Skip to main content
ACS AuthorChoice logoLink to ACS AuthorChoice
. 2023 Apr 27;145(18):10364–10375. doi: 10.1021/jacs.3c02256

Cationic Phosphinidene as a Versatile P1 Building Block: [LC–P]+ Transfer from Phosphonio–Phosphanides [LC–P–PR3]+ and Subsequent LC Replacement Reactions (LC = N-Heterocyclic Carbene)

Philipp Royla , Kai Schwedtmann , Zeyu Han , Jannis Fidelius , Derek P Gates ‡,*, Rosa M Gomila §, Antonio Frontera §, Jan J Weigand †,*
PMCID: PMC10177976  PMID: 37105536

Abstract

graphic file with name ja3c02256_0014.jpg

Cationic imidazoliumyl(phosphonio)-phosphanides [LC–P–PR3]+ (1a–e+, LC = 4,5-dimethyl-1,3-diisopropylimidazolium-2-yl; R = alkyl, aryl) are obtained via the nucleophilic fragmentation of tetracationic tetraphosphetane [(LC–P)4][OTf]4 (2[OTf]4) with tertiary phosphanes. They act as [LC–P]+ transfer reagents in phospha-Wittig-type reactions, when converted with various thiocarbonyls, giving unprecedented cationic phosphaalkenes [LC–P=CR2]+ (5a-f[OTf]) or phosphanides [LC–P–CR(NR2)]+ (6a-d[OTf]). Theoretical calculations suggest that three-membered cyclic thiophosphiranes are crucial intermediates of this reaction. To test this hypothesis, treatment of [LC–P–PPh3]+ with phosphaalkenes, that are isolobal to thioketones, permits the isolation of diphosphirane salts 11a,b[OTf]. Furthermore, preliminary studies suggest that the cationic phosphaalkene [LC–P=CPh2]+ may be employed to access rare examples of η2–P=C π-complexes with Pd0 and Pt0 when treated with [Pd(PPh3)4] and [Pt(PPh3)3] for which analogous complexes of neutral phosphaalkenes are scarce. The versatility of [LC–P]+ as a valuable P1 building block was showcased in substitution reactions of the transferred LC-substituent using nucleophiles. This is demonstrated through the reactions of 5a[OTf] and 6c[OTf] with Grignard reagents and KNPh2, providing a convenient, high-yielding access to MesP=CPh2 (16) and otherwise difficult-to-synthesize 1,3-diphosphetane 17 and P-aminophosphaalkenes.

Introduction

Simple phosphorus-containing functionalities can represent important tools for the construction of novel structural architectures with application in areas such as catalysis, polymers, and materials. For instance, phosphinidenes [R–P] are considered valuable and simple P1 building blocks for the synthesis of organophosphorus substrates or as diverse ligands in transition-metal complexes.1 Despite their synthetic utility and fundamental curiosity, phosphinidenes display exceedingly high reactivity, and the first isolable “free” phosphinidene, reported in 2016, remains the only example.2 Thus, researchers have designed a variety of more applicable precursors, phosphinidenoids, which can be employed in phosphinidene transfer reactions. Of particular importance were the early investigations of transition-metal-supported phosphinidenoid reagents as [R–P] building blocks, thereby affording otherwise difficult to access organophosphorus compounds.1,3 In a few cases, the release of [R–P] from metal-free precursors has also been described, for example, from so-called inversely polarized phosphaalkenes R–Pδ−=Cδ+R2.4

More recently, a new generation of phosphinidene chemistry has evolved with exciting breakthroughs involving isolable, neutral, and metal-free singlet [R–P] transfer reagents that are tolerable of a variety of substituents (Figure 1a).

Figure 1.

Figure 1

(a) Examples of neutral phosphinidene [R–P] transfer reagents; (b) synthesis of cationic [LC–P]+ transfer reagents reported here. Mes* = 2,4,6-tBuC6H2, Dipp = 2,6-iPr2C6H3.

For instance, amino-phosphinidene, [R2N–P], transfer has been enabled from precursor I with concomitant formation of anthracene.5 Importantly, the first transfers of the parent phosphinidene, [H–P], were observed from II.6 Carbene-phosphinidene adducts III and phosphanylidenephosphoranes IV (or “phospha-Wittig reagents”7) have been shown to transfer aryl- and alkyl-substituted phosphinidenes to a wide range of substrates [e.g., organic electrophiles,8 aldehydes,9 NHCs (N-heterocyclic carbenes),10 isonitriles,11 ammonia,12 and AlI species.13 Diphosphadiboretane V has been utilized as a [Mes*–P] transfer agent to ketones, amides, and esters in the unprecedented phospha-bora-Wittig reaction.14 Despite advances in the field, the development of a single phosphinidene transfer reagent capable of transferring phosphinidenes [R–P] with a multitude of different substituents R is still desired. In a recent study, we demonstrated the versatility of cationically substituted phosphorus compounds for the formation of P–C, P–N, and P–O bonds by easily replacing the cationic substituent using commercially available reagents.15 This inspired us to explore the potential of employing cationic substituents for phosphinidenes to create unprecedented cationic phosphinidene transfer reagents, namely, [LC–P]+ (Figure 1b). This could enable further functionalization at the P atom after the transfer reaction and render [LC–P]+ a versatile P1 building block.

We recently discovered that the tetracationic tetraphosphetane 24+ (Figure 1), formally a tetramer of [LC–P]+, may be conveniently obtained in good yields (86%) as its triflate salt from the reduction of 3[OTf] with 1,4-bis(trimethylsilyl)-1,4-dihydropyrazine (4, Scheme 1).16 Computational studies on 24+ suggest a high electrophilicity due to the four imidazoliumyl substituents. We therefore hypothesized that nucleophilic cleavage with tertiary phosphanes R3P might provide suitable access to phosphonio–phosphanides 1+ as potential [LC–P]+ transfer reagents.

Scheme 1. Synthesis of 2[OTf]4 and Its Nucleophilic Fragmentation with Tertiary Phosphanes R3P (R = Me, Et, Cy, and Ph).

Scheme 1

Reagents and conditions: (i) +4 Ph3P, CD3CN, rt, 16 h; (ii) +4 R3P, CH3CN, rt, 4–16 h, 88–93%.

We now report a straightforward route to simple [LC–P]+ transfer agents (1a-d+) from readily available starting reagents. Their utility is demonstrated by cationic phosphinidene transfer to substrates, including thioketones, thioamides, thiourea, thioesters, and phosphaalkenes R–Pδ+=Cδ−R2. Unprecedented phosphonio–phosphanides, 1b-d+, have been characterized crystallographically as triflate salts, along with a series of hitherto unknown cationic phosphaalkenes, phosphanides, diphosphiranes, and metal complexes, including very rare η2–P=C–Pd0 and Pt0 complexes. In addition, we demonstrate the ability to perform substitution reactions of the transferred LC-substituent in selected substrates using widely applied nucleophilic aryl and alkyl Grignard reagents RMgBr (R = Mes, Me), as well as amide KNPh2. This results in the formation of differently P-functionalized organophosphorus compounds.

Results and Discussion

Preparation of Phosphonio–Phosphanides

Upon adapting our published synthesis of 2[OTf]4 to a larger scale (ca. 50 g, see Supporting Information S2.1), we opted to investigate the reaction of 2[OTf]4 with Ph3P (Scheme 1). Thus, isolated 2[OTf]4 was treated with Ph3P (4 equiv.) in CD3CN. Subsequent analysis of an aliquot removed from the reaction mixture revealed a new AX spin system [δ(31PA) = −168.6 ppm, δ(31PX) = 31.3 ppm, 1J(PP) = −519 Hz] in its 31P NMR spectrum assigned to phosphonio-phosphanide 1a+ (Figure 2). In addition, the spectrum showed signals assigned to the starting materials suggestive of equilibrium. In comparison with phosphanylidenephosphorane DmpP–PPh3 (Dmp = 2,6-Mes2C6H3, Table 1),17 the phosphanide (PA) moiety in 1a+ is further upfield, and the magnitude of 1J(PP) is significantly lower. In related triphosphenium cations {e.g., [(Ph3P)2P]+}18 the high field chemical shift and smaller coupling constant have been attributed to the dominance of the bis(ylidic) canonical structure.19 Further investigation of the reaction mixture by means of 31P–31P EXSY NMR experiments confirmed the underlying thermodynamic equilibrium (Figure 2). Notably, nucleophilic fragmentation of pentaphospholane (PhP)54a,20 and the more electrophilic tetraphosphetane [(CF3)P)4]21 has been described previously, although stronger nucleophiles, that is, NHCs or Me3P, respectively, are required.

Figure 2.

Figure 2

31P NMR spectrum of an aliquot of the reaction mixture of 2[OTf]4 with four equivalents of Ph3P in CD3CN after 16 h (top, CD3CN, 300 K) and zoom in of a 31P–31P-EXSY NMR spectrum (bottom, CD3CN, 300 K) displaying spin polarization exchange between the phosphonio moiety in 1a+ and Ph3P.

Table 1. Comparison of 31P NMR Chemical Shifts and Coupling Constants in 1a-e+ and Selected Related Compounds9,17,23.

compound PA (in ppm) PX (in ppm) 1J(PP) (in Hz)
1a[OTf] (R = Ph) –168.6 31.3 –519.0
1b[OTf] (R = Me) –167.0 12.0 –472.0
1c[OTf] (R = Et) –202.0 36.0 –492.0
1d[OTf] (R = Cy) –208.8 38.1 –545.0
1e[OTf] (R = Ph2(CH2PPh2)) –164.3 38.1 –519.0
DmpP–PPh317 –138.8 25.2 –639.0
DmpP–PMe39 –114.7 –2.8 –582.0
[(Ph3P)2P][AlCl4]23 –174.0 30.0 –502.0

In an effort to prepare isolable phosphonio–phosphanides, the reaction of 2[OTf]4 with more nucleophilic trialkyl-substituted tertiary phosphanes (R3P: R = Me, Et, and Cy) was conducted in CH3CN solution. The complete formation of the corresponding phosphonio–phosphanides 1b-d+ was observed after 4–16 h, and they could be isolated by precipitation with Et2O in excellent yields as their triflate salts (88–92%, Scheme 1). Their respective 31P NMR spectra show the expected characteristic AX spin systems (see Table 1), in accordance with reported values for the related phosphanylidenephosphoranes ArP–PMe39,10 and with the expected group contribution effects.18,19,22 Alternatively, 1b-d[OTf] can be synthesized directly from the reduction of 3[OTf] using an excess of R3P (Scheme S2, Figure S6); however, isolation is best achieved using the procedure described above. Vapor diffusion of Et2O into saturated CH3CN solutions of 1b-d[OTf] at −30 °C afforded colorless crystals suitable for single crystal X-ray analysis. The molecular structures of 1b[OTf] and 1d[OTf] are shown in Figure 3 and that of 1c[OTf] is shown in the Figure S14.

