.; Hansen, A.; Grimme, S. Best-Practice DFT Protocols for Basic Molecular Computational Chemistry. Angew. Chem. Int. Ed. Engl. 2022, 61, e202205735. [Google Scholar] [CrossRef]
  62. Hübler, C. Conradhuebler/Curcuma: Curcuma. 2020. Available online: https://zenodo.org/record/4302723 (accessed on 6. June 2023). [CrossRef]
  63. Lu, T.; Chen, F. Multiwfn: A multifunctional wavefunction analyzer. J. Comput. Chem. 2011, 33, 580–592. [Google Scholar] [CrossRef] [PubMed]
  64. Rodrigues, R.R.; Gabbaï, F.P. Structural Evidence for Pnictogen-Centered Lewis Acidity in Cationic Platinum-Stibine Complexes Featuring Pendent Amino or Ammonium Groups. Molecules 2021, 26, 1985. [Google Scholar] [CrossRef]
Figure 1. Synthesized compounds.
Figure 1. Synthesized compounds.
Molecules 28 04959 g001
Figure 2. Molecule structure of 3 with atomic numbering scheme, shown with 50% probability ellipsoids (The same atomic numbering was applied in the crystal structures of 1, 2, and 4).
Figure 2. Molecule structure of 3 with atomic numbering scheme, shown with 50% probability ellipsoids (The same atomic numbering was applied in the crystal structures of 1, 2, and 4).
Molecules 28 04959 g002
Figure 3. Dimer formation in the crystal structure of 3. The disorder at the ethyl group was omitted for clarity (Symmetry equivalent atoms are generated by symmetry operation 1 − x, 1 − y, 1 − z).
Figure 3. Dimer formation in the crystal structure of 3. The disorder at the ethyl group was omitted for clarity (Symmetry equivalent atoms are generated by symmetry operation 1 − x, 1 − y, 1 − z).
Molecules 28 04959 g003
Figure 4. Distances Sb…E for compounds 14, and 6, sum of atomic radii and sum of Van der Waals radii (all values in Å). The dashed lines illustrate the trends.
Figure 4. Distances Sb…E for compounds 14, and 6, sum of atomic radii and sum of Van der Waals radii (all values in Å). The dashed lines illustrate the trends.
Molecules 28 04959 g004
Figure 5. Molecule structure of 6 with atomic numbering scheme, shown with 50% probability ellipsoids.
Figure 5. Molecule structure of 6 with atomic numbering scheme, shown with 50% probability ellipsoids.
Molecules 28 04959 g005
Figure 6. Molecule structure of 7 with atomic numbering scheme, shown with 50% probability ellipsoids.
Figure 6. Molecule structure of 7 with atomic numbering scheme, shown with 50% probability ellipsoids.
Molecules 28 04959 g006
Figure 7. Intermolecular interactions between oxygen and antimony in the crystal structure of 7 (Symmetry equivalent molecule A is generated by symmetry operation 1 − x, 1 − y, 1 − z. Symmetry equivalent molecule B is generated by symmetry operation 1 − x, 2 − y, 1 − z).
Figure 7. Intermolecular interactions between oxygen and antimony in the crystal structure of 7 (Symmetry equivalent molecule A is generated by symmetry operation 1 − x, 1 − y, 1 − z. Symmetry equivalent molecule B is generated by symmetry operation 1 − x, 2 − y, 1 − z).
Molecules 28 04959 g007
Figure 8. Graphical representation of the topological analysis of 3. Bond critical points are marked in orange; ring critical points in yellow. The numbering of the atoms results from the line number of the input structure (see Supplementary Materials). (Remark: There is an additional BCP between H16 and O3 with a corresponding RCP in the Sb1-O4-C14-H16-O3 ring, which overlaps with the RCP of the Sb1-O4-C14-C11-Se2 ring in this figure).
Figure 8. Graphical representation of the topological analysis of 3. Bond critical points are marked in orange; ring critical points in yellow. The numbering of the atoms results from the line number of the input structure (see Supplementary Materials). (Remark: There is an additional BCP between H16 and O3 with a corresponding RCP in the Sb1-O4-C14-H16-O3 ring, which overlaps with the RCP of the Sb1-O4-C14-C11-Se2 ring in this figure).
Molecules 28 04959 g008
Figure 9. Electron density (a) and Laplacian of the electron density (b) of 3 in the Sb4-Se9-O7 plane. BCPs are marked in blue and RCPs in orange. Positive values of 2ρ are drawn with solid lines and represent regions of charge depletion; negative values are drawn with purple dashed lines and represent regions of charge concentration. Contour values in atomic units are 0.001, 0.002, 0.004, 0.008, 0.02, 0.04, 0.08, 0.2, 0.4, 0.8, 2, 4, 8, 20, 40, 80, 200, 400, 800.
Figure 9. Electron density (a) and Laplacian of the electron density (b) of 3 in the Sb4-Se9-O7 plane. BCPs are marked in blue and RCPs in orange. Positive values of 2ρ are drawn with solid lines and represent regions of charge depletion; negative values are drawn with purple dashed lines and represent regions of charge concentration. Contour values in atomic units are 0.001, 0.002, 0.004, 0.008, 0.02, 0.04, 0.08, 0.2, 0.4, 0.8, 2, 4, 8, 20, 40, 80, 200, 400, 800.
Molecules 28 04959 g009
Figure 10. Sb-lone pair IBO (a), Intrinsic bond orbitals of the energetically lowest unoccupied IBO (b) and Sb…N interaction (c) in compound 6.
Figure 10. Sb-lone pair IBO (a), Intrinsic bond orbitals of the energetically lowest unoccupied IBO (b) and Sb…N interaction (c) in compound 6.
Molecules 28 04959 g010
Figure 11. The molecule (a) and the electrostatic potential map (b) of 6.
Figure 11. The molecule (a) and the electrostatic potential map (b) of 6.
Molecules 28 04959 g011
Figure 12. Electron density (a) and Laplacian of the electron density (b) of 3 dimers in the Sb-O-Sb-O-plane. BCPs are marked in blue and RCPs in orange. Positive values of ∇2ρ are drawn with solid lines and represent regions of charge depletion; negative values are drawn with purple dashed lines and represent regions of charge concentration. Contour values in atomic units are 0.001, 0.002, 0.004, 0.008, 0.02, 0.04, 0.08, 0.2, 0.4, 0.8, 2, 4, 8, 20, 40, 80, 200, 400, 800.
Figure 12. Electron density (a) and Laplacian of the electron density (b) of 3 dimers in the Sb-O-Sb-O-plane. BCPs are marked in blue and RCPs in orange. Positive values of ∇2ρ are drawn with solid lines and represent regions of charge depletion; negative values are drawn with purple dashed lines and represent regions of charge concentration. Contour values in atomic units are 0.001, 0.002, 0.004, 0.008, 0.02, 0.04, 0.08, 0.2, 0.4, 0.8, 2, 4, 8, 20, 40, 80, 200, 400, 800.
Molecules 28 04959 g012aMolecules 28 04959 g012b
Figure 13. Atomic numbering scheme from quantum chemical calculation of the dimer of 3. Only one molecule and O28 and Sb25 of the second molecule are shown (Remark: There is a BCP between H16 and O3 and a corresponding RCP in the Sb1-O4-C14-H16-O3 ring, which are quite near to the RCP of Sb1-O3-C5-C8-Se2 ring in this figure).
Figure 13. Atomic numbering scheme from quantum chemical calculation of the dimer of 3. Only one molecule and O28 and Sb25 of the second molecule are shown (Remark: There is a BCP between H16 and O3 and a corresponding RCP in the Sb1-O4-C14-H16-O3 ring, which are quite near to the RCP of Sb1-O3-C5-C8-Se2 ring in this figure).
Molecules 28 04959 g013
Figure 14. IBOs of a dimer of 3 for visualization of the intramolecular Sb…Se (a), and the intermolecular Sb…O (b) interactions, as well as the Sb lone pairs (c).
Figure 14. IBOs of a dimer of 3 for visualization of the intramolecular Sb…Se (a), and the intermolecular Sb…O (b) interactions, as well as the Sb lone pairs (c).
Molecules 28 04959 g014aMolecules 28 04959 g014b
Scheme 1. Reaction of 6 with acetic anhydride. Traces of acetic acid were present in the reaction mixture.
Scheme 1. Reaction of 6 with acetic anhydride. Traces of acetic acid were present in the reaction mixture.
Molecules 28 04959 sch001
Scheme 2. Reaction of 6 with acetic acid.
Scheme 2. Reaction of 6 with acetic acid.
Molecules 28 04959 sch002
Table 1. Comparison between calculated (in brackets) and experimental Raman spectra. The index “cy” indicates inner-cyclic moieties. Frequencies are scaled as reported in [29].
Table 1. Comparison between calculated (in brackets) and experimental Raman spectra. The index “cy” indicates inner-cyclic moieties. Frequencies are scaled as reported in [29].
Compoundν(Sb-OEt)δ(O-Sb-O)cyνasym(O-Sb-O)cyνsym(O-Sb-O)cyν(Sb…E)νasym(C-O)cyνsym(C-O)cy
1576 (617)188 (209)510 (500)546 (525)159 (180)1059 (1074)1075 (1079)
2591 (609)199 (198)485 (496)521 (504)165 (171)1085 * (1074)1085 * (1077)
3589 (608)199 (194)469 (493)500 (512)117 (143)1050 (1063)1083 (1076)
4599 (605)190 (202)469 (488)495 (512)143 (129)1079 (1058)1175 (1072)
5531 * (605)115 (191)531 * (500)531 * (518)205 (217)1063 (1064)1096 (1088)
6560 (604)190 (201)500 (501)522 (517)138 (163)1050 (1071)1071 (1084)
7541 (614)303 (316) 1570 (559) 2350 (365) 3
1 δ(S-Sb-O)cy; 2 ν(Sb-O)cy; 3 ν(Sb-S)cy; * overlapping with other bands.
Table 2. Distances Sb…E for compounds 14, and 6 from the molecule structures, sum of atomic radii and sum of Van der Waals radii (all values in Å).
Table 2. Distances Sb…E for compounds 14, and 6 from the molecule structures, sum of atomic radii and sum of Van der Waals radii (all values in Å).
CompoundEDistance Sb…ESum of Atomic Radii aSum of Van der Waals Radii b
6N2.523(2)2.113.61
1O2.776(7)2.073.58
2S3.182(1)2.453.86
3Se3.2856(6)2.583.96
4Te3.5955(4)2.784.12
a Values for atomic radii taken from [39]: r(Sb) = 1.41, r(O) = 0.66, r(S) = 1.04, r(Se) = 1.17, r(Te) = 1.37 Å; b consistent values for Van der Waals-radii taken from [40]: r(Sb) = 2.06, r(O) = 1.52, r(S) = 1.80, r(Se) = 1.90, r(Te) = 2.06 Å.
Table 3. Angles C2-E-C3 (values in °).
Table 3. Angles C2-E-C3 (values in °).
CompoundEValue
6N113.50 (16)
1O113.8 (8)
2S102.4 (2)
3Se99.43 (15)
4Te99.59 (16)
Table 4. Results of the topological analysis of 3 (all values in atomic units). Parameters are: charge density (ρ), its Laplacian (∇2ρ), kinetic energy density (G), total energy density (H), and electronic potential energy density (V). The atom numbering is the same as in Figure 8, resulting from the quantum chemical calculation file (see Supplementary Materials).
Table 4. Results of the topological analysis of 3 (all values in atomic units). Parameters are: charge density (ρ), its Laplacian (∇2ρ), kinetic energy density (G), total energy density (H), and electronic potential energy density (V). The atom numbering is the same as in Figure 8, resulting from the quantum chemical calculation file (see Supplementary Materials).
ρ2ρGG/ρHV
Bond critical points
Sb1-Se20.0190.0900.0190.9770.004−0.014
Sb2-O30.2040.8170.2601.279−0.056−0.317
Sb1-O40.2070.8010.2601.254−0.060−0.320
Sb1-O170.2220.7720.2631.183−0.070−0.333
Ring critical points
Sb1-O3-C5-C8-Se20.0130.0550.0131.0240.001−0.012
Sb1-O4-C14-C11-Se20.0120.0540120.9470.002−0.010
Table 6. Crystallographic data from data collection and refinement processes for 13.
Table 6. Crystallographic data from data collection and refinement processes for 13.
Compound123
FormulaC6H13O4SbC6H13O3SSbC6H13O3SbSe
Mr270.91286.97333.87
T (K)173173 K193 K
λ (Å)0.710730.71073 A0.71073 A
Crystal systemTriclinicTriclinicTriclinic
Space groupP-1P-1P-1
a (Å)6.8212(14)6.8546(9)6.8941(9)
b (Å)8.1159(17)8.9585(11)9.0062(11)
c (Å)9.1607(18)8.9782(11)9.1442(11)
α (°)109.984(15)111.722(9)112.040(9)
β (°)106.005(16)103.710(10)103.694(10)
γ (°)100.526(16)101.051(10)100.022(10)
V (Å3)436.07(16)473.14(11)489.04(11)
Z222
ρcalc (g·cm−3)2.0632.0142.267
μ (mm−1)3.1313.0976.504
F(000)264280316
θmax (°)27.72827.36227.494
Reflections collected/unique [Rint]5772/57727414/2110
[R(int) = 0.0355]
7114/2125
[R(int) = 0.0171]
Completeness to θ = 25.242°100.0%99.9%93.7%
Absorption correctionIntegrationIntegrationIntegration
Max. and min. transmission0.6891 and 0.39100.8413 and 0.46360.6084 and 0.3435
Data/restraints/parameters5772/0/1042110/3/1212125/9/122
GoF on F21.1081.1351.179
Final R indices [I>2sigma(I)]R1 = 0.0428,
wR2 = 0.1036
R1 = 0.0253,
wR2 = 0.0625
R1 = 0.0208,
wR2 = 0.0523
R indices (all data)R1 = 0.0473,
wR2 = 0.1115
R1 = 0.0292,
wR2 = 0.0648
R1 = 0.0220,
wR2 = 0.0537
Extinction coefficient0.012(2)n/a0.033(4)
Largest peak and hole (e·Å−3)2.018 and
−1.429
0.465 and
−1.130
0.416 and
−0.703
Table 7. Crystallographic data from data collection and refinement processes for 4, 6, and 7.
Table 7. Crystallographic data from data collection and refinement processes for 4, 6, and 7.
Compound46·HCCl37
FormulaC12H26O6Sb2Te2C7H16NO3Sb·HCCl3 C4H9O2SSb
Mr765.03403.32242.92
T (K)173183153
λ (Å)0.710730.710730.71073
Crystal systemMonoclinicTriclinicMonoclinic
Space groupI2/aP-1P2(1)/n
a (Å)12.8769(8)7.1133(6)6.9188(6)
b (Å)12.8026(8)9.8959(8)6.7334(4)
c (Å)13.7344(10)11.0079(9)15.7931(17)
α (°)90107.308(6)90
β (°)115.085(5)99.987(6)95.307(8)
γ (°)9094.568(6)90
V (Å3)2050.7(2)721.44(11)732.60(11)
Z424
ρcalc (g·cm−3)2.4781.8572.202
μ (mm−1)5.4442.4603.969
F(000)1408396464
θmax (°)27.49427.497 deg.27.406 deg.
Reflections collected/unique [Rint]16987/2352
[R(int) = 0.0208]
12482/3264
[R(int) = 0.0228]
6249/1616
[R(int) = 0.0184]
Completeness to θ = 25.242°99.5%99.5%97.1%
Absorption correctionIntegrationIntegrationIntegration
Max. and min. transmission0.5982 and 0.40410.8195 and 0.57370.6661 and 0.3293
Data/restraints/parameters2352/0/1013264/0/1481616/0/75
GoF on F21.2461.1251.230
Final R indices [I>2sigma(I)]R1 = 0.0232,
wR2 = 0.0523
R1 = 0.0189,
wR2 = 0.0485
R1 = 0.0236,
wR2 = 0.0547
R indices (all data)R1 = 0.0265,
wR2 = 0.0545
R1 = 0.0202,
wR2 = 0.0492
R1 = 0.0295,
wR2 = 0.0604
Extinction coefficient-0.0238(12)0.0108(10)
Largest peak and hole (e·Å−3)1.077 and
−0.885
0.698 and
−0.540
0.575 and
−0.621
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Böhme, U.; Herbig, M. New Complexes of Antimony(III) with Tridentate O,E,O-Ligands (E = O, S, Se, Te, NH, NMe) Derived from N-Methyldiethanolamine. Molecules 2023, 28, 4959. https://doi.org/10.3390/molecules28134959

AMA Style

Böhme U, Herbig M. New Complexes of Antimony(III) with Tridentate O,E,O-Ligands (E = O, S, Se, Te, NH, NMe) Derived from N-Methyldiethanolamine. Molecules. 2023; 28(13):4959. https://doi.org/10.3390/molecules28134959

Chicago/Turabian Style

Böhme, Uwe, and Marcus Herbig. 2023. "New Complexes of Antimony(III) with Tridentate O,E,O-Ligands (E = O, S, Se, Te, NH, NMe) Derived from N-Methyldiethanolamine" Molecules 28, no. 13: 4959. https://doi.org/10.3390/molecules28134959

Article Metrics

Back to TopTop