Borole

Last updated
Borole
Borole-2D-skeletal.png
Borole-3D-balls.png
Borole-3D-spacefill.png
Names
Preferred IUPAC name
1H-Borole
Identifiers
3D model (JSmol)
ChemSpider
PubChem CID
  • InChI=1S/C4H5B/c1-2-4-5-3-1/h1-5H X mark.svgN
    Key: XQIMLPCOVYNASM-UHFFFAOYSA-N X mark.svgN
  • InChI=1/C4H5B/c1-2-4-5-3-1/h1-5H
    Key: XQIMLPCOVYNASM-UHFFFAOYAW
  • C1=CC=CB1
Properties
C4H5B
Molar mass 63.89 g·mol−1
Except where otherwise noted, data are given for materials in their standard state (at 25 °C [77 °F], 100 kPa).
X mark.svgN  verify  (what is  Yes check.svgYX mark.svgN ?)

Boroles represent a class of molecules known as metalloles, which are heterocyclic 5-membered rings. As such, they can be viewed as structural analogs of cyclopentadiene, pyrrole or furan, with boron replacing a carbon, nitrogen and oxygen atom respectively. They are isoelectronic with the cyclopentadienyl cation C5H+5 or abbreviated as Cp+ and comprise four π electrons. Although Hückel's rule cannot be strictly applied to borole, it is considered to be antiaromatic due to having 4 π electrons. [1] As a result, boroles exhibit unique electronic properties not found in other metalloles.

Contents

The parent unsubstituted compound with the chemical formula C 4 H 4 BH has yet to be isolated outside a coordination sphere of transition metals. [2] Substituted derivatives, which have been synthesized, can have various substituents at the 4 carbons and boron. [3] The high electron deficiency leads to various reactivities such as metal free hydrogen activation and rearrangements upon cycloaddition which are unobserved in other structural analogues like pyrrole or furan.

Once reduced to the dianion, the borolediide complex gains aromaticity and can then participate in similar reactions as the Cp anion, including forming sandwich complexes.

Electronic properties

Hückel analysis

According to Hückel's rule which states that a cyclic molecule is aromatic if it has (4n + 2) π electrons and antiaromatic if there are 4n electrons, boroles represent antiaromatic molecules. In agreement with chemical intuition, ab initio calculations on the parent borole C4H4BH predict it to have an antiaromatic singlet ground state. [3] Its backbone structure features strongly alternating bond lengths, consistent with localised electrons in the π system. This characteristic is preserved in almost all structurally characterised borole derivatives except those derived from [PhBC4Ph4]. [4] This discrepancy was attributed to intermolecular phenyl→boron π donation within dimeric subunits.

In addition, theoretical studies also suggest that borole is significantly destabilised by the delocalisation of its four π electrons (NICS 17.2; ASE 19.3 kcal mol−1). UV-Vis spectroscopy and reactivity studies have been conducted to assess the consequences of antiaromaticity in boroles. Their antiaromatic character entails strong electrophilicity of the boron center resulting in even weak donors such as ethers or nitriles being capable of forming stable Lewis acid–base adducts. Moreover, boroles' highly activated carbon backbone readily participates in Diels–Alder reactions and is prone to two-electron reductions affording borolediides.

A simple Hückel model can be used to compare the spectroscopic properties and observed reactivity of boroles against the isoelectronic cyclopentadienyl cation [C5H5]+. Unlike [C5H5]+ which has a doubly degenerate pair of HOMOs, the introduction of a boron center lifts their degeneracy by increasing the energy of the antisymmetric molecular orbital somewhat and the symmetric molecular orbital significantly. As a result, the HOMO in boroles is doubly occupied and no biradical character is observed, in line with a singlet ground state and the diamagnetic character of boroles. Furthermore, boroles exhibit a small HOMO–LUMO gap (principally consisting of a π–π* transition) and the lowest-energy electronic absorption of boroles is considerably red shifted in the UV-Vis spectra (e.g.[PhBC4Ph4]: λmax = 560 nm). Accordingly, boroles exhibit a characteristic blue color.