Figure 3.

Figure 3

Molecular structures of phosphonio–phosphanides 1b,d+ in 1b,d[OTf]; hydrogen atoms and anions are omitted for clarity, and thermal ellipsoids are displayed at 50% probability; selected bond lengths (Å) and angles(°): for 1b+: P1–P2 2.1162(4), P1–C1 1.8306(13), C1–P1–P2 97.98(4); 1c+ (Supporting Information, Figure S14): P1–P2 2.1195(4), P1–C1 1.8299(11), C1–P1–P2 99.29(4); 1d+: P1–P2 2.1446(4), P1–C1 1.8280(11), C1–P1–P2 105.93(4).

The observed P–P bond lengths [for 1b+: P1–P2 2.1162(4) Å, 1c+: P1–P2 2.1195(4) Å, and 1d+: P1–P2 2.1446(4) Å] match values for related triphosphenium cations19,23 and range between a typical P–P single24 and P=P double bond.25 This shortening has previously been attributed to result from ylidic-type negative hyperconjugation between the lone pairs at the phosphanide moiety and the σ*(P–R) orbitals.19 Phosphonio–phosphanides 1b-d[OTf] can be stored indefinitely under an inert atmosphere at ambient temperature, whereas neutral derivatives of phosphanylidenephosphoranes ArP–PMe3 have a tendency to decompose with respect to the formation of (ArP)n (n = 2,3) under concomitant release of PMe3.9,26 In an effort to rationalize this apparent high stability, density functional theory (DFT) calculations were performed on the 1a+, 1b+, and 1c+ cations using CH3CN as solvent (details are provided in the Supporting Information). As the energy of the HOMO–LUMO gap in 1a+ (Egap = 2.472 eV; 1b+: R = Me: Egap = 2.791 eV; 1c+: R = Et: Egap = 2.760 eV) is still slightly higher than in DmpP–PMe3 (Egap = 2.443 eV), even larger HOMO–LUMO gaps can be achieved through the introduction of alkyl substituents at the phosphonio moiety (Table S9, Figure S124).

Notably, reaction of 2[OTf]4 with ditopic phosphane 1,1-bis(diphenylphosphino)methane (dppm) leads to the formation of 1e[OTf] (Table 1) instead of the corresponding bis(phosphonio-phosphanide). Changing the ditopic phosphane to bis(diphenylphosphino)ethane (dppe) gives rise to a mixture of previously reported cyclic triphosphenium cation [(Ph2PC2H4PPh2)P]+ and di(imidazoliumyl)phosphanide [(LC)2P]+ as evidenced by means of 31P NMR spectroscopy (Scheme S3 and Figure S8).18,23,27

Phosphonio–Phosphanides as Cationic Phosphinidene Transfer Agents

We continued to investigated the ability of compounds 1a-d[OTf] to transfer [LC–P]+ in phospha-Wittig-type reactions. An initial effort to treat 1b[OTf] with 4-methoxybenzaldehyde resulted in encouraging 31P NMR spectra (see Supporting Information S2.14) of the reaction mixture, showing a small downfield signal at 178.3 ppm along with resonances assigned to free Me3P (δ = −61.5 ppm) and Me3PO (δ = 36.2 ppm). We speculated that the downfield signal observed was consistent with that anticipated for an unprecedented cationic phosphaalkene {i.e., [LCP=CH(C6H4OMe)]+}. However, the conversion to phosphaalkene was very low (<5%), thus we turned our attention to more reactive thiocarbonyls. The latter can typically be accessed directly by thionation of the respective ketone, for example, via conversion with H2S, P4S10, or Lawesson’s reagent.28

For our following studies, compound 1a+ was selected to investigate its [LC–P]+ transfer capability, as it holds the greatest synthetic value compared with 1b-d[OTf], owing to its ease of handling and the comparatively low cost of its starting material PPh3 compared with the other alkyl-substituted tertiary phosphines. When treated with equimolar amounts of selected thioketones in CH3CN, in situ generated 1a+ completely converts into the respective phosphane sulfide R3PS and cationic phosphaalkenes 5a-e+ within 16 h at room temperature (Scheme 2), as evidenced by 31P NMR spectroscopy. The resonances of 5a-e[OTf] in CD3CN (Figure 4) are significantly upfield shifted relative to neutral phosphaalkenes [e.g., MesP=CPh2: δ(31P) = 233 ppm].29 For heteroleptic 5d[OTf], both configurational diastereomers (E/Z) are observed in a near 1:1 ratio. The formation of (+)-camphor-derived 5e[OTf] requires heating of the reaction mixture to 80 °C for 3 h in a microwave reactor. The title compounds can be isolated as analytically pure solids as their triflate salts in very good to excellent yields by precipitation from the respective reaction mixture by addition of Et2O (63–91%, Figure 4).

Scheme 2. Reactions of 1a,d+ with Thiocarbonyls Yield Phosphaalkenes 5a-f[OTf] (R = Aryl, Alkyl; R′ = Aryl, Alkyl, OMe) or Phosphanides 6a-d[OTf] (R = H, Aryl, NR2; R′ = NR2).

Scheme 2

Reagents and conditions: (i) for 5a-d[OTf] and 6a,b[OTf]: −R3PS, CH3CN, rt, 16 h, 63–91%; for 5e[OTf] and 6c,d[OTf]: −R3PS, CH3CN, 80 °C, 3 h, 76–77%; 5a,d,e[OTf] and 6a-d[OTf] were prepared using in situ generated 1a+, 5b-c[OTf] were prepared using isolated 1d[OTf].

Figure 4.

Figure 4

Synthesized phosphaalkenes 5a-f[OTf] and phosphanides 6a-d[OTf] (left); molecular structure of phosphaalkenes 5a+ in 5a[OTf] and phosphanide 6a+ in 6a[OTf] (right); hydrogen atoms and anions are omitted for clarity, and thermal ellipsoids are displayed at 50% probability; selected bond lengths (Å) and angles (°): for 5a+: P1–C1 1.707(3), P1–C2 1.834(33), C1–P1–C2 104.55(16); 6a+: P1–C1 1.7417(13), P1–C2 1.8382(12), C1–P1–C2 93.77(6).

A second set of cationic phosphinidene transfer reactions were explored by treating in situ generated 1a+ with thioamides [R(NMe2)C=S (R = H, Ph, NMe2)] and LC=S. In each case, analysis of the reaction mixtures by means of 31P NMR spectroscopy showed only a signal assigned to Ph3PS (δ = 42.4 ppm) along with a new singlet resonance {δ(31P) = −8.8 ppm (br), R = H; 7.9 ppm, R = Ph; −60.2 ppm (br), R = NMe2; −124.6 ppm, cf. known [(LC)2P]+30}. Remarkably, each were shifted considerably upfield compared to those of 5a-e+. A similar trend to higher field shifts is observed in the 31P NMR spectra of inversely polarized phosphaalkenes bearing C-amino substituents when compared to conventional phosphaalkenes.4e,31 Given this apparent higher shielding/increased electron density at the P atoms, the products were formulated with the cationic phosphanide canonical form (i.e., 6a-d+ in Figure 4) rather than cationic phosphaalkene (cf. 5+).

A supporting trend for this observation was found in the molecular structures of 5a-f[OTf] and 6a-c[OTf] (5a[OTf] and 6a[OTf] in Figure 4; 5b-f[OTf] and 6b,c[OTf] in Supporting Information). The P=C bond length in cationic phosphaalkene 5a[OTf] [P=C 1.707(3) Å, Figure 4] is only slightly elongated compared to the related MesP=CPh2 [P=C 1.692(3) Å].32 Likewise, the P=C bonds of 5b,d+ [1.700(4), 1.702(4) Å, respectively] are in the range typical of phosphaalkenes. In contrast, the camphor-substituted 5e+ has a shorter P=C bond length [1.675(4) Å], presumably due to reduced delocalization of the P=C bond. The introduction of donating amino groups leads to significant elongation of the P–C bond [5c+: 1.7247(19) Å, 6a+: 1.7417(13) Å, 6c+: 1.7838(11) Å], consistent with increased contribution of the phosphanide canonical form, and comparable to values for reported inversely polarized phosphaalkenes.4e,8a,20,33 Generally, the LC–P=C bond angles are more acute for those formulated as phosphanides [C1–P1–C2 (°) = 93.77(6), 6a+; 99.30(15), 6b+; 100.81(5), 6c+] when compared to those formulated a phosphaalkenes [C1–P1–C2 (°) = 104.55(16), 5a+; 104.68(19), 5b+; 103.86(9), 5c+; 104.56(16), 5d+; 99.3(3), 5e+]. The reaction with O-methyl benzothioate proceeds likewise but results in multiple products, as evidenced by the 31P NMR spectra of the reaction mixture (Figure S74). Single crystal analysis of some crystalline material obtained by vapor diffusion of Et2O into the reaction mixture, however, confirms the formation of methoxy-substituted phosphaalkene 5f[OTf] [δ(31P) = 53.0 ppm, see Supporting Information S2.25].

In general, Wittig-type conversions are thought to proceed via open-chain betaine-type structures or four-membered oxaphosphetanes as intermediates.14,34 One report proposes an oxadiphosphetane intermediate for the related phospha-Wittig-Horner reaction.35 Next to this, detailed investigations on the mechanism of the phospha-Wittig reaction specifically are scarce. Therefore, we employed DFT calculations [RI-BP86-D3/def2-TZVP (acetonitrile)] to gain further insight into the reaction of 1a+ with thiobenzophenone (Figure 5). The reaction profile reveals an overall exergonic transformation (−13.0 kcal/mol) that is initiated by the formation of a supramolecular complex INT-1 (see Figure S126), which is 13.3 kcal/mol more stable than the isolated reactants and possibly results from interaction of the lone pairs at the phosphanide moiety of 1a+ with the π*-orbital of the C=S double bond. This pre-organization is followed by a [LC–P]+ transfer onto the C=S double bond via TS-1 with concomitant release of Ph3P to give the three-membered thiophosphirane36INT-2, which is more stable than TS-1 by 5.8 kcal/mol. In an attempt to identify the formation of INT-2 during the course of the reaction, we monitored the reaction between isolated 1d[OTf] (in favor of 1a+ to rule out possible [LC–P]+ transfer from 2[OTf]4 present in the reaction mixture of in situ generated 1a+) and thiobenzophenone by means of time-resolved 31P NMR spectroscopy. The spectrum of an aliquot of the reaction mixture after 10 min at room temperature reveals the formation of a new singlet resonance at δ(31P) = −89.3 ppm (Figure S1), which is within the margin of error37 for the calculated 31P NMR shift of INT-2 [δ(31P) = −78.4 ppm]. We therefore assign this resonance to thiophosphirane 7a+. Notably, the formation of thiophosphirane 7a+ can also be observed in other transformations (see Figures S2, S4, and S7), as evidenced by its characteristic chemical shift in the 31P NMR spectra.

