Pnictogen-substituted tetrahedranes

Last updated

Pnictogen-substituted tetrahedranes are pnictogen-containing analogues of tetrahedranes with the formula RxCxPn4-x (Pn = N, P, As, Sb, Bi). [1] Computational work has indicated that the incorporation of pnictogens to the tetrahedral core alleviates the ring strain of tetrahedrane. [2] Although theoretical work on pnictogen-substituted tetrahedranes has existed for decades, only the phosphorus-containing species have been synthesized. [3] [4] [5] These species exhibit novel reactivities, most often through ring-opening and polymerization pathways. Phosphatetrahedranes are of interest as new retrons for organophosphorus chemistry. Their strain also make them of interest in the development of energy-dense compounds.

Contents

History

The first synthetic tetrahedral molecule, tetra-tert-butyltetrahedrane (tBu4C4) was reported in 1978 by Maier and coworkers [6] following the synthesis of other Platonic solid species, like cubane and dodecahedrane. The tert-butyl substituents were used to encumber the tetrahedral core and quell the radical-mediated ring-opening of an otherwise kinetically stable but thermodynamically strained molecule via the corset effect. As of 2023, the unencumbered tetrahedrane (H4C4) has yet to be synthesized.

The substitution of carbons in the tetrahedral core of tetrahedrane with pnictogens to stabilize the highly strained molecule has been suggested due to the known existence of elemental tetrahedral pnictogens. Notably, white phosphorus, the most stable allotrope of phosphorus, is tetrahedral with the molecular formula P4. [7] Arsenic can also exist as a metastable tetrahedral allotrope, As4, known as yellow arsenic. Furthermore, mixed tetrahedral pnictogen molecules have been synthesized, such as AsP3 [8] and, more recently, (PbBi3)-. [9] Elements on the extreme ends of the pnictogen family have not yet been observed in a tetrahedral Pn4 configuration, however. Nitrogen's orbitals lack diffusivity, and bismuth’s orbitals undergo minimal hybridization due to relativistic contraction.

Computational studies into mixed pnictogen-tetrel tetrahedranes have suggested that pnictogen-substituted tetrahedranes are more stable than their all tetrel counterparts decades before their first synthesis. In 1990, calculations on azatetrahedranes suggested positive correlation between the number of nitrogens in the tetrahedral core and thermodynamic stability. [2] In the same vein, calculations done in 2010 on pnictacubanes suggested positive correlation between the number of phosphoruses in the cuboidal core and the thermodynamic stability. [10]

All currently synthesized pnictogen-substituted tetrahedranes in the scientific literature as of 2023. All Currently Synthesized Phosphatetrahedranes.png
All currently synthesized pnictogen-substituted tetrahedranes in the scientific literature as of 2023.

In 2019, Wolf and coworkers synthesized the first pnictogen-substituted tetrahedrane: di-tert-butyldiphosphatetrahedrane (tBu2C2P2), produced from the reaction of nickel catalyst with phosphaalkynes. [4] Shortly thereafter, in 2020, Cummins and coworkers announced that they had synthesized tri-tert-butylmonophosphatetrahedrane (tBu3C3P) from a phosphorus-containing anthracene derivative. [3] In 2021, Cummins and coworkers published the synthesis of triphosphatetrahedrane (HCP3), completing the set of tetrahedral molecules with carbon- and phosphorus-containing cores. [5]

Phosphatetrahedrane Synthesis

Despite the presentation of phosphatetrahedranes as a series of incrementally changing tetrahedrane derivatives, their syntheses are vastly different. [7]

Tri-tert-butylmonophosphatetrahedrane

The original synthesis of tBu3C3P reported by Cummins and coworkers in 2020 begins with the phosphorus derivative of anthracene. The addition of sodium hexamethyldisilazide deprotonates the phosphorus followed by triphenylborane then bonding to the phosphorus. The anionic phosphorus-containing intermediate then forms an ionic interaction with the sodium cation in solution. The addition of tri-tert-butyl cyclopropenium ion produces the thermally stable cyclopropenyl phosphine intermediate. Upon irradiation with 254 nm light in the presence of triflic acid and either tetra-n-butyl ammonium chloride or tetramethylammonium fluoride, the anthracene leaving group is forced out, leaving a halogenated phosphine. Upon addition of lithium tetramethylpiperidide with heating, tBu3C3P and lithium halide salt is generated. [3]

Synthesis of tri-tert-butylmonophosphatetrahedrane with anthracene leaving group. Synthesis of Monophosphatetrahedrane.png
Synthesis of tri-tert-butylmonophosphatetrahedrane with anthracene leaving group.

An improved version of tri-tert-butylmonophosphatetrahedrane synthesis, where the anthracene is replaced by two trimethylsilyl groups, was reported by Cummins and coworkers a year later. To replace one of the trimethylsilyl groups with a chloride, hexachloroethane is added, generating trimethylsilyl chloride and tetrachloroethylene as byproducts. Tetramethylammonium fluoride is then added to remove the remaining trimethylsilyl group and the chloride, generating the tBu3C3P. This route has a tBu3C3P yield of 33% [11] as compared to the original route's 19%. [3]

Di-tert-butyldiphosphatetrahedrane

In 2019, Wolf and coworkers reported the synthesis of tBu2C2P2 through the use of a metal catalyst. Ni(IPr)(CO)3, upon addition of 1 equivalent of tert-butylphosphaacetylene (tBuCP), loses two carbon monoxide ligands. The addition of a second equivalent of tBuCP generates the 1,3-diphosphacyclobutadiene ligand, now binding with η4 hapticity. Density functional theory calculations into the catalytic cycle suggest that the 1,3-diphosphacyclobutadiene isomerizes into the desired tetrahedrane. Upon addition of a final tBuCP, (tBu2C2P2) is released and the catalytic cycle can begin again. [4]

Synthesis of di-tert-diphosphatetrahedrane. Diphosphatetrahedrane Synthesis.png
Synthesis of di-tert-diphosphatetrahedrane.