Qualitative comparison of molecular orbitals. Left: Cyclopentadienyl cation. Right: Borole. The green electrons represent the electronic state after coordinating to a Lewis base Borole vs Cp+.png
Qualitative comparison of molecular orbitals. Left: Cyclopentadienyl cation. Right: Borole. The green electrons represent the electronic state after coordinating to a Lewis base

By contrast, introducing two electrons into the vacant LUMO either by reduction or adduct formation with Lewis bases significantly increases the HOMO–LUMO gap. Consequently, a dramatic blue shift of the lowest-energy excitation is observed (e.g.[PhBC4Ph4]·pyridine: λmax = 340 nm) and the resulting species are usually yellow to red in color. A qualitative drawing is presented to the left. The small HOMO-LUMO gap also makes boroles excellent participants in Diels–Alder type reactions either with themselves or with a variety of alkenes and alkynes.

Natural bond orbitals

Natural bond orbital (NBO) analysis of C4H4BH has been performed in order to understand the bonding of borole in the familiar Lewis picture. [5] According to the computational results, the occupancy of the two C−C π orbitals is about 1.9, with a tiny amount of electronic charge (an occupancy of 0.13) delocalised on the out-of-plane boron p orbital, illustrated below. The standard Lewis structure of borole captures more than 50% of the overall electronic structure according to Natural Resoanance Theory analysis. As delocalisation of the 4π electrons is prevented by antiaromaticity, the unsaturated boron atom has low occupancy of its vacant pz orbital and is highly Lewis acidic. Along with the low energy LUMO, boroles show an inherent propensity to form Lewis acid–base adducts even with substrates of low donor strengths.

Natural Bonding Orbitals of Borole. Structure optimised using ORCA BP86-D3BJ and def2- TZVPP basis set. The calculated occupencies of the obitals going from left to right are 0.13, 1.9 and 1.9 respectively. NBOborole.png
Natural Bonding Orbitals of Borole. Structure optimised using ORCA BP86-D3BJ and def2- TZVPP basis set. The calculated occupencies of the obitals going from left to right are 0.13, 1.9 and 1.9 respectively.
Frontier molecular orbitals of Borole optimised using ORCA BP86-D3BJ and def2- TZVPP basis set. Top: HOMO Bottom: LUMO Borole homolumo.png
Frontier molecular orbitals of Borole optimised using ORCA BP86-D3BJ and def2- TZVPP basis set. Top: HOMO Bottom: LUMO
Bond lengths in Borole optimised at the CCSD(T)/6-311G(2df,p) level. Bond lengths are in A. Energy1801328688 1.png
Bond lengths in Borole optimised at the CCSD(T)/6-311G(2df,p) level. Bond lengths are in Å.

As the p orbital of boron is virtually vacant and nonbonding (as indicated by its NBO energy level), borole is regarded as a good Lewis acid or electron acceptor. The figure to the right shows the lack of involvement of boron's p orbital in the HOMO and the substantial Lewis acidic character at boron in the LUMO. Chemically, borole is reactive and unstable in ambient conditions. The pentaphenylborole analog is a highly reactive green solid; it readily undergoes oxidation, partial protolysis, and Diels−Alder reaction with dienophiles. Borole, even in perarylated form, is still very labile. Due to its reactive nature, the structural parameters and thermochemical data of borole are not known.

In the optimized structure of borole shown to the left, the B1C, 1C2C, and 2C2C bond lengths are approximately 1.58, 1.338, and 1.518 Å respectively, as shown to the left. The longer 2C2C bond in C4H4BH agree with NBO analysis that the π-electron delocalizations are mainly confined on the methine carbons, supporting the antiaromatic nature of the neutral borole.

Borole dianion (borolediides)

As the boron has an empty p orbital, boroles tend to accept 2 electrons in order to have 6 π electrons and gain aromaticity. This reduction to borolediides was demonstrated in the early 1980s by Herberich et al. with the isolation of K2[PhBC4Ph4]. [7] Because the atomic orbitals of boron are different from those from carbon in terms of energy, not all atoms contribute equally to the π system in the borole dianion. A Natural Resonance Theory (NRT) calculation [6] shows that there are 3 dominant resonance structures for the isolated [C4BH5]2− dianion as illustrated below.

Resonance forms in the Borole dianion. Contributions to overall structure are 30.30%, 30.71% and 13.04% respectively going from left to right as determed using NRT. Analyses were performed using NBO7 on a [C4BH5] structure optimised using BP86-D3BJ and def2- TZVPP basis set. Resonance in borole 2-.png
Resonance forms in the Borole dianion. Contributions to overall structure are 30.30%, 30.71% and 13.04% respectively going from left to right as determed using NRT. Analyses were performed using NBO7 on a [C4BH5] structure optimised using BP86-D3BJ and def2- TZVPP basis set.

