Skip to main content
Wiley Open Access Collection logoLink to Wiley Open Access Collection
. 2022 Jan 20;28(7):e202104135. doi: 10.1002/chem.202104135

Borane Adducts of Aromatic Phosphorus Heterocycles: Synthesis, Crystallographic Characterization and Reactivity of a Phosphinine‐B(C6F5)3 Lewis Pair

Jinxiong Lin 1,+, Friedrich Wossidlo 1,+, Nathan T Coles 1,2, Manuela Weber 1, Simon Steinhauer 1, Tobias Böttcher 3,, Christian Müller 1,
PMCID: PMC9303379  PMID: 34967480

Abstract

A phosphinine‐borane adduct of a Me3Si‐functionalized phosphinine and the Lewis acid B(C6F5)3 has been synthesized and characterized crystallographically for the first time. The reaction strongly depends on the nature of the substituents in the α‐position of the phosphorus heterocycle. In contrast, the reaction of B2H6 with various substituted phosphinines leads to an equilibrium between the starting materials and the phosphinine–borane adducts that is determined by the Lewis basicity of the phosphinine. The novel phosphinine borane adduct (6‐B(C6F5)3) shows rapid and facile insertion and [4+2] cycloaddition reactivity towards phenylacetylene. A hitherto unknown dihydro‐1‐phosphabarrelene is formed with styrene. The reaction with an ester provides a new, facile and selective route to 1‐R‐phosphininium salts. These salts then undergo a [4+2] cycloaddition in the presence of Me3Si−C≡CH and styrene to cleanly form unprecedented derivatives of 1‐R‐phosphabarrelenium salts.

Keywords: crystallography, density functional calculations, heterocycles, phosphinine, phosphorus


Roll out the barrelene: The synthesis and crystallographic characterization of a phosphinine–borane adduct has been achieved for the first time. The novel Lewis adduct shows rapid insertion and [4+2] cycloaddition reactivity towards phenylacetylene, whereas a hitherto unknown dihydro‐1‐phosphabarrelene is formed with styrene. Reaction with an ester provides a new, facile and selective route to 1‐R‐phosphininium salts. These subsequently undergo a [4+2] cycloaddition in the presence of alkenes and alkynes to cleanly form unprecedented derivatives of 1‐R‐phosphabarrelenium salts.

graphic file with name CHEM-28-0-g019.jpg

Introduction

Tertiary phosphines (PR3) readily react with various electrophilic species. For instance, they can be protonated and also form phosphine‐borane adducts upon reaction with boranes (BR3). [1] Boranes are particularly important as protecting groups for phosphines, especially for the synthesis of P‐stereogenic compounds, as they can easily be removed afterwards by reaction with amines or more basic phosphines. [2] In recent years, the interaction of phosphines with boranes has been widely explored in the context of “frustrated Lewis pairs” (FLPs).3 The reactivity of phosphines towards boranes is typically related to the pronounced basicity and nucleophilicity of phosphines, but steric factors on both the donor and the acceptor site also play a significant role.[ 3 , 4 ]

In contrast to classical phosphines, λ3‐phosphinines (phosphabenzenes) are extremely weak bases and very poor nucleophiles. This can be attributed to the rather high 3 s character of the phosphorus lone pair in C5H5P, which is much higher than the value found for the nitrogen lone pair in pyridine C5H5N (64 % 3 s vs. 29 % 2 s character). [5] Nevertheless, phosphinines are very good π‐acceptors due to an energetically low‐lying LUMO. [6] Thus, they readily form coordination compounds particularly with late transition metals in low oxidation states.[ 5 , 7 ] On the other hand, little is known about the reactivity of phosphinines towards main group‐based Lewis acids. Reed and co‐worker succeeded in protonating the phosphorus atom in 2,4,6‐tris‐tBu‐λ3‐phosphinine with the in situ generated, non‐oxidizing superacid H[CHB11Me5Cl6] to afford the salt [H(C5H2 t Bu3P)][CHB11Me5Cl6] (1‐H, Figure 1). [8]

Figure 1.

Figure 1

Reactivity of phosphinines towards Lewis acids and selenium, and a brief summary of this work.

Recently we have shown direct methylation of phosphinine 2 by using the strong alkylating reagent [(CH3)2Cl]+[Al(OTeF5)4] (2‐CH3). [9] According to DFT calculations by Erhardt and Frenking, the bond dissociation energy of a phosphinine–borane adduct is about 70 % that of a pyridine‐borane adduct (25.8 vs. 35.6 kcal mol−1). [10] Based on this calculated value, the formation of phosphinine‐borane adducts should consequently be possible. As a first indication, Nöth and Deberitz observed a decrease in the vapor pressure upon adding B2H6 to 2,4,6‐triphenylphosphinine (3) at low temperatures. [11] This indicated the formation of the adduct 3‐BH3 (Figure 1). At room temperature, however, the equilibrium shifted again to the starting materials. Even though this observation was mentioned in a footnote almost 50 years ago, no additional reports on adduct formations of phosphinines with boranes, or other main group compounds have been published since, even though borane adducts with five‐membered aromatic phosphorus heterocycles are known. [12] In this respect, we recently found that the basicity and nucleophilicity of phosphinines can be increased significantly by introducing σ‐donating Me3Si substituents to the aromatic phosphorus heterocycle. [13] This allowed for the synthesis and characterization of the first phosphinine selenides (4=Se, 5=Se). [14] These results prompted us to investigate the reaction between phosphinines and boranes in more detail and in this work we report the first preparation, crystallographic characterization and reactivity of a phosphinine‐borane adduct.