Triphosphatetrahedrane

Cummins and coworkers reported the synthesis of HCP3 in 2021. Due to the similarity of HCP3 to AsP3, the [NbII(ODipp)3(P3)]- previously shown to be a retron for AsP3 was used for the synthesis of HCP3. To add a -CH group to [P3]3-, bromodichloromethane undergoes halogen abstraction, leaving a carbon-centered radical. The niobium complex then undergoes P3 transfer to yield HCP3. The use of bromodichloromethyl trimethylsilane instead of bromodichloromethane in this process yields trimethylsilyl triphosphatetrahedrane ((Me3S)CP3). [5] [12]

Synthesis of triphosphatetrahedrane. Triphosphatetrahedrane Synthesis.png
Synthesis of triphosphatetrahedrane.

Reactivity

Tri-tert-butylmonophosphatetrahedrane

Lewis Acid-Induced Reactions

Addition of W(CO)5(THF) to tBu3C3P generates a phosphorus-containing housene analogue. [3]

The addition of 0.2 equivalents of triphenylborane in benzene can produce several cycloadducts. In the absence of exogenous reagents, tBu3C3P dimerizes into a ladderane-like compound with a P-P bond. In the presence of excess styrene or an atmosphere of ethylene, [4 + 2] cycloadditions occur to give 1-phosphabicyclo[2.2.0]hexenes. [11]

Lewis acid-induced reactions of triphosphatetrahedranes. Triphosphatetrahedrane Lewis Acid-Induced Reactions.png
Lewis acid-induced reactions of triphosphatetrahedranes.

Silylene Reaction

The cage opening of tBu3C3P can be induced by PhC(NtBu)2SiN(SiMe3)2 over the course of 24 hours to generate the dark red phosphasilene PhC(NtBu)2Si=P(tBu3C3). [13]

Monophosphatetrahedrane reaction with silylene. Monophosphatetrahedrane Silylene Reaction.png
Monophosphatetrahedrane reaction with silylene.

Ylide Reaction

Reaction of tBu3C3P with the ylide Ph3P=CH2 over 48 hours and with heat induces cage opening in the same manner as the silylene reaction to generate H2C=P(tBuC)3. Reaction of this product with tBu3C3P generates the symmetric product (tBuC)3P(C)P(tBuC)3. [13]

Monophosphatetrahedrane reaction with Ph3P=CH2 ylide. Monophosphatetrahedrane Ylide Reaction.png
Monophosphatetrahedrane reaction with Ph3P=CH2 ylide.

Formation of Phosphirane

tBu3C3P is a retron for phosphirane synthesis. Upon reaction with Ni(COD)2 (COD = cycloocta-1,5-diene) catalyst in triisopropylphosphine, cage opening occurs. Like the silylene and ylide reactions, the phosphorus bridges the (tBuC)3 and the alkene components. The phosphate undergoes cycloaddition with the double bond to form the phosphirane moiety. This reaction pathway has been demonstrated for styrene, ethylene, and neohexene. Furthermore, this reaction pathway is also capable of synthesizing vinyl-substituted phosphirane as evidenced by tBu3C3P and cyclohexa-1,3-diene. [13]

Formation of phosphiranes from monophosphatetrahedrane. Monophosphatetrahedrane Phosphirane Formation Reactions.png
Formation of phosphiranes from monophosphatetrahedrane.

Ligand Substitution

tBu3C3P can be used to replace the ethylene ligand of (Ph3P)Pt(C2H4) in melting THF. [11]

Monophosphatetrahedrane substitutes an ethylene ligand. Monophosphatetrahedrane Ligand Substitution with Ethylene.png
Monophosphatetrahedrane substitutes an ethylene ligand.

Di-tert-butyldiphosphatetrahedrane

Dimerization Reactions

Above the melting point of tBu2C2P2 (–32 °C), tBu2C2P2 dimerizes into another ladderane-like structure but it is prone to decomposition. This reaction can be hampered by keeping tBu2C2P2 under its melting point and/or by keeping the tBu2C2P2 concentration low. [4]

tBu2C2P2 can also be dimerized using nickel complexes to form a variety of exotic structures. tBu2C2P2 reacted with 1 equivalent of Ni(CpR)(IPr) (IPr = 1,3-bis(2,6-diisopropylphenyl)imidazolin-2-ylidene, R = H, CH3, 4-(CH3CH2)-C6H4) generates 0.5 equivalent of a tetracyclo-compound. Upon addition of another equivalent of the same nickel complex, a butterfly-like geometry is adopted, with two nickel atoms coordinated to opposite phosphorus atoms and two coordinated to adjacent phosphorus atoms on different four membered rings. This butterfly-structured compound is a dark red color. The reaction to the butterfly structure is believed to depend on kinetic access to the middle P-P bond. Bulky substituents on CpR kinetically hinder the P-P bond cleavage and transformation into the butterfly-structured product.