Since borole dianions are isoelectronic to the ubiquitous cyclopentadienyl anion, aromatic delocalisation of the 6π electrons should cause bond lengths assimilation within the BC4 backbone. This is exactly what was found for several characterised structures, namely K2[PhBC4Ph4], K2[(4-Me3Si-C6H4)BC4Ph4] and K2[ClBC4Ph4]. Thus, the observed B–C bonds are rather short (1.505–1.543 Å) and all C–C bond lengths lie within a narrow array (1.409–1.456 Å). [8] [9] [7]

Synthesis

The first borole derivative to be isolated was pentaphenylborole [PhBC4Ph4], synthesised by Eisch et al. in 1969 as a deep blue solid. [10] Referring to the figure below, the practical synthesis of [PhBC4Ph4] (1) was initially accomplished in two different ways: [10] [7] [11] (a) by direct reaction of 1,4-dilithio-1,2,3,4-tetraphenylbutadiene with PhBBr2 which gives a Lewis base adduct of pentaphenylborole (1·OEt2) in diethylether, and subsequent removal of the solvent. (b) By boron–tin exchange between 2,3,4,5-tetraphenylstannole derivatives with PhBCl2. [12] Eisch et al. have demonstrated that the latter method can be expanded to other borole derivatives, even though these species have only been generated in situ.

Synthesis of substituted boroles. Reaction conditions: (i) R2'SnCl2,THF/Et2O; (ii) PhBCl2, PhMe; (iii) PhBBr2, Et2O; (iv) -Et2O Synthesis of boroles.png
Synthesis of substituted boroles. Reaction conditions: (i) R2'SnCl2,THF/Et2O; (ii) PhBCl2, PhMe; (iii) PhBBr2, Et2O; (iv) -Et2O

Accounting for the drawbacks of each method, the boron–tin exchange is the current method of choice and has been widely accepted for the synthesis of numerous differently substituted borole derivatives. The other approach has yet to play a significant role in the further development of borole chemistry.
Besides the development of a synthetic pathway to the perfluorinated version of (1), [13] the substituents on the BC4 backbone is largely still limited to phenyl substituents. However, substituents beside H has been attached to the boron atom, such as halide, aryl and amino functionalities. Ferrocenyl, cymantrenyl and platinum complex fragments have also been successfully attached.

Depending on the boron-bound substituent, the electron density at boron can be altered. Hence, substituents can exert strong influences on the spectroscopic properties of the whole borole system. For instance, significant π-back bonding interactions from nitrogen in [(Me3Si)2NBC4Ph4] [14] raises the borole LUMO energy and a resulting blue shift of the lowest energy absorption in UV-Vis spectra (λmax = 478 nm) compared to [PhBC4Ph4] (λmax = 560 nm).

In order to synthesis less sterically congested boroles, a zirconacycle transfer strategy was adopted by Fagan et al. [15] [16] Reaction of [Cp2ZrC4Me4] with PhBCl2 was expected to result in the formation of [PhBC4Me4]. However, the product was too reactive and only its Diels–Alder dimer has been isolated. Evidence for the intermediate [PhBC4Me4] before dimerisation was shown through trapping experiments with 2-butyne and reactivity studies using a variety of unactivated alkenes.

In 2018, Lee et al. successfully transformed a borapyramidane into a stable halogen substituted planar borole dianion which was stabilized by Li+ ions positioned above and below the plane of the borole ring, [17] revealing a direct synthetic path to borolediides from borapyramidane. [18]

Reactions

Pentaphenylborole is known to show a wide range of reactivity, owing to its antiaromatic and highly Lewis acidic nature.

Lewis acid-base adducts

Borole Adduct formation.png

As suggested previously, the high Lewis acidity of boroles allows the ready formation of Lewis acid–base adducts with a variety of different donor molecules. This simplest case of reactivity has already been realised in the early days and has frequently been used to highlight the antiaromatic nature of boroles. [10] [19] [20] Pyridines, ethers, phosphines, and different carbene species have been successfully attached to the unsaturated boron center. [21] [22] [23]

In general, such reactions are facile and proceed quantitatively, facilitating their isolation in high yield. Upon Lewis base coordination, the former vacant p orbital at boron becomes occupied and cyclic delocalization of the π electron system is no longer feasible, corresponding to the loss of antiaromaticity. However, strong bond length alternation in the BC4 backbone is still observed and remain almost unaffected by these fundamental electronic changes. In contrast, spectroscopic measurements are much more sensitive to adduct formation. Unlike the respective borole precursors which are intensely coloured, the adducts are pale yellow solids with characteristic UV-Vis excitations at λmax = 350–380 nm which agrees with an increase in the HOMO-LUMO gap.

