Phosphasilene

Last updated
Schematic illustration of a generic phosphasilene with a three-coordinate silicon atom and a two-coordinate phosphorus atom. Generic phosphasilene.tif
Schematic illustration of a generic phosphasilene with a three-coordinate silicon atom and a two-coordinate phosphorus atom.

Phosphasilenes or silylidenephosphanes are a class of compounds with silicon-phosphorus double bonds. [1] Since the electronegativity of phosphorus (2.1) is higher than that of silicon (1.9), the "Si=P" moiety of phosphasilene is polarized. [2] [3] The degree of polarization can be tuned by altering the coordination numbers of the Si and P centers, or by modifying the electronic properties of the substituents. [2] The phosphasilene Si=P double bond is highly reactive, yet with the choice of proper substituents, it can be stabilized via donor-acceptor interaction or by steric congestion. [3]

Contents

Structure of the first phosphasilene prepared by Bickelhaupt et al. Structure of the first phosphasilene prepared by Bickelhaupt et al.tif
Structure of the first phosphasilene prepared by Bickelhaupt et al.

The landmark discovery of the first phosphasilene by NMR spectroscopy was made in 1984 by Bickelhaupt et al. [4] The first phosphasilene came with bulky aryl substituents at the phosphorus and silicon atoms. [4] Almost a decade after this spectroscopic observation, the first structural characterization of phosphasilene was achieved in 1993 by Niecke et al. [5] The successful isolation of phosphasilenes with silicon-phosphorus double bonds represents one of the discoveries that challenged and disproved the "double-bond rule".

Synthesis

Synthesis via β-elimination

The synthetic pathway towards the first metastable arylphosphasilenes developed by Bickelhaupt et al. Synthetic scheme of arylphosphasilenes employed by Bickelhaupt et al.tif
The synthetic pathway towards the first metastable arylphosphasilenes developed by Bickelhaupt et al.

An important synthetic pathway towards phosphasilene is the 1,2-elimination reactions of silylphosphane derivatives. The first metastable arylphosphasilenes were accessed by Bickelhaupt et al. via the deprotonation of in situ formed (chlorosilyl)phosphanes ArP(H)–(Cl)SiAr2 using organolithium bases. [2] [4] [6] [7] As shown in the scheme on the left, it is also possible to use pre-formed lithium phosphanides Ar'P(H)Li as both a phosphorus source and a base. However, the latter synthetic pathway involves formation of primary phosphines ArPH2, which can be difficult of be separated from phosphasilenes. [2] Despite such a drawback, this strategy has been successfully applied by Niecke et al. to obtain a series of 1,3-diphospha-2-silaallyl anions, which serve as precursors for 2-phosphanylphosphasilenes. [2] [8]

The preparation of 2-phosphanylphosphasilenes via a 1,3-diphospha-2-silaallyl anion intermediate developed by Niecke et al. Preparation of phosphasilenes via a 1,3-diphospha-2-silaallyl anion intermediate.tif
The preparation of 2-phosphanylphosphasilenes via a 1,3-diphospha-2-silaallyl anion intermediate developed by Niecke et al.

Applying an analogous strategy, Driess and coworkers developed an effective approach for synthesizing P-silyl phosphasilenes via the thermal elimination of LiF from corresponding lithium (flouorosilyl)phosphanides. [9] [10] [11]

The synthesis of phosphasilenes via the thermal elimination of LiF from lithium (flouorosilyl)phosphanides developed by Driess et al. Phosphasilene synthesis via LiF elimination.tif
The synthesis of phosphasilenes via the thermal elimination of LiF from lithium (flouorosilyl)phosphanides developed by Driess et al.

Synthesis based on reactivities of stable silyene

Synthetic scheme of phosphasilene with 4-coordinate silicon based on reactivities of stable silyene complexes developed by Inoue, Driess and coworkers. Synthesis of phosphasilene with 4-coordinate silicon.jpg
Synthetic scheme of phosphasilene with 4-coordinate silicon based on reactivities of stable silyene complexes developed by Inoue, Driess and coworkers.

Phosphasilenes with 4-coordinate silicon, which can also be viewed as silylene-stabilized phosphinidene, can be synthesized based on the reactivities of stable silyene complexes. [2] For example, Inoue and coworkers demonstrated that benzamidinate-stabilized phosphanylsilylene can give rise to corresponding Si- and P-trimethylsilyl-substituted phosphasilene via thermal rearrangement, while the reaction can also yield 4-disila-1,3-diphosphacyclobuta-diene with the addition of a mild chlorinating agent Ph3PCl2. [12] [2] In these phosphasilenes with four-coordinate silicon, even though formally five bonds are drawn around silicon, the Si–P π bonds are calculated to be strongly polarized towards the P atoms. The contribution from phosphorus and silicon to the π orbitals are calculated to be 87.53% and 12.47%, respectively.