Stepwise dimerization of diphosphatetrahedrane with Ni(Cp)(IPr). Diphosphatetrahedrane Dimerization.png
Stepwise dimerization of diphosphatetrahedrane with Ni(Cp)(IPr).

tBu2C2P2 can also be reacted with Ni(IMes)2 (IMes = 1,3-bis(2,4,6-trimethylphenyl)imidazolin-2-ylidene) in toluene to produce 0.5 equivalents of an asymmetric compound with two hexacoordinate Ni atoms, a Ni-Ni bond, and two weak P-P interactions. This product is an intermediate for further chemistry. Heating the product at 60 °C for 3 hours causes the expulsion of a di-tert-butylacetylene and the reformation of P-P bonds. Another reaction pathway involves the addition of 3 equivalents of CO to the product, leading to the production of Ni(IMes)(CO)3 and the ladderane-like compound described at the beginning of this section. A third reaction pathway involves the addition of hexachloroethane. This produces a 1,2-diphosphocyclobutadiene ring (vide supra) that is coordinated to both nickel atoms. This third reaction pathway also produces the ladderane analogue. [14]

Di-tert-butyl-diphosphatetrahedrane reaction with Ni(IMes)2 and further reactions. Diphosphatetrahedrane Dimerization with NHCs.png
Di-tert-butyl-diphosphatetrahedrane reaction with Ni(IMes)2 and further reactions.

Ligand Substitution

In solution with coordination complexes, tBu2C2P2 can cause ligand substitution. Most of these reactions cause cage-opening. The reaction of tBu2C2P2 with [K([18]crown-6)][Fe(anthracene)2] in toluene and THF causes the expulsion of one anthracene and the cage opening of tBu2C2P2 to form the replacement 1,2-diphosphacyclobutadiene ligand. This cage opening is due to P-C bond cleavage. Mono-ligand substitution is also observed in the reaction of tBu2C2P2 with [(DippBIAN)M(COD)] (Dipp = 2,6-diisopropylphenyl, BIAN = bis(arylimino)acenaphthene, M = Fe, Co). The cobalt product, upon reaction with Cy2PCl (Cy = cyclohexyl), forms a 1,2,3-triphospholium ligand. tBu2C2P2 can also displace the toluene in Ni(toluene)(IPr). Toluene is replaced by the aforementioned ladderane analogue in a η2-fashion. [15]

Diphosphatetrahedrane ligand substitution of anthracene forms 1,2-diphosphacyclobutadiene analogue. Diphosphatetrahedrane Ligand Substitution of Anthracene Forming 1,2-Diphosphacyclobutadiene Analogue.png
Diphosphatetrahedrane ligand substitution of anthracene forms 1,2-diphosphacyclobutadiene analogue.
Diphosphatetrahedrane ligand substitution of cycloocta-1,5-diene forms triphosphacyclopentane analogue. Diphosphatetrahedrane Ligand Substitution of Cycloocta-1,5-diene Forming Triphosphacyclopentane Analogue.png
Diphosphatetrahedrane ligand substitution of cycloocta-1,5-diene forms triphosphacyclopentane analogue.
Diphosphatetrahedrane ligand substitution of toluene forms ladderane analogue. Diphosphatetrahedrane Ligand Substitution of Toluene Forming Ladderane Analogue.png
Diphosphatetrahedrane ligand substitution of toluene forms ladderane analogue.

Double-ligand substitution is seen in the reaction of tBu2C2P2 with [Co(COD)2][K(THF)0.2]. The major product has its ligands doubly substituted by 1,2-diphosphacyclobutadiene. A minor product with double substitution by 1,3-diphosphacyclobutadiene ligand is also observed. However, the cobalt complex with one 1,2-diphosphacyclobutadiene ligand and one 1,3-diphosphacyclobutadiene ligand is not observed; this is likely due to steric clash between the tert-butyl substituents. The preference for 1,2-diphosphacyclobutadiene makes tBu2C2P2 a potentially valuable retron as phosphalkynes are known to produce 1,3-diphosphacyclobutadiene. [15]

Diphosphatetrahedrane ligand substitution of cycloocta-1,5-diene forms diphosphacyclobutadiene analogues. Diphosphatetrahedrane Ligand Substitution of Cycloocta-1,5-diene Forming Diphosphacyclobutadiene Analogues.png
Diphosphatetrahedrane ligand substitution of cycloocta-1,5-diene forms diphosphacyclobutadiene analogues.

Only two of tBu2C2P2's ligand substitution reactions are known to preserve the tetrahedral cage. Reacting (pftb)[Ag(CH2Cl2)2] (pftb = Al[PFTB]- = Al[OC(CF3)3]4-) with tBu2C2P2 in lightless conditions leads to the generation of a disilver complex wherein each of the two tBu2C2P2 ligates to one silver atom and each of the two ladderane analogues (vida infra) ligates to both silver atoms. [4] By reacting Ni(CO)4 in THF at –80 °C and in the absence of light, three tBu2C2P2 molecules coordinate to Ni in an η2-fashion. Calculations by Intrinsic Bond Orbital (IBO) theory suggest that the coordination occurs through a 3-center-2-electron bond. [16]

Diphosphatetrahedrane ligand substitution with [Ag(C2H2Cl2)2](pftb). Diphosphatetrahedrane Ligand Substitution with Silver Complex.png
Diphosphatetrahedrane ligand substitution with [Ag(C2H2Cl2)2](pftb).
Reaction of diphosphatetrahedrane with Ni(CO)4 yields intact diphosphatetrahedrane ligands. Intact Diphosphatetrahedrane Ligand Substitution.png
Reaction of diphosphatetrahedrane with Ni(CO)4 yields intact diphosphatetrahedrane ligands.

Reactions with N-Heterocyclic Carbenes

tBu2C2P2 can be used a retron to form phosphirenes or phosphaalkenes with the addition of 1 equivalent or 2 equivalents of N-heterocyclic carbenes (NHC), respectively. Upon the addition of 1 equivalent of IPr, IMes, or MesDAC (1,3-bis(2,4,6-trimethylphenyl)diamidocarbene), tBu2C2P2 undergoes ring opening at one phosphorus atom's P-C bonds, creating structures with a bridging P-P bond between the NHC and the phosphirene. IBO calculations and crystallographic evidence support the assignment of double bonding to the P=C bond to resultant molecule. This reaction is very slow, taking several weeks to reach completion.

Upon the addition of 2 equivalents of TMC (2,3,4,5-tetramethylimidazolin-2-ylidene) in benzene, both phosphorus atoms bond to THC. The P-P bond is broken. A double bond also forms between the two carbons of tBu2C2P2, generating a phosphaalkene. This reaction happens significantly faster, with a reported speed of 1 hour.

Selectivity between the two reactions is suggested to be achieved by changing the steric bulk of the NHC used. A bulky NHC should prefer generating a phosphirene, whereas a smaller NHC should prefer generating a phosphaalkene. [17]

Diphosphatetrahedrane reactivity with NHCs is controlled by sterics. Diphosphatetrahedrane Reactions with NHCs.png
Diphosphatetrahedrane reactivity with NHCs is controlled by sterics.

Triphosphatetrahedrane

Reaction with (dppe)Fe(Cp*)Cl

HCP3, upon addition of [(dppe)Fe(Cp*)]Cl (dppe = 1,2-bis-(diphenylphosphino)ethane) in sodium tetraphenylborate and THF, undergoes salt metathesis and produces [(dppe)Fe(Cp*)(HCP3)][BPh4]. This product is crystallizable, producing purple crystals. This product is prone to decomposition back to HCP3. [5]

Triphosphatetrahedrane reaction with (dppe)Fe(Cp*)Cl. Triphosphatetrahedrane Reaction with Fe Complex.png
Triphosphatetrahedrane reaction with (dppe)Fe(Cp*)Cl.

Theoretical Work

Azatetrahedranes

Bonding Parameters

Ring and cage strain results in poor angular overlap of orbitals, leading to non-linear bonding. Due to the interest in tetrahedrane and their azatetrahedral analogues as highly strained molecules, Politzer and Seminario introduced the "bond deviation index" to determine the deviation of the bond path — defined as the path following maxima between nuclei — and the linear bond between the nuclei.

The strain of experienced by H4C4 is calculated to be partially alleviated upon the substitution of carbon atoms by nitrogen atoms. [18] The bond deviation index decreases with the number of nitrogens in azatetrahedranes from 0.114 to 0.087 from 0 to 2 nitrogen atoms. [19] The engendered stability is countered by the propensity of the N-N bond in the strained system to escape as dinitrogen. This is evidenced by the calculated bond length: 1.59 Å, which is much higher than that of aromatic N-N bonds: 1.21-1.36 Å. [18] [19]

On the basis of the higher electronegativity of nitrogen than carbon, azatetrahedranes have less negative electrostatic potentials at their C-C bonds than H4C4, leading to greater stability against electrophilic attacks. [18] C-C and N-C bonds both contract in azatetrahedranes. [18] This contraction leads to smaller bond deviations and stronger bonding, [20] although the latter is dependent on the distance from the nitrogen atom(s). [19] Azatetrahedranes also seem to distort from an ideal tetrahedral symmetry. These trends directly correlates with the number of nitrogen atoms. Nitration also contracts the tetrahedral core. However, essentially thermodynamically neutral nitro group rotation leads to small amounts of C-N lengthening, [21] weaking the interaction and localizing the electrons.

Ionization of all azatetrahedranes cores in the series led to cage-opening at the G4MP2 and G4 levels of theory. [22]

Thermodynamics

Early work on azatetrahedranes has also utilized isodesmic reactions — aphysical reactions where compounds are changing but bond types are not — to understand molecular stability. [2] [23] The isodesmic reaction energy is thus a metric of the stabilization/destabilization relative to the starting reagents. Overall, as the number of nitrogens increase, the stability of the system increases. The isodesmic reaction energy goes from 151.8 kcal/mol to 81.8 kcal/mol for 0 to 4 nitrogen atoms. Nitration destabilizes the energy dramatically, as exemplified by 2,4-dinitro-1,3-diazatetrahedrane ((O2NC)2N2) having an isodesmic reaction energy of 152.9 kcal/mol. [23]

Due to the instability of many azatetrahedranes, isodesmic comparisons to azacyclobutadiene analogues have been used to determine which core structures are the most synthetically feasible. Alkorta, Elguero, and Rozas reported that every member of the azatetrahedrane core series is always slightly more unstable than their azacyclobutadiene analogue(s). [2] Jursic's calculations suggest that the energetic differential between the azatetrahedrane and the azacyclobutadiene starts off large and decreases as the number of nitrogen atoms increase in the cage. Furthermore, Jursic's calculations suggest that tetraazatetrahedrane may be slightly more stable (difference of 4.4 kcal/mol at 0 K and the CBSQ level of theory) than its azacyclobutadiene analogue. Despite the instability, non-sterically hindered azatetrahedranes may still be detectable in gaseous matrices. [24]

Phosphatetrahedranes

Intrinsic bond orbital analysis of monophosphatetrahedrane cage closing. Note that the intrinsic bond orbital corresponding P-Cl s bond is localized on the chloride anion upon cage closing. Monophosphatetrahedrane Ring Opening Orbitals.png
Intrinsic bond orbital analysis of monophosphatetrahedrane cage closing. Note that the intrinsic bond orbital corresponding P-Cl σ bond is localized on the chloride anion upon cage closing.

Bonding Parameters

Molecular graph of tri-tert-butylmonophosphatetrahedrane optimized at the B3LYP-D3/6-31G** level of theory shows significant bond deviations. Tri-tert-butylmonophosphatetrahedrane Molecular Graph.png
Molecular graph of tri-tert-butylmonophosphatetrahedrane optimized at the B3LYP-D3/6-31G** level of theory shows significant bond deviations.

Much like azatetrahedranes, phosphatetrahedranes show bond deviation. The molecular graph of tBu3C3P shows significantly warped bond paths; the bond critical points lie far from the linear representation. [3]

Investigations into the bonds of phosphatetrahedranes used quasi-atomic orbital analysis with isodesmic reactions. In contrast to azatetrahedranes, the lower electronegativity of phosphorus relative to carbon leads to a localization of negative charge in the carbons and the localization of positive charge in the phosphorus(es). This leads to greater occupancy in the C-C-C ring from 4.17 to 4.28 electrons with the addition of one phosphorus atom. Overall, the extent of charge transfer increases with the number of phosphorus atoms in the tetrahedral core. [25]

Of the phosphatetrahedranes, only the triphosphatetrahedrane core did not show evidence of cage-opening upon ionization. [22]

Thermodynamics

Depiction of the corset effect in tri-tert-butylmonophosphatetrahedrane where the attractive interaction between hydrogen atoms of neighboring tert-butyl groups are highlighted in red. Monophosphatetrahedrane Corset Effect.png
Depiction of the corset effect in tri-tert-butylmonophosphatetrahedrane where the attractive interaction between hydrogen atoms of neighboring tert-butyl groups are highlighted in red.

Ivanov, Bozhenko, and Boldyrev studied the energetic landscape of the phosphatetrahedrane series ((HC)xP4-x). Their calculations suggest that substitution of phosphorus for carbon increasingly favors the tetrahedral structure. The tetrahedral structure is the absolute minima starting for triphosphatetrahedrane, but diphosphatetrahedrane is only 2.3 kcal/mol higher in energy than the absolute minima. They attribute the stabilization of the tetrahedral structure to phosphorus' amiability towards more acute bond angles. The more diffuse orbitals of phosphorus versus carbon also favor the tetrahedral structure's σ-interactions over the planar phosphacyclobutadiene's π-interactions. [1]

Riu, Ye, and Cummins report similar computational findings. Their calculations show decreasing strain energy with the number of phosphorus atoms in the tetrahedral cage. They also attribute the stabilization to the diffusitivity of phosphorus orbitals. They also note that the accumulation of p-character on the bond orbitals leads to greater s-character on the lone pairs. [5]

The isolability of tBu3C3P was attributed to the controversial hydrogen-hydrogen bonds (HHB), which some chemists have argued may not exist. Each HHB of the tert-butyl network were calculated (in absence of steric repulsion) to contribute 0.7 kcal/mol of stabilization. Calculations with one of the tert-butyl substituents with a methyl, ethyl, or isopropyl group result in net repulsion due to the loss of HHBs. [26] In total, this forms the basis of the corset effect. [5] Non-Lewis donation of electron density from the tetrahedral core to the tert-butyl substituents also stabilizes tBu3C3P according to natural bond orbital theory. [5] This effect was also demonstrated in silico for unsubstituted monophosphatetrahedrane. [27]

Substitution by Heavier Congeners

Schaefer and coworkers, in light of the synthesis of tBu3C3P, ran calculations on the mono-pnictogen-substituted tetrahedrane series, represented by R3C3Pn (R = H, tBu, Pn = N, P, As, Sb, Bi). These calculations yielded a series of well correlated trends.

Consistent with a perturbation of the pnictogen residing above a C-C-C ring, the C-Pn bonds elongate from 1.493 Å to 2.289 Å, and C-Pn-C angle decreases from 58.0° to 37.1° as heavier congeners are used. This is due to the larger atomic radius of the heavier pnictogens. The H-[C-C-C plane] angle increases from 9.1° to 31.1°, which is also attributed to the diffusivity of the heavier congener's orbitals.

As noted above with the aza- and phosphatetrahedranes, the change in pnictogen electronegativity changes the interaction between the Pn atom and the C-C-C ring. The C-C-C ring becomes increasingly more negatively charged with the heavier pnictogens.

Tabulated bond angles and lengths of the mono-pnictogen-substituted tetrahedrane series calculated at the CCSD(T)/aug-cc-pwCVTZ(-PP) level of theory. Monopnictatetrahedrane Bond Angles and Lengths.png
Tabulated bond angles and lengths of the mono-pnictogen-substituted tetrahedrane series calculated at the CCSD(T)/aug-cc-pwCVTZ(-PP) level of theory.

Isodesmic reactions show greater stabilization of the cage structure due to the diffusivity of the pnictogen's orbitals, although even with bismuth, the mono-pnictogen-substituted tetrahedrane is unstable. Delocalization plays a large part in the stabilization of the heavier analogues. For example, electron density is increasingly transferred from the Pn-C bonds into the Pn lone pair in the heavier congeners. These lone pairs are also noted to follow Bent's rule.

As noted above with tBu3C3P, non-Lewis interactions stabilize the tetrahedral core. These effects also become more pronounced with the heavier pnictogens. Second order perturbations suggest that the key non-Lewis interactions are C-C to C-R* and C-Pn to C-H* (i.e., cage-opening), as well as interactions to Pn-C*. The former set of interactions stabilize the tetrahedral core most when the substituent is an electron-withdrawing group (e.g., fluoride), although decreased electron density in C-C and C-Pn can facilitate cage-opening as well. [27]

See also

Related Research Articles

An ylide or ylid is a neutral dipolar molecule containing a formally negatively charged atom (usually a carbanion) directly attached to a heteroatom with a formal positive charge (usually nitrogen, phosphorus or sulfur), and in which both atoms have full octets of electrons. The result can be viewed as a structure in which two adjacent atoms are connected by both a covalent and an ionic bond; normally written X+–Y. Ylides are thus 1,2-dipolar compounds, and a subclass of zwitterions. They appear in organic chemistry as reagents or reactive intermediates.

<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 2023. 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.

In chemistry, a hypervalent molecule is a molecule that contains one or more main group elements apparently bearing more than eight electrons in their valence shells. Phosphorus pentachloride, sulfur hexafluoride, chlorine trifluoride, the chlorite ion, and the triiodide ion are examples of hypervalent molecules.

<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.

Diphosphorus is an inorganic chemical with the chemical formula P
2
. Unlike nitrogen, its lighter pnictogen neighbor which forms a stable N2 molecule with a nitrogen to nitrogen triple bond, phosphorus prefers a tetrahedral form P4 because P-P pi-bonds are high in energy. Diphosphorus is, therefore, very reactive with a bond-dissociation energy (117 kcal/mol or 490 kJ/mol) half that of dinitrogen. The bond distance has been measured at 1.8934 Å.

<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.

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

Tetranitrogen is a neutrally charged polynitrogen allotrope of the chemical formula N
4
and consists of four nitrogen atoms. The tetranitrogen cation is the positively charged ion, N+
4
, which is more stable than the neutral tetranitrogen molecule and is thus more studied.

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

Phosphorus mononitride is an inorganic compound with the chemical formula PN. Containing only phosphorus and nitrogen, this material is classified as a binary nitride. From the Lewis structure perspective, it can be represented with a P-N triple bond with a lone pair on each atom. It is isoelectronic with N2, CO, P2, CS and SiO.

<span class="mw-page-title-main">Decamethylsilicocene</span> Chemical Compound

Decamethylsilicocene, (C5Me5)2Si, is a group 14 sandwich compound. It is an example of a main-group cyclopentadienyl complex; these molecules are related to metallocenes but contain p-block elements as the central atom. It is a colorless, air sensitive solid that sublimes under vacuum.

A phosphetane is a 4-membered organophosphorus heterocycle. The parent phosphetane molecule, which has the formula C3H7P, is one atom larger than phosphiranes, one smaller than phospholes, and is the heavy-atom analogue of azetidines. The first known phosphetane synthesis was reported in 1957 by Kosolapoff and Struck, but the method was both inefficient and hard to reproduce, with yields rarely exceeding 1%. A far more efficient method was reported in 1962 by McBride, whose method allowed for the first studies into the physical and chemical properties of phosphetanes. Phosphetanes are a well understood class of molecules that have found broad applications as chemical building blocks, reagents for organic/inorganic synthesis, and ligands in coordination chemistry.

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

Phosphasilenes or silylidenephosphanes are a class of compounds with silicon-phosphorus double bonds. Since the electronegativity of phosphorus (2.1) is higher than that of silicon (1.9), the "Si=P" moiety of phosphasilene is polarized. 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. 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.

<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">Nontrigonal pnictogen compounds</span>

Nontrigonal pnictogen compounds refer to tricoordinate trivalent pnictogen compounds that are not of typical trigonal pyramidal molecular geometry. By virtue of their geometric constraint, these compounds exhibit distinct electronic structures and reactivities, which bestow on them potential to provide unique nonmetal platforms for bond cleavage reactions.

<span class="mw-page-title-main">Allotropes of arsenic</span>

Arsenic in the solid state can be found as gray, black, or yellow allotropes. These various forms feature diverse structural motifs, with yellow arsenic enabling the widest range of reactivity. In particular, reaction of yellow arsenic with main group and transition metal elements results in compounds with wide-ranging structural motifs, with butterfly, sandwich and realgar-type moieties featuring most prominently.

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

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

Bismuthinidenes are a class of organobismuth compounds, analogous to carbenes. These compounds have the general form R-Bi, with two lone pairs of electrons on the central bismuth(I) atom. Due to the unusually low valency and oxidation state of +1, most bismuthinidenes are reactive and unstable, though in recent decades, both transition metals and polydentate chelating Lewis base ligands have been employed to stabilize the low-valent bismuth(I) center through steric protection and π donation either in solution or in crystal structures. Lewis base-stabilized bismuthinidenes adopt a singlet ground state with an inert lone pair of electrons in the 6s orbital. A second lone pair in a 6p orbital and a single empty 6p orbital make Lewis base-stabilized bismuthinidenes ambiphilic.

<span class="mw-page-title-main">Stable phosphorus radicals</span>

Stable and persistent phosphorus radicals are phosphorus-centred radicals that are isolable and can exist for at least short periods of time. Radicals consisting of main group elements are often very reactive and undergo uncontrollable reactions, notably dimerization and polymerization. The common strategies for stabilising these phosphorus radicals usually include the delocalisation of the unpaired electron over a pi system or nearby electronegative atoms, and kinetic stabilisation with bulky ligands. Stable and persistent phosphorus radicals can be classified into three categories: neutral, cationic, and anionic radicals. Each of these classes involve various sub-classes, with neutral phosphorus radicals being the most extensively studied. Phosphorus exists as one isotope 31P (I = 1/2) with large hyperfine couplings relative to other spin active nuclei, making phosphorus radicals particularly attractive for spin-labelling experiments.

Phosphiranes are organic compounds with the phosphirane functional group – a three-membered ring with two atoms of carbon and one atom of phosphorus that has lots of ring angle strain. Phosphiranes are usually synthesized by double substitution reactions or pericyclic pathways. Phosphiranes can also be oxidized into phosphirane oxides, undergo SN2 substitution reactions, or decompose into different units.

<span class="mw-page-title-main">Tris(silox)tantalum</span> Chemical compound

Tris(silox)tantalum, Ta(SiOtBu)3, is an organotantalum complex bound with three siloxide (silox="tBu3SiO-") ligands,. The tantalum center has a d-electron count of 2 and an oxidation state of III. The complex is trigonal planar whose point group is assigned as D3h. It is a crystalline light blue solid which forms blue-green solutions in THF.

References

  1. 1 2 Ivanov, Alexander S.; Bozhenko, Konstantin V.; Boldyrev, Alexander I. (2012-01-10). "Peculiar Transformations in the C x H x P 4– x ( x = 0–4) Series". Journal of Chemical Theory and Computation. 8 (1): 135–140. doi:10.1021/ct200727z. ISSN   1549-9618.
  2. 1 2 3 4 Alkorta, Ibon; Elguero, Jose; Rozas, Isabel; Balaban, Alexandru T. (August 1990). "Theoretical studies on aza-analogs of platonic hydrocarbons". Journal of Molecular Structure. 208 (1–2): 63–77. doi:10.1016/0166-1280(92)80008-A.
  3. 1 2 3 4 5 6 Riu, Martin-Louis Y.; Jones, Rebecca L.; Transue, Wesley J.; Müller, Peter; Cummins, Christopher C. (2020-03-27). "Isolation of an elusive phosphatetrahedrane". Science Advances. 6 (13). doi:10.1126/sciadv.aaz3168. ISSN   2375-2548. PMC   7096166 . PMID   32232162.
  4. 1 2 3 4 5 Hierlmeier, Gabriele; Coburger, Peter; Bodensteiner, Michael; Wolf, Robert (2019-11-18). "Di‐ tert ‐butyldiphosphatetrahedrane: Catalytic Synthesis of the Elusive Phosphaalkyne Dimer". Angewandte Chemie International Edition. 58 (47): 16918–16922. doi:10.1002/anie.201910505. ISSN   1433-7851. PMC   6899750 . PMID   31591760.
  5. 1 2 3 4 5 6 7 Riu, Martin-Louis Y.; Ye, Mengshan; Cummins, Christopher C. (2021-10-13). "Alleviating Strain in Organic Molecules by Incorporation of Phosphorus: Synthesis of Triphosphatetrahedrane". Journal of the American Chemical Society. 143 (40): 16354–16357. doi:10.1021/jacs.1c07959. ISSN   0002-7863.
  6. Maier, Günther; Pfriem, Stephan; Schäfer, Ulrich; Matusch, Rudolf (July 1978). "Tetra‐ tert ‐butyltetrahedrane". Angewandte Chemie International Edition in English. 17 (7): 520–521. doi:10.1002/anie.197805201. ISSN   0570-0833.
  7. 1 2 Jupp, Andrew R.; Slootweg, J. Chris (2020-06-26). "Mixed Phosphatetrahedranes". Angewandte Chemie International Edition. 59 (27): 10698–10700. doi:10.1002/anie.202004514. hdl: 11245.1/1c4a49bd-da63-4c0e-b0b0-1a2f79f44634 . ISSN   1433-7851.
  8. Cossairt, Brandi M.; Diawara, Mariam-Céline; Cummins, Christopher C. (2009-01-30). "Facile Synthesis of AsP 3". Science. 323 (5914): 602–602. doi:10.1126/science.1168260. ISSN   0036-8075.
  9. Pan, Fuxing; Guggolz, Lukas; Dehnen, Stefanie (July 2022). "Capture the missing: formation of (PbBi 3 ) − and {[AuPb 5 Bi 3 ] 2 } 4− via atom exchange or reorganization of the pseudo‐ tetrahedral Zintl anion (Pb 2 Bi 2 ) 2−". Natural Sciences. 2 (3). doi: 10.1002/ntls.202103302 . ISSN   2698-6248.
  10. Rayne, Sierra; Forest, Kaya (2010-11-01). "Structures, enthalpies of formation, and ionization energies for the parent and binary mixed carbon, silicon, nitrogen, and phosphorus cubane derivatives: A G4MP2 theoretical study". Nature Precedings. doi: 10.1038/npre.2010.5154.1 . ISSN   1756-0357.
  11. 1 2 3 Riu, Martin-Louis Y.; Eckhardt, André K.; Cummins, Christopher C. (2021-08-25). "Dimerization and Cycloaddition Reactions of Transient Tri- tert -butylphosphacyclobutadiene Generated by Lewis Acid Induced Isomerization of Tri- tert -butylphosphatetrahedrane". Journal of the American Chemical Society. 143 (33): 13005–13009. doi:10.1021/jacs.1c06840. ISSN   0002-7863.
  12. Riu, Martin-Louis Y. (September 2022). Synthesis and Reactivity of Phosphorus-Containing Heterocycles and Tetrahedranes (Thesis thesis). Massachusetts Institute of Technology.
  13. 1 2 3 Riu, Martin-Louis Y.; Eckhardt, André K.; Cummins, Christopher C. (2022-05-04). "Reactions of Tri- tert -Butylphosphatetrahedrane as a Spring-Loaded Phosphinidene Synthon Featuring Nickel-Catalyzed Transfer to Unactivated Alkenes". Journal of the American Chemical Society. 144 (17): 7578–7582. doi:10.1021/jacs.2c02236. ISSN   0002-7863.
  14. Hierlmeier, Gabriele; Wolf, Robert (2021-03-15). "Activation of Di‐ tert ‐butyldiphosphatetrahedrane: Access to ( t BuCP) n ( n= 2, 4) Ligand Frameworks by P−C Bond Cleavage". Angewandte Chemie International Edition. 60 (12): 6435–6440. doi:10.1002/anie.202015680. ISSN   1433-7851. PMC   7986217 . PMID   33403771.
  15. 1 2 Hierlmeier, Gabriele; Coburger, Peter; Scott, Daniel J.; Maier, Thomas M.; Pelties, Stefan; Wolf, Robert; Pividori, Daniel M.; Meyer, Karsten; van Leest, Nicolaas P.; de Bruin, Bas (2021-10-25). "Di ‐tert ‐butyldiphosphatetrahedrane as a Source of 1,2‐Diphosphacyclobutadiene Ligands". Chemistry – A European Journal. 27 (60): 14936–14946. doi:10.1002/chem.202102335. ISSN   0947-6539. PMC   8596834 . PMID   34424579.
  16. Uttendorfer, Maria K.; Hierlmeier, Gabriele; Wolf, Robert (2022-10-14). "A Homoleptic Diphosphatetrahedrane Nickel(0) Complex". Zeitschrift für anorganische und allgemeine Chemie. 648 (19). doi:10.1002/zaac.202200124. ISSN   0044-2313.
  17. Hierlmeier, Gabriele; Uttendorfer, Maria K.; Wolf, Robert (2021). "Di- tert -butyldiphosphatetrahedrane as a building block for phosphaalkenes and phosphirenes". Chemical Communications. 57 (19): 2356–2359. doi: 10.1039/D0CC07103J . ISSN   1359-7345.
  18. 1 2 3 4 Politzer, Peter; Seminario, Jorge M. (January 1989). "Computational determination of the structures and some properties of tetrahedrane, prismane, and some of their aza analogs". The Journal of Physical Chemistry. 93 (2): 588–592. doi:10.1021/j100339a019. ISSN   0022-3654.
  19. 1 2 3 Politzer, Peter; Seminario, Jorge M. (January 1990). "Relative bond strengths in tetrahedrane, prismane, and some of their aza analogs". Structural Chemistry. 1 (1): 29–32. doi:10.1007/BF00675781. ISSN   1040-0400.
  20. Seminario, Jorge M.; Politzer, P. (June 1989). "Analysis of different computational treatments of highly strained molecules". Chemical Physics Letters. 159 (1): 27–31. doi:10.1016/S0009-2614(89)87447-0.
  21. Politzer, Peter; Seminario, Jorge M. (June 1989). "Computational analysis of the structures, bond properties, and electrostatic potentials of some nitrotetrahedranes and nitroazatetrahedranes". The Journal of Physical Chemistry. 93 (12): 4742–4745. doi:10.1021/j100349a013. ISSN   0022-3654.
  22. 1 2 Rayne, Sierra; Forest, Kaya (2011-08-29). "Theoretical studies in the molecular Platonic solids: Pure and mixed carbon, nitrogen, phosphorus, and silicon tetrahedranes". Nature Precedings. doi: 10.1038/npre.2011.6299.1 . ISSN   1756-0357.
  23. 1 2 Murray, Jane S.; Seminario, Jorge M.; Lane, Pat; Politzer, Peter (June 1990). "Anomalous energy effects associated with the presence of aza nitrogens and nitro substituents in some strained systems". Journal of Molecular Structure. 207 (3–4): 193–200. doi:10.1016/0166-1280(90)85023-G.
  24. Jursic, B.S. (February 2001). "Structures and properties of nitrogen derivatives of tetrahedrane". Journal of Molecular Structure. 536 (2–3): 143–154. doi:10.1016/S0166-1280(00)00619-9.
  25. Del Angel Cruz, Daniel; Galvez Vallejo, Jorge L.; Gordon, Mark S. (2023). "Analysis of the bonding in tetrahedrane and phosphorus-substituted tetrahedranes". Physical Chemistry Chemical Physics. 25 (40): 27276–27292. doi: 10.1039/D3CP03619G . ISSN   1463-9076.
  26. Riu, Martin-Louis Y.; Bistoni, Giovanni; Cummins, Christopher C. (2021-07-22). "Understanding the Nature and Properties of Hydrogen–Hydrogen Bonds: The Stability of a Bulky Phosphatetrahedrane as a Case Study". The Journal of Physical Chemistry A. 125 (28): 6151–6157. doi:10.1021/acs.jpca.1c04046. ISSN   1089-5639.
  27. 1 2 Wolf, Mark E.; Doty, Elizabeth A.; Turney, Justin M.; Schaefer III, Henry F. (2021-04-01). "Highly Strained Pn(CH) 3 (Pn = N, P, As, Sb, Bi) Tetrahedranes: Theoretical Characterization". The Journal of Physical Chemistry A. 125 (12): 2612–2621. doi:10.1021/acs.jpca.1c01022. ISSN   1089-5639.