Addition reactions with unsaturated carbon compounds

As a result of only having 4 electrons in the planar π system, boroles experience a large destabilizing effect and thus exhibits high reactivity, such as in dimerisations [24] and cycloadditions. [25] This is why stable borole derivatives can only be achieved by employing bulky substituents around the reactive BC4 core.

Reaction conditions: (i)
(C6F5)CC(C6F5), toluene, 110 degC, 7 d; (ii) EtCCEt, CH2Cl2, RT; (iii) PhCCPh, CH2Cl2, RT. Cycloadditions of boroles.png
Reaction conditions: (i) (C6F5)CC(C6F5), toluene, 110 °C, 7 d; (ii) EtCCEt, CH2Cl2, RT; (iii) PhCCPh, CH2Cl2, RT.

The small HOMO-LUMO gap of boroles makes them eminently suitable as Diels–Alder participants. The reaction of [PhBC4Ph4] (1) with diphenylacetylene to afford aromatic heptaphenylborepine has been reported by Eisch et al. in the “paradigm of pericyclic reactions”. [26] Recently, the interest in this reaction pathway was revived by Piers et al.., who studied the reactivity of perfluorinated [PhBC4Ph4] (2 in the figure above) towards alkynes in great detail. [27] Of particular relevance was the question of whether the enhanced Lewis acidity of 2 compared 1 affects its reactivity and whether that entails novel reactivity patterns. In that study, 2 was reacted with three differently substituted alkynes that greatly differed in their electronic properties, namely (C6F5)CC(C6F5), (C6H5)CC(C6H5) and Et–CC–Et.

The reaction with the electron-poor alkyne (C6F5)CC(C6F5) required rather harsh reaction conditions (110 °C, 7 days) and the only observable species was the expected Diels–Alder product (3), whose formation presumably follows a mechanism similar to that proposed for 5. The same pathway was found for the more electron-rich Et–CC–Et (3-hexyne). However, the mild reaction conditions (spontaneous at room temperature) enabled the isolation of the direct Diels–Alder cycloaddition product 1,2-Et2-4, which is considered the thermodynamically favored isomer of the two possible 7-borabicyclo[2.2.1]hepta-dienes. Prolonged heating of solutions of 4 resulted in the formation of various products, some of which have been identified as borepine isomers of Et2-5 based on 11B NMR data. This provides for the first time conclusive experimental evidence for the proposed mechanism of borepine formation from boroles and alkynes.

However, the reaction of 2 with diphenylacetylene gave unexpected results. Similar to the observations described for 3-hexyne, the electron rich PhCCPh facilitated a spontaneous reaction. Surprisingly, the diphenyl-substituted heptaarylborepine isomers 6 formed by the common Diels–Alder reactivity represented only minor components of the product mixture (25%). The potent Lewis acidity of 12 revealed a novel reaction pathway whereby the alkyne first adds to the borole nucleophilically, followed by subsequent aryl migration and ring expansion to afford the boracyclohexadiene 7 as the predominant species (75%).

Both reaction pathways likely compete with each other and as a result, the nature of the alkyne (nucleophilicity, substitution pattern) strongly dictates which reactivity is favored.

Ring expansion

Reaction scheme borole ring expansion.png

The highly reactive nature of boroles also leads to them participating in ring expansion [28] reactions upon reaction with azides. [29] In the reaction of 1‐(2,3,4,5‐tetraphenylborolyl)ferrocene [4] with 4‐azido‐N,N‐dimethylaniline to generate a new species (8 in the figure above). Cyclic voltametry studies indicated that the 1,2,3‐diazaborinine exhibits similar electronic effects to that of pyridine, which makes (8) weakly electron‐withdrawing. Analyses of the frontier molecular orbitals using DFT shows that the HOMO mainly comprises the iron dx2-y2 orbital while the LUMO resides largely on the BN pyridine ring as an antibonding π* orbital. As (8) contains two nucleophilic nitrogen lone pairs, NBO calculations were performed to assess the charges at nitrogen and determine which might be the more reactive. Results show that the partial charges of the pyridine and dimethylamino nitrogen atoms are -0.232 and -0.446 respectively, suggesting a larger accumulation of electron density on the latter group which should make it the more reactive nucleophile.