Structure and bonding

The parent phosphasilene H2Si=PH

Laplacian map of the parent phosphasilene H2Si=PH in the plane of the p bond. Selected critical points and bond paths are shown. The structure of H2Si=PH was optimized at the B97-D/6-311+G** level of theory. Multiwfn was used to carry out the AIM analysis. Laplacian map for the parent phosphasilene.png
Laplacian map of the parent phosphasilene H2Si=PH in the plane of the π bond. Selected critical points and bond paths are shown. The structure of H2Si=PH was optimized at the B97-D/6-311+G** level of theory. Multiwfn was used to carry out the AIM analysis.

The unstable parent phosphasilene H2Si=PH has been generated in the gas phase by the reacting atomic silicon with phosphine PH3, and identified via matrix isolation spectroscopy methods. [14] Density functional theory (DFT) calculations suggests that in the ground state, H2Si=PH exists in a singlet spin state, with Cs-symmetric planar geometry. [14] At the B3LYP/6-311+G** level of theory, the Si=P bond length and the Si-P-H bond angle are calculated to be 2.084 Å and 90.7o. [14] The Si=P bond dissociation energy is 75.0 kcal mol−1 at the B97-D/6-31G(d) level of theory; while the π-bond energy, Dπ(Si=P) is 36.6/35.9 kcal mol−1. [13] The frontier orbitals of the parent phosphasilene consists of the π bonding and π anti-bonding orbitals: π(Si=P) and π*(Si=P) correspond to the HOMO and LUMO, respectively. HOMO-1 was calculated to be the lone pair on phosphorus n(P). [15]

"Half-Parent" Phosphasilene R2Si=PH

Driess and coworkers prepared thermally stable "half"-parent phosphasilene R2Si=PH (R2Si = (tBu3Si)(iPr3C6H2)Si), which is the first example of phosphasilene with a terminal PH group. [16] This species was obtained as a mixture of E/Z isomers, thus its 31P NMR spectrum featured two doublets with 29Si satellites (δ=123.1, 1J (P, H)=123 Hz, 1J (P, Si)=157 Hz and δ=134.2 ppm, 1J (P, H)=131 Hz, 1J (P, Si)=130 Hz). [16] These 1J (P, H) coupling constants are much smaller compared to those of secondary phosphanes (R2PH) and phosphaalkenes with a PH group, which indicates that the phosphasilene phosphorus atom possesses more 3p character. [16] X-ray crystallography of this "half"-parent phosphasilene species shows that the silicon atom occupies a trigonal-planar coordination environment. [16] The Si=P bond distance was reported to be 2.094(5) Å, which is about 7% shorter than a typical silicon-phosphorus single bond, but only slightly longer than that of P-silyl-substituted phosphasilenes, [16] which suggests that the potential of Si=P bond on a potential energy surface is relatively shallow. [16] [17]

π-conjugated phosphasilenes

Structures of p-conjugated phosphasilenes prepared by Tamao et al. Pi-conjugated phosphasilenes.tif
Structures of π-conjugated phosphasilenes prepared by Tamao et al.

Tamao et al. reported a series of π-conjugated phosphasilenes stabilized by Eind groups (Eind=1,1,3,3,5,5,7,7-octaethyl-s-hydrindacen-4-yl). [18] These systems feature Si=P units that are highly coplanar with the aromatic ring, allowing strong ππ* absorptions. The coplanarity is made possible by the rigidity of the two Eind groups that are oriented trans and perpendicular with respect to the Si=P bond. [18] The Si=P bond length observed by X-ray crystallography are ca. 2.09-2.10 Å, which are typical for phosphasilenes. [18]

Frontier orbitals of a p-conjugated phosphasilene reported by Tamao et al. The orbital calculations correspond to the "complex a" shown in the figure above). Frontier orbitals of p conjugated phosphasilene.tif
Frontier orbitals of a π-conjugated phosphasilene reported by Tamao et al. The orbital calculations correspond to the "complex a" shown in the figure above).

The bonding of π-conjugated phosphasilenes has been probed by DFT calculations at the B3LYP/6-31G** level. The HOMO was calculated to represent mostly the 3pπ(Si–P), while the LUMO featured significant contribution from the 3pπ*(Si–P)–2pπ*(phenyl) conjugation. The HOMO-1 orbital involves the 3n–2pπ conjugation, which originate from the presence of lone pair on the phosphorus atom and the π-orbital on the Eind benzene ring. [18]

"Push-pull" phosphasilene

By installing electron-donating substituents on silicon and electron-withdrawing substituents on phosphorus, the Si=P bond polarization can be decreased and even reversed through the "push-pull" interaction of the substituents with opposing electronic effects. [19] Applying this design strategy, Escudié et al. prepared stable "push-pull" phosphasilene (tBu2MeSi)2Si=PMes* (Mes* = 2,4,6-tri-tertbutylphenyl) with electron-donating silyl groups on Si and an electron-withdrawing aryl group on P. [19] Computations on the model compound (Me3Si)2Si=PMes (Mes = 2,4,6-trimethylphenyl) demonstrate that n(P) and π*(Si=P) correspond to the HOMO and LUMO, respectively. [19] The relatively small energy gap between the interacting occupied (n(P)) and vacant (π*(Si=P)) molecular orbitals gives rise to a large paramagnetic contribution, [20] which explains the extreme deshielding of the doubly bonded Si and P atoms, as well as the red shift in the UV spectrum that are observed in (tBu2MeSi)2Si=PMes*. [19] In the "push-pull" phosphasilene prepared by Escudié and co-workers, the Si=P bond length was reported to be 2.1114(7) Å, which is longer than what was observed in most other reported phosphasilenes (2.062(1)–2.094(5) Å). [19]