Results and Discussion

As mentioned above, the presumed existence of a phosphinine‐borane adduct (3‐BH3) is based solely on the observed reduction of the vapor pressure when adding B2H6 to 2,4,6‐triphenylphosphinine 3 (Figure 1). In order to obtain more conclusive data and to verify the assumption, we first decided to repeat this experiment. Diborane was condensed into a solution of 3 in dichloromethane at T=−70 °C. In order to shift the equilibrium more to the side of the product, an excess of diborane was used and the solution was investigated by means of low‐temperature NMR spectroscopy. At T=−70 °C, the 31P{1H} NMR spectrum of the reaction mixture shows a second, yet rather small, signal at δ=163.1 ppm (cf. δ=179.0 ppm for 3). The equilibrium of the reaction is clearly located on the side of the starting materials, even though a large excess of diborane was used. This is also confirmed by the 11B NMR spectrum of the solution. Even if the experiment corroborates the evidence for the existence of a phosphinine‐borane adduct, any subsequent isolation and characterization of the phosphinine–borane adduct remained unsuccessful due to its low concentration (ratio 3/3‐BH3≈20 : 1). We anticipated that the higher basicity and nucleophilicity of Me3Si‐substituted phosphinines could shift the equilibrium in favor of the sought‐after products. Therefore, phosphinines 4 and 5 were reacted with a slight excess of diborane (Scheme 1).[ 13 , 15 ] Analogous to the reaction of 3 with B2H6, the reaction mixture was kept at temperatures below T=−70 °C and investigated by means of low‐temperature NMR spectroscopy. For both reactions a new, broad signal can be observed in the 31P{1H} NMR spectra, which is slightly shifted to higher field when compared to the starting material (4: δ=230.7 ppm; 4‐BH3: δ=209.4 ppm; 5: δ=256.4 ppm; 5‐BH3: δ=226.0 ppm).

Scheme 1.

Scheme 1

Reaction of phosphinines 4 and 5 with diborane.

Much to our delight, phosphinine 5 shows the highest conversion (5‐BH3/5=98 : 2), while phosphinine 4 forms considerably less of the phosphinine–borane adduct (4‐BH3/4=88 : 12, Figure 2a).

Figure 2.

Figure 2

a) 31P{1H} NMR spectrum of the reaction of 4 with B2H6. b) Enlargement of the signal of 4‐BH3. c) 11B NMR spectrum of 4‐BH3.

This is perfectly in line with our assumption that the more basic phosphinine 5 should shift the equilibrium more to the right side of the reaction equation (see above and Scheme 1). We noticed, however, that increasing the temperature to room temperature only had a marginal influence on the position of the equilibrium. In the case of the reaction of 4 with B2H6, a splitting of the new signal in the 31P{1H} NMR spectrum into a doublet (1 J (P,B)=18 Hz) can be observed (Figure 2b). As 11B has a nuclear spin of 3/2 , splitting of the 31P{1H} NMR signal into a quartet is to be expected. Our observation might thus be due to a significant broadening of these signals. In the 11B NMR spectra of the reaction mixtures, a new signal at δ=−34.6 (5‐BH3) or −35.3 ppm (4‐BH3) can be observed in addition to the signal for excess diborane. The chemical shifts are in good agreement with values typically found for phosphine‐borane adducts. Moreover, the 11B signal for 4‐BH3 shows a quartet, caused by the coupling to the three 1H atoms. This signal is additionally split into a doublet by coupling to the 31P atom with 1 J (B,P)=18 Hz. The same coupling constant is also found in the 31P{1H} NMR spectrum (Figure 2c). This confirms the existence of a phosphinine–borane adduct. The 1H NMR spectra of the reaction mixtures show that the aromatic protons are somewhat more strongly shielded than in the free phosphinines. The signals of the borane protons can be found at δ=1.5 ppm.

As B2H6 is highly reactive, we also attempted to use the commercially available BH3 ⋅ SMe2 adduct as a borane source for the reaction with 4. In this case, we could also observe the formation of 4‐BH3, however, the equilibrium of the reaction was clearly located on the side of the starting material, even when using a large excess of BH3⋅SMe2. This can most likely be attributed to a competition between dimethylsulfide and the phosphinine to form an adduct with the Lewis acidic borane. Once again, this clearly shows the low basicity and nucleophilicity of phosphinines compared to classical phosphines.