HOMO and LUMO of 8. Left: LUMO Right: HOMO. Structures optimised at the B3LYP level and 6- 31G(d) basis set. Diazaborole.png
HOMO and LUMO of 8. Left: LUMO Right: HOMO. Structures optimised at the B3LYP level and 6- 31G(d) basis set.

Metal-free H2-activation

Borole H2 activate.png

Given the many studies on frustrated Lewis pairs (FLP) that point them towards high potential small molecule activators, Piers and coworkers set out to investigate whether the strong electrophilicity of antiaromatic boroles might entail a comparable reactivity.

In a seminal paper published in 2010, they successfully demonstrated the metal-free activation of dihydrogen by [PhBC4Ph4] and its perfluorinated analog. [30] Their studies initially focused on perfluorinated [PhBC4Ph4] due to its exceptionally high Lewis acid strength, which readily reacted with H2 both in solution and in the solid state to form two possible isomers as shown above. DFT calculations show that the trans product is thermodynamically favoured by 6.2 kcal mol−1, but no interconversion between isomers was observed under thermal conditions. A plausible reaction mechanism involving a borole H2 adduct was proposed on the basis of the observed isomers ratios and theoretical studies. The authors suggested that the driving force is most likely provided by the disruption of antiaromaticity in the borole ring through ring opening. Even the less Lewis acidic borole 1 was shown to be capable of facile H–H bond cleavage to give similar reaction products (trans/cis: 1 : 4.3). Most likely, the unique combination of antiaromaticity and high Lewis acidity enables the metal-free activation of H2 by boroles.

Sandwich complexes

Since the borole dianion is isoelectronic to the cyclopentadienide anion, it also exhibits capability to from ferrocene-like sandwich compounds [31] through an η5 coordination.

Boroles have been experimentally shown to form stable complexes other group 13 elements, such as aluminium and galium. [6] Both were synthesised from the neutral substituted borole and the corresponding metal (I)-Cp* complex in benzene as summarised in below.

Synthesis of the Al and Ga sandwich complexes. Synthetic Scheme.png
Synthesis of the Al and Ga sandwich complexes.

In the Aluminium complex, an Al (III) ion is sandwiched between a Cp* and a substituted borolediide dianion, giving an overall neutral structure. Sindlinger [6] reported that examining the orbitals of the simpler hypothetical model complex (C4BH5)Al(C5H5) revealed similar features to the fully substituted compound. The successful transfer of two electrons onto the borole ring becomes apparent from the borole‐based HOMO essentially being identical with the LUMO in free borole. The oxidation of Al is further in line with Bader charges of +2.29 at Al and -0.78 on the borole (C4B) unit. However, this charge resides on the butadiene backbone as opposed to on boron (Cβ −0.24; Cα −0.99; B +1.68). In contrast, the charge accumulated on the central (C5)‐Cp* moiety amounting to −1.17 is equally distributed among the five carbon atoms. Topology analysis was performed using QTAIM, which revealed no bond critical point between Al and B. In line with a strong localisation of electron density on Cα, bond critical points are only found for the between Al and Cα (delocalization index, DI=0.25) but not between Al and Cβ (DI=0.11) as shown below.

Bond Critical Point analysis on
(C4BH5)Al(C5H5). Note that bond critical points are represented by the orange dots. Ca and Cb are numbered (3, 12 ) and (18, 20) respectively. Left: Back view. Right: Side view. Molecules were optimised using BP86 functional and def2-TZVPP basis sets and analyses performed with multiwfn. Alborole sandwich.png
Bond Critical Point analysis on (C4BH5)Al(C5H5). Note that bond critical points are represented by the orange dots. Cα and Cβ are numbered (3, 12 ) and (18, 20) respectively. Left: Back view. Right: Side view. Molecules were optimised using BP86 functional and def2-TZVPP basis sets and analyses performed with multiwfn.