Figure 5.

Figure 5

Reaction profile for the conversion of 1a+ to 5a+ at the RI-BP86-D3/def2-TZVP (acetonitrile) level of theory; optimized geometries of the transition states (TS) with distances (red) in Å. Y axis shows ΔG in kcal/mol.

The formation of phosphanides 6a-d+, exemplified for the conversion of 1a+ with tetramethylthiourea, can be described by the same mechanism, although higher energy barriers are calculated (Figure S125). However, the corresponding thiophosphiranes 8c+ could not be observed spectroscopically.

Following the initially reported protocols for the phospha-Wittig reaction,9,38 phosphaalkenes 5a-e+ and phosphanides 6a-c+ can also be synthesized using one-pot reactions of dichlorophosphane 3[OTf] with thiocarbonyls in the presence of Ph3P and Zn, yet isolation cannot always be achieved satisfyingly (see Supporting Information S2.15).

Cationic Disphosphiranes from [LC–P]+ Transfer to Phosphaalkenes

We further explored the [LC–P]+ transfer reactivity of 1a+ toward phosphaalkenes,39 which are isolobal to thioketones. Indeed, reacting 2[OTf]4 with a phosphaalkene 9(29) (4 equiv) or 1,2-diphosphetane 10(40) (2 equiv) in the presence of catalytic Ph3P (0.1 equiv) afforded diphosphiranes 11a,b[OTf] in very good or excellent yield (83 and 90%, respectively, Scheme 3). 31P NMR spectroscopic investigations on the isolated compounds showed the expected AB spin systems (11a[OTf]: δ(31PA) = −127.7 ppm, δ(31PB) = −100.8 ppm, 1J(PP) = 146 Hz, 11b[OTf]: δ(31PA) = −139.5 ppm, δ(31PB) = −107.9 ppm, 1J(PP) = 133 Hz).

Scheme 3. Synthesis of Diphosphiranes 11a,b [OTf] (R = Mes, Ph) via [LC–P]+ Transfer from 1a+ onto Phosphaalkenes.

Scheme 3

Reagents and conditions: (i) +0.1 Ph3P, +0.25 2[OTf]4, CH3CN, rt, 2 h, 83%; (ii) +0.1 Ph3P, +0.5 2[OTf]4, CH3CN, 60 °C, 2 h, 90%.

The observed high field-shifted resonances for both phosphorus nuclei are characteristic for phosphorus-containing three-membered ring {e.g., [(LC)3P3][OTf]3, (tBuP)3}.16,41 The molecular structures of 11a-b+ in 11a[OTf] (Figure 6) and 11b[OTf]·o-C6H4F2 (Figure S86) reveal shortened P–P bond lengths [11a+: 2.1905(7) Å, 11b+: 2.1817(4) Å] and acute P1–C1–P2 bond angles [11a+: 70.50(9)°, 11b+: 70.82(15)°] in the range of other diphosphiranes42 and diphosphiranium cations [RP(R(CH2tBu)P)((tBu)HC)]+ (R = tBu, Ad).43

Figure 6.

Figure 6

Molecular structure of 11a+ in 11a[OTf]; hydrogen atoms and the anion are omitted for clarity, and thermal ellipsoids are displayed at 50% probability selected bond lengths (Å) and angles (°): P1–C1 1.908(2), P1–P2 2.1905(7), P2–C1 1.887(2), P2–C2 1.852(2), P1–C1–P2 70.50(9), C1–P1–P2 54.30(7), P1–P2–C2 109.68(7), C1–P2–C2 106.65(9).

Reactivity of Cationic Phosphaalkenes

Motivated by the successful isolation of 5a-e[OTf], we further probed the reactivity of 5a[OTf], as a model compound for cationic phosphaalkenes, due to its structural similarity to the well-studied MesP=CPh2, toward typical conversions of phosphaalkenes (Scheme 4). Treating 5a[OTf] with [Pd(PPh3)4] or [Pt(PPh3)3] gave the metallaphosphiranes 12a[OTf] and 12b[OTf], respectively, under concomitant release of Ph3P as evidenced by 31P NMR spectroscopy. The η2-coordination of the phosphaalkenes was indicated by a strong high field shift [12a+: δ(31PA) = −5.6 ppm (br), 12b+: δ(31PA) = −55.3 ppm] compared to the resonance of 5a[OTf] and the modest coupling constant to the Pt atom in 12b[OTf] [1J(PAPt) = 564 Hz].44 The 195Pt NMR spectrum of 12b[OTf] showed the expected doublet of doublet of doublet resonance at δ(195Pt) = −4822 ppm [1J(PtP) = 3571 Hz, 1J(PtP) = 3207 Hz, and 1J(PtP) = 563 Hz, Figure S102].

Scheme 4. Reactions of Phosphaalkenes 5a[OTf] with Low Oxidation State Transition Metal Complexes [M(Ph3P)n] (M = Pd: n = 4, M = Pt: n = 3) toward Metallaphosphiranes 12a,b[OTf], with [Fe2(CO)9] toward Iron Complex 13[OTf] and with 3,4,5,6-Tetrachloro-1,2-benzoquinone (14) toward 15[OTf].

Scheme 4

Reagents and conditions: (i) +[M(PPh3)n], −(n – 2) Ph3P, C6H5F (12a[OTf]), toluene (12b[OTf]), rt, 1–4 h, 90–97%; (ii) +[Fe2(CO)9], −[Fe(CO)5], THF, rt, 16 h, 86%; (iii) +14, C6H5F, rt, 1 h, 85%.

In general, the chemical shifts for 12a,b+ are comparable to the reported values for some related η2-diphosphene complexes16,45 and the η2-phosphaalkene complex [Pt(Ph3P)22-MesP=CPh2)].46 Notably, the latter phosphaalkene complex was structurally characterized as the η1-complex with the η2-complex only being observed by 31P NMR spectroscopy at −70 °C in solution. Remarkably, the 31P NMR spectra of solutions 12a-b[OTf] in toluene-d8 did not show evidence for the formation of η1-complex upon heating to 100 °C. The latter might be a result of an increased π-acceptor ability in 12a,b+ due to the presence of the imidazoliumyl-substituent, thereby favoring the η2-coordination mode.47 After workup, analytically pure red 12a[OTf] and yellow 12b[OTf] were obtained in excellent yields (90 and 97%, respectively).

Subsequently, single crystals of each were obtained and characterized by X-ray crystallography (Figures 7 and S96). Both complexes show the expected trigonal core, including the P–C bond of the [LCP=CPh2]+ ligand and the metal centers. Therein, the P–M–C bond angles are relatively acute [12a+ (M = Pd): 45.89(12)–46.46(9)°, 12b+ (M = Pt): 47.93(5)°]. In line with the η2-coordination mode, the P=C bond in both metallaphosphiranes is elongated to the extent of a P–C single bond [12a+: 1.782(4)–1.787(5) Å, 12b+: 1.8260(18) Å] presumably due to electron-backdonation from the metal centers into the π*-orbitals of the P=C bond.48 Similar structural features have been observed for other transition-metal complexes involving η2-phosphaalkene or η4-phosphabutadiene ligands.49

Figure 7.

Figure 7

Molecular structures of metallaphosphirane 12b+ in 12b[OTf], iron complex 13+ in 13[OTf] (left) and of 15+ in 15[OTf]·0.5C6H5n-pentane (right); hydrogen atoms and anions are omitted for clarity and thermal ellipsoids are displayed at 50% probability; selected bond lengths (Å) and angles(°): 12b+ (Pt): P1–C1 1.8260(18), Pt–P1 2.3127(5), Pt–C1 2.1718(16), Pt···P1–C1 2.0418(10), P1–Pt–C1 47.93(5); 13+: P–C1 1.6846(18), P–Fe 2.1697(5), C–P–C 106.51(8), ⌀C≡Oax.: 1.1405, ⌀C≡Oeq.: 1.1435; 15+: P–C1 1.897(3), P–O1 1.657(2), C1–O2 1.458(3); C1–P–C2 104.01(12)°.

We further reacted phosphaalkene 5a[OTf] (1 equiv) with [Fe2(CO)9] (1 equiv) in THF, which led to a red precipitate after 16 h at room temperature. Analysis of a CD3CN solution of the red product by means of 31P NMR spectroscopy showed only one singlet resonance at δ(31P) = 154.5 ppm. The slight downfield shift compared to that for 5a[OTf] [δ(31P) = 152.8 ppm] suggests η1-phosphaalkene complex 13[OTf] (Scheme 4). Crystallographic analysis of single crystals confirmed that the phosphaalkene ligand binds in a η1-fashion in the equatorial position of the Fe0 center (Figure 7). The P–Fe–COeq [119.62(6)° and 122.44(7)°] and P–Fe–COax angles [88.29(6)° and 89.69(6)°] are typical of [FeL(CO)4] complexes, where L is a π-acceptor.50 As a result, the bonding parameters of the phosphaalkenes ligand are only slightly affected compared to uncoordinated 5a+.51

For instance, the P=C bond in 13[OTf] is marginally shortened [1.6846(18) Å versus 5a+: 1.703(3) Å] and the C–P–C bond angle is widened [106.51(8)° versus 5a+: 104.55(16)°], which may indicate a strengthening of the double bond character. Compound 13[OTf] could be isolated in 86% yield. The IR stretching frequencies of the CO ligands are found at 2073, 2013, 1993, and 1965 cm–1, which renders the ligand properties of 5a+ similar to those observed in phosphites, according to Tolman analysis.52

To evaluate the potential of the P=C double bond in 5a[OTf] to be involved in cycloaddition reactions, we performed its conversion with 3,4,5,6-tetrachloro-1,2-benzoquinone (14). Upon dropwise addition of the red solution of 14 in C6H5F to yellow 5a[OTf], a colorless reaction mixture is obtained immediately. X-ray analysis of single crystals obtained by vapor diffusion of n-pentane into the reaction mixture confirms 15[OTf] as the product resulting from a [4 + 2] cycloaddition reaction (Figure 7). While the C1–P–C2 bond angle is barely affected [104.01(12)°], the P–C bond length is significantly elongated [1.897(3) Å] as a result of the conversion, and the P atom takes on a trigonal pyramidal geometry. The loss of the double bond character is consistent with the high field shift in the 31P NMR spectrum of 15+ [δ(31P) = 110.3 ppm], which has been observed in the reaction of related phosphaalkenes with the same and other ortho-quinones.53

Replacement Reactions of Imidazoliumyl Substituent LC

We further assumed that, due to the cationic charge in synthesized 5a-f+ and 6a-c+, their interaction with nucleophiles may lead to substitution at the P atom. Based on this, we hypothesized that the reaction of 5a[OTf] with MesMgBr could provide a practical route to obtaining MesP=CPh2 (16), a significant monomer for the creation of P-containing polymers54 that is typically synthesized via the phospha-Peterson reaction.55 As confirmed by 31P NMR spectroscopy, treating the cationic phosphaalkene 5a[OTf] with 1 equiv of MesMgBr in THF at −78 °C resulted in its complete conversion to MesP=CPh2 (16) within 30 min (Figure 8a).

Figure 8.

Figure 8

(a) Substitution reactions of the cationic imidazoliumyl substituent in 5a[OTf] using Grignard reagents; reagents and conditions: (i) +MesMgBr (1 M in THF), −0.5 [Mg(OTf)2(THF)4], −0.5 [MgBr2(LC)2], THF, −78 °C, 30 min, 93%; (ii) +MeMgBr (1 M in nBu2O), THF, −78 °C, 30 min, quenched with 2 M HCl in Et2O, 49%; (b) reactions of 5a[OTf] and 6c[OTf] with KNPh2; reagents and conditions: (iii) +KNPh2, THF, rt, 30 min, quenched with Me3SiOTf, 90% (for 18), 80% (for 19); (c) molecular structures of 1,3-diphosphetane 17 and aminophosphaalkene 18; hydrogen atoms are omitted for clarity, and thermal ellipsoids are displayed at 50% probability; selected bond lengths (Å) and angles(°): for 17: P1–C1 1.8373(12), P1–C2 1.9085(11), P1–C4 1.9096(12), C1–P1–C2 103.44(5), C2–P1–C4 88.30(5), P1–C2–P2 91.70(5), C1–P1–P2–C3 180.00(9); 18: P–C1 1.726(2), P–N 1.7132(17), C1–P–N 103.28(9).

The product was conveniently obtained in 93% yield after extraction with n-hexane and subsequent recrystallization. One of the side products of the reaction was identified crystallographically as [MgBr2(LC)2].56

The successful conversion of 5a[OTf] to MesP=CPh2 prompted us to expand our investigation to screening reactions with other nucleophiles that would result in the substitution of the imidazoliumyl substituent. Without adequate steric protection at the P atom, phosphaalkenes have a tendency to dimerize into either head-to-head (1,2-diphosphetanes) or head-to-tail (1,3-diphosphetane) dimers.40,55,57 Consequently, the 31P NMR spectrum of an aliquot removed from the reaction mixture of 5a[OTf] with MeMgBr at −78 °C after quenching with HCl and warming to room temperature shows two new major resonances at δ(31P) = 35.6 ppm (70% integral ratio) and δ(31P) = −34.8 ppm (18% integral ratio), which we assigned to the head-to-tail and head-to-head dimers of MeP=CPh2, respectively. 1,3-Diphosphetane (17) was isolated from the reaction mixture in 49% yield (Figure 8a) and structurally characterized by single crystal X-ray analysis (Figure 8c). The nearly square planar P2C2 core of 17 features P–C bond distances of approximately 1.909 Å and bond angles C2–P1–C4 and P1–C2–P2 of 88.30(5)° and 91.70(5)°, respectively. The two methyl substituents are arranged in a trans position, forming a dihedral angle C1–P1–P2–C3 of 180.00(9)°.57

In some cases, thermodynamic equilibria between phosphaalkenes and their dimers have been reported.40,57a,57c,57e Remarkably, no meaningful change within the 1H and 31P NMR spectra of isolated 17 in toluene-d8 was observed upon stepwise heating to 80 °C.

In a separate experiment, we investigated the reaction of two model compounds, 5a[OTf] and 6c[OTf], which are representative of cationic phosphaalkenes and phosphanides, respectively, toward KNPh2. The reaction was carried out by adding 1 equiv of KNPh2 to solutions of either 5a[OTf] or 6c[OTf] in THF at ambient temperature. After 30 min, the formation of free Me/iPrNHC was observed, as confirmed by 1H and 13C NMR spectroscopic analysis of a sample removed from the reaction mixture. The 31P NMR spectra displayed low-field shifted resonances in both cases, indicating the exchange of the LC-substituent with an amino group, resulting in the formation of two new examples of aminophosphaalkenes [18: δ(31P) = 231.0 ppm; and 19: δ(31P) = 106.1 ppm; Figure 8b]. Since both resulting compounds had similar solubility to free Me/iPrNHC, 1 equiv of Me3SiOTf was added to the respective reaction mixture, resulting in the formation of [LCSiMe3][OTf], which is a precursor in the synthesis of 3[OTf].58 Subsequently, 18 and 19 were isolated via extraction with n-hexane, with yields of 90 and 80%, respectively. Single crystals of both compounds were obtained through recrystallization from n-hexane or n-pentane and subjected to X-ray analysis.

In the molecular structures of 18 [P–C1 1.7132(17) Å, Figure 8 bottom], the P–C1 bond length is slightly elongated compared to a typical P=C double bond. This elongation is attributed to the donating effect of the P-amino substituent. The P–C1 bond length in C-amino substituted 19 [P–C1 1.754(1) Å, Figure S123] is even further elongated. The P–N bonds in both compounds [18: P–N 1.726(2) Å; 19: P–N 1.7501(12) Å] are comparable with other structurally related compounds.59

Conclusions

In summary, we showed the synthesis of cationic imidazoliumyl(phosphonio)-phosphanides 1a-d+ via the nucleophilic fragmentation of tetracationic tetraphosphetane 2[OTf]4 with tertiary phosphanes R3P (R = Ph, Me, Et, Cy). We tested their ability to undergo the hitherto unknown transfer of a cationic phosphinidene, that is, [LC–P]+. Employing in situ generated 1a+ or isolated 1d[OTf] in phospha-Wittig-type reactions with thiocarbonyls allowed the isolation and characterization of a series of novel cationic phosphaalkenes 5a-f+ as well as phosphanides 6a-d+ bearing a wide variety of substituents as their triflate salts. As evidenced spectroscopically and by DFT calculations [RI-BP86-D3/def2-TZVP (acetonitrile) level of theory], the mechanism of the formation of phosphaalkenes proceeds via the intermediary three-membered thiophosphiranes as a result of a [LC–P]+ transfer from 1a+ onto the C=S double bond. Although calculations show a similar pathway for the formation of phosphanides, energy barriers are found to be significantly higher. Furthermore, when in situ generated 1a+ is reacted with phosphaalkenes that are isolobal to thioketones, [LC–P]+ transfer is also observed, leading to the isolation of heteroleptic diphosphiranes 11a,b[OTf].

In order to evaluate the reactivity of the formed cationic phosphaalkenes, we subjected 5a[OTf] to reactions with low oxidation state transition-metal complexes [Pd(PPh3)4], [Pt(PPh3)3], and [Fe2(CO)9]. While the conversion with the latter gave iron complex 13[OTf], in which the phosphaalkenes are in an equatorial position and have a η1-coordination mode, metallaphosphiranes 12a,b[OTf] are formed in the reaction with the former two complexes, including 5a+ in a η2-coordination mode. Lastly, we showed the potential to use the P=C double bond in 5a[OTf] for cycloaddition reactions by its conversion with ortho-quinone 14 giving 15[OTf].

We furthermore exemplified the possibility of exchanging the transferred LC-substituent in 5a[OTf] by reacting it with MesMgBr. This reaction allowed for the convenient and high-yield synthesis of MesP=CPh2, a compound that is typically obtained through reactions involving malodorous primary phosphines and silylphosphines. Moreover, our work enables access to unprecedented 1,3-diphosphetane 17 as well as aminophosphaalkenes 18 and 19 through the conversions of 5a[OTf] and 6c[OTf] with MeMgBr or KNPh2. These conversions demonstrate the versatility of [LC–P]+ as a P1 building block.

Acknowledgments

This work was supported by the German Science Foundation (DFG; WE 4621/6-1; WE4621/3-1; WE4621/3-2) and the Natural Sciences and Engineering Research Council (NSERC) of Canada (RGPIN-2020-06506 to D.P.G.). A.F. thanks the MICIU/AEI of Spain (project PID2020-115637GB-I00 FEDER funds) for financial support. P.R. thanks the Fonds der Chemischen Industrie (Kekulé scholarship). P.R. and Z.H. thank the DAAD for financial support within the ERA+ program. Philipp Lange is acknowledged for performing elemental analyses. TUD is also thanked for financial support.

Supporting Information Available

The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/jacs.3c02256.

  • Experimental and characterization details for all new compounds, including spectroscopic data and NMR spectra, molecular structures, crystallographic data, and computational details (PDF)

The authors declare no competing financial interest.

Supplementary Material

ja3c02256_si_001.pdf (13.4MB, pdf)

References

  1. Mathey F. The Development of a Carbene-like Chemistry with Terminal Phosphinidene Complexes. Angew. Chem., Int. Ed. Engl. 1987, 26, 275–286. 10.1002/anie.198702753. [DOI] [Google Scholar]; Angew. Chem. 1987, 99, 285 - 296.10.1002/ange.19870990404
  2. a Hansmann M. M.; Bertrand G. Transition-Metal-like Behavior of Main Group Elements: Ligand Exchange at a Phosphinidene. J. Am. Chem. Soc. 2016, 138, 15885–15888. 10.1021/jacs.6b11496. [DOI] [PubMed] [Google Scholar]; b Liu L.; Ruiz D. A.; Munz D.; Bertrand G. A Singlet Phosphinidene Stable at Room Temperature. Chem 2016, 1, 147–153. 10.1016/j.chempr.2016.04.001. [DOI] [Google Scholar]
  3. a Marinetti A.; Mathey F. The carbene-like behavior of terminal phosphinidene complexes toward olefins. A new access to the phosphirane ring. Organometallics 1984, 3, 456–461. 10.1021/om00081a021. [DOI] [Google Scholar]; b Marinetti A.; Mathey F.; Fischer J.; Mitschler A. Generation and trapping of terminal phosphinidene complexes. Synthesis and x-ray crystal structure of stable phosphirene complexes. J. Am. Chem. Soc. 1982, 104, 4484–4485. 10.1021/ja00380a029. [DOI] [Google Scholar]; c Aktaş H.; Slootweg J. C.; Lammertsma K. Nucleophilic phosphinidene complexes: access and applicability. Angew. Chem., Int. Ed. Engl. 2010, 49, 2102–2113. 10.1002/anie.200905689. [DOI] [PubMed] [Google Scholar]; Angew. Chem. 2010, 122, 2148 - 2159.10.1002/ange.200905689; d Breen T. L.; Stephan D. W. Phosphinidene Transfer Reactions of the Terminal Phosphinidene Complex Cp2Zr(:PC6H2-2,4,6-t-Bu3)(PMe3). J. Am. Chem. Soc. 1995, 117, 11914–11921. 10.1021/ja00153a013. [DOI] [Google Scholar]; e Cowley A. H. Terminal Phosphinidene and Heavier Congeneric Complexes. The Quest Is Over. Acc. Chem. Res. 1997, 30, 445–451. 10.1021/ar970055e. [DOI] [Google Scholar]; f Cowley A. H.; Barron A. R. The quest for terminal phosphinidene complexes. Acc. Chem. Res. 1988, 21, 81–87. 10.1021/ar00146a006. [DOI] [Google Scholar]; g Cummins C. C.; Schrock R. R.; Davis W. M. Phosphinidenetantalum(V) Complexes of the Type[(N3N)Ta=PR] as Phospha-Wittig Reagents. Angew. Chem., Int. Ed. Engl. 1993, 32, 756–759. 10.1002/anie.199307561. [DOI] [Google Scholar]; Angew. Chem. 1993, 105, 758 - 761.10.1002/ange.19931050522; h Mathey F.; Huy N. T.; Marinetti A. Electrophilic Terminal-Phosphinidene Complexes: Versatile Phosphorus Analogues of Singlet Carbenes. Helv. Chim. Acta 2001, 84, 2938–2957. . [DOI] [Google Scholar]; i Marinetti A.; Mathey F.; Fischer J.; Mitschler A. Stabilization of 7-phosphanorbornadienes by complexation; X-ray crystal structure of 2,3-bis(methoxycarbonyl)-5,6-dimethyl-7-phenyl-7-phosphanorbornadiene(pentacarbonyl)-chromium. J. Chem. Soc., Chem. Commun. 1982, 12, 667. 10.1039/c39820000667. [DOI] [Google Scholar]
  4. a Arduengo A. J. III; Carmalt C. J.; Clyburne J. A. C.; Cowley A. H.; Pyati R. Nature of the bonding in a carbene–phosphinidene: a main group analogue of a Fischer carbene complex: Isolation and characterisation of a bis(borane) adduct. Chem. Commun. 1997, 981–982. 10.1039/a700296c. [DOI] [Google Scholar]; b Fritz G.; Vaahs T.; Fleischer H.; Matern E. Nachweis des Phosphino-phosphinidens (Me3C)2P–P bei der Umsetzung von [(Me3C)2P]2PLi mit 1,2-Dibromethan durch Abfangreaktionen. Z. Anorg. Allg. Chem. 1989, 570, 54–66. 10.1002/zaac.19895700104. [DOI] [Google Scholar]; c Kovacs I.; Fritz G. Die Bildung von [(tBu)2P]2P–P[P(tBu)2]2 aus LiP[P(tBu)2]2 und 1,2-Dibromethan. Die Thermolyse von tBu2P–P–P(Br)tBu2. Z. Anorg. Allg. Chem. 1994, 620, 1–3. 10.1002/zaac.19946200102. [DOI] [Google Scholar]; d Matern E.; Fritz G.; Pikies J. Der thermische Zerfall des tBu2P–P–P(Me)tBu2. Z. Anorg. Allg. Chem. 1997, 623, 1769–1773. 10.1002/zaac.19976231118. [DOI] [Google Scholar]; e Weber L. Phosphaalkenes with Inverse Electron Density. Eur. J. Inorg. Chem. 2000, 2000, 2425–2441. . [DOI] [Google Scholar]; f Weber L.; Lassahn U.; Stammler H.-G.; Neumann B. Inversely Polarized Phosphaalkenes as Phosphinidene- and Carbene-Transfer Reagents. Eur. J. Inorg. Chem. 2005, 2005, 4590–4597. 10.1002/ejic.200500454. [DOI] [Google Scholar]
  5. a Transue W. J.; Velian A.; Nava M.; García-Iriepa C.; Temprado M.; Cummins C. C. Mechanism and Scope of Phosphinidene Transfer from Dibenzo-7-phosphanorbornadiene Compounds. J. Am. Chem. Soc. 2017, 139, 10822–10831. 10.1021/jacs.7b05464. [DOI] [PubMed] [Google Scholar]; b Velian A.; Cummins C. C. Facile synthesis of dibenzo-7λ3-phosphanorbornadiene derivatives using magnesium anthracene. J. Am. Chem. Soc. 2012, 134, 13978–13981. 10.1021/ja306902j. [DOI] [PubMed] [Google Scholar]
  6. Hansen K.; Szilvási T.; Blom B.; Inoue S.; Epping J.; Driess M. A fragile zwitterionic phosphasilene as a transfer agent of the elusive parent phosphinidene (:PH). J. Am. Chem. Soc. 2013, 135, 11795–11798. 10.1021/ja4072699. [DOI] [PubMed] [Google Scholar]
  7. a Protasiewicz J. D. Coordination-Like Chemistry of Phosphinidenes by Phosphanes. Eur. J. Inorg. Chem. 2012, 2012, 4539–4549. 10.1002/ejic.201200273. [DOI] [Google Scholar]; b Shah S.; Simpson M. C.; Smith R. C.; Protasiewicz J. D. Three different fates for phosphinidenes generated by photocleavage of phospha-Wittig reagents ArP=PMe3. J. Am. Chem. Soc. 2001, 123, 6925–6926. 10.1021/ja015767l. [DOI] [PubMed] [Google Scholar]; c Shah S.; Protasiewicz J. ‘Phospha-variations’ on the themes of Staudinger and Wittig: phosphorus analogs of Wittig reagents. Coord. Chem. Rev. 2000, 210, 181–201. 10.1016/s0010-8545(00)00311-8. [DOI] [Google Scholar]
  8. a Krachko T.; Bispinghoff M.; Tondreau A. M.; Stein D.; Baker M.; Ehlers A. W.; Slootweg J. C.; Grützmacher H. Facile Phenylphosphinidene Transfer Reactions from Carbene-Phosphinidene Zinc Complexes. Angew. Chem., Int. Ed. Engl. 2017, 56, 7948–7951. 10.1002/anie.201703672. [DOI] [PubMed] [Google Scholar]; Angew. Chem. 2017, 129, 8056 - 8059.; b Shah S.; Yap G. P.; Protasiewicz J. D. Crystal structure of the phosphanylidene-σ4-phosphorane DmpP=PMe3 (Dmp=2,6-Mes2C6H3) and reactions with electrophiles. J. Org. Chem. 2000, 608, 12–20. 10.1016/s0022-328x(00)00284-9. [DOI] [Google Scholar]
  9. Shah S.; Protasiewicz J. D. ‘Phospha-Wittig’ reactions using isolable phosphoranylidenephosphines ArP=PR3 (Ar = 2,6-Mes2C6H3 or 2,4,6-tBu3C6H2). Chem. Commun. 1998, 1585–1586. 10.1039/a802722f. [DOI] [Google Scholar]
  10. Gupta P.; Siewert J.-E.; Wellnitz T.; Fischer M.; Baumann W.; Beweries T.; Hering-Junghans C. Reactivity of Phospha-Wittig reagents towards NHCs and NHOs. Dalton Trans. 2021, 50, 1838–1844. 10.1039/d1dt00071c. [DOI] [PubMed] [Google Scholar]
  11. Fischer M.; Hering-Junghans C. On 1,3-phosphaazaallenes and their diverse reactivity. Chem. Sci. 2021, 12, 10279–10289. 10.1039/d1sc02947a. [DOI] [PMC free article] [PubMed] [Google Scholar]
  12. Dankert F.; Siewert J.-E.; Gupta P.; Weigend F.; Hering-Junghans C. Metal-Free N-H Bond Activation by Phospha-Wittig Reagents. Angew. Chem., Int. Ed. Engl. 2022, 61, e202207064 10.1002/anie.202207064. [DOI] [PMC free article] [PubMed] [Google Scholar]; Metallfreie N−H-Bindungsaktivierung mit Phospha-Wittig Reagenzien. Angew. Chem. 2022, 134, e20220706410.1002/ange.202207064
  13. Fischer M.; Nees S.; Kupfer T.; Goettel J. T.; Braunschweig H.; Hering-Junghans C. Isolable Phospha- and Arsaalumenes. J. Am. Chem. Soc. 2021, 143, 4106–4111. 10.1021/jacs.1c00204. [DOI] [PubMed] [Google Scholar]
  14. Borys A. M.; Rice E. F.; Nichol G. S.; Cowley M. J. The Phospha-Bora-Wittig Reaction. J. Am. Chem. Soc. 2021, 143, 14065–14070. 10.1021/jacs.1c06228. [DOI] [PMC free article] [PubMed] [Google Scholar]
  15. Donath M.; Schwedtmann K.; Schneider T.; Hennersdorf F.; Bauzá A.; Frontera A.; Weigand J. J. Direct conversion of white phosphorus to versatile phosphorus transfer reagents via oxidative onioation. Nat. Chem. 2022, 14, 384–391. 10.1038/s41557-022-00913-4. [DOI] [PubMed] [Google Scholar]
  16. Schwedtmann K.; Haberstroh J.; Roediger S.; Bauzá A.; Frontera A.; Hennersdorf F.; Weigand J. J. Formation of an imidazoliumyl-substituted (LC)4P44+ tetracation and transition metal mediated fragmentation and insertion reaction (LC = NHC). Chem. Sci. 2019, 10, 6868–6875. 10.1039/c9sc01701a. [DOI] [PMC free article] [PubMed] [Google Scholar]
  17. Smith R. C.; Shah S.; Protasiewicz J. D. A role for free phosphinidenes in the reaction of magnesium and sterically encumbered ArPCl2 in solution at room temperature. J. Org. Chem. 2002, 646, 255–261. 10.1016/s0022-328x(01)01384-5. [DOI] [Google Scholar]
  18. Schmidpeter A.; Lochschmidt S.; Sheldrick W. S. Symmetrical and Unsymmetrical Triphosphenium Ions. Angew. Chem., Int. Ed. Engl. 1985, 24, 226–227. 10.1002/anie.198502261. [DOI] [Google Scholar]; Symmetrsiche und Unsymmetrische Triphosphenium-Ionen. Angew. Chem. 1985, 97, 214 - 215.10.1002/ange.19850970313
  19. Ellis B. D.; Macdonald C. L. B. Phosphorus(I) iodide: a versatile metathesis reagent for the synthesis of low oxidation state phosphorus compounds. Inorg. Chem. 2006, 45, 6864–6874. 10.1021/ic060186o. [DOI] [PubMed] [Google Scholar]
  20. Arduengo A. J. III; Dias H. V. R.; Calabrese J. C. A Carbene·Phosphinidene Adduct: “Phosphaalkene”. Chem. Lett. 1997, 26, 143–144. 10.1246/cl.1997.143. [DOI] [Google Scholar]
  21. Burg A. B.; Mahler W. Bicoordinate Phosphorus: Base-PCF3. J. Am. Chem. Soc. 1961, 83, 2388–2389. 10.1021/ja01471a037. [DOI] [Google Scholar]
  22. a Baker R. J.; Jones C.; Mills D. P.; Murphy D. M.; Hey-Hawkins E.; Wolf R. The reactivity of gallium-(I), -(II) and -(III) heterocycles towards Group 15 substrates: attempts to prepare gallium-terminal pnictinidene complexes. Dalton Trans. 2006, 64–72. 10.1039/b511451a. [DOI] [PubMed] [Google Scholar]; b Grim S. O.; McFarlane W. Phosphorus-31 Chemical Shifts in Secondary and Tertiary Phosphines. Nature 1965, 208, 995–996. 10.1038/208995a0. [DOI] [Google Scholar]; c Grim S. O.; McFarlane W.; Davidoff E. F. Group contributions to phosphorus-31 chemical shifts of tertiary phosphines. J. Org. Chem. 1967, 32, 781–784. 10.1021/jo01278a059. [DOI] [Google Scholar]; d Grim S. O.; McFarlane W.; Davidoff E. F.; Marks T. J. Phosphorus-31 Chemical Shifts of Quaternary Phosphonium Salts. J. Phys. Chem. 1966, 70, 581–584. 10.1021/j100874a502. [DOI] [Google Scholar]; e van Wazer J. R.; Callis C. F.; Shoolery J. N.; Jones R. C. Principles of Phosphorus Chemistry. II. Nuclear Magnetic Resonance Measurements 1. J. Am. Chem. Soc. 1956, 78, 5715–5726. 10.1021/ja01603a002. [DOI] [Google Scholar]
  23. Schmidpeter A.; Lochschmidt S.; Sheldrick W. S. A diphos-Complex of P+: The 1,1,3,3-Tetraphenyl-1λ5,2λ3,3λ5-triphospholenyl Cation. Angew. Chem., Int. Ed. Engl. 1982, 21, 63–64. 10.1002/anie.198200631. [DOI] [Google Scholar]; Ein diphos-Komplex von P+: das 1,1,3,3-Tetraphenyl-1λ5,2λ3,3λ5-triphospholenyl-Kation. Angew. Chem. 1982, 94, 172. 10.1002/ange.19820940120
  24. a Dyker C. A.; Burford N. catena-phosphorus cations. Chem.—Asian J. 2008, 3, 28–36. 10.1002/asia.200700229. [DOI] [PubMed] [Google Scholar]; b Wiberg N.; Fischer G.; Holleman A. F.; Wiberg E.. Lehrbuch der Anorganischen Chemie; de Gruyter, 2008, 102. stark umgearb. und verb. Aufl.. [Google Scholar]
  25. Yoshifuji M.; Shima I.; Inamoto N.; Hirotsu K.; Higuchi T. Synthesis and structure of bis(2,4,6-tri-tert-butylphenyl)diphosphene: isolation of a true phosphobenzene. J. Am. Chem. Soc. 1981, 103, 4587–4589. 10.1021/ja00405a054. [DOI] [Google Scholar]
  26. Schumann A.; Reiß F.; Jiao H.; Rabeah J.; Siewert J.-E.; Krummenacher I.; Braunschweig H.; Hering-Junghans C. A selective route to aryl-triphosphiranes and their titanocene-induced fragmentation. Chem. Sci. 2019, 10, 7859–7867. 10.1039/c9sc02322d. [DOI] [PMC free article] [PubMed] [Google Scholar]
  27. Dionisi E. M.; Binder J. F.; LaFortune J. H. W.; Macdonald C. L. B. Triphosphenium salts: air-stable precursors for phosphorus(I) chemistry. Dalton Trans. 2020, 49, 12115–12127. 10.1039/d0dt02219e. [DOI] [PubMed] [Google Scholar]
  28. a Cava M. P.; Levinson M. I. Thionation reactions of Lawesson’s reagents. Tetrahedron 1985, 41, 5061–5087. 10.1016/s0040-4020(01)96753-5. [DOI] [Google Scholar]; b Pedersen B. S.; Scheibye S.; Clausen K.; Lawesson S.-O. Studies on organophosphorus compounds XXII The dimer of p-methoxyphenylthionophosphine sulfide as thiation reagent. a new route to O-substituted thioesters and dithioesters. Bull. Soc. Chim. Belg. 2010, 87, 293–297. 10.1002/bscb.19780870407. [DOI] [Google Scholar]; c Polshettiwar V.; Kaushik M. P. A new, efficient and simple method for the thionation of ketones to thioketones using P4S10/Al2O3. Tetrahedron Lett. 2004, 45, 6255–6257. 10.1016/j.tetlet.2004.06.091. [DOI] [Google Scholar]; d Polshettiwar V.; Nivsarkar M.; Paradashani D.; Kaushik M. P. Thionation of carbonyl compounds using phosphorus pentasulfide and hexamethyldisiloxane under microwave irradiations. J. Chem. Res. 2004, 2004, 474–476. 10.3184/0308234042037149. [DOI] [Google Scholar]; e Scheibye S.; Kristensen J.; Lawesson S.-O. Studies on organophosphorus compounds—XXVII. Tetrahedron 1979, 35, 1339–1343. 10.1016/0040-4020(79)85027-9. [DOI] [Google Scholar]
  29. Klebach T. C.; Lourens R.; Bickelhaupt F. Synthesis of Mesityldiphenylmethylenephosphine: a stable compound with a localized P=C bond. J. Am. Chem. Soc. 1978, 100, 4886–4888. 10.1021/ja00483a041. [DOI] [Google Scholar]
  30. Ellis B. D.; Dyker C. A.; Decken A.; Macdonald C. L. B. The synthesis, characterisation and electronic structure of N-heterocyclic carbene adducts of P(I) cations. Chem. Commun. 2005, 1965–1967. 10.1039/b500692a. [DOI] [PubMed] [Google Scholar]
  31. a Krachko T.; Slootweg J. C. N-Heterocyclic Carbene–Phosphinidene Adducts: Synthesis, Properties, and Applications. Eur. J. Inorg. Chem. 2018, 2018, 2734–2754. 10.1002/ejic.201800459. [DOI] [Google Scholar]; b Schwedtmann K.; Zanoni G.; Weigand J. J. Recent Advances in Imidazoliumyl-Substituted Phosphorus Compounds. Chem.—Asian J. 2018, 13, 1388–1405. 10.1002/asia.201800199. [DOI] [PubMed] [Google Scholar]
  32. van der Knaap T.; Klebach T.; Visser F.; Bickelhaupt F.; Ros P.; Baerends E. J.; Stam C. H.; Konijn M. Synthesis and structure of aryl-substituted phospha-alkenes. Tetrahedron 1984, 40, 765–776. 10.1016/s0040-4020(01)91106-8. [DOI] [Google Scholar]
  33. Adhikari A. K.; Grell T.; Lönnecke P.; Hey-Hawkins E. Formation of a Carbene–Phosphinidene Adduct by NHC-Induced P–P Bond Cleavage in Sodium Tetramesityltetraphosphanediide. Eur. J. Inorg. Chem. 2016, 2016, 620–622. 10.1002/ejic.201500952. [DOI] [Google Scholar]
  34. a Schlosser M.; Christmann K. F. Trans-Selective Olefin Syntheses. Angew. Chem., Int. Ed. Engl. 1966, 5, 126. 10.1002/anie.196601261. [DOI] [Google Scholar]; Trans-selektive Olefinsynthesen. Angew. Chem. 1966, 78, 115. 10.1002/ange.19660780118; b Vedejs E.; Marth C. F. Mechanism of Wittig reaction: evidence against betaine intermediates. J. Am. Chem. Soc. 1990, 112, 3905–3909. 10.1021/ja00166a026. [DOI] [Google Scholar]; c Yamataka H.; Nagase S. Theoretical Calculations on the Wittig Reaction Revisited. J. Am. Chem. Soc. 1998, 120, 7530–7536. 10.1021/ja974237f. [DOI] [Google Scholar]
  35. Arkhypchuk A. I.; Svyaschenko Y. V.; Orthaber A.; Ott S. Mechanism of the Phospha-Wittig-Horner reaction. Angew. Chem., Int. Ed. Engl. 2013, 125, 6612–6615. 10.1002/ange.201301469. [DOI] [PMC free article] [PubMed] [Google Scholar]; Mechanism of the Phospha-Wittig–Horner Reaction. Angew. Chem. 2013, 125, 6612 - 6615.10.1002/ange.201301469 [DOI] [PMC free article] [PubMed]
  36. a Espinosa Ferao A.; Streubel R. Thiaphosphiranes and Their Complexes: Systematic Study on Ring Strain and Ring Cleavage Reactions. Inorg. Chem. 2016, 55, 9611–9619. 10.1021/acs.inorgchem.6b01322. [DOI] [PubMed] [Google Scholar]; b Toyota K.; Shimura K.; Takahashi H.; Yoshifuji M. Preparation and X-Ray Structure Analysis of 3,3-Diphenyl-2-(2,4,6-tri- t -butylphenyl)-1,2-thiaphosphirane. Chem. Lett. 1994, 23, 1927–1930. 10.1246/cl.1994.1927. [DOI] [Google Scholar]; c Turcheniuk K. V.; Rozhenko A. B. (σ35)-Phosphoranes versus (σ33)-thiaphosphiranes: quantum chemical investigation of products of phosphaalkene sulfurization. J. Comput. Chem. 2012, 33, 1023–1028. 10.1002/jcc.22932. [DOI] [PubMed] [Google Scholar]; d Turcheniuk K. V.; Rozhenko A. B.; Shevchenko I. V. Synthesis and Some Chemical Properties of a 1,2λ 3 σ 3 -Thiaphosphirane. Eur. J. Inorg. Chem. 2011, 2011, 1762–1767. 10.1002/ejic.201001192. [DOI] [Google Scholar]
  37. a Latypov S. K.; Polyancev F. M.; Yakhvarov D. G.; Sinyashin O. G. Quantum chemical calculations of 31P NMR chemical shifts: scopes and limitations. Phys. Chem. Chem. Phys. 2015, 17, 6976–6987. 10.1039/c5cp00240k. [DOI] [PubMed] [Google Scholar]; b Robert V.; Petit S.; Borshch S. A.; Bigot B. Theoretical Analysis of 31P NMR Chemical Shifts in Vanadium Phosphorus Oxides. J. Phys. Chem. A 2000, 104, 4586–4591. 10.1021/jp994400o. [DOI] [Google Scholar]
  38. Marinetti A.; Bauer S.; Ricard L.; Mathey F. The ″phospha-Wittig″ reaction: a new method for building phosphorus-carbon double and single bonds from carbonyl compounds. Organometallics 1990, 9, 793–798. 10.1021/om00117a040. [DOI] [Google Scholar]
  39. a Niecke E.; Rüger R.; Lysek M.; Schoeller W. W. Aminophosphinidene derivatives. Phosphorus, Sulfur Silicon Relat. Elem. 1983, 18, 35–38. 10.1080/03086648308075961. [DOI] [Google Scholar]; b Streubel R.; Huy N. H. T.; Mathey F. The Reaction of Terminal Phosphinidene Complexes with Phosphaalkenes: Thermal Symmetrisation of Diphosphiranes with Asymmetric Substitution Patterns. Synthesis 1993, 1993, 763–764. 10.1055/s-1993-25932. [DOI] [Google Scholar]; c Treubel R.; Niecke E. Halogen(Silyl)phosphane – Synthese und Eigenschaften. Chem. Ber. 1990, 123, 1245–1251. 10.1002/cber.19901230606. [DOI] [Google Scholar]
  40. Wang S.; Samedov K.; Serin S. C.; Gates D. P. PhP=CPh2 and Related Phosphaalkenes: A Solution Equilibrium between a Phosphaalkene and a 1,2-Diphosphetane. Eur. J. Inorg. Chem. 2016, 2016, 4144–4151. 10.1002/ejic.201600599. [DOI] [Google Scholar]
  41. a Baudler M.; Gruner C. Beiträge zur Chemie des Phosphors, 69 Über ein einfaches Verfahren zur Darstellung von Tri-t-butyl-cyclotriphosphan. Z. Naturforsch., B: Anorg. Chem., Org. Chem. 1976, 31, 1311–1312. 10.1515/znb-1976-1002. [DOI] [Google Scholar]; b Baudler M.; Hahn J.; Dietsch H.; Fürstenberg G. Beiträge zur Chemie des Phosphors, 68 Tri-t-butyl-cyclotriphosphan. Z. Naturforsch., B: Anorg. Chem., Org. Chem. 1976, 31, 1305–1310. 10.1515/znb-1976-1001. [DOI] [Google Scholar]
  42. a Etemad-Moghadam G.; Koenig M.. Three-membered Rings. Phosphorus-Carbon Heterocyclic Chemistry; Elsevier, 2001; pp 57–86. [Google Scholar]; b Riu M.-L. Y.; Eckhardt A. K.; Cummins C. C. Reactions of Tri-tert-Butylphosphatetrahedrane as a Spring-Loaded Phosphinidene Synthon Featuring Nickel-Catalyzed Transfer to Unactivated Alkenes. J. Am. Chem. Soc. 2022, 144, 7578–7582. 10.1021/jacs.2c02236. [DOI] [PubMed] [Google Scholar]; c Streubel R.; Nieger M.; Niecke E. Synthese, Struktur und Reaktivität von Diphosphiranen. Chem. Ber. 1993, 126, 645–648. 10.1002/cber.19931260314. [DOI] [Google Scholar]
  43. a Bates J. I.; Gates D. P. Diphosphiranium (P2C) or diphosphetanium (P2C2) cyclic cations: different fates for the electrophile-initiated cyclodimerization of a phosphaalkene. J. Am. Chem. Soc. 2006, 128, 15998–15999. 10.1021/ja0667662. [DOI] [PubMed] [Google Scholar]; b Bates J. I.; Gates D. P. P=C bonds as building blocks for three- and four-membered heterocyclic cations: synthesis, structures and mechanistic studies. Chem.—Eur. J. 2012, 18, 1674–1683. 10.1002/chem.201101535. [DOI] [PubMed] [Google Scholar]
  44. a Kroto H. W.; Klein S. L.; Meidine M. F.; Nixon J. F.; Harris R. K.; Packer K. J.; Reams P. η1- and η2-coordination in phosphaalkeneplatinum(0) complexes. High resolution solid state 31P NMR spectrum of mesityl(diphenylmethylene)phosphinebis(triphenylphosphine)platinum(0). J. Org. Chem. 1985, 280, 281–287. 10.1016/0022-328x(85)88100-6. [DOI] [Google Scholar]; b Floch P. Phosphaalkene, phospholyl and phosphinine ligands: New tools in coordination chemistry and catalysis. Coord. Chem. Rev. 2006, 250, 627–681. 10.1016/j.ccr.2005.04.032. [DOI] [Google Scholar]; c Pidcock A.; Richards R. E.; Venanzi L. M. 195Pt–31P nuclear spin coupling constants and the nature of the trans-effect in platinum complexes. J. Chem. Soc. A 1966, 1707–1710. 10.1039/j19660001707. [DOI] [Google Scholar]
  45. a Chatt J.; Hitchcock P. B.; Pidcock A.; Warrens C. P.; Dixon K. R. Synthesis and 31P n.m.r. spectroscopy of platinum and palladium complexes containing side-bonded diphenyldiphosphene. The X-ray crystal and molecular structure of [Pd(PhP=pph){bis(diphenyl-phosphino)ethane}]. J. Chem. Soc., Chem. Commun. 1982, 16, 932–933. 10.1039/c39820000932. [DOI] [Google Scholar]; b Chatt J.; Hitchcock P. B.; Pidcock A.; Warrens C. P.; Dixon K. R. The nature of the co-ordinate link. Part 11. Synthesis and phosphorus-31 nuclear magnetic resonance spectroscopy of platinum and palladium complexes containing side-bonded (E)-diphenyldiphosphene. X-Ray crystal and molecular structures of [Pd{(E)-PhP=PPh}(Ph2PCH2CH2PPh2)] and [Pd{[(E)-PhP=PPh] [W(CO)5]2}(Ph2PCH2CH2PPh2)]. J. Chem. Soc., Dalton Trans. 1984, 2237–2244. 10.1039/dt9840002237. [DOI] [Google Scholar]; c Phillips I. G.; Ball R. G.; Cavell R. G. Reactions of perfluoromethyl-substituted cyclopolyphosphines with zerovalent group 10 metal complexes. Crystal and molecular structure of a [palladium] complex with a coordinated diphosphene, [Pd(.eta.2-CF3P=PCF3)(PPh3)2]. Inorg. Chem. 1992, 31, 1633–1641. 10.1021/ic00035a022. [DOI] [Google Scholar]; d Yoshifuji M. Coordination Chemistry of Some Low Coordinate Organophosphorus Compounds of Coordination Number 2. Bull. Chem. Soc. Jpn. 1997, 70, 2881–2893. 10.1246/bcsj.70.2881. [DOI] [Google Scholar]
  46. a van der Knaap T. A.; Bickelhaupt F.; Kraaykamp J. G.; van Koten G.; Bernards J. P. C.; Edzes H. T.; Veeman W. S.; De Boer E.; Baerends E. J. The η1- and η2-coordination in a (phosphaalkene)platinum(0) complex. Organometallics 1984, 3, 1804–1811. 10.1021/om00090a005. [DOI] [Google Scholar]; b van der Knaap T. A.; Bickelhaupt F.; van der Poel H.; van Koten G.; Stam C. H. Synthesis and structural investigation of [mesityl(diphenylmethylene)phosphine]bis(triphenylphosphine)platinum(0). J. Am. Chem. Soc. 1982, 104, 1756–1757. 10.1021/ja00370a060. [DOI] [Google Scholar]
  47. a Alcarazo M. α-Cationic phosphines: synthesis and applications. Chem.—Eur. J. 2014, 20, 7868–7877. 10.1002/chem.201402375. [DOI] [PubMed] [Google Scholar]; b Alcarazo M. Synthesis, Structure, and Applications of α-Cationic Phosphines. Acc. Chem. Res. 2016, 49, 1797–1805. 10.1021/acs.accounts.6b00262. [DOI] [PubMed] [Google Scholar]
  48. a Mathey F. Phospha-organic chemistry: panorama and perspectives. Angew. Chem., Int. Ed. Engl. 2003, 34, 1578–1604. 10.1002/chin.200329207. [DOI] [PubMed] [Google Scholar]; b Mathey F. Phosphaorganische Chemie: Panorama und Perspektiven. Angew. Chem. 2003, 115, 1616–1643. 10.1002/ange.200200557. [DOI] [Google Scholar]
  49. a Huy N. H. T.; Fischer J.; Mathey F. Synthesis and x-ray crystal structure analysis of a .eta.4-1-phosphabutadienetetracarbonyltungsten complex. J. Am. Chem. Soc. 1987, 109, 3475–3477. 10.1021/ja00245a057. [DOI] [Google Scholar]; b Knoll K.; Huttner G.; Wasiucionek M.; Zsolnai L. Synthesis, Structure, and Reactions of Cluster-Stabilized Phosphaalkenes. Angew. Chem., Int. Ed. Engl. 1984, 23, 739–740. 10.1002/anie.198407391. [DOI] [Google Scholar]; Synthese, Struktur und Reaktionen Clusterstabilisierter Phosphaalkene. Angew. Chem. 1984, 96, 708 - 709.10.1002/ange.19840960922; c Marinetti A.; Bauer S.; Ricard L.; Mathey F. Preliminary investigations on the chemistry of (η4-1-phosphabutadiene)tricarbonyliron complexes. Synthesis and X-ray crystal structure analysis of a (η4-2-phosphabutadiene)tricarbonyliron complex. J. Chem. Soc., Dalton Trans. 1991, 1991, 597–602. 10.1039/dt9910000597. [DOI] [Google Scholar]
  50. a Bauer D. P.; Ruff J. K. Novel iron tetracarbonyl fluorophosphine complexes. Inorg. Chem. 1983, 22, 1686–1689. 10.1021/ic00153a026. [DOI] [Google Scholar]; b Binder J. F.; Kosnik S. C.; Macdonald C. L. B. Assessing the Ligand Properties of NHC-Stabilised Phosphorus(I) Cations. Chem.—Eur. J. 2018, 24, 3556–3565. 10.1002/chem.201705224. [DOI] [PubMed] [Google Scholar]; c Cowley A. H.; Kemp R. A.; Ebsworth E.; Rankin D. W.; Walkinshaw M. D. Structure/reactivity relationships for cationic (phosphenium)iron tetracarbonyl complexes. J. Org. Chem. 1984, 265, c19–c21. 10.1016/0022-328x(84)80078-9. [DOI] [Google Scholar]
  51. Klebach T. C.; Lourens R.; Bickelhaupt F.; Stam C. H.; van Herk A. Synthesis and structure of pentacarbonyl(mesityldiphenylmethylenephosphine)-chromium(0). J. Org. Chem. 1981, 210, 211–221. 10.1016/s0022-328x(00)82233-0. [DOI] [Google Scholar]
  52. Tolman C. A. Steric effects of phosphorus ligands in organometallic chemistry and homogeneous catalysis. Chem. Rev. 1977, 77, 313–348. 10.1021/cr60307a002. [DOI] [Google Scholar]
  53. a Bansal R. K.; Kumawat S. K. Diels–Alder reactions involving CP– functionality. Tetrahedron 2008, 64, 10945–10976. 10.1016/j.tet.2008.09.037. [DOI] [Google Scholar]; b Gust T.; Du Mont W.-W.; Schmutzler R.; Hrib C. G.; Wismach C.; Jones P. G. Diastereoselective Reactions of Sulfur- and Selenium-Bridged Bisphosphaalkenes with Tetrachloro-o-benzoquinone. Phosphorus, Sulfur Silicon Relat. Elem. 2009, 184, 1599–1611. 10.1080/10426500902947948. [DOI] [Google Scholar]; c van der Knaap T.; Bickelhaupt F. The reaction of a phosphaalkene with orthoquinones. Tetrahedron 1983, 39, 3189–3196. 10.1016/s0040-4020(01)91565-0. [DOI] [Google Scholar]
  54. a Gates D. P. Expanding the Analogy between P=C and C=C Bonds to Polymer Science. Top. Curr. Chem. 2005, 250, 107–126. 10.1007/b100983. [DOI] [Google Scholar]; b Noonan K. J. T.; Gates D. P. Ambient-temperature living anionic polymerization of phosphaalkenes: homopolymers and block copolymers with controlled chain lengths. Angew. Chem., Int. Ed. Engl. 2006, 45, 7271–7274. 10.1002/anie.200602955. [DOI] [PubMed] [Google Scholar]; Angew. Chem. 2006, 118, 7429 - 7432.10.1002/ange.200602955; c Siu P. W.; Serin S. C.; Krummenacher I.; Hey T. W.; Gates D. P. Isomerization polymerization of the phosphaalkene MesP=CPh2: an alternative microstructure for poly(methylenephosphine)s. Angew. Chem., Int. Ed. Engl. 2013, 52, 6967–6970. 10.1002/anie.201301881. [DOI] [PubMed] [Google Scholar]; Angew. Chem. 2013, 125, 7105 - 7108.10.1002/ange.201301881
  55. a Becker G.; Uhl W.; Wessely H.-J. Acyl- und Alkylidenphosphane. XVI. (Dimethylaminomethyliden)- und (Diphenylmethyliden)phosphane. Z. Anorg. Allg. Chem. 1981, 479, 41–56. 10.1002/zaac.19814790805. [DOI] [Google Scholar]; b Yam M.; Chong J. H.; Tsang C.-W.; Patrick B. O.; Lam A. E.; Gates D. P. Scope and limitations of the base-catalyzed phospha-peterson P=C bond-forming reaction. Inorg. Chem. 2006, 45, 5225–5234. 10.1021/ic060236p. [DOI] [PubMed] [Google Scholar]
  56. Wong Y. O.; Freeman L. A.; Agakidou A. D.; Dickie D. A.; Webster C. E.; Gilliard R. J. Two Carbenes versus One in Magnesium Chemistry: Synthesis of Terminal Dihalide, Dialkyl, and Grignard Reagents. Organometallics 2019, 38, 688–696. 10.1021/acs.organomet.8b00866. [DOI] [Google Scholar]
  57. a Arkhypchuk A. I.; D’Imperio N.; Wells J. A. L.; Ott S. [2 + 2] Cycloaddition of phosphaalkenes as a key step for the reductive coupling of diaryl ketones to tetraaryl olefins. Chem. Sci. 2022, 13, 12239–12244. 10.1039/d2sc03073j. [DOI] [PMC free article] [PubMed] [Google Scholar]; b Bates J. I.; Patrick B. O.; Gates D. P. A Lewis acid-mediated synthesis of P-alkyl-substituted phosphaalkenes. New J. Chem. 2010, 34, 1660. 10.1039/c0nj00188k. [DOI] [Google Scholar]; c D’Imperio N.; Arkhypchuk A. I.; Mai J.; Ott S. Triphenylphosphaalkenes in Chemical Equilibria. Eur. J. Inorg. Chem. 2019, 2019, 1562–1566. 10.1002/ejic.201801322. [DOI] [Google Scholar]; d Becker G.; Gresser G.; Uhl W. Acyl-und Alkylidenphosphane, XV 2.2-Dimethylpropylidinphosphan, eine stabile Verbindung mit einem Phosphoratom der Koordinationszahl 1/Acyl- and Alkylidenephosphines, XV 2,2-Dimethylpropvlidynephosphine, a Stable Compound with a Phosphorus Atom of Coordination Number 1. Z. Naturforsch., B: Anorg. Chem., Org. Chem. 1981, 36, 16–19. 10.1515/znb-1981-0105. [DOI] [Google Scholar]; e Becker G.; Massa W.; Mundt O.; Schmidt R. Acyl- und Alkylidenphosphane. XIX. Molekül- und Kristallstruktur des 2,4-Bis(dimethylamino)-1,3-diphenyl-1,3-diphosphetans. Z. Anorg. Allg. Chem. 1982, 485, 23–35. 10.1002/zaac.19824850104. [DOI] [Google Scholar]; f Becker G.; Mundt O. Bildung und Eigenschaften von Acylphosphanen. IX. Reaktion von Phenylbis(trimethylsilyl)phosphan mit Formaldehyd und Dimethylformamid. Z. Anorg. Allg. Chem. 1980, 462, 130–142. 10.1002/zaac.19804620114. [DOI] [Google Scholar]; g Grützmacher H.; Pritzkow H. Three Independent Molecules in the Unit Cell of a Phosphorus Ylide; Changes in the Molecular Geometry on Rotation about the Phosphorus–Carbon Ylide Bond. Angew. Chem., Int. Ed. Engl. 1992, 31, 99–101. 10.1002/anie.199200991. [DOI] [Google Scholar]; Drei unabhängige Moleküle in der Elementarzelle eines Phosphor-Ylids; Änderungen der Molekülgeometrie bei der Rotation um die Phosphor-Kohlenstoff-Ylidbindung. Angew. Chem. 1992, 104, 92 - 94.10.1002/ange.19921040131; h Schrödel H.-P.; Jochem G.; Schmidpeter A.; Nöth H. Ylidediyl Halophosphane Oligomers, (Ph3PCPX)n—A Structurally Diverse Class of Compounds. Angew. Chem., Int. Ed. 1995, 34, 1853–1856. 10.1002/anie.199518531. [DOI] [Google Scholar]; Yliddiyl-halogenphosphan-Oligomere (Ph3PCPX)n – eine strukturell vielfältige Verbindungsfamilie. Angew. Chem. 1995, 107, 2006 - 2010.10.1002/ange.19951071714; I Becker G.; Uhl W. Acyl- und Alkylidenphosphane. XIII [1]. Molekül- und Kristallstruktur des 2,4-Di(tert.butyl)-2,4-bis-(trimethylsiloxy)-1,3-diphosphetans. Z. Anorg. Allg. Chem. 1981, 475, 35–44. 10.1002/zaac.19814750406. [DOI] [Google Scholar]; j Becker G.; Becker W.; Uhl G.; Uhl W.; Wessely H.-J. 1,2- and 1,3-diphosphetanes from alkylidenephosphines. Phosphorus Sulfur Relat. Elem. 1983, 18, 7–10. 10.1080/03086648308075954. [DOI] [Google Scholar]
  58. Weigand J. J.; Feldmann K.-O.; Henne F. D. Carbene-stabilized phosphorus(III)-centered cations LPX2+ and L2PX2+ (L = NHC; X = Cl, CN, N3). J. Am. Chem. Soc. 2010, 132, 16321–16323. 10.1021/ja106172d. [DOI] [PubMed] [Google Scholar]
  59. a Bîrzoi R. M.; Bugnariu D.; Gimeno R. G.; Riecke A.; Daniliuc C.; Jones P. G.; Könczöl L.; Benkõ Z.; Nyulászi L.; Bartsch R.; Du Mont W.-W. Access to Metal Complexes of the Elusive Imidobis(phosphaalkene) Anion by N–Si Bond Cleavage of a N -Silylimino-Bridged Bis(phosphaalkene). Eur. J. Inorg. Chem. 2010, 2010, 29–33. 10.1002/ejic.200900982. [DOI] [Google Scholar]; b Bîrzoi R. M.; Jones P. G.; Bartsch R.; Du Mont W.-W. P -Aminophosphaalkenes with C -Isopropyldimethylsilyl Groups. Z. Anorg. Allg. Chem. 2019, 645, 712–722. 10.1002/zaac.201900049. [DOI] [Google Scholar]; c Bîrzoi R. M.; Lungu D.; Jones P. G.; Bartsch R.; Du Mont W.-W. Sterically Tuned P -Phosphanylamino Phosphaalkenes (Me3Si)2C=PN(R)PPh2 and (iPrMe2Si)2C=PN(R)PPh2. Z. Anorg. Allg. Chem. 2018, 644, 381–390. 10.1002/zaac.201700446. [DOI] [Google Scholar]

Associated Data

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

Supplementary Materials

ja3c02256_si_001.pdf (13.4MB, pdf)

Articles from Journal of the American Chemical Society are provided here courtesy of American Chemical Society

RESOURCES