During the course of our NMR investigations, we noticed that a subsequent reaction took place with the initially formed phosphinine–borane adducts 4‐BH3 and 5‐BH3, even at T=−70 °C, which made crystallization and further characterization of 4‐BH3 and 5‐BH3 impossible. While warming the NMR samples up to room temperature, a colorless, viscous oil formed inside the NMR tube and within a few weeks, a slightly yellow solid formed. Additionally, when the adducts were stored at low temperatures, a colorless solid formed over time, which turned into a pale yellow solid as the final product after a few months. In the 11B NMR spectrum only a large number of very broad signals can be detected while the 31P{1H} NMR spectra showed mainly a very broad signal at δ=21.4 ppm. Boranes are frequently used in the hydroboration of double bonds and the hydroboration of phospha‐alkenes has previously been reported. [16] It is therefore reasonable to assume that the subsequent hydroboration of the P=C double bond in 4‐BH3 and 5‐BH3 slowly takes place, while insoluble polymers are finally formed (Scheme 2). According to quantum chemical calculations by Ermolaeva and Ionkin as well as experimental work by Yoshifuji et al. on the reaction of phospha‐alkenes with BH3, we suggest that the regioisomer 5(H)‐BH2, shown in Scheme 2, is initially formed.[ 17 , 18 ] The chemical shift of δ=−21.4 ppm, observed during the further conversion of 5‐BH3, is in line with the chemical shift of δ=−6.4 ppm reported by Yoshifuji et al. for the product of the hydroboration of a phospha‐alkene, which subsequently dimerized. [18] Moreover, Arbuzov and co‐workers reported polymerization reactions with hydroborated phospha‐alkenes. [16] This finding could explain the formation of the insoluble solid in the reaction mixtures of 4‐BH3 and 5‐BH3, which could not be further characterized thus far.

Scheme 2.

Scheme 2

Potential hydroboration of the P=C double bond in 5‐BH3 and subsequent polymerization.

In order to avoid hydroboration of the P=C double bond, we decided to use the strong Lewis acid B(C6F5)3 in combination with phosphinines 35 in CH2Cl2 at room temperature. However, no reaction could be observed as judged by 31P{1H} NMR spectroscopy (Scheme 3a).

Scheme 3.

Scheme 3

Reaction of phosphinines 35 with B(C6F5)3.

Apparently, formation of the adduct between phosphinines 35 and the sterically demanding B(C6F5)3 is hindered by the bulky phenyl and Me3Si substituents at the α‐position of the phosphinine. We therefore considered the reaction of phosphinine 6, which contains Me3Si substituents at the 3,5‐positions of the ring, with B(C6F5)3 in CH2Cl2 at room temperature. Compound 6 was recently reported by us and its gas phase basicity was calculated to be higher than that of unsubstituted phosphinine C5H5P, but lower than for 5, which contains Me3Si groups at the 2,6‐positions. [14]

Nevertheless, we were delighted to detect only one signal in the 31P{1H} NMR spectrum of the reaction mixture, which is shifted upfield with respect to the starting material (δ=176.6 vs. 206.4 ppm). The 1H, 19F and 11B NMR spectra also indicate the successful and quantitative conversion of 6 to the phosphinine–borane adduct 6‐B(C6F5)3 (Scheme 3b).

We considered that the high Lewis acidity of B(C6F5)3 might also allow the formation of adducts with less basic phosphinines, such as 3,5‐diphenylphosphinine (7). Gratifyingly, the quantitative formation of 7‐B(C6F5)3 was also observed upon reaction of 7 with B(C6F5)3 in CH2Cl2 at room temperature (31P{1H} NMR: δ=182.7 ppm). As already anticipated, these results clearly show that steric factors play a significant role in the interaction between phosphinines and boranes with strong Lewis acidic properties.

Crystals of 6‐B(C6F5)3, suitable for X‐ray diffraction, were obtained by slow evaporation of a solution of the phosphinine–borane adduct in a mixture of dichloromethane and n‐pentane and the molecular structure of the first crystallographically characterized phosphinine‐borane adduct, along with selected bond lengths, angles and distances are depicted in Figure 3.

Figure 3.

Figure 3

a) Top and b) side views of the Molecular structure of 6‐B(C6F5)3 in the crystal. Displacement ellipsoids are shown at the 50 % probability level. Selected bond lengths [Å] and angles [°]: P(1)−B(1): 2.0415(12); P(1)−C(1): 1.7046(11); P(1)−C(5) 1.1.7082(10); P(1)⋅⋅⋅F(5) 2.8679(8); C(1)−P(1)−C(5): 106.87(5); B(1)−P(1)−C(3): 168.36(4); Σ(C−B−C): 338.14(9).

The P(1)−B(1) distance of 2.042 Å is shorter than in the corresponding triphenylphosphine adduct Ph3P→B(C6F5)3 (2.181 Å). [19] Figure 3 clearly shows that sterically demanding substituents in the α‐position of the phosphorus heterocycle, such as Me3Si or Ph groups, would indeed prevent any adduct formation. Interestingly, a closer look at the solid‐state structure of 6‐B(C 6F5)3 reveals that F(5) is located directly above the phosphorus atom. One of the electron lone pair of F(5) points directly towards the LUMO of the phosphinine, while the distance between F(5) and P(1) is 2.868 Å, which is shorter than the sum of the van der Waals radii (3.34 Å). [20] Additionally, the boron‐phosphorus bond is slightly tilted out of the plane of the phosphinine ring (Figure 3b). These observations could indicate a significant fluorine–phosphorus interaction, which might stabilize the adduct further. However, the NMR spectra do not show such an interaction at room temperature nor at T=−70 °C, as all ortho‐fluorine atoms remain magnetically equivalent.