Metallophosphasilenes R2Si=PM

Driess et al. demonstrated that stable metallophosphasilenes of the type R2Si=PM can be prepared from metalation reaction of "half"-parent phosphasilenes R2Si=PH (R2Si = (tBu3Si)(iPr3C6H2)Si). [16] NMR spectroscopic studies demonstrated that metalation led to the deshielding of the 31P nucleus, while the low coordinate 29Si atom in the metallophosphasilene became more shielded. Driess and coworkers explained this observation by proposing that the stabilization of the non-bonding orbital at phosphorus through n(P) → σ*(Si-Si) hyperconjugation is more effective after metalation. [16] This is due to the higher negative partial charge at the phosphorus atom in the metallophosphasilene. As shown in the scheme below, this shielding effects is analogous to what has been observed for related alkali-metal-substituted disilenylides of the type [M(R3Si)Si=Si(SiR3)2]. [16] [21] [22] The Si=P distance in the metallophosphasilene that Driess et al. synthesized was reported to be 2.064(1) Å, which is significantly shorter than that of the "half"-parent compound (R2Si=PH) from which R2Si=PM was derived. [16] This contraction of the Si=P bond, together with a slight elongation of the Si–Si bond and a shrinkage of the P-Si-Si angle, has been rationalized by the increased hyperconjugative interactions in the R2Si=PM system. [16]

The proposed n(P) - s*(Si-Si) negative hyperconjugation in P-zinciophosphasilene by Driess et al. (left) and analogous n(Si-M) - s*(Si-Si) hyperconjugation in isoelectronic silyl-substituted alkali-metal disilenylides. Negative hyperconjugation proposed by Driess et al.tif
The proposed n(P) → σ*(Si–Si) negative hyperconjugation in P-zinciophosphasilene by Driess et al. (left) and analogous n(Si–M) → σ*(Si–Si) hyperconjugation in isoelectronic silyl-substituted alkali-metal disilenylides.

Phosphasilene with amino-substituents at Si

Resonance structures of phosphasilene with amino-substituents at Si. Amino-substituted phosphasilene.tif
Resonance structures of phosphasilene with amino-substituents at Si.

Due to the n(N)–π*(Si=P) orbital interaction, there exists strong delocalization of electron density from nitrogen to phosphorus in phosphasilene with amino-substituents at silicon. [2] Among the resonance structures shown in the figure on the left, the structure in the middle with zwitterionic structure was calculated to have a significant contribution. [2] Therefore, compared to the parent phosphasilene, the double-bond character of the Si–P is significantly reduced in phosphasilene with amino-substituents at Si, giving rise to longer Si–P bond distance.

Reactivity

Reaction with Lewis acids and bases

Due to the higher electronegativity of phosphorus and the polarized nature of the Si=P moiety, phosphasilene tend to react with Lewis acids and bases at the phosphorus atom and silicon atom, respectively. The reaction of phosphasilene with Lewis acids, which usually occurs at the Lewis-basic two-coordinate phosphorus atom, is also dependent upon the nature of the reactants. [2]

Lewis bases have been shown to react with both three-coordinate and four-coordinate silicon atoms. The coordination of Lewis bases to the three-coordinate silicon atom of phosphasilene can lead to effective stabilization of the Si=P moiety. [2] For example, it has been demonstrated that unstable phosphasilene can react with DMAP and small N-heterocyclic carbene (NHC) to afford the corresponding stable complexes. [23] In phosphasilenes stabilized by NHC, the Si=P bonds are elongated and the negative charge gets localized on the phosphorus atom. Among the resonance structures shown in the figure below, Natural Resonance Theory (NRT) indicates that the third structure with a Si–P single bond is predominant, carrying a resonance weight of 76.3%. [2] [24] Therefore, these stabilized phosphasilene can also be interpreted as a silyene-phosphinidene adduct. [2]

Resonance structures of NHC stabilized phosphasilene. The third structure with a Si-P single bond is shown to be predominant by Natural Resonance Theory calculations. Resonance structures of NHC stabilized phosphasilene.tif
Resonance structures of NHC stabilized phosphasilene. The third structure with a Si–P single bond is shown to be predominant by Natural Resonance Theory calculations.

For donor-stabilized phosphasilene with four coordinate silicon, the reaction with Lewis bases may lead to ligand exchange at the silicon atom. [25]

Metalation

Metalation of phosphasilenes gives rise to either complexes featuring the coordination of the phosphorus lone pair to a metal center [26] [3] [24] or P-metalated phosphasilenes. [27] [16] [28] In the former case, the binding of the phosphasilenes to transition metals via the phosphorus lone pair reduces the double-bond character of the Si–P bond. [2] Some examples of this type of phosphasilene transition metal complexes are shown below.

Examples of complexes with phosphasilene coordinated to transition metals via the phosphorus lone pair. Examples of complexes with phosphasilene coordinated to transition metals via the phosphorus lone pair.tif
Examples of complexes with phosphasilene coordinated to transition metals via the phosphorus lone pair.

Driess and coworkers first observed the formation of P-metalated phosphasilenes of the latter case: P-ferrio-substituted phosphasilene R2Si=P[Fe(CO)25-C5H5)] (R = 2,4,6-iPr3C6H2). [27] They further demonstrated that P-metalated phosphasilenes R2Si=PM can be obtained by metalating "half"-parent phosphasilenes, which substitutes the R2Si=PH hydrogen atom with transition metal-containing fragments. [3] [16]

Phosphinidene transfer

Phosphasilenes with significant zwitterionic characters undergoes facile hemolytic cleavage of the fragile Si=P bond. This can be utilized for the liberation and transfer of phosphinidene (:PH) to unsaturated organic molecules. [29] Driess et al. demonstrated that a fragile "half"-parent phosphasilene LSi=PH (L = CH[(C=CH2)CMe(NAr)2]; Ar = 2,6-iPr2C6H3) with highly shielded PH moiety is capable of transferring :PH to NHC. [29]

Phosphasilene as a phosphinidene (:PH) transfer agent. Phosphasilene as a Phosphinidene transfer reagent.tif
Phosphasilene as a phosphinidene (:PH) transfer agent.
The polarized Si=P p orbital (HOMO-1) of a zwitterionic "half"-parent phosphasilene prepared by Driess et al. HOMO-1 orbital of phosphasilene by Driess et al.tif
The polarized Si=P π orbital (HOMO-1) of a zwitterionic "half"-parent phosphasilene prepared by Driess et al.

Theoretical investigation by DFT (B3LYP/6-31G(d) level) revealed that this phosphasilene bears two highly localized lone pairs on the phosphorus atom due to the LSi=PH ↔ LSi–P+H resonance. Based on natural bond orbital (NBO) analysis, the σ bond of Si=P involves even contributions from Si and P, while the π bond (HOMO-1) is strongly polarized to the phosphorus atom. This indicates that the π bond between silicon and phosphorus is not predominant, supporting the significance of the zwitterionic resonance structure in the description of Si–P bonding. [29]

Small molecule activation

P4 and S8 activation by phosphasilenes reported in literature. Small molecule activation reactions of phosphasilene.tif
P4 and S8 activation by phosphasilenes reported in literature.

The Si=P moiety of phosphasilene has been reported to demonstrate small molecule activation reactivities analogous to those observed in Si=Si, P=C, and other heavier alkene analogues. For example, phosphasilene with silyl substituents has been shown to activate white phosphorus (P4) under relatively mild reaction conditions to form 1,2,3-triphospha-4-silabicyclo[1.1.0]butanes, [9] which is similar to disilenes' reactivity. [31] [2] Analogous to the behavior of phosphaalkenes, [32] phosphasilene can also activate chalcogens such as S8 and Te to form unstable three-membered ring compounds. [7]

N-H bond activation reaction of phosphasilene reported by Driess et al. Reaction of phosphasilene with NH3.tif
N-H bond activation reaction of phosphasilene reported by Driess et al.

Using a fragile zwitterionic "half-parent" phosphasilene L'Si=PH (L' = CH[(C=CH2)CMe(NAr)2], Ar = 2,6-iPr2C6H3), Driess and coworkers demonstrated that an unusual N-H activation reactivity can be achieved by the Si=P moiety in ammonia, affording L'Si(NH2)PH2 species. [23] This represent a rare example of catalyst-free 1,2-hydroamination reaction that's been reported in heavier alkene analogues. [2] [33] [34]

Related Research Articles

<span class="mw-page-title-main">Tetrahedrane</span> Hypothetical organic molecule with a tetrahedral structure

Tetrahedrane is a hypothetical platonic hydrocarbon with chemical formula C4H4 and a tetrahedral structure. The molecule would be subject to considerable angle strain and has not been synthesized as of 2021. However, a number of derivatives have been prepared. In a more general sense, the term tetrahedranes is used to describe a class of molecules and ions with related structure, e.g. white phosphorus.

<span class="mw-page-title-main">Silabenzene</span> Chemical compound

A silabenzene is a heteroaromatic compound containing one or more silicon atoms instead of carbon atoms in benzene. A single substitution gives silabenzene proper; additional substitutions give a disilabenzene, trisilabenzene, etc.

<span class="mw-page-title-main">Silylene</span> Chemical compound

Silylene is a chemical compound with the formula SiH2. It is the silicon analog of methylene, the simplest carbene. Silylene is a stable molecule as a gas but rapidly reacts in a bimolecular manner when condensed. Unlike carbenes, which can exist in the singlet or triplet state, silylene (and all of its derivatives) are singlets.

<span class="mw-page-title-main">Organotin chemistry</span> Branch of organic chemistry

Organotin chemistry is the scientific study of the synthesis and properties of organotin compounds or stannanes, which are organometallic compounds containing tin–carbon bonds. The first organotin compound was diethyltin diiodide, discovered by Edward Frankland in 1849. The area grew rapidly in the 1900s, especially after the discovery of the Grignard reagents, which are useful for producing Sn–C bonds. The area remains rich with many applications in industry and continuing activity in the research laboratory.