In contrast, the gallium analogue forms a Lewis‐base adduct with a dative Ga−B bond rather than the neutral heteroleptic case for Aluminium, suggesting that Ga still retains the +1 oxidation state. [6]

Model complex used for quantum chemical calculations Model Germanium borole complex.png
Model complex used for quantum chemical calculations

Furthermore, synthetic evidence for a neutral η5‐aminoborole germanium(II) complex half sandwich has been reported [32] as illustrated to the right. Tholen et al. performed DFT calculations at the M06‐2X/Def2‐TZVP level of theory and determined that the model complex matched up closely with experimental data determined experimentally by X-Ray Diffraction (XRD), with a maximum of 1.6% deviation in atomic distances and angles. Using the model compound, its frontier orbitals were calculated and they reveal covalent bonding interactions between the apical germanium atom and the borole base. As shown in the figure to the right, the HOMO is mainly located at the nitrogen atom while HOMO‐1 and HOMO‐2 are bonding combinations of germanium 4px and 4py orbitals with the π orbitals of the borole ring. HOMO‐3 is the antibonding combination of mainly the germanium 4s and 4pz orbitals and the lowest π orbital of the borole ring.

Isosurfaces of Germanium borole complex at 0.05. Top left to right: HOMO-3, HOMO-2, HOMO-1. Bottom left to right: HOMO, LUMO LUMO+1 Gebor.png
Isosurfaces of Germanium borole complex at 0.05. Top left to right: HOMO-3, HOMO-2, HOMO-1. Bottom left to right: HOMO, LUMO LUMO+1

Performing natural bond analysis (NBO) on the model pointed to significant orbital interactions between the apical germanium atom and the borole basis as reflected by Wiberg bond indices (WBIs) between the ring carbon and boron atoms and the apical germanium atom (between 0.31 and 0.42). For reference, these values are very similar to those of the germanium (II) Cp* cation complex (WBI(GeC)=0.29) and significantly larger than those between lithium ions and the ring atoms of the borole ring in the inverse sandwich complex Li2[H2NBC4H4]. Ionic interactions prevail (WBI(LiC)=WBI(LiB)=0.02) in the latter complex. While the calculated charge distribution for Li2[H2NBC4H4] has an expected highly negative charge (−1.90 a.u.) on the aminoborole unit, the same unit has a much less negative charge in the aminoborole germanium complex (−0.83 a.u.). The increased charge transfer from the aminoborole unit to the germanium atom is also reflected by the uniformly smaller WBI indices between the ring atoms compared those in Li2[H2NBC4H4]. In line with the NBO analysis results, quantum theory of atoms in molecules (QTAIM) analysis predicted a cage‐like molecular topology for the model compound, with bond paths between all ring carbon atoms and the germanium atom as shown below.

Topology Analysis of the model germanium half sandwich complex. Bond CP (3, -1): Orange. Ring CP (3, +1): Yellow. Cage CP (3, +3): Green. Left: Side view. Right: Top view. Geborbcpf.png
Topology Analysis of the model germanium half sandwich complex. Bond CP (3, -1): Orange. Ring CP (3, +1): Yellow. Cage CP (3, +3): Green. Left: Side view. Right: Top view.

Further theoretical studies have also been conducted at the M06-2X/Def2-TZVP level of theory to investigate the stability of half sandwich complexes between C4H4BNH2 with other group 14 elements (C, Si, Sn, Pb), [17] where the borole ring binds to the divalent metal cation in η5 mode. AIM analysis was performed and the laplacian for C is < 0, while the other group 14 elements had values > 0. The former indicates a substantial amount of covalency, while the latter confirms the presence of a polar covalent interaction.