In order to gain more insight into a possible interaction between P(1) and F(5), the electron density distributions were calculated for 6‐B(C6F5)3, and the results are depicted in Figure 4 as Laplacian plots through the atoms P(1), B(1) and F(5). This indicates that a bond‐critical point (red arrow) can be detected between P(1) and F(5), however, the value of 0.098 e Å−3 is very low. In comparison, the value for the bond critical point between P(1) and B(1) is 0.678 e Å−3. We therefore suggest that the presence of a bond‐critical point between P(1) and F(5) should not be over interpreted.

Figure 4.

Figure 4

2D contour plot of the Laplacian of the electron density ▿2 ρ(r) in the P(1)−B(1)−F(5) plane of 6‐B(C6F5)3. Red dashed contours indicate regions of local charge accumulation (▿2 ρ(r)<0); Blue contours indicate regions of local charge depletion (▿2 ρ(r)>0). Bond paths are shown as gray lines. Bond critical points (BCPs) shown as blue dots. The red arrows indicate the electron density ρ(r) [e A−3] for the BCPs.

Instead, we propose on the basis of the calculations and the charges of the atoms that the fluorine–phosphorus interaction can rather be attributed as an electrostatic effect. This is best visualized in the electrostatic potential plot of 6‐B(C6F5)3, in which the green, negatively polarized fluorine atom (F5) is directly positioned above the red, positively polarized phosphorus atom (Figure 5, left). For comparison, the electrostatic potential plot of the pyridine–B(C6F5)3 adduct was also calculated (Figure 5 right). In this case, a higher electron density can be observed at the heteroatom of the ring due to the higher electronegativity of nitrogen, compared to phosphorus. Consequently, the fluorine atom is positioned above the C−N bond. The B(C6F5)3 moiety shows the expected propeller‐like geometry with no unusually short N−F contacts.

Figure 5.

Figure 5

Electrostatic potential of 6‐B(C6F5)3 (left) and pyridine‐B(C6F5)3 (right) calculated at the B3LYP‐D3(BJ)/def2‐TZVPP level of theory. The electrostatic potential [a.u.] is mapped onto electron density isosurfaces of 0.02 e au−3.

As expected, we found that the phosphorus‐boron interaction in such Lewis pairs is rather weak, due to the low basicity and nucleophilicity of the aromatic phosphorus heterocycle (see above). The calculated bond dissociation enthalpy for the reaction LB+B(C6F5)3→LB−B(C6F5)3 (LB=Lewis base) is ΔH°=−54.9 kJ ⋅ mol−1 for 6‐B(C6F5)3, which is almost half the value obtained for the triphenylphosphine adduct Ph3P−B(C6F5)3H°=−110.2 kJ ⋅ mol−1).

Based on our observations and due to the strong polarization of the phosphorus heterocycle caused by the interaction with the Lewis acid, we anticipated that 6‐B(C6F5)3 might exhibit a pronounced reactivity particularly towards unsaturated substrates. Previous studies have shown that activated alkynes can react with certain phosphinines, phosphinine–metal complexes, 1‐methylphosphininium salts, phosphinine‐sulfides and phosphinine‐selenides to form 1‐phosphabarrelene derivatives.[ 14 , 21 , 22 , 23 ] This orbital controlled [4+2] cycloaddition reaction proceeds by a 1,4‐addition of the dienophile across the phosphorus heterocycle. Again, this process is facilitated by increasing the polarization of the aromatic phosphinine, for example, by P coordination to a metal center or by oxidation to a formal P(V) derivative. In this respect, we first considered the reaction between 6‐B(C6F5)3 and the less reactive phenylacetylene. So far, no 1‐phosphabarrelene derivatives, generated from this dienophile and the corresponding phosphorus heterocycle, have been reported to date. Much to our delight, upon addition of excess PhC≡CH, we could observe a quantitative and selective reactivity at room temperature to a new compound (9) after 100 min. The product shows a single resonance at δ=−18.3 ppm in the 31P{1H} NMR spectrum. However, its analysis by high resolution mass spectrometry indicates that 9 is not only a simple borane adduct of a 1‐phosphabarrelene. The crystallographic characterization reveals that next to a regioselective cycloaddition of phenylacetylene to the phosphinine ring, a further insertion of PhC≡CH into the P→B bond occurred and the zwitterionic alkenyl‐phosphabarrelenium borate salt 9 had been formed (Scheme 4 and Figure 6).

Scheme 4.

Scheme 4

Reaction of 6‐B(C6F5)3 with excess phenylacetylene.

Figure 6.

Figure 6

Molecular structure of 9 in the crystal. Displacement ellipsoids are shown at the 50 % probability level for one orientation of the 1‐phosphabarrelene moiety. Selected bond lengths [Å]: P(1)−C(8): 1.767(4); C(8)−C(9): 1.344(6); C(9)−B(1): 1.626(6).

The regioselective activation of terminal alkynes both by frustrated and classical Lewis acid/phosphine pairs has been reported in literature before. Nevertheless, the formation of phosphonium alkynyl borate salts (R3PH+ PhC≡C−B(C6F5)3 ) is also observed in this case, resulting from deprotonation of the alkyne by the basic phosphine. [24] By performing the reaction of 6‐B(C6F5)3 with PhC≡CH in a 1 : 1 ratio, additional information on the mechanism of the product formation could be obtained. Figure 7 shows the 31P{1H} NMR spectra of the reaction mixture at t=10 and 80 min after addition of the alkyne to the phosphinine–borane adduct.

Figure 7.

Figure 7

31P{1H} NMR spectra for the reaction of 6‐B(C6F5)3 with PhC≡CH at a 1 : 1 ratio.

Interestingly, a transient intermediate (8) can be observed during the course of the reaction, which fully converts to 9 when access phenylacetylene is applied. In the 1 : 1 reaction, the ratio between the starting material 6‐B(C6F5)3, 8 and the product 9 remains constant after t=80 min. The chemical shift of 8 at δ=140.0 ppm in the 31P{1H} NMR spectrum is characteristic for a phosphininium salt (see below). We therefore anticipate that intermediate 8 is a zwitterionic alkenyl‐phosphininium borate salt, which is first generated by insertion of PhC≡CH into the dative P→B bond (Scheme 5). Intermediate 8 further reacts rapidly in a [4+2] cycloaddition reaction with PhC≡CH to form the final product 9 (Scheme 5). As a matter of fact, 1‐R‐phosphininium salts readily react with alkynes to 1‐R‐phosphabarrelenium salts even at low temperature.[ 22d , 25 ] Cycloaddition reactions at highly polarized Te/B‐heterocycles have also been observed by Stephan et al. [26] By means of 31P NMR spectroscopy we can further rule out that 8 is the 1‐H‐phosphininium salt, which might be generated by deprotonation of phenylacetylene.

Scheme 5.

Scheme 5

Reaction of 6‐B(C6F5)3 with excess phenylacetylene.

As cycloaddition reactions of C=C double bonds to six‐membered aromatic phosphorus heterocycles are unprecedented in the literature, we subsequently treated 6‐B(C6F5)3 with styrene (Scheme 6).

Scheme 6.

Scheme 6

Reaction of 6‐B(C6F5)3 with styrene.

Again, we could observe the rapid, quantitative and selective conversion of 6‐B(C6F5)3 to a new product, which shows a single resonance at δ=−38.0 ppm in the 31P{1H} NMR spectrum. Crystals of the product, suitable for X‐ray diffraction were obtained by slow evaporation of the solvent. The result of the X‐ray crystal structure analysis is depicted in Figure 8 along with selected bond lengths and angles and unambiguously confirms that a racemic mixture of the cycloaddition product 10‐B(C6F5)3 had been formed. This compound represents the first example of a dihydro‐1‐phosphabarrelene derivative. It is interesting to note that only one regioisomer is formed in the conversion of 6‐B(C6F5)3 with styrene. In 10‐B(C6F5)3, the phenyl group of the former styrene substrate is pointing away from the B(C6F5)3 moiety, most likely due to steric effects during the [4+2] cycloaddition step, as this reaction is orbital controlled.[ 22 , 23 ]

Figure 8.

Figure 8

Molecular structure of 10‐B(C6F5)3 in the crystal. Displacement ellipsoids are shown at the 50 % probability level for one enantiomer. Selected bond lengths [Å] and angles [°]: P(1)−B(1): 2.030(3); P(1)−C(1): 1.806(7); P(1)−C(5): 1.793(3); P(1)−C(6): 1.840(3); C(1)−C(2): 1.340(3); C(4)−C(5): 1.341(4); C(6)−C(7): 1.559(4). C(1)−P(1)−C(5): 100.21(12); C(1)−P(1)−C(6): 98.04(12); C(5)−P(1)−C(6): 100.54(12).

Motivated by these initials results, we attempted to apply 6‐B(C6F5)3 in typical frustrated Lewis‐pair reactions. [27] As styrene derivatives react with diaryl‐substituted esters in the presence of the FLP system R3P/B(C6F5)3 to form substituted olefins, we decided to treat 6‐B(C6F5)3 first with ester 11 (Scheme 7). [28]

Scheme 7.

Scheme 7

Study of the reactivity of 6‐B(C6F5)3 towards ester 11.

Much to our surprise, the reaction of 6‐B(C6F5)3 with 11 in CDCl3 proceeds instantaneously. The NMR spectroscopic analysis of the reaction mixture reveals again the selective and quantitative formation of a single, new species. This compound shows a resonance at δ=156.8 ppm in the 31P NMR spectrum, which is corroborated in the corresponding 1H NMR spectrum with a coupling constant of 2 J P‐H=27.0 Hz to the alkyl CH. This is in line with the presence of the 1‐R‐phosphininium salt 12, as the 31P{1H} NMR signal of the related 1‐methyl‐phosphininium cation can be observed at δ=160.2 ppm.[ 9 , 21 ] Most intriguingly, phosphininum salts can currently only be prepared in a multistep synthesis, or alternatively, with the very strong methylation reagent [(CH3)2Cl]+[Al(OTeF5)4] due to the low nucleophilicity of the phosphorus atom (see above, Figure 1).[ 9 , 21 , 29 ] Our first results, depicted in Scheme 8, thus impressively show that various novel phosphininium salts might be easily formed by making use of phosphinine–borane adducts in combination with aromatic alkyl esters. To this end, 1‐R‐phosphininium salts can be generated and further explored, which are otherwise synthetically inaccessible. However, at this point, we do not have any insight into the reaction mechanism of this surprisingly fast quaternization reaction, which can occur either through a single‐electron (radical) or a two‐electron transfer mechanism during the C−O bond scission reaction. [28]

Scheme 8.

Scheme 8

Reaction of 6 with B(C6F5)3, ester 11, styrene and TMS‐acetylene.

Subsequently, we added styrene to 11 and observed again the spontaneous formation of a single species, which shows a signal at δ=−6.4 ppm in the 31P{1H} NMR spectrum. The same immediate reactions occurs, when 6 was mixed with B(C6F5)3, 11 and styrene in an equimolar ratio with CHCl3 as solvent. However, we could not detect any signals of the substituted olefin 13 by means of 1H NMR spectroscopy (Scheme 8).

Crystals, suitable for X‐ray diffraction were obtained of the reaction product 14 by slow evaporation of the solvent. The result of the X‐ray crystal structure analysis is depicted in Figure 8 along with selected bond lengths and angles.

Interestingly, crystallographic characterization of 14 (Figure 9) reveals the first example of a dihydro‐1‐R‐phosphabarrelenium salt. The compound forms by a fast and regioselective [4+2] cycloaddition of styrene to the aromatic 1‐R‐phosphininium heterocycle. It should be mentioned that the reaction of a related 1‐methyl‐1‐phospha‐7‐bora‐norbornadiene with phenylacetylene has recently been described by us. [30]

Figure 9.

Figure 9

Molecular structure of 14 in the crystal. Displacement ellipsoids are shown at the 50 % probability level. Hydrogen atoms and co‐crystallized THF are omitted for clarity. Selected bond lengths [Å]: P(1)−C(1): 1.783(2); P(1)−C(8): 1.8199(19); P(1)−C(5): 1.779(2); P(1)−C(6): 1.821(2); C(1)−C(2): 1.347(3); C(2)−C(3): 1.525(3); C(3)−C(4): 1.533(3); C(4)−C(5): 1.341(3); B(1)−O(1): 1.514(2).

Finally, we also converted rapidly, quantitatively and selectively a mixture of 6, ester 11 and Me3Si−C≡CH towards the corresponding 1‐R‐phosphabarrelenium salt 15 (Scheme 8), which we could also characterize crystallographically (Figure 10).

Figure 10.

Figure 10

Molecular structure of 15 in the crystal. Displacement ellipsoids are shown at the 50 % probability level. Hydrogen atoms are omitted for clarity. Selected bond lengths [Å] and angles [°]: P(1)−C(17): 1.8235(19); P(1)−C(1): 1.7951(19); P(1)−C(5): 1.7873(19); P(1)−C(6): 1.792(2); C(1)−C(2): 1.342(3); C(2)−C(3): 1.558(3); C(4)−C(5): 1.339(3); C(6)−C(7): 1.338(3); B(1)−O(1): 1.510(2). C(5)−P(1)−C(6): 100.35(9); C(5)−P(1)−C(1): 99.68(9); C(1)−P(1)−C(6): 101.00(9).

Conclusion

In conclusion, we could demonstrate that the equilibrium of the reaction between 2,4,6‐triphenylphosphinine and B2H6 is almost exclusively located on the side of the starting materials, even at low temperature. In contrast, more basic Me3Si‐substituted phosphinines lead to a shift of the equilibrium towards the products, although complete conversion to phosphinine–borane adducts is still not achieved. This demonstrates that Me3Si‐substituted phosphinines are still rather weak Lewis bases. Moreover, hydroboration of the P=C double bond is observed with B2H6, which presumably leads to the formation of polymeric species. The use of the much stronger, yet more sterically demanding, Lewis acid B(C6F5)3 reveals a strong influence of the α substituents of the aromatic phosphorus heterocycle on the formation of a discrete adduct. Phosphinines with Ph or Me3Si substituents in the ortho position of the phosphinine prevent any reaction with B(C6F5)3, whereas 3,5‐bis(trimethylsilyl)phosphinine as well as 3,5‐diphenylphosphinine undergo complete conversion to the phosphinine–borane adduct. In the case of 3,5‐bis(trimethylsilyl)phosphinine, the product could be characterized crystallographically. The solid‐state structure reveals an interaction between the phosphorus atom and one of the fluorine atoms of the Lewis acid; this is supported by theoretical calculations. We could further highlight that the novel phosphinine–borane adduct shows distinct insertion and subsequent [4+2] cycloaddition reactivity towards phenylacetylene. This results in a zwitterionic alkenyl‐phosphabarrelenium borate salt that is formed selectively and quantitatively. With styrene, the borane adduct of a dihydrophosphabarrelene is formed. In the presence of an ester, the Lewis pair forms 1‐R‐phosphininium salts. This route significantly improves upon the previous multistep synthesis, as this procedure is performed at room temperature in a fast, facile, and selective manner. The cationic heterocycle was then shown to readily undergo a [4+2] cycloaddition reaction with styrene ultimately forming the first example of a dihydro‐1‐R‐phosphabarrelenium salt. The [4+2] cycloaddition of the 1‐R‐phosphininium salt generated in situ with TMS‐acetylene quantitatively affords the corresponding 1‐R‐phosphabarrelenium salt. The results presented herein highlight how the phosphinine–borane adduct 6‐B(C6F5)3 apparently mimics frustrated Lewis‐pair reactivity. Our results provide fascinating new perspectives for the future, particularly with respect to the activation of small molecules and the synthesis of adducts of phosphinines with other main group elements and compounds. Experiments in this direction are currently being pursued in our laboratories.

Experimental Section

Experimental details are given in the Supporting Information.

Deposition Numbers 2067304 (for 6‐B(C6F5)3), 2121299 (for 9), 2067303 (for 13) and 2121300 (for 14) contain the supplementary crystallographic data for this paper. These data are provided free of charge by the joint Cambridge Crystallographic Data Centre and Fachinformationszentrum Karlsruhe Access Structures service.

Conflict of interest

The authors declare no conflict of interest.

1.

Supporting information

As a service to our authors and readers, this journal provides supporting information supplied by the authors. Such materials are peer reviewed and may be re‐organized for online delivery, but are not copy‐edited or typeset. Technical support issues arising from supporting information (other than missing files) should be addressed to the authors.

Supporting Information

Acknowledgements

Funding by the Freie Universität Berlin and the Deutsche Forschungsgemeinschaft DFG (project no. 2100302201) is gratefully acknowledged. The authors thank the Scientific Computing Service of the Freie Universität Berlin (https://doi.org/10.17169/refubium‐26754) for the use of high‐performance computing resources. Dr. Stephen Argent is thanked for his input on the solution of compound 9. Open Access funding enabled and organized by Projekt DEAL.

J. Lin, F. Wossidlo., N. T. Coles, M. Weber, S. Steinhauer, T. Böttcher, C. Müller, Chem. Eur. J. 2022, 28, e202104135.

In memory of Professor Paul C. J. Kamer.

Contributor Information

Priv.‐Doz. Dr. Tobias Böttcher, Email: tobias.boettcher@ac.uni-freiburg.de.

Prof. Dr. Christian Müller, Email: c.mueller@fu-berlin.de.

Data Availability Statement

The data that support the findings of this study are available in the supplementary material of this article.

References

  • 1. Brunel J. M., Faure B., Maffei M., Coord. Chem. Rev. 1998, 178–180, 665. [Google Scholar]
  • 2. 
  • 2a. Burg A. B., Wagner R. I., J. Am. Chem. Soc. 1953, 75, 3872; [Google Scholar]
  • 2b. Ohff M., Holz J., Quirmbach M., Börner A., Synthesis 1998, 1391. [Google Scholar]
  • 3.See for example:
  • 3a. Stephan D. W., Erker G., Angew. Chem. Int. Ed. 2015, 54, 6400; [DOI] [PubMed] [Google Scholar]
  • 3b. Stephan D. W., Science 2016, 354, aaf7229-1;27940818 [Google Scholar]
  • 3c. Li N., Zhang W.-X., Chin. J. Chem. 2020, 38, 1360. [Google Scholar]
  • 4. Staubitz A., Robertson A. P. M., Sloan M. E., Manners I., Chem. Rev. 2010, 110, 4023. [DOI] [PubMed] [Google Scholar]
  • 5. Le Floch P., Coord. Chem. Rev. 2006, 250, 627. [Google Scholar]
  • 6. 
  • 6a. Waluk J., Klein H.-P., Ashe A. J., Michl J., Organometallics 1989, 8, 2804; [Google Scholar]
  • 6b. Batich C., Heilbronner E., Hornung V., Ashe A. J., Clark D. T., Cobley U. T., Kilcast D., Scanlan I., J. Am. Chem. Soc. 1973, 99, 928. [Google Scholar]
  • 7. 
  • 7a. Le Floch P., Mathey F., Coord. Chem. Rev. 1998, 178–180, 771; [Google Scholar]
  • 7b. Mézailles N., Mathey F., Le Floch P., Prog. Inorg. Chem. 2001, 455; [Google Scholar]
  • 7c. Le Floch P., Coord. Chem. Rev. 2006, 250, 627–681; [Google Scholar]
  • 7d. Coles N. T., Abels A. S., Leitl J., Wolf R., Grützmacher H., Müller C., Coord. Chem. Rev. 2021, 433, 213729. [Google Scholar]
  • 8. Zhang Y., Tham F. S., Nixon J. F., Taylor C., Green J. C., Reed C. A., Angew. Chem. Int. Ed. 2008, 47, 3801. [DOI] [PubMed] [Google Scholar]
  • 9. Fischer L., Wossidlo F., Frost D. S., Coles N. T., Steinhauer S., Riedel S., Müller C., Chem. Commun. 2021, 57, 9522. [DOI] [PubMed] [Google Scholar]
  • 10. Erhardt S., Frenking G., Chem. Eur. J. 2006, 12, 4620. [DOI] [PubMed] [Google Scholar]
  • 11. Dimroth K., Top. Curr. Chem. 1973, 38, 1. [Google Scholar]
  • 12. 
  • 12a. Scheibitz M., Bats J. W., Bolte M., Wagner M., Eur. J. Inorg. Chem. 2003, 2049; [Google Scholar]
  • 12b. Frison G., Mathey F., Sevin A., J. Phys. A 2002, 106, 5653; [Google Scholar]
  • 12c. Panova Y., Khristolyubova A., Zolotareva N., Sushev V., Galperin V., Rumyantcev R., Fukin G., Kornev A., Dalton Trans. 2021, 50, 5890. [DOI] [PubMed] [Google Scholar]
  • 13. 
  • 13a. Habicht M. H., Wossidlo F., Weber M., Müller C., Chem. Eur. J. 2016, 22, 12877; [DOI] [PubMed] [Google Scholar]
  • 13b. Habicht M. H., Wossidlo F., Bens T., Pidko E. A., Müller C., Chem. Eur. J. 2018, 24, 944. [DOI] [PubMed] [Google Scholar]
  • 14. Wossidlo F., Coles N. T., Steinhauer S., Böttcher T., Müller C., Chem. Eur. J. 2021, 27, 12788. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15. Avarvari N., Le Floch P., Mathey F., J. Am. Chem. Soc. 1996, 118, 11978. [Google Scholar]
  • 16. Ionkin A. S., Ignatév S. N., Nekhoroshkov V. M., Efremov J. J., Arbuzov B. A., Phosphorus Sulfur Silicon Relat. Elem. 1990, 53, 1. [Google Scholar]
  • 17. Ermolaeva L. V., Ionkin A. S., J. Mol. Struct. 1992, 276, 25. [Google Scholar]
  • 18. Yoshifuji M., Takahashi H., Toyota K., Heteroat. Chem. 1999, 10, 187. [Google Scholar]
  • 19. Jacobsen H., Berke H., Döring S., Kehr G., Erker G., Fröhlich R., Meyer O., Organometallics 1999, 18, 1724. [Google Scholar]
  • 20. Mantina M., Chamberlin A. C., Valero R., Cramer C. J., Truhlar D. G., J. Phys. Chem. A 2009, 113, 5806. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21. Moores A., Ricard L., Le Floch P., Angew. Chem. 2003, 115, 5090. [DOI] [PubMed] [Google Scholar]
  • 22. 
  • 22a. Märkl G., Lieb F., Angew. Chem. Int. Ed. 1968, 7, 733; [Google Scholar]
  • 22b. Mézailles N., Ricard L., Mathey F., Le Floch P., Eur. J. Inorg. Chem. 1999, 2233; [Google Scholar]
  • 22c. Fuchs E., Keller M., Breit B., Chem. Eur. J. 2006, 12, 6930; [DOI] [PubMed] [Google Scholar]
  • 22d. Moores A., Cantat T., Ricard L., Mézailles N., Le Floch P., New J. Chem. 2007, 31, 1493; [Google Scholar]
  • 22e. Rigo M., Habraken E. R. M., Bhattacharyya K., Weber M., Ehlers A. W., Mézailles N., Slootweg J. C., Müller C., Chem. Eur. J. 2019, 25, 8769. [DOI] [PubMed] [Google Scholar]
  • 23. Bruce M., Papke M., Ehlers A. W., Weber M., Lentz D., Mézailles N., Slootweg J. C., Müller C., Chem. Eur. J. 2019, 25, 14332. [DOI] [PubMed] [Google Scholar]
  • 24. 
  • 24a. Dureen M. A., Stephan D. W., J. Am. Chem. Soc. 2009, 131, 8396; [DOI] [PubMed] [Google Scholar]
  • 24b. Dureen M. A., Brown C. C., Stephan D. W., Organometallics 2010, 29, 6594. [Google Scholar]
  • 25.F. Wossidlo, Ph.D. thesis, Freie Universität Berlin (Germany), 2021.
  • 26.For cycloaddition/cycloreversions at Te/B-heterocycles see: Tsao F. A., Cao L., Grimme S., Stephan D. W., J. Am. Chem. Soc. 2015, 137, 13264. [DOI] [PubMed] [Google Scholar]
  • 27. 
  • 27a. Stephan D. W., Acc. Chem. Res. 2015, 48, 306; [DOI] [PubMed] [Google Scholar]
  • 27b. Johnstone T. C., Wee G. N. J. H., Stephan D. W., Angew. Chem. Int. Ed. 2018, 57, 5881. [DOI] [PubMed] [Google Scholar]
  • 28. 
  • 28a. Soltani Y., Dasgupta A., Gazis T. A., Ould D. M. C., Richards E., Slater B., Stefkova K., Vladimirov V. Y., Wilkins L. C., Willcox D., Melen R. L., Cell Rep. Phys. Sci. 2020, 100016, 1; [Google Scholar]
  • 28b. Dasgupta A., Stefkova K., Rabaahmadi R., Yates B. F., Buurma N. J., Ariafard A., Richards E., Melen R. L., J. Am. Chem. Soc. 2021, 143, 4451. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29. Dave T. N., Kaletsch H., Dimroth K., Angew. Chem. Int. Ed. Engl. 1984, 23, 989. [Google Scholar]
  • 30. Leitl J., Jupp A. R., Habraken E. R. M., Streitferdt V., Coburger P., Scott D. J., Gschwind R. M., Müller C., Slootweg J. C., Wolf R., Chem. Eur. J. 2020, 26, 7788. [DOI] [PMC free article] [PubMed] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

As a service to our authors and readers, this journal provides supporting information supplied by the authors. Such materials are peer reviewed and may be re‐organized for online delivery, but are not copy‐edited or typeset. Technical support issues arising from supporting information (other than missing files) should be addressed to the authors.

Supporting Information

Data Availability Statement

The data that support the findings of this study are available in the supplementary material of this article.


Articles from Chemistry (Weinheim an Der Bergstrasse, Germany) are provided here courtesy of Wiley

RESOURCES