A transition metal carbene complex is an organometallic compound featuring a divalent organic ligand. The divalent organic ligand coordinated to the metal center is called a carbene. Carbene complexes for almost all transition metals have been reported. Many methods for synthesizing them and reactions utilizing them have been reported. The term carbene ligand is a formalism since many are not derived from carbenes and almost none exhibit the reactivity characteristic of carbenes. Described often as M=CR2, they represent a class of organic ligands intermediate between alkyls (−CR3) and carbynes (≡CR). They feature in some catalytic reactions, especially alkene metathesis, and are of value in the preparation of some fine chemicals.

Diphosphene is a type of organophosphorus compound that has a phosphorus–phosphorus double bond, denoted by R-P=P-R'. These compounds are not common but are of theoretical interest. Normally, compounds with the empirical formula RP exist as rings. However, like other multiple bonds between heavy main-group elements, P=P double bonds can be stabilized by a large steric hindrance from the substitutions. The first isolated diphosphene bis(2,4,6-tri-tert-butylphenyl)diphosphene was exemplified by Masaaki Yoshifuji and his coworkers in 1981, in which diphosphene is stabilized by two bulky phenyl group.

<span class="mw-page-title-main">Phosphaalkyne</span>

In chemistry, a phosphaalkyne is an organophosphorus compound containing a triple bond between phosphorus and carbon with the general formula R-C≡P. Phosphaalkynes are the heavier congeners of nitriles, though, due to the similar electronegativities of phosphorus and carbon, possess reactivity patterns reminiscent of alkynes. Due to their high reactivity, phosphaalkynes are not found naturally on earth, but the simplest phosphaalkyne, phosphaethyne (H-C≡P) has been observed in the interstellar medium.

In chemistry, the double bond rule states that elements with a principal quantum number (n) greater than 2 for their valence electrons (period 3 elements and higher) tend not to form multiple bonds (e.g. double bonds and triple bonds). The double bonds, when they exist, are often weak due to poor orbital overlap between the n>2 orbitals of the two atoms. Although such compounds are not intrinsically unstable, they instead tend to polymerize. An example is the rapid polymerization that occurs upon condensation of disulfur, the heavy analogue of O2. Numerous violations to the rule exist.

Carbene analogs in chemistry are carbenes with the carbon atom replaced by another chemical element. Just as regular carbenes they appear in chemical reactions as reactive intermediates and with special precautions they can be stabilized and isolated as chemical compounds. Carbenes have some practical utility in organic synthesis but carbene analogs are mostly laboratory curiosities only investigated in academia. Carbene analogs are known for elements of group 13, group 14, group 15 and group 16.

<span class="mw-page-title-main">Germylene</span> Class of germanium (II) compounds

Germylenes are a class of germanium(II) compounds with the general formula :GeR2. They are heavier carbene analogs. However, unlike carbenes, whose ground state can be either singlet or triplet depending on the substituents, germylenes have exclusively a singlet ground state. Unprotected carbene analogs, including germylenes, has a dimerization nature. Free germylenes can be isolated under the stabilization of steric hindrance or electron donation. The synthesis of first stable free dialkyl germylene was reported by Jutzi, et al in 1991.

<span class="mw-page-title-main">Phosphinidene</span> Type of compound

Phosphinidenes are low-valent phosphorus compounds analogous to carbenes and nitrenes, having the general structure RP. The "free" form of these compounds is conventionally described as having a singly-coordinated phosphorus atom containing only 6 electrons in its valence level. Most phosphinidenes are highly reactive and short-lived, thereby complicating empirical studies on their chemical properties. In the last few decades, several strategies have been employed to stabilize phosphinidenes, and researchers have developed a number of reagents and systems that can generate and transfer phosphinidenes as reactive intermediates in the synthesis of various organophosphorus compounds.

<i>tert</i>-Butylphosphaacetylene Chemical compound

tert-Butylphosphaacetylene is an organophosphorus compound. Abbreviated t-BuCP, it was the first example of an isolable phosphaalkyne. Prior to its synthesis, the double bond rule had suggested that elements of Period 3 and higher were unable to form double or triple bonds with lighter main group elements because of weak orbital overlap. The synthesis of t-BuCP discredited much of the double bond rule and opened new studies into the formation of unsaturated phosphorus compounds.

Hydrophosphination is the insertion of a carbon-carbon multiple bond into a phosphorus-hydrogen bond forming a new phosphorus-carbon bond. Like other hydrofunctionalizations, the rate and regiochemistry of the insertion reaction is influenced by the catalyst. Catalysts take many forms, but most prevalent are bases and free-radical initiators. Most hydrophosphinations involve reactions of phosphine (PH3).

<span class="mw-page-title-main">Digermyne</span> Class of chemical compounds

Digermynes are a class of compounds that are regarded as the heavier digermanium analogues of alkynes. The parent member of this entire class is HGeGeH, which has only been characterized computationally, but has revealed key features of the whole class. Because of the large interatomic repulsion between two Ge atoms, only kinetically stabilized digermyne molecules can be synthesized and characterized by utilizing bulky protecting groups and appropriate synthetic methods, for example, reductive coupling of germanium(II) halides.

<span class="mw-page-title-main">Diphosphagermylene</span> Class of compounds

Diphosphagermylenes are a class of compounds containing a divalent germanium atom bound to two phosphorus atoms. While these compounds resemble diamidocarbenes, such as N-heterocyclic carbenes (NHC), diphosphagermylenes display bonding characteristics distinct from those of diamidocarbenes. In contrast to NHC compounds, in which there is effective N-C p(π)-p(π) overlap between the lone pairs of planar nitrogens and an empty p-orbital of a carbene, systems containing P-Ge p(π)-p(π) overlap are rare. Until 2014, the geometry of phosphorus atoms in all previously reported diphosphatetrylenes are pyramidal, with minimal P-Ge p(π)-p(π) interaction. It has been suggested that the lack of p(π)-p(π) in Ge-P bonds is due to the high energetic barrier associated with achieving a planar configuration at phosphorus, which would allow for efficient p(π)-p(π) overlap between the phosphorus lone pair and the empty P orbital of Ge. The resulting lack of π stabilization contributes to the difficulty associated with isolating diphosphagermylene and the Ge-P double bonds. However, utilization of sterically encumbering phosphorus centers has allowed for the isolation of diphosphagermylenes with a planar phosphorus center with a significant P-Ge p(π)-p(π) interaction.

<span class="mw-page-title-main">Trisilaallene</span> Class of silicon chemical compounds

Trisilaallene is a subclass of silene derivatives where a central silicon atom forms double bonds with each of two terminal silicon atoms, with the generic formula R2Si=Si=SiR2. Trisilaallene is a silicon-based analog of an allene, but their chemical properties are markedly different.

<i>N</i>-heterocyclic silylene Chemical compound

An N-Heterocyclic silylene (NHSi) is an uncharged heterocyclic chemical compound consisting of a divalent silicon atom bonded to two nitrogen atoms. The isolation of the first stable NHSi, also the first stable dicoordinate silicon compound, was reported in 1994 by Michael Denk and Robert West three years after Anthony Arduengo first isolated an N-heterocyclic carbene, the lighter congener of NHSis. Since their first isolation, NHSis have been synthesized and studied with both saturated and unsaturated central rings ranging in size from 4 to 6 atoms. The stability of NHSis, especially 6π aromatic unsaturated five-membered examples, make them useful systems to study the structure and reactivity of silylenes and low-valent main group elements in general. Though not used outside of academic settings, complexes containing NHSis are known to be competent catalysts for industrially important reactions. This article focuses on the properties and reactivity of five-membered NHSis.

<span class="mw-page-title-main">Plumbylene</span> Divalent organolead(II) analogues of carbenes

Plumbylenes (or plumbylidenes) are divalent organolead(II) analogues of carbenes, with the general chemical formula, R2Pb, where R denotes a substituent. Plumbylenes possess 6 electrons in their valence shell, and are considered open shell species.

1-Phosphaallenes is are allenes in which the first carbon atom is replaced by phosphorus, resulting in the structure: -P=C=C<.

Negative hyperconjugation is a theorized phenomenon in organosilicon compounds, in which hyperconjugation stabilizes or destabilizes certain accumulations of positive charge. The phenomenon explains corresponding peculiarities in the stereochemistry and rate of hydrolysis.

References

  1. Driess, M. (1994). "Some Aspects of the Chemistry of Silylidene-Phosphanes and -Arsanes". Coord. Chem. Rev. 145: 1–25. doi:10.1016/0010-8545(95)90212-0.
  2. 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 Nesterov, V.; Breit, N.; Inoue, S. (2017). "Advances in Phosphasilene Chemistry". Chem. Eur. J. 23 (50): 12014–12039. doi: 10.1002/chem.201700829 . PMID   28379639.
  3. 1 2 3 4 5 Hansen, K.; Szilvási, T.; Blom, B.; Driess, M. (2015). "Transition Metal Complexes of a "Half-Parent" Phosphasilene Adduct Representing Silylene→Phosphinidene→Metal Complexes". Organometallics. 34 (24): 5703–5708. doi:10.1021/acs.organomet.5b00772.
  4. 1 2 3 4 5 Smit, C. N.; Lock, F. M.; Bickelhaupt, F. (1984). "2,4,6-Tri-tert-butylphenylphosphene-Dimesitylsilene; The First Phosphasilaalkene". Tetrahedron Lett. 25 (28): 3011–3014. doi:10.1016/S0040-4039(01)81351-4.
  5. Bender, H. R. G.; Niecke, E.; Nieger, M. (1993). "The First X-ray Structure of a Phosphasilene: 1,3,4-Triphospha-2-silabutene-(1)". J. Am. Chem. Soc. 115 (8): 3314–3315. doi:10.1021/ja00061a034.
  6. 1 2 Smit, C. N.; Bickelhaupt, F. (1987). "Phosphasilenes: synthesis and spectroscopic characterization". Organometallics. 6 (6): 1156–1163. doi:10.1021/om00149a004.
  7. 1 2 3 van den Winkel, Y.; Bastiaans, H. M. M.; Bickelhaupt, F. (1991). "Phosphasilene synthesis and reactivity: an improved route to 1-(2,4,6-tri-tert-butylphenyl)-2-tert-butyl-2-(2,4,6-tri-isopropylphenyl)phosphasilene". J. Organomet. Chem. 405 (2): 183–194. doi:10.1016/0022-328X(91)86271-Q.
  8. 1 2 Lange, D.; Klein, E.; Bender, H.; Niecke, E.; Nieger, M. Pietschnig, R. (1998). "1,3-Diphospha-2-silaallylic Lithium Complexes and Anions: Synthesis, Crystal Structures, Reactivity, and Bonding Properties". Organometallics. 17 (12): 2425–2432. doi:10.1021/om9707937.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  9. 1 2 3 Driess, M. (1991). "1,2,3‐Triphospha‐4‐silabicyclo[l.1.0]butanes from Activated, Stable Phosphasilenes and White Phosphorus". Angew. Chem. Int. Ed. Engl. 30 (8): 1022–1024. doi:10.1002/anie.199110221.
  10. 1 2 Driess, M.; Pritzkow, H.; Rell, S.; Winkler, U. (1996). "Synthesis and Unusual Reactivity of Compounds Containing Silicon−Phosphorus and Silicon−Arsenic Double Bonds: New Silylidenephosphanes and -arsanes of the Type R2SiE(SiR3) (E = P, As)". Organometallics. 15 (7): 1845–1855. doi:10.1021/om9508851.
  11. Driess, M.; Rell, S.; Pritzkow, H. (1995). "First structural characterization of silicon–arsenic and silicon–phosphorus multiple bonds in silylated silylidene-arsanes and -phosphanes; X-ray structure of a tellura-arsasilirane derivative". J. Chem. Soc., Chem. Commun. 15 (2): 253–254. doi:10.1039/C39950000253.
  12. 1 2 Inoue, S.; Wang, W.; Präsang, C.; Asay, M.; Irran, E.; Driess, M. (2011). "An Ylide-like Phosphasilene and Striking Formation of a 4π-Electron, Resonance-Stabilized 2,4-Disila-1,3-diphosphacyclobutadiene". J. Am. Chem. Soc. 133 (9): 2868–2871. doi:10.1021/ja200462y. PMID   21322569.
  13. 1 2 Avakyan, V. G.; Sidorkin, V. F.; Belogolova, E. F.; Guselnikov, S. L.; Gusel’nikov, L. E. (2006). "AIM and ELF Electronic Structure/G2 and G3 π-Bond Energy Relationship for Doubly Bonded Silicon Species, H2SiX (X = E14H2, E15H, E16)". Organometallics. 25 (26): 6007–6013. doi:10.1021/om0605478.
  14. 1 2 3 Glatthaar, J.; Maier, G. (2004). "Reaktion von atomarem Silicium mit Phosphan: eine matrix‐spektroskopische Studie". Angew. Chem. 116 (10): 1314–1317. Bibcode:2004AngCh.116.1314G. doi:10.1002/ange.200353111.
  15. Driess, M.; Janoschek, R. (1994). "A comparison of silylidene-amines, -phosphanes and -arsanes: syntheses and quantum chemical calculations". J. Mol. Struct. 313 (1): 129–139. doi:10.1016/0166-1280(94)85036-4.
  16. 1 2 3 4 5 6 7 8 9 10 11 12 13 14 Driess, M.; Block, S.; Brym, M.; Gamer, M. T. (2006). "Synthesis of a "half"-parent phosphasilene R2Si=PH and its metalation to the corresponding P zinciophosphasilene [R2Si=PM]". Angew. Chem. Int. Ed. 45 (14): 2293–2296. doi:10.1002/anie.200504145. PMID   16518791.
  17. Dykema, K. J.; Truong, T. N.; Gordon, M. S. (1985). "Studies of silicon-phosphorus bonding". J. Am. Chem. Soc. 107 (15): 4535–4541. doi:10.1021/ja00301a026.
  18. 1 2 3 4 5 6 7 Li, B.; Matsuo, T.; Hashizume, D.; Fueno, H.; Tanaka, K.; Tamao, K. (2009). "π-Conjugated Phosphasilenes Stabilized by Fused-Ring Bulky Groups". J. Am. Chem. Soc. 131 (37): 13222–13223. doi:10.1021/ja9051153. PMID   19715271.
  19. 1 2 3 4 5 Lee, V. Y.; Kawai, M.; Sekiguchi, A.; Ranaivonjatovo, H.; Escudié, J. (2009). "A "Push-Pull" Phosphasilene and Phosphagermene and Their Anion-Radicals". Organometallics. 28 (15): 4262–4265. doi:10.1021/om900310u.
  20. Ramsey, N. F. (1950). "Magnetic Shielding of Nuclei in Molecules". Phys. Rev. 78 (6): 699–703. Bibcode:1950PhRv...78..699R. doi:10.1103/PhysRev.78.699.
  21. 1 2 Inoue, S.; Ichinohe, M.; Sekiguchi, A. (2005). "Disilenyl Anions Derived from Reduction of Tetrakis(di-tert-butylmethylsilyl)disilene with Metal Naphthalenide through a Disilene Dianion Intermediate: Synthesis and Characterization". Chem. Lett. 34 (11): 1564–1565. doi:10.1246/cl.2005.1564.
  22. 1 2 Ichinohe, M.; Sanuki, K.; Inoue, S.; Sekiguchi, A. (2004). "Disilenyllithium from Tetrasila-1,3-butadiene: A Silicon Analogue of a Vinyllithium". Organometallics. 23 (13): 3088–3090. doi:10.1021/om040056s.
  23. 1 2 3 Hansen, K.; Szilvási, T.; Blom, B.; Irran, E. Driess, M. (2014). "A Donor-Stabilized Zwitterionic "Half-Parent" Phosphasilene and its Unusual Reactivity towards Small Molecules". Chem. Eur. J. 20 (7): 1947–1956. doi:10.1002/chem.201303906. PMID   24436015.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  24. 1 2 3 4 Hansen, K.; Szilvási, T.; Blom, B.; Driess, M. (2015). "A Persistent 1,2‐Dihydrophosphasilene Adduct". Angew. Chem. Int. Ed. 54 (50): 15060–15063. doi: 10.1002/anie.201508149 . PMID   26483284.
  25. Dhara, D.; Mandal, D.; Maiti, A.; Yildiz, C. B.; Kalita, P.; Chrysochos, N.; Schulzke, C.; Chandrasekhar, V.; Jana, A. (2016). "Assembly of NHC-stabilized 2-hydrophosphasilenes from Si(IV) precursors: a Lewis acid–base complex". Dalton Trans. 45 (48): 19290–19298. doi:10.1039/c6dt04321f. PMID   27872933.
  26. 1 2 Li, B.; Matsuo,, T.; Fukunaga, T.; Hashizume, D.; Fueno, H.; Tanaka, K.; Tamao, K. (2011). "Neutral and Cationic Gold(I) Complexes with π-Conjugated Phosphasilene Ligands". Organometallics. 30 (13): 3453–3456. doi:10.1021/om2003442.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  27. 1 2 Driess, M.; Pritzkow, H.; Winkler, U. (1997). "Synthesis of (fluorosilyl)phosphido iron and nickel complexes and ferriosilyl-substituted (fluorosilyl)phosphanes and evidence for the formation of a metallo phosphasilene of the type [M]-P=SiR2". J. Organomet. Chem. 529 (1–2): 313–321. doi:10.1016/S0022-328X(96)06483-2.
  28. Yao, S.; Block, S.; Brym, M.; Driess, M. (2007). "A new type of heteroleptic complex of divalent lead and synthesis of the P-plumbyleniophosphasilene, R2Si=P–Pb(L): (L = β-diketiminate)". Chem. Commun. 37 (37): 3844–3846. doi:10.1039/b710888e. PMID   18217666.
  29. 1 2 3 4 5 Hansen, K.; Szilvási, T.; Blom, B.; Inoue, S.; Epping, J.; Driess, M. (2013). "A Fragile Zwitterionic Phosphasilene as a Transfer Agent of the Elusive Parent Phosphinidene (:PH)". J. Am. Chem. Soc. 135 (32): 11795–11798. doi:10.1021/ja4072699. PMID   23895437.
  30. van den Winkel, Y.; Bastians, H. M. M.; Bickelhaupt, F. (1991). "Phosphasilene synthesis and reactivity: an improved route to 1-(2,4,6-tri-tert-butylphenyl)-2-tert-butyl-2- (2,4,6-tri-isopropylphenyl)phosphasilene". J. Organomet. Chem. 405 (2): 183–194. doi:10.1016/0022-328X(91)86271-Q.
  31. Scheer, M.; Balázs, G.; Seitz, A. (2010). "P4 Activation by Main Group Elements and Compounds". Chem. Rev. 110 (7): 4236–4256. doi:10.1021/cr100010e. PMID   20438122.
  32. Yoshifuji, M.; Toyota, K.; Ando, K.; Inamoto, N. (1984). "Isolation and Characterization of a Stable Dithioxophosphorane". Chem. Lett. 13 (3): 317–318. doi: 10.1246/cl.1984.317 .
  33. Zhu, Z.; Wang, X.; Peng, Y.; Lei, H.; Fettinger, J. C.; Rivard, E.; Power, P. P. (2009). "Addition of Hydrogen or Ammonia to a Low‐Valent Group 13 Metal Species at 25 °C and 1 Atmosphere". Angew. Chem. Int. Ed. 48 (11): 2031–2034. doi:10.1002/anie.200805982. PMID   19180617.
  34. Boomgaarden, S.; Saak, W.; Weidenbruch, M.; Marsmann, H. (2001). "Ammonia and Chlorine Additions to a Tetrasilabuta‐1,3‐diene: Conglomerate versus Racemate Crystallization". Z. Anorg. Allg. Chem. 627 (3): 349–352. doi:10.1002/1521-3749(200103)627:3<349::AID-ZAAC349>3.0.CO;2-R.