See also

Papapishu-Lab-icon-6.svg  Chemistryportal

Related Research Articles

<span class="mw-page-title-main">Diels–Alder reaction</span> Chemical reaction

In organic chemistry, the Diels–Alder reaction is a chemical reaction between a conjugated diene and a substituted alkene, commonly termed the dienophile, to form a substituted cyclohexene derivative. It is the prototypical example of a pericyclic reaction with a concerted mechanism. More specifically, it is classified as a thermally-allowed [4+2] cycloaddition with Woodward–Hoffmann symbol [π4s + π2s]. It was first described by Otto Diels and Kurt Alder in 1928. For the discovery of this reaction, they were awarded the Nobel Prize in Chemistry in 1950. Through the simultaneous construction of two new carbon–carbon bonds, the Diels–Alder reaction provides a reliable way to form six-membered rings with good control over the regio- and stereochemical outcomes. Consequently, it has served as a powerful and widely applied tool for the introduction of chemical complexity in the synthesis of natural products and new materials. The underlying concept has also been applied to π-systems involving heteroatoms, such as carbonyls and imines, which furnish the corresponding heterocycles; this variant is known as the hetero-Diels–Alder reaction. The reaction has also been generalized to other ring sizes, although none of these generalizations have matched the formation of six-membered rings in terms of scope or versatility. Because of the negative values of ΔH° and ΔS° for a typical Diels–Alder reaction, the microscopic reverse of a Diels–Alder reaction becomes favorable at high temperatures, although this is of synthetic importance for only a limited range of Diels-Alder adducts, generally with some special structural features; this reverse reaction is known as the retro-Diels–Alder reaction.

<span class="mw-page-title-main">Conjugated system</span> System of connected p-orbitals with delocalized electrons in a molecule

In theoretical chemistry, a conjugated system is a system of connected p-orbitals with delocalized electrons in a molecule, which in general lowers the overall energy of the molecule and increases stability. It is conventionally represented as having alternating single and multiple bonds. Lone pairs, radicals or carbenium ions may be part of the system, which may be cyclic, acyclic, linear or mixed. The term "conjugated" was coined in 1899 by the German chemist Johannes Thiele.

In organic chemistry, a cycloaddition is a chemical reaction in which "two or more unsaturated molecules combine with the formation of a cyclic adduct in which there is a net reduction of the bond multiplicity". The resulting reaction is a cyclization reaction. Many but not all cycloadditions are concerted and thus pericyclic. Nonconcerted cycloadditions are not pericyclic. As a class of addition reaction, cycloadditions permit carbon–carbon bond formation without the use of a nucleophile or electrophile.

The 1,3-dipolar cycloaddition is a chemical reaction between a 1,3-dipole and a dipolarophile to form a five-membered ring. The earliest 1,3-dipolar cycloadditions were described in the late 19th century to the early 20th century, following the discovery of 1,3-dipoles. Mechanistic investigation and synthetic application were established in the 1960s, primarily through the work of Rolf Huisgen. Hence, the reaction is sometimes referred to as the Huisgen cycloaddition. 1,3-dipolar cycloaddition is an important route to the regio- and stereoselective synthesis of five-membered heterocycles and their ring-opened acyclic derivatives. The dipolarophile is typically an alkene or alkyne, but can be other pi systems. When the dipolarophile is an alkyne, aromatic rings are generally produced.

In chemistry, frontier molecular orbital theory is an application of molecular orbital theory describing HOMO–LUMO interactions.

<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.

The inverse electron demand Diels–Alder reaction, or DAINV or IEDDA is an organic chemical reaction, in which two new chemical bonds and a six-membered ring are formed. It is related to the Diels–Alder reaction, but unlike the Diels–Alder reaction, the DAINV is a cycloaddition between an electron-rich dienophile and an electron-poor diene. During a DAINV reaction, three pi-bonds are broken, and two sigma bonds and one new pi-bond are formed. A prototypical DAINV reaction is shown on the right.

A metal-centered cycloaddition is a subtype of the more general class of cycloaddition reactions. In such reactions "two or more unsaturated molecules unite directly to form a ring", incorporating a metal bonded to one or more of the molecules. Cycloadditions involving metal centers are a staple of organic and organometallic chemistry, and are involved in many industrially-valuable synthetic processes.

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

A borylene is the boron analogue of a carbene. The general structure is R-B: with R an organic moiety and B a boron atom with two unshared electrons. Borylenes are of academic interest in organoboron chemistry. A singlet ground state is predominant with boron having two vacant sp2 orbitals and one doubly occupied one. With just one additional substituent the boron is more electron deficient than the carbon atom in a carbene. For this reason stable borylenes are more uncommon than stable carbenes. Some borylenes such as boron monofluoride (BF) and boron monohydride (BH) the parent compound also known simply as borylene, have been detected in microwave spectroscopy and may exist in stars. Other borylenes exist as reactive intermediates and can only be inferred by chemical trapping.

<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">Phosphirenium ion</span> Series of chemical compounds

Phosphirenium ions are a series of organophosphorus compounds containing unsaturated three-membered ring phosphorus (V) heterocycles and σ*-aromaticity is believed to be present in such molecules. Many of the salts containing phosphirenium ions have been isolated and characterized by NMR spectroscopy and X-ray crystallography.

<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.

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

Tellurophenes are the tellurium analogue of thiophenes and selenophenes.

<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">Triboracyclopropenyl</span>

The triboracyclopropenyl fragment is a cyclic structural motif in boron chemistry, named for its geometric similarity to cyclopropene. In contrast to nonplanar borane clusters that exhibit higher coordination numbers at boron (e.g., through 3-center 2-electron bonds to bridging hydrides or cations), triboracyclopropenyl-type structures are rings of three boron atoms where substituents at each boron are also coplanar to the ring. Triboracyclopropenyl-containing compounds are extreme cases of inorganic aromaticity. They are the lightest and smallest cyclic structures known to display the bonding and magnetic properties that originate from fully delocalized electrons in orbitals of σ and π symmetry. Although three-membered rings of boron are frequently so highly strained as to be experimentally inaccessible, academic interest in their distinctive aromaticity and possible role as intermediates of borane pyrolysis motivated extensive computational studies by theoretical chemists. Beginning in the late 1980s with mass spectrometry work by Anderson et al. on all-boron clusters, experimental studies of triboracyclopropenyls were for decades exclusively limited to gas-phase investigations of the simplest rings (ions of B3). However, more recent work has stabilized the triboracyclopropenyl moiety via coordination to donor ligands or transition metals, dramatically expanding the scope of its chemistry.

<span class="mw-page-title-main">9-Borafluorene</span> Class of chemical compounds

9-borafluorenes are a class of boron-containing heterocycles consisting of a tricyclic system with a central BC4 ring with two fused arene groups. 9-borafluorenes can be thought of as a borole with two fused arene rings, or as a trigonal planar boron atom with an empty p orbital bridging two biphenyl rings. However, 9-borafluorenes are generally less reactive than boroles due to less antiaromatic character and Lewis acidity. Containing highly conjugated π systems, 9-borafluorenes possess interesting photophysical properties. In addition, 9-borafluorenes are good Lewis acids. This combination of properties enables potential uses such as in light-emitting materials, solar cells, and sensors for some molecules.

<span class="mw-page-title-main">Superelectrophilic anion</span> Superelectrophilic anions

Superelectrophilic anions are a class of molecular ions that exhibit highly electrophilic reaction behavior despite their overall negative charge. Thus, they are even able to bind the unreactive noble gases or molecular nitrogen at room temperature. The only representatives known so far are the fragment ions of the type [B12X11] derived from the closo-dodecaborate dianions [B12X12]2–. X represents a substituent connected to a boron atom (cf. Fig. 1). For this reason, the following article deals exclusively with superelectrophilic anions of this type.

<span class="mw-page-title-main">Boraacenes</span> Boron containing acene compounds

Boraacenes are polycyclic aromatic hydrocarbons containing at least one boron atom. Structurally, they are related to acenes, linearly fused benzene rings. However, the boron atom is electron deficient and may act as a Lewis Acid when compared to carbon. This results in slightly less negative charge within the ring, smaller HOMO-LUMO gaps, as well as differences in redox chemistry when compared to their acene analogues. When incorporated into acenes, Boron maintains the planarity and aromaticity of carbon acenes, while adding an empty p-orbital, which can be utilized for the fine tuning of organic semiconductor band gaps. Due to this empty p orbital, however, it is also highly reactive when exposed to nucleophiles like water or normal atmosphere, as it will readily be attacked by oxygen, which must be addressed to maintain its stability.

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

1,3-Diphospha-2,4-diboretanes, or B2P2, is a class of 4-member cyclic compounds of alternating boron and phosphorus atoms. They are often found as dimers during the synthesis of boraphosphenes (RB=PR'). Compounds can exhibit localized singlet diradical character (diradicaloid) between the boron atoms in the solution and solid state.

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

Aluminylenes are a sub-class of aluminium(I) compounds that feature singly-coordinated aluminium atoms with a lone pair of electrons. As aluminylenes exhibit two unoccupied orbitals, they are not strictly aluminium analogues of carbenes until stabilized by a Lewis base to form aluminium(I) nucleophiles. The lone pair and two empty orbitals on the aluminium allow for ambiphilic bonding where the aluminylene can act as both an electrophile and a nucleophile. Aluminylenes have also been reported under the names alumylenes and alanediyl.

References

  1. Alan R. Katritