Next Article in Journal
Preparation and Characterization of Graphene from Refined Benzene Extracted from Low-Rank Coal: Based on the CVD Technology
Next Article in Special Issue
Insertion of Phosphenium Ions into a Bicyclo[1.1.0]Tetraphosphabutane Iron Complex
Previous Article in Journal
Chemical Diversity and Biological Activity of Secondary Metabolites Isolated from Indonesian Marine Invertebrates
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

C2-Symmetric P-Stereogenic Ferrocene Ligands with Heavier Chalcogenophosphinous Acid Ester Donor Sites †

Institute for Chemistry and CINSaT, University of Kassel, Heinrich Plett-Straße 40, 34132 Kassel, Germany
*
Author to whom correspondence should be addressed.
Dedicated to Prof. Dr. Ulrich Siemeling on the occasion of his 60th birthday.
Molecules 2021, 26(7), 1899; https://doi.org/10.3390/molecules26071899
Submission received: 12 March 2021 / Revised: 23 March 2021 / Accepted: 24 March 2021 / Published: 27 March 2021
(This article belongs to the Special Issue Recent Advances in Phosphorus Chemistry)

Abstract

:
tert-Butyl-substituted diphospha[2]ferrocenophane was used as a stereochemically confined diphosphane to investigate the addition of various dichalcoganes (R2Ch2; Ch = S, Se, Te and R = Me, Ph). Bischalcogenophosphinous acid esters bearing four soft donor sides were obtained as a mixture of rac and meso diastereomers and characterized by means of multinuclear NMR and X-ray analysis. The coordination chemistry of multidentate ligand 3b was explored toward d10 coinage metal centers (Cu(I), Ag(I), and Au(I)), yielding various bimetallic complexes.

1. Introduction

Phosphinite ligands are well established in coordination chemistry and catalysis, featuring useful properties between phosphanes and phosphites in terms of donor–acceptor characteristics [1]. Similarly, thio-, seleno-, and tellurophosphinous acid esters have been reported, which can be formally regarded as isomers of the related thio-, seleno-, and tellurophosphoranes [2,3].
In general, chalcogeno phosphinous acid esters are prepared from diorganohalophosphanes by reaction with the corresponding metalchalcogenolate [1,4]. As an alternative, oxidative cleavage of a diphosphane with dichalcoganes (alias dichalcogenides) has also been reported [2,3]. The latter approach was employed recently by Hey-Hawkins and coworkers in preparing a set of carbaborane-bridged bischalcogenophosphinous acid esters with sulfur or selenium [5]. In the context of our ongoing investigation of > P-P < bond activation of phosphorus-rich ferrocenophanes [6,7], we wondered whether a similar ferrocene-bridged system may be accessible starting from [2]ferrocenophane 1. Given the relevance of ferrocene-based ligand systems [8,9,10], we set out to explore the preparation of 1,1′-ferrocenylene-bridged bischalcogenophosphinous acid esters with sulfur, selenium, and tellurium. Owing to the vicinity of two adjacent soft donor sites, phosphorus and chalcogen, we explored the coordination behavior of these ligands toward d10 coinage metal centers.

2. Results and Discussion

2.1. Ligand Preparation and Properties

To study the addition of dichalcoganes to stereochemically predefined diphosphane, we chose [2]ferrocenophane 1, which occurs exclusively in the rac form as the meso form is disfavored owing to repulsive steric interaction of the tert-butyl substituents at phosphorus [6]. Reaction of 1 with diphenyldisulfide, -selenide, and -telluride afforded the respective bischalcogenophosphinous acid esters 3a–c (a = S, b = Se, and c = Te) via formal addition to the central P-P bond (Scheme 1). Despite the stereochemically preorganized relative configuration of the P-stereogenic centers in 1, the stereoinformation got lost during the reaction, and 3ac were obtained as mixtures of their rac and meso diastereomers. The same products were obtained when starting from lithium bisphosphanide 2 (and diphenyldisulfide or diphenyldiselenide), in which the phosphorus atoms adopted a prochiral situation owing to doubly bridged metal coordination, as previously established [11,12]. In the latter reaction, two equivalents of the dichalcogenide were required, making it less efficient in terms of chalcogene transfer as one equivalent was eliminated as lithium chalcogenolate (Scheme 1). The loss of stereoinformation might be attributed to a stepwise transfer of the chalcogenyl units. Moreover, variable temperature (VT)-NMR investigations on closely related ferrocenophanes revealed low inversion barriers at phosphanyl units adjacent to ferrocene [7,13]. The latter explanation would be consistent with findings for carbaborane-bridged analogs, where a preference for the rac diastereomer was observed [5], which might be a consequence of the different electronic properties of the bridging unit affecting the inversion barrier at phosphorus. The stepwise transfer of chalcogenyl units could also explain the formation of [2]ferrocenophane 1 after the elimination of lithium phenyl tellurolate when 2 was treated with diphenylditelluride.
Apart from the chalcogeno phosphinous acid phenylesters described above, we were also interested in exploring the corresponding methyl esters as the smallest alkyl derivatives. To this end, we reacted [2]ferrocenophane 1 with dimethyl disulfide and dimethyldiselenide. Again, the corresponding thio- and selenophosphinous acid esters 4ab (a = S and b = Se) were obtained in ca. 80% yield as mixtures of rac and meso diastereomers (Scheme 2). For compounds 3ac and 4ab, the molecular structure in solid state was determined using single-crystal X-ray crystallography (Figure 1 and Figure 2). While chalcogenophosphinous acid esters have been known for several decades, to our knowledge, no solid-state structure of a telluro phosphinous acid ester has yet been reported in the literature.
Thio phosphinous acid ester 3a crystallized with two molecules in the asymmetric unit in the monoclinic space group P21/c, while heavier phosphinous acid esters 3b and 3c crystallized in the triclinic space group P 1 - . While 3a showed staggered (torsion angle τ between Cp-Cp(centroid)-Cp(centroid)-Cp was about 158°) conformation of the Cp units, the Cp rings of 3b (τ = 72°) and 3c (τ = 70°) were oriented in the eclipsed form. The sum of C-P-C and C-P-Ch angles around P1 slightly increased in the series 3ac from 301.28(8)° in 3a to 305.80(4)° in 3c. The Ch–P distances in 3ac (3a: S1-P1 2.1236(8) Å, 3b: Se1-P1 2.2595(11) Å, and 3c: Te1-P1 2.4660(2) Å) were typical for chalcogenophosphinous acid esters containing three-coordinate phosphorus [5,14,15,16,17]. Bisthiophosphinous acid methylester 4a and bisselenophosphinous acid methylester 4b crystallized in the orthorhombic space group Pbcn with a staggered conformation (τ = 90° in 4a and τ = 88° in 4b) of the ferrocene moiety. The Ch–P distances (4a: S1-P1 2.1119(13) Å and 4b: Se1-P1 2.2581(17) Å) in 4a and 4b and the sum of C-P-C and C-P-Ch angles around P1 (4a: 302.31(13)° and 4b: 303.30(20)°) were almost identical to those observed in 3a and 3b.
The chalcogeno phosphinous acid esters described above featured characteristic multinuclear NMR spectra. Upon ring cleavage of the [2]ferrocenophane scaffold, the 31P NMR chemical shift of 20.6 ppm in 1 [18] changed to lower field for the chalcogeno phosphinous acid esters 3ac and 4ab (Table 1). Two sets of signals (Δδ(31P) = 0.3–1.0 ppm) could be observed in the NMR spectra due to the fact that both diastereomers (rac and meso) were formed in nearly equal amounts starting from sterically confined 1 or prochiral 2.
An interesting feature was observed for chalcogeno phosphinous acid esters 3b, 3c, and 4b, which showed signal splitting of the selenium or tellurium satellites in the case of one diastereomer (Figure 3). This signal splitting might indicate intramolecular Ch···Ch interactions in solution, which, for steric reasons, should only be present in the meso form. The latter observation is also consistent with 77Se NMR spectra of 3b and 4b showing two sets of signals (3b: 216.2 (d) and 216.4 ppm (dd) and 4b: −21.5 (d) and −21.8 ppm (dd)) for the rac and meso diastereomers, which split into a doublet (3b: 1JSeP = 225 Hz and 4b: 1JSeP = 218 Hz) and a doublet of doublets (3b: 1JSeP = 225 Hz, nJSeP = 5 Hz and 4b: 1JSeP = 218 Hz, nJSeP = 6 Hz), indicating 1JSeP and 2JSeP coupling. An analogous situation was found in 125Te NMR of 3c, where two resonances were found at 288.2 (d, 1JTeP = 497 Hz) and 284.2 (dd, 1JTeP = 497 Hz, 1JTeP = 5 Hz) ppm, which showed coupling to just one or both phosphorus nuclei depending on the nature of the isomer. Moreover, an intermolecular Ch···Ch interaction in 3b could be excluded by a diffusion ordered spectroscopy NMR (DOSY-NMR) experiment (Figure S1), which showed no aggregate formation in solution. In the literature, homo- and heteronuclear interactions involving heavier chalcoganes have both been described even for the gas phase [19]. We cannot unambiguously conclude whether a direct P···Ch or a Ch···Ch interaction was operative in the case of esters 3b, 3c, and 4b, but based on the unpeculiar 31P NMR chemical shifts of the meso forms of these compounds, a Ch···Ch interaction definitely seems more likely.
Besides NMR spectroscopic characterization, the identity and purity of all newly prepared compounds were further established with mass spectrometry and elemental analysis.

2.2. Coinage Metal Complexes

To explore the ligand properties of the seleno derivative 3b, its coordination behavior toward d10 coinage metal centers (Cu(I), Ag(I), and Au(I)) was explored, which should allow characterization of the resulting products with NMR spectroscopy. Despite the presence of two adjacent soft donor centers (P and Se) in the investigated ligand, coordination at the phosphorus lone pair was preferred exclusively in all cases (Scheme 3). In the case of copper(I) and silver (I), 3b acted as a chelating ligand, forming metalla-bridged [3]ferrocenophanes 5 and 6, where one acetonitrile completed the coordination sphere of the d10 metal. Forming such a structural motif, in principle, two diastereomers might be anticipated, where the lone pairs of the phosphorus atoms point to the same side (meso form) or opposite sides (racemic form) of the ring. In the case of 5 and 6 exclusively, the racemic form was observed, indicating epimerization of a stereocenter in the meso ligand during complexation, which proceeded quantitatively in the case of 6. When ligand 3b was reacted with two equivalents of gold(I) chloride, a dinuclear gold complex is formed as a mixture of rac and meso diastereomers. In the latter motif, the repulsive steric interaction of the substituents at phosphorus could be avoided more easily than in the mononuclear complexes 5 and 6.
Upon coordination to copper(I), the 31P NMR chemical shift of the set of signals at 57.1 and 57.8 ppm in 3b changed to higher field. Chemical equivalence of the 31P nuclei in 5 resulted in only one broad signal at 51.6 ppm. Contrary to that, the coordination of phosphinous acid ester 3b to silver(I) changed the 31P NMR chemical shift to lower field (75.6 ppm) in 6 and signal splitting owing to coupling to 107Ag (1JAgP = 401 Hz) and 109Ag (1JAgP = 462 Hz) nuclei was observed. The 77Se NMR spectra of the copper(I) and silver(I) complexes showed higher-order spectra with a chemical shift at 230.5 ppm for 5 and 227.1 ppm for 6 resonating at lower field compared to the free ligand. To determine the phosphorus–selenium coupling constants, the 77Se NMR spectra were simulated (Figure 4and Figure S2). The values of the coupling constants to the directly bonded phosphorus nuclei (1JSeP = 276 Hz in 5 and 1JSeP = 286 Hz in 6) were similar and slightly higher than in 3b (1JSeP = 225 Hz).
The coordination of 3b to gold(I) changed the 31P NMR shift to lower field and resulted in two singlet signals at 85.9 and 85.2 ppm in 7 with selenium satellites (1JSeP = 329 Hz and 1JSeP = 332 Hz, respectively). 77Se NMR spectra of compound 7 showed a set of two doublets at 271.1 (1JSeP = 329 Hz) and 268.9 (1JSeP = 332 Hz) ppm for the Se atoms of the two different rac and meso diastereomers, which were deshielded compared to 5 and 6.
Copper(I) complex 5 and silver(I) complex 6 crystallized with two molecules in the asymmetric unit in the triclinic space group P 1 - , while the gold(I) complex 7 crystallized in the monoclinic space group P21/c. The X-ray crystal structures of 5 and 6 (Figure 5) revealed a trigonal planar coordination geometry (5: ΣCu1 = 358.9(1)° and 6: ΣAg1 = 357.6(1)°) at the coinage metal atom with d10-metal atoms bonded to two phosphorus units with bite angles between 113.82(2)° in 6 and 115.36(5)° in 5, which are comparable to those found in related dppf analogs [20,21,22]. The tilt angle (α) of the ferrocene unit showed values of 2.3(2)° (5) and 2.25(14)° (6), which is the typical range for phosphorus-rich [3]ferrocenophanes [6,7,11,12,13,23]. The torsion angle τ increased from 49° in 5 to 55° in 6, which indicated a high distortion of the ferrocene unit. The preferred linear coordination of gold(I) prevented the formation of a gold(I)-bridged [3]ferrocenophane. The resulting dinuclear complex 7 (Figure 6) entailed two Au-Cl units that were almost linear (P1-Au1-Cl1 179.12(4)° and P2-Au2-Cl2 176.83(3)°) with P-Au distances of 2.2277(90) Å (P1-Au1) to 2.2307(9) Å (P2-Au2). The P-Se bond lengths remained unchanged after coordination to the transition metal centers (5: 2.2556(14) Å, 6: 2.2523(7) Å, and 7: 2.2541(10) Å) compared to the free ligand 3b (P1-Se1 2.2595(10) Å).
To investigate the possibility of achieving the same coordination motif as for copper(I) and silver(I) bisselenophosphinous acid phenylester, 3b was reacted with one equivalent (tht)AuCl, which resulted in the formation of various undefined products besides unreacted free ligand. If the same reaction was repeated in the presence of aluminum chloride as Lewis acid, dimeric ionic gold(I) complex 8 (Scheme 4) was formed in good yield. Dimerization entails a linear coordination and a stabilization of both gold(I) centers in the ferrocenophane.
The dinuclear gold(I) complex 8 described above featured characteristic multinuclear NMR spectra. Upon coordination to gold(I), the 31P NMR chemical shift of the set of signals at 57.1 and 57.8 ppm in 3b changed to lower field. Chemical equivalence of the 31P nuclei in 8 resulted in only one signal at 91.0 ppm, which was almost unchanged compared to the gold(I) complex 7.
The corresponding 77Se NMR resonance of 8 at 275.3 ppm was deshielded by more than 50 ppm upon coordination to gold(I) compared to 3b, while the 1JSeP coupling dropped to 177 Hz. The resulting 77Se NMR spectrum was again a higher-order spectrum due to the presence of four selenium nuclei in compound 8. Gold(I) complex 8 crystallized in the triclinic space group P 1 - . The X-ray crystal structure of 8 (Figure 6) revealed that the gold atoms were bonded to two phosphorus units. The tilt angles α showed values of 3.9(1)° at Fe1 and 5.8(1)° at Fe2 for both ferrocene units and indicated that almost no ring strain was present. The Au-P contacts between 2.305(10) Å and 2.337(9) Å were slightly longer than those observed in 7 but in the typical range for P-Au bond lengths [24,25]. For the solid-state structures of compounds 7 and 8, no inter- or intramolecular Au···Au interactions were observed (Au···Au > 6.7 Å in 8). Nevertheless, for related dinuclear Au(I) complexes of ferrocene-bridged bisphosphanes, Au···Au distances were found to be quite variable in different polymorphs [26].

3. Materials and Methods

3.1. Experimental

All reactions were carried out by means of standard Schlenk or glovebox techniques under inert gas atmosphere (argon). Solvents were dried over Na/K alloy before use and were freshly distilled under inert gas. Deuterated solvents for NMR spectroscopy were dried and stored over molecular sieves. [Fe(C5H4P(tBu)Li)2] and [Fe(C5H4P(tBu)2] were prepared according to procedures in the literature [6,11], while other reagents were used as received without further purification. 1H-, 13C-, 31P-, 77Se-, 125Te-, and DOSY-NMR data were recorded on Varian VNMRS-500 MHz or MR-400 MHz spectrometers at 25 °C. Chemical shifts were referenced to residual protic impurities in the solvent (1H) or the deuterio solvent (13C) and reported relative to external SiMe4 (1H, 13C), H3PO4 (31P), Me2Se (77Se), or Me2Te (125Te). For molecular weight estimation, external calibration curves together with a correction factor dependent on the molar density were used in the CC-MW estimation software v1.3 [27,28]. APCI-DIP (atmospheric pressure chemical ionization-direct inlet probe) mass determinations were performed on a Finnigan LCQ Deca (ThermoQuest). ESI (electrospray ionization) mass spectra were recorded on a microTOF (Bruker Daltonics, Bremen, Germany). Mass calibration was carried out immediately before sample measurement on sodium formate clusters or by the ESI-Tune Mix standard (Agilent, Waldbronn, Germany). Elemental analyses were performed with a HEKAtech Euro EA CHNS elemental analyzer (Wegberg, Germany). Samples were prepared in a Sn cup and analyzed with added V2O5.

3.2. X-ray Diffraction Measurements

Crystallographic measurements were carried out on a Stoe IPDS2 or a Stoe StadiVari diffractometer with a STOE image plate detector and a Mo-Kα (λ = 0.71073 Å) monochromator or a Stoe StadiVari diffractometer with a Pilatus 200K image plate detector and Cu-Kα (λ = 1.54186 Å) radiation. Direct methods were used to solve the measurements and refined by “least-square” cycles (SHELXL-2014) [29]. All nonhydrogen atoms were anisotropically refined without restriction. Evaluation of the data sets and graphical preparation of the structures were carried out using Olex2 [30] and Mercury [31]. Details of the structure determination and refinement are summarized in Tables S1–S3 (supporting information).

3.3. Synthetic Protocols and Characterization

Synthesis of 3a
To a solution of 1 (36 mg, 0.1 mmol) in 0.6 mL C6D6, 24 mg (0.1 mmol) of diphenyldisulfide was added at room temperature. This mixture was stirred at room temperature for 2 h. All volatile components were removed in vacuo, and the residue was subjected to flash column separation on silica using a mixture of pentane and methylene chloride (gradient) as eluent, which afforded 85% yield (49 mg, 0.09 mmol) as orange crystals.
Alternative: To a mixture of 2 (374 mg, 1 mmol) in 20 mL toluene, 437 mg (2 mmol) of diphenyldisulfide was added as a solution in 10 mL toluene at 0 °C. Under vigorous stirring, a few drops of tetrahydrofuran (THF) was added until all solid was dissolved. All volatile components were removed in vacuo, and the residue was subjected to flash column separation on silica using a mixture of pentane and methylene chloride (9/1) (Rf = 0.17) as eluent, which afforded 79% yield (458 mg, 0.79 mmol) as orange crystals. The following NMR chemical shifts, marked with an asterisk, indicate two distinct signals due to rac/meso diastereomers.
1H-NMR (400 MHz, C6D6): δ 7.87 (m, 4H, Ph), 7.09 (m, 4H, Ph), 6.95 (m, 2H, Ph), 4.52 (m, 1H, Cp), 4.42 (m, 1H, Cp), 4.31 (m, 1H, Cp), 4.27 (m, 1H, Cp), 4.21 (m, 3H, Cp), 4.19 (m, 1H, Cp), 1.05* and 1.05* (d, 3JPH = 12.9 Hz and 12.8 Hz, 18H, tBu CH3). 13C-NMR (101 MHz, C6D6): δ 137.3 (m, Ph Cq), 131.4 (m, Ph), 129.2 (m, Ph), 126.3 (m, Ph), 77.8–77.3 (m, Cp Cipso), 77.0 (m, Cp), 76.5 (m, Cp), 76.5 (m, Cp), 72.7–72.2 (m, Cp), 72.1 (m, Cp), 33.4 (m, tBu Cq), 27.3 (m, tBu CH3). 31P-NMR (202 MHz, C6D6): δ 51.2* (s), 50.2* (s). Elemental analysis (%): calculated: C 62.28, H 6.27, S 11.08, found: C 62.56, H 6.46, S 11.12. MS (APCI-DIP-HR) m/z: 579.115627 ([M + H]+ 100%), calculated: 579.115584.
Synthesis of 3b
To a solution of 1 (64 mg, 0.18 mmol) in 1 mL toluene, 64 mg (0.2 mmol) of diphenyldiselenide was added at room temperature. This mixture was stirred at room temperature for 30 min. All volatile components were removed in vacuo, and the product was crystallized from a concentrated pentane solution at −20 °C, which afforded 95% yield (113 mg, 0.17 mmol) as orange crystals.
Alternative: To a mixture of 2 (374 mg, 1 mmol) in 20 mL toluene, 624 mg (2 mmol) of diphenyldisulfide was added as a solution in 10 mL toluene at 0 °C. Under vigorous stirring, a few drops of THF was added until all solid was dissolved. All volatile components were removed in vacuo, and the residue was subjected to flash column separation on silica using a mixture of pentane and methylene chloride (9/1) (Rf = 0.18) as eluent, which afforded 82% yield (548 mg, 0.82 mmol) as orange crystals. The following NMR chemical shifts, marked with an asterisk, indicate two distinct signals due to rac/meso diastereomers.
1H-NMR (400 MHz, C6D6): δ 7.94 (m, 4H, Ph), 7.05 (m, 4H, Ph), 6.98 (m, 2H, Ph), 4.42 (m, 1H, Cp), 4.36 (m, 1H, Cp), 4.31 (m, 1H, Cp), 4.28 (m, 2H, Cp), 4.24 (m, 3H, Cp), 1.08* and 1.07* (d, 3JPH = 12.9 Hz and 12.8 Hz, 18H, tBu CH3). 13C-NMR (101 MHz, C6D6): δ 137.3 (m, Ph Cq), 131.4 (m, Ph), 129.2 (m, Ph), 126.3 (m, Ph), 77.4 (m, Cp), 76.9 (d, 1JCP = 32 Hz, Cp Cipso), 76.9 (m, Cp), 76.7 (m, 1JCP = 33 Hz, Cp Cipso), 73.8–73.6 (m, Cp), 73.0–72.7 (m, Cp), 33.0 (m, tBu Cq), 27.8 (m, tBu CH3). 31P-NMR (202 MHz, C6D6): δ 57.8* (s), 57.1* (s). 77Se-NMR (95 MHz, C6D6): 216.4* (d, 1JPSe = 225 Hz, nJPSe = 5 Hz), 215.2* (d, 1JPSe = 224 Hz). Elemental analysis (%): calculated: C 53.59, H 5.40, found: C 53.61, H 5.42. MS (APCI-DIP-HR) m/z: 675.005211 ([M + H]+ 100%), calculated: 675.004486.
Synthesis of 3c
To a solution of 1 (360 mg, 1 mmol) in 10 mL of THF diphenylditelluride (410 mg, 1 mmol) was added at room temperature, and the mixture was stirred at 90 °C overnight in a pressure Schlenk tube. All volatile components were removed in vacuo, and the product was crystallized from a concentrated pentane solution at −20 °C, which afforded the crude product as orange crystals. The product contained a significant amount of starting material (diphenylditelluride and diphospha-[2]ferrocenophane 1). The following NMR chemical shifts, marked with an asterisk, indicate two distinct signals due to rac/meso diastereomers.
1H-NMR (400 MHz, C6D6): δ 8.04 (m, 4H, Ph), 6.99 (m, 6H, Ph), 4.44 (m, 1H, Cp) 4.37 (m, 1H, Cp), 4.34 (m, 1H, Cp), 4.30 (m, 1H, Cp), 4.25 (m, 3H, Cp), 4.00 (m, 1H, Cp), 1.12* and 1.11* (d, 3JPH = 12.6 Hz and 12.3 Hz, 18H, tBu CH3). 13C-NMR (101 MHz, C6D6): δ 138.3 (m, Ph), 129.6 (m, Ph), 127.5 (m, Ph), 112.1 (m, Ph Cq), 78.0 (m, Cp), 77.5 (m, Cp), 77.1 (m, Cp), 76.9 (m, Cp), 76.2 (m, 1JCP = 25 Hz, Cp Cipso), 75.9 (m, 1JCP = 26 Hz, Cp Cipso), 73.5 (m, Cp), 32.3 (m, tBu Cq), 28.8 (m, tBu CH3). 31P-NMR (202 MHz, C6D6): δ 45.8* (s), 45.3* (s). 125Te-NMR (158 MHz, C6D6): δ 288.2* (dd, 1JPTe = 497 Hz, nJPTe = 14 Hz), 284.2* (d, 1JPTe = 497 Hz). MS (APCI-DIP-HR) m/z: 806.974229 ([M + H + 2O]+ 17%), calculated: 806.973718.
Synthesis of 4a
To a solution of 1 (32 mg, 0.09 mmol) in 0.6 mL C6D6, 0.1 mL (1.1 mmol) of dimethyldisulfide was added at room temperature. This mixture was stirred overnight at 90 °C. All volatile components were removed in vacuo, and the residue was subjected to flash column separation on silica using a mixture of pentane and methylene chloride as eluent, which afforded 87% yield (35 mg, 0.08 mmol) as orange crystals. The following NMR chemical shifts, marked with an asterisk, indicate two distinct signals due to rac/meso diastereomers.
1H-NMR (400 MHz, C6D6): δ 4.46 (m, 2H, Cp), 4.38 (m, 6H, Cp), 2.23 (m, 3JPH = 13.1 Hz 6H, S-CH3), 1.07 (d, 3JPH = 12.7 Hz 18H, tBu CH3). 13C-NMR (101 MHz, C6D6): δ 78.2–77.6 (m, Cp Cipso), 76.5 (m, Cp), 76.1 (m, Cp), 72.5–72.2 (m, Cp), 33.2 (m, tBu Cq), 27.1 (m, tBu CH3), 17.8 (m, 2JPC = 32 Hz, S-CH3), 17.7 (d, 2JPC = 32 Hz, S-CH3). 31P-NMR (202 MHz, C6D6): δ 56.8* (s), 56.5* (s). Elemental analysis (%): calculated: C 52.87, H 7.10, S 14.11, found: C 53.75, H 7.57, S 13.81. MS (APCI-DIP-HR) m/z: 455.084305 ([M + H]+ 100%), calculated: 455.084284.
Synthesis of 4b
To a solution of 1 (72 mg, 0.2 mmol) in 2 mL toluene, 0.1 mL (1.0 mmol) of dimethyldiselenide were added at room temperature. This mixture was stirred at room temperature for 3 h. All volatile components were removed in vacuo, and the residue was subjected to flash column separation on silica using a mixture of pentane and methylene chloride as eluent affording 73% yield (80 mg, 0.15 mmol) as an orange oil, which transformed over time to a yellow solid. The following NMR chemical shifts, marked with an asterisk, indicate two distinct signals due to rac/meso diastereomers.
1H-NMR (400 MHz, C6D6): δ 4.44–4.37 (m, 7H, Cp), 4.33 (m, 1H, Cp), 2.09 (m, 3JPH = 10.0 Hz 6H, Se-CH3), 1.10 (d, 3JPH = 12.5 Hz 18H, tBu CH3). 13C-NMR (101 MHz, C6D6): 77.1 (m, 1JPC = 30 Hz, Cp Cipso), 77.0 (m, Cp), 76.5 (m, Cp), 73.9–73.7 (m, Cp), 72.9 (m, Cp), 72.8–72.6 (m, Cp), 32.6 (m, tBu Cq), 27.5 (m, tBu CH3), 6.5 (d, 2JPC = 30 Hz, Se-CH3), 6.4 (d, 2JPC = 30 Hz, Se-CH3). 31P-NMR (202 MHz, C6D6): δ 58.7* (s), 58.4* (s). 77Se-NMR (95 MHz, C6D6): −21.5* (d, 1JPSe = 218 Hz), −21.8* (dd, 1JPSe = 218 Hz, nJPSe = 6 Hz). Elemental analysis (%): calculated: C 43.82, H 5.88, found: C 43.68, H 6.20. MS (APCI-DIP-HR) m/z: 550.973667 ([M + H]+ 100%), calculated: 550.973186.
Synthesis of 5
To a solution of 3b (67 mg, 0.1 mmol) in 1 mL DCM, 31 mg (0.1 mmol) of [Cu(MeCN)4][BF4] was added at room temperature. This mixture was placed in the ultrasonic bath for 1 h. All volatile components were removed in vacuo, and the residue was dissolved in 10 mL toluene. The solid was removed by centrifugation, and the product was crystallized from a concentrated toluene solution at −20 °C, which afforded 35% yield (30 mg, 0.03 mmol) as orange crystals.
1H-NMR (400 MHz, CD3CN): δ 7.87 (m, 4H, Ph), 7.43 (m, 6H, Ph),4.76 (m, 2H, Cp), 4.54 (m, 2H, Cp), 4.51 (m, 2H, Cp), 4.39 (m, 2H, Cp), 1.96 (s, 3H, MeCN), 0.95 (m, 18H, tBu CH3). 13C-NMR (101 MHz, CD3CN): δ 137.1 (m, Ph), 131.3 (s, Ph), 130.4 (m, Ph), 127.1 (m, Ph Cq), 118.3 (s, Me-CN), 78.1 (m, Cp), 75.3 (m, Cp), 75.2 (s, Cp), 74.0 (m, Cp Cipso), 72.8 (s, Cp), 35.4 (m, tBu Cq), 26.8 (m, tBu CH3), 1.7 (s, H3C-CN). 31P-NMR (202 MHz, CD3CN): δ 51.2 (s). 77Se-NMR (95 MHz, CD3CN): δ 233.4–228.1 (m). Elemental analysis (%): calculated: C 44.50, H 4.55, N 1.62 found: C 43.58, H 4.64, N 1.72. MS (ESI-HR) m/z: 736.926803 ([M − (BF4) − (MeCN)]+ 100%), calculated: 736.926262.
Synthesis of 6
To a solution of 3b (67 mg, 0.1 mmol) in 1 mL MeCN, 20 mg (0.1 mmol) of AgBF4 was added at room temperature. This mixture was placed in the ultrasonic bath for 1 h. All volatile components were removed in vacuo, and the residue was washed with 10 mL toluene. The product was crystallized from a concentrated MeCN solution at −20 °C, which afforded quantitative yield (91mg, 0.1 mmol) as yellow crystals.
1H-NMR (400 MHz, CD3CN): δ 7.84 (m, 4H, Ph), 7.44 (m, 6H, Ph), 4.76 (m, 2H, Cp), 4.53 (m, 2H, Cp), 4.51 (m, 2H, Cp), 3.97 (m, 2H, Cp), 1.94 (m, MeCN), 0.93 (m, 18H, tBu CH3). 13C-NMR (101 MHz, CD3CN): δ 136.3 (m, Ph), 131.6 (m, Ph), 130.3 (m, Ph), 130.1 (m, Ph Cq), 118.4 (s, Me-CN), 78.7 (m, Cp), 75.0 (m, Cp), 75.0 (m, Cp), 72.8 (s, Cp), 35.4 (m, tBu Cq), 27.0 (m, tBu CH3), 1.5 (m, H3C-CN). 31P-NMR (202 MHz, CD3CN): δ 75.6 (m, 1JP-(109)Ag = 462 Hz, 1JP-(107)Ag = 401 Hz). 77Se-NMR (95 MHz, CD3CN): δ 230.1–225.1 (m, 1JPSe = 286 Hz). Elemental analysis (%): calculated: C 42.33, H 4.33, N 1.54 found: C 42.16, H 4.33, N 1.37. MS (ESI-HR) m/z: 780.902890 ([M − (BF4) − (MeCN)]+, calculated: 780.901754.
Synthesis of 7
To a solution of 3b (67 mg, 0.1 mmol) in 2 mL toluene, 64 mg (0.2 mmol) of (tht)AuCl was added at room temperature. The mixture was stirred for 1 h at room temperature. All volatile components were removed in vacuo, and the residue was washed with 5 mL toluene. The product was crystallized from a concentrated toluene/THF solution at −20 °C, which afforded 92% yield (105 mg, 0.09 mmol) as yellow crystals. The following NMR chemical shifts, marked with an asterisk, indicate two distinct signals due to rac/meso diastereomers.
1H-NMR (400 MHz, CD2Cl2): δ 8.01 (m, 2H, Ph), 7.50 (m, 2H, Ph), 7.45 (m, 4H, Ph), 4.81–4.75 (m, 6H, Cp), 4.63 (m, 1H, Cp), 4.60 (m, 1H, Cp), 1.18* and 1.17* (d, 3JPH = 19.3 Hz and 19.2 Hz, 18H, tBu CH3). 13C-NMR (101 MHz, CD2Cl2): δ 137.6 (m, Ph), 130.6 (m, Ph), 130.3 (m, Ph), 127.2 (m, Ph Cq), 77.5 (m, Cp), 76.0 (m, Cp), 75.6 (m, Cp), 74.4 (m, Cp), 72.7 (m, 1JCP = 23 Hz, Cp Cipso), 38.3 (m, tBu Cq), 26.9 (m, tBu CH3). 31P-NMR (202 MHz, CD2Cl2): δ 85.9* (s), 85.2* (s). 77Se-NMR (95 MHz, CD2Cl2): δ 271.1* (d, 1JPSe = 330 Hz), 268.9* (d, 1JPSe = 332 Hz). Elemental analysis (%): calculated: C 31.69, H 3.19 found: C 31.73, H 3.21. MS (ESI-HR) m/z: 1102.898926 ([M − Cl]+ 100%), calculated: 1102.898618.
Synthesis of 8
For this, 67 mg of 3b (0.1 mmol) and 32 mg (0.1 mol) of (tht)AuCl were dissolved in 0.6 mL σ-difluorobenzene. The solution was added to 14 mg (0.1 mmol) AlCl3 at room temperature, and the resulting suspension was placed into the ultrasonic bath until all solid was dissolved. It was crystallized overnight at room temperature. The crystals were separated from the solution and washed with 3 mL of toluene, which afforded 46% yield (96 mg, 0.05 mmol) as yellow crystals.
1H-NMR (400 MHz, CD2Cl2): δ 7.85 (m, 8H, Ph), 7.55 (m, 4H, Ph), 7.46 (m, 8H, Ph), 5.25 (m, 4H, Cp), 4.84 (m, 4H, Cp), 4.81 (m, 4H, Cp), 4.77 (m, 4H, Cp), 1.03 (m, 36H, tBu CH3). 13C-NMR (101 MHz, CD2Cl2): δ 136.9 (s, Ph), 131.3 (s, Ph), 131.0 (s, Ph), 124.4 (s, Ph Cq), 77.7 (m, Cp), 75.6 (m, Cp), 74.8 (m, Cp), 74.0 (m, Cp), 72.3 (m, Cp Cipso), 38.8 (m, tBu Cq), 26.9 (m, tBu CH3). 27Al-NMR (104 MHz, CD2Cl2): δ 103.7 (s). 31P-NMR (202 MHz, CD2Cl2): δ 91.0 (s). 77Se-NMR (95 MHz, CD2Cl2): δ 275.3 (m, 1JPSe = 177 Hz). Elemental analysis (%): calculated: C 34.71, H 3.50 found: C 34.71, H 3.53. MS (ESI-HR) m/z: 870.963213 ([1/2(M − AlCl4)]+ 76%), calculated: 870.963937.

4. Conclusions

In summary, we report on a series of unprecedented ferrocene-bridged bischalcogenophosphinous acid esters with the chalcogens S, Se, and Te. For tellurophosphinous acid, this is the first example of a molecular structure established via single X-ray diffraction. The compounds are generally obtained as mixtures of rac and meso diastereomer where the ratio is not affected by the stereochemical orientation of the precursor, for instance, starting from chiral [2]ferrocenophane 1 or prochiral bisphosphanide 2. Investigation of the coordination behavior of bisselenophosphinous acid ester toward d10 coinage metal centers revealed exclusive metal coordination at the lone pair at phosphorus based on X-ray crystallography. Interestingly, the coordination shift of the 77Se NMR resonance was equally or more sensitive to the nature of the metal than the corresponding 31P NMR resonance although the latter atom was directly attached to the metal center. For 77Se, coordination of M(I) (M = Cu, Ag, and Au) deshielding was observed in all cases (Cu: +14 ppm, Ag: +11 ppm, and Au: +55 ppm). For 31P, shielding was observed for Cu (−6 ppm), as in related cases [26,32,33], but deshielding was observed for Ag and Au coordination (Ag: +18 ppm and Au: +28 ppm). An intriguing feature of meso bischalcogenophosphinous acid esters with Se and Te was the observation of a resolved coupling of 77Se and 125Te to both chemically equivalent but magnetically inequivalent phosphorus nuclei within the molecule. This finding is interpreted as a consequence of attractive interactions of the chalcogenophenyl or chalcogenomethyl units in solution, most likely directly via the chalcogen atoms (Ch···Ch).

Supplementary Materials

The following are available online, multinuclear NMR spectra and crystallographic details. Table S1: Summary of structure determinations and refinement for 3a, 3b, 3c and 4a. Table S2: Summary of structure determinations and refinement for 4b, 5 and 6. Table S3: Summary of structure determinations and refinement for 7 and 8. Figure S1: DOSY-nmr in benzene-d6 as solvent and tetramethylsilane (TMS) as internal standard for calibration. ECC C 6 d 6 Merge was used to determine MWdet, cor = 700 g/mol (MWdif, cor = −4%). Figure S2: Measured (top) and simulated (bottom) 77Se-NMR spectrum of silver(I) complex 6 (MeCN-d3, 300 K). Following coupling constants were used for the simulation: 1JPSe = 276 Hz, 2JPP = 145 Hz, 1JP-(63)Cu = 100 Hz, 1JP-(65)Cu = 150 Hz, 3JPSe = 0 Hz, 2JSe-(63)Cu = 2JSe-(65)Cu = 0 Hz.

Author Contributions

R.F. performed the experiments, analyzed the spectroscopic data, and wrote significant parts of the manuscript; C.B. performed the X-ray interpretation; R.P. provided the funding and materials and wrote significant parts of the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This research was partially funded by the German Science Fund (DFG) within CRC 1319 (ELCH).

Data Availability Statement

The data presented in this study are available in the Supplementary Material.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Not available.

References

  1. Rajanbabu, T.V. Phosphinite and Phosphonite Ligands. In Phosphorus(III) Ligands in Homogeneous Catalysis: Design and Synthesis; Kamer, P.C.J., van Leeuwen, P.W.N.M., Eds.; John Wiley & Sons, Ltd.: West Sussex, UK, 2012; pp. 159–232. [Google Scholar]
  2. DuMont, W.-W.; Kubiniok, S.; Severengiz, T. Eigenschaften von Tellur-Tellur-Bindungen. IV. Dismutationsreaktionen von Di-p-tolylditellurid mit Tetra-tert-butyldiphosphan und Tetra-isopropyldiphosphan. Z. Anorg. Allg. Chem. 1985, 531, 21–25. [Google Scholar] [CrossRef]
  3. Ashe, A.J., III; Ludwig, E.G., Jr. The exchange reaction of tetramethyldipnictogens with dimethyldichalcogenides. J. Organomet. Chem. 1986, 308, 289–296. [Google Scholar] [CrossRef] [Green Version]
  4. Medena, C.; Calogero, F.; Lemoine, Q.; Aubert, C.; Derat, E.; Fensterbank, L.; Gontard, G.; Khaled, O.; Vanthuyne, N.; Moussa, J.; et al. A HELIXOL-Derived Bisphosphinite Ligand: Synthesis and Application in Gold-Catalyzed Enynes Cycloisomerization. Eur. J. Org. Chem. 2019, 2019, 2129–2137. [Google Scholar] [CrossRef]
  5. Coburger, P.; Schulz, J.; Klose, J.; Schwarze, B.; Sárosi, M.B.; Hey-Hawkins, E. C2-Symmetric P,N Ligands Derived from Carborane-Based Diphosphetanes: Synthesis and Coordination Chemistry. Inorg. Chem. 2017, 56, 292–304. [Google Scholar] [CrossRef]
  6. Franz, R.; Nasemann, S.; Bruhn, C.; Kelemen, Z.; Pietschnig, R. Chalcogen transfer rearrangement: Exploring inter- vs. intramolecular P-P bond activation. Chem. Eur. J. 2021, 27, 641–648. [Google Scholar]
  7. Isenberg, S.; Weller, S.; Kargin, D.; Valić, S.; Schwederski, B.; Kelemen, Z.; Bruhn, C.; Krekić, K.; Maurer, M.; Feil, C.; et al. Bis-[3]ferrocenophanes with central >E-E’< bonds (E, E’ = P, SiH): Preparation, properties, and thermal activation. Chem. Open 2019, 8, 1235–1243. [Google Scholar] [CrossRef]
  8. Togni, A.; Hayashi, T. Ferrocenes—Homogeneous Catalysis, Organic Synthesis, Materials Science; VCH Verlagsgesellschaft: Weinheim, Germany, 1995. [Google Scholar]
  9. Štěpnička, P. Ferrocenes: Ligands, Materials and Biomolecules; John Wiley and Sons, Ltd.: West Sussex, UK, 2008. [Google Scholar]
  10. Dey, S.; Pietschnig, R. Chemistry of sterically demanding dppf-analogs. Coord. Chem. Rev. 2021, accepted. [Google Scholar]
  11. Kargin, D.; Kelemen, Z.; Krekić, K.; Maurer, M.; Bruhn, C.; Nyulászi, L.; Pietschnig, R. [3]Ferrocenophanes with bisphosphanotetryl bridge: Inorganic rings on the way to tetrylenes. Dalton Trans. 2016, 45, 2180–2189. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Isenberg, S.; Frenzel, L.-M.; Bruhn, C.; Pietschnig, R. Metallated [3]ferrocenophanes containing P3M bridges (M = Li, Na, K). Inorganics 2018, 6, 67. [Google Scholar] [CrossRef] [Green Version]
  13. Lik, A.; Kargin, D.; Isenberg, S.; Kelemen, Z.; Pietschnig, R.; Helten, H. PBP bridged [3]ferrocenophane: A bisphosphanylborane with a redox trigger Chem. Commun. 2018, 54, 2471–2474. [Google Scholar] [CrossRef]
  14. Koner, A.; Kelemen, Z.; Schnakenburg, G.; Nyulászi, L.; Streubel, R. 1,4-Additions of tricyclic 1,4-diphosphinines—A novel system to study σ-bond activation and π–π dispersion interactions. Chem. Commun. 2018, 54, 1182–1184. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Coburger, P.; Aures, R.; Schulz, P.; Hey-Hawkins, E. Exploiting the Ring Strain of Diphosphetanes: A Synthetic and Computational Approach towards 1,2,5-Selenadiphospholanes. ChemPlusChem 2018, 83, 1057–1064. [Google Scholar] [CrossRef] [PubMed]
  16. Ebels, J.; Pietschnig, R.; Kotila, S.; Dombrowski, A.; Niecke, E.; Nieger, M.; Schiffner, H.M. Bis(pentamethylcyclopentadienyl)-Substituted Phosphanes: Synthesis and Structure. Eur. J. Inorg. Chem. 1998, 1998, 331–337. [Google Scholar] [CrossRef]
  17. Graham, C.M.E.; Macdonald, C.L.B.; Boyle, P.D.; Wisner, J.A.; Ragogna, P.J. Addressing the Nature of Phosphinidene Sulfides via the Synthesis of P−S Heterocycles. Chem. Eur. J. 2018, 24, 743–749. [Google Scholar] [CrossRef]
  18. Tanimoto, Y.; Ishizu, Y.; Kubo, K.; Miyoshi, K.; Mizuta, T. Synthesis of [2]ferrocenophanes containing trivalent diphosphine units. J. Organomet. Chem. 2012, 713, 80–88. [Google Scholar] [CrossRef]
  19. Glodde, T.; Vishnevskiy, Y.V.; Zimmermann, L.; Stammler, H.-G.; Neumann, B.; Mitzel, N.W. The Nature of Chalcogen-Bonding-Type Tellurium–Nitrogen Interactions: A First Experimental Structure from the Gas Phase. Angew. Chem. Int. Ed. 2021, 60, 1519–1523. [Google Scholar] [CrossRef] [PubMed]
  20. Dey, S.; Buzsáki, D.; Bruhn, C.; Kelemen, Z.; Pietschnig, R. Bulky 1,1′-bisphosphanoferrocenes and their coordination behaviour towards Cu(i). Dalton Trans. 2020, 49, 6668–6681. [Google Scholar] [CrossRef]
  21. Trivedi, M.; Ujjain, S.K.; Singh, G.; Kumar, A.; Dubey, S.K.; Rath, N.P. Syntheses, characterization, and electrochemistry of compounds containing 1-diphenylphosphino-1′-(di-tert-butylphosphino)ferrocene (dppdtbpf). J. Organomet. Chem. 2014, 772-773, 202–209. [Google Scholar] [CrossRef]
  22. Trivedi, M.; Nagarajan, R.; Kumar, A.; Rath, N.P.; Valerga, P. Synthesis, characterization, crystal structures and photophysical properties of copper(I) complexes containing 1,1′-bis(diphenylphosphino)ferrocene (B-dppf) in doubly-bridged mode. Inorg. Chim. Acta 2011, 376, 549–556. [Google Scholar] [CrossRef]
  23. Borucki, S.; Kelemen, Z.; Maurer, M.; Bruhn, C.; Nyulaszi, L.; Pietschnig, R. Stereochemical Alignment in Triphospha[3]ferrocenophanes. Chem. Eur. J. 2017, 23, 10438–10450. [Google Scholar] [CrossRef] [Green Version]
  24. Laube, J.; Jäger, S.; Thöne, C. Synthesis and Structural Studies of Pyridine-2-selenolates − Reactions with Electrophilic Phosphorus(III) Compounds and Related Complex Chemistry. Eur. J. Inorg. Chem. 2001, 2001, 1983–1992. [Google Scholar] [CrossRef]
  25. Bowmaker, G.A.; Dyason, J.C.; Healy, P.C.; Engelhardt, L.M.; Pakawatchai, C.; White, A.H. Lewis-base adducts of Group 11 metal(I) compounds. Part 27. Solid-state phosphorus-31 cross-polarization magic-angle spinning nuclear magnetic resonance, far-infrared, and structural studies on the mononuclear 2: 1 adducts of triphenylphosphine with copper(I) and gold(I) halides. J. Chem. Soc. Dalton Trans. 1987. [Google Scholar] [CrossRef] [Green Version]
  26. Orthaber, A.; Borucki, S.; Shen, W.; Réau, R.; Lescop, C.; Pietschnig, R. Coordination behavior of a hexadentate 1,1’-ferrocenylene bridged bisphosphole towards coinage metal centers. Eur. J. Inorg. Chem 2014, 1751–1759. [Google Scholar] [CrossRef] [Green Version]
  27. Bachmann, S.; Gernert, B.; Stalke, D. Solution structures of alkali metal cyclopentadienides in THF estimated by ECC-DOSY NMR-spectroscopy (incl. software). Chem. Commun. 2016, 52, 12861–12864. [Google Scholar] [CrossRef] [Green Version]
  28. Kreyenschmidt, A.-K.; Bachmann, S.; Niklas, T.; Stalke, D. Molecular Weight Estimation of Molecules Incorporating Heavier Elements from van-der-Waals Corrected ECC-DOSY. ChemistrySelect 2017, 2, 6957–6960. [Google Scholar] [CrossRef]
  29. Sheldrick, G. Crystal structure refinement with SHELXL. Acta. Crystallogr. C. 2015, 71, 3–8. [Google Scholar] [CrossRef]
  30. Dolomanov, O.V.; Bourhis, L.J.; Gildea, R.J.; Howard, J.A.K.; Puschmann, H. OLEX2: A complete structure solution, refinement and analysis program. J. Appl. Cryst. 2009, 42, 339–341. [Google Scholar] [CrossRef]
  31. Macrae, C.F.; Edgington, P.R.; McCabe, P.; Pidcock, E.; Shields, G.P.; Taylor, R.; Towler, M.; van de Streek, J. Mercury: Visualization and analysis of crystal structures. J. Appl. Crystallogr. 2006, 39, 453–457. [Google Scholar] [CrossRef] [Green Version]
  32. Orthaber, A.; Belaj, F.; Pietschnig, R. Synthesis, Structure and π-Delocalization of a Phosphaalkenyl Based Neutral PNP-Pincer. Inorg. Chim. Acta 2011, 374, 211–215. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Orthaber, A.; Fuchs, M.; Belaj, F.; Rechberger, G.N.; Kappe, O.C.; Pietschnig, R. Bis(diethylamino)(pentafluorophenyl)phosphane—A Push–Pull Phosphane Available for Coordination. Eur. J. Inorg. Chem 2011, 2011, 2588–2596. [Google Scholar] [CrossRef] [Green Version]
Scheme 1. Synthesis of bischalcogenophosphinous acid phenylesters 3ac from racemic 1 or prochiral 2. Yields marked with an asterisk refer to the conversion of 2 as starting material (n.i. = not isolated).
Scheme 1. Synthesis of bischalcogenophosphinous acid phenylesters 3ac from racemic 1 or prochiral 2. Yields marked with an asterisk refer to the conversion of 2 as starting material (n.i. = not isolated).
Molecules 26 01899 sch001
Scheme 2. Synthesis of bischalcogenophosphinous acid methylesters 4ab.
Scheme 2. Synthesis of bischalcogenophosphinous acid methylesters 4ab.
Molecules 26 01899 sch002
Figure 1. Molecular structure of 3a (top left), 3b (top right), and 3c (bottom) in solid state. Ellipsoids are shown at 30% probability, and all hydrogen atoms have been omitted for better overview. For 3a, two molecules were found within the unit cell from which one is depicted. Selected bond lengths and angles of 3a: S1-P1 2.1236(8) Å, S2-P2 2.1292(8) Å, ΣP1 = 301.28(8)°, ΣP2 = 303.70(8)°; 3b: Se1-P1 2.2595(11) Å, Se2-P2 2.2599(9) Å, ΣP1 = 304.70(14)°, ΣP2 = 303.92(13)°; and 3c: Te1-P1 2.4660(2) Å, Te2-P2 2.4880(3) Å, ΣP1 = 305.80(4)°, ΣP2 = 306.90(4)°.
Figure 1. Molecular structure of 3a (top left), 3b (top right), and 3c (bottom) in solid state. Ellipsoids are shown at 30% probability, and all hydrogen atoms have been omitted for better overview. For 3a, two molecules were found within the unit cell from which one is depicted. Selected bond lengths and angles of 3a: S1-P1 2.1236(8) Å, S2-P2 2.1292(8) Å, ΣP1 = 301.28(8)°, ΣP2 = 303.70(8)°; 3b: Se1-P1 2.2595(11) Å, Se2-P2 2.2599(9) Å, ΣP1 = 304.70(14)°, ΣP2 = 303.92(13)°; and 3c: Te1-P1 2.4660(2) Å, Te2-P2 2.4880(3) Å, ΣP1 = 305.80(4)°, ΣP2 = 306.90(4)°.
Molecules 26 01899 g001
Figure 2. Molecular structure of 4a (left) and 4b (right) in solid state. Ellipsoids are shown at 30% probability, and all hydrogen atoms have been omitted for better overview. The phosphorus and selenium atoms in 4b showed a small disorder (ca. 7%) over two positions, which is not shown in the figure. Selected bond lengths and angles of 4a: S1-P1 2.1119(13) Å, ΣP1 = 302.31(13)° and 4b: Se1-P1 2.2581(17) Å, ΣP1 = 303.30(20)°.
Figure 2. Molecular structure of 4a (left) and 4b (right) in solid state. Ellipsoids are shown at 30% probability, and all hydrogen atoms have been omitted for better overview. The phosphorus and selenium atoms in 4b showed a small disorder (ca. 7%) over two positions, which is not shown in the figure. Selected bond lengths and angles of 4a: S1-P1 2.1119(13) Å, ΣP1 = 302.31(13)° and 4b: Se1-P1 2.2581(17) Å, ΣP1 = 303.30(20)°.
Molecules 26 01899 g002
Figure 3. 31P-NMR spectrum (left) and 77Se-NMR spectrum (right) of 3b (C6D6, 300 K).
Figure 3. 31P-NMR spectrum (left) and 77Se-NMR spectrum (right) of 3b (C6D6, 300 K).
Molecules 26 01899 g003
Scheme 3. Formation of coinage metal complexes with bisselenophosphinous acid phenylester 3b as ligand. i: 1 equiv. Cu(MeCN)4BF4, rt, 1 h; ii: 1 equiv. AgBF4, rt, 1 h; and iii: 2 equiv. (tht)AuCl, rt, 1 h.
Scheme 3. Formation of coinage metal complexes with bisselenophosphinous acid phenylester 3b as ligand. i: 1 equiv. Cu(MeCN)4BF4, rt, 1 h; ii: 1 equiv. AgBF4, rt, 1 h; and iii: 2 equiv. (tht)AuCl, rt, 1 h.
Molecules 26 01899 sch003
Figure 4. Measured (top) and simulated (bottom) 77Se-NMR spectrum of silver(I) complex 6 (MeCN-d3, 300 K). The following coupling constants were used for the simulation: 1JPSe = 286 Hz, 2JPP = 152 Hz, 1JP-(107)Ag = 401 Hz, 1JP-(109)Ag = 462 Hz, 3JPSe = 0 Hz, 2JSe-(107)Ag = 2JSe-(109)Ag = 0 Hz.
Figure 4. Measured (top) and simulated (bottom) 77Se-NMR spectrum of silver(I) complex 6 (MeCN-d3, 300 K). The following coupling constants were used for the simulation: 1JPSe = 286 Hz, 2JPP = 152 Hz, 1JP-(107)Ag = 401 Hz, 1JP-(109)Ag = 462 Hz, 3JPSe = 0 Hz, 2JSe-(107)Ag = 2JSe-(109)Ag = 0 Hz.
Molecules 26 01899 g004
Figure 5. Molecular structure of 5 and 6 in solid state. Ellipsoids are shown at 30% probability, and all hydrogen atoms have been omitted for clarity. The phenyl substituents at Se1 and Se2 are indicated as wireframe for better overview. For 5 and 6, two molecules were found within the unit cell from which one is depicted. Selected bond lengths and angles of 5: Se1-P1 2.2556(14) Å, Se2-P2 2.2408(14) Å, Cu1-P1 2.2465(14) Å, Cu1-P2 2.2611(14) Å, Cu1-N1 1.944(5) Å, tilt angle α =2.3(2)°, torsion angle τ = 47.97° and 6: Se1-P1 2.2523(7) Å, Se2-P2 2.2542(7) Å, Ag1-P1 2.4758(7) Å, Ag1-P2 2.4438(7) Å, Ag1-N1 2.233(3) Å, tilt angle α = 2.25(14)°, torsion angle τ = 54.97°.
Figure 5. Molecular structure of 5 and 6 in solid state. Ellipsoids are shown at 30% probability, and all hydrogen atoms have been omitted for clarity. The phenyl substituents at Se1 and Se2 are indicated as wireframe for better overview. For 5 and 6, two molecules were found within the unit cell from which one is depicted. Selected bond lengths and angles of 5: Se1-P1 2.2556(14) Å, Se2-P2 2.2408(14) Å, Cu1-P1 2.2465(14) Å, Cu1-P2 2.2611(14) Å, Cu1-N1 1.944(5) Å, tilt angle α =2.3(2)°, torsion angle τ = 47.97° and 6: Se1-P1 2.2523(7) Å, Se2-P2 2.2542(7) Å, Ag1-P1 2.4758(7) Å, Ag1-P2 2.4438(7) Å, Ag1-N1 2.233(3) Å, tilt angle α = 2.25(14)°, torsion angle τ = 54.97°.
Molecules 26 01899 g005
Scheme 4. Synthesis of an ionic gold(I) metal complex with 3b as ligand. i: 1 equiv. (tht)AuCl and AlCl3, rt, 30 min.
Scheme 4. Synthesis of an ionic gold(I) metal complex with 3b as ligand. i: 1 equiv. (tht)AuCl and AlCl3, rt, 30 min.
Molecules 26 01899 sch004
Figure 6. Molecular structure of 7 (top) and 8 (bottom) in solid state. Ellipsoids are shown at 30% probability, and all hydrogen atoms have been omitted for clarity. The phenyl substituents at Se1A, Se2, Se3A, and Se4 in 8 are indicated as wireframe for better overview. The phosphorus (P1) and selenium atoms (Se1 and Se2) as well as the tert-butyl- and phenyl substituents at these atoms showed a disorder over two positions in 8, which is not shown in the figure. Selected bond lengths and angles of 7: Se1-P1 2.2541(10) Å, Se2-P2 2.2546(10) Å, Au1-P1 2.2277(9) Å, Au2-P2 2.2307(9) Å, Au1-Cl1 2.2741(9) Å, Au2-Cl2 2.2840(9) Å, torsion angle τ = 177.44° and 8: Se1A-P1A 2.277(8)Å, Se2-P2 2.234(2) Å, Se3A-P3 2.135(4) Å, Se4-P4 2.2332(19) Å, Au1-P1A 2.337(9) Å, Au1-P3 2.314(2) Å Au2-P2 2.3104(17) Å, Au2-P4 2.3131(17) Å, P3-Au1-P1A 173.95(18)°, P2-Au2-P4 174.91(6)°.
Figure 6. Molecular structure of 7 (top) and 8 (bottom) in solid state. Ellipsoids are shown at 30% probability, and all hydrogen atoms have been omitted for clarity. The phenyl substituents at Se1A, Se2, Se3A, and Se4 in 8 are indicated as wireframe for better overview. The phosphorus (P1) and selenium atoms (Se1 and Se2) as well as the tert-butyl- and phenyl substituents at these atoms showed a disorder over two positions in 8, which is not shown in the figure. Selected bond lengths and angles of 7: Se1-P1 2.2541(10) Å, Se2-P2 2.2546(10) Å, Au1-P1 2.2277(9) Å, Au2-P2 2.2307(9) Å, Au1-Cl1 2.2741(9) Å, Au2-Cl2 2.2840(9) Å, torsion angle τ = 177.44° and 8: Se1A-P1A 2.277(8)Å, Se2-P2 2.234(2) Å, Se3A-P3 2.135(4) Å, Se4-P4 2.2332(19) Å, Au1-P1A 2.337(9) Å, Au1-P3 2.314(2) Å Au2-P2 2.3104(17) Å, Au2-P4 2.3131(17) Å, P3-Au1-P1A 173.95(18)°, P2-Au2-P4 174.91(6)°.
Molecules 26 01899 g006
Table 1. 31P NMR chemical shifts and coupling constants of 3ac and 4ab.
Table 1. 31P NMR chemical shifts and coupling constants of 3ac and 4ab.
δ(31P) [ppm]JPCh [Hz]
3a51.2 (s); 50.2 (s)---
3b57.8 (s); 57.1 (s)1J = 225; 2J = 5
3c45.8 (s); 45.3 (s)1J = 497; 2J = 5
4a56.8 (s); 56.5 (s)---
4b58.7 (s); 58.4 (s)1J = 218; 2J = 6
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Franz, R.; Bruhn, C.; Pietschnig, R. C2-Symmetric P-Stereogenic Ferrocene Ligands with Heavier Chalcogenophosphinous Acid Ester Donor Sites. Molecules 2021, 26, 1899. https://doi.org/10.3390/molecules26071899

AMA Style

Franz R, Bruhn C, Pietschnig R. C2-Symmetric P-Stereogenic Ferrocene Ligands with Heavier Chalcogenophosphinous Acid Ester Donor Sites. Molecules. 2021; 26(7):1899. https://doi.org/10.3390/molecules26071899

Chicago/Turabian Style

Franz, Roman, Clemens Bruhn, and Rudolf Pietschnig. 2021. "C2-Symmetric P-Stereogenic Ferrocene Ligands with Heavier Chalcogenophosphinous Acid Ester Donor Sites" Molecules 26, no. 7: 1899. https://doi.org/10.3390/molecules26071899

Article Metrics

Back to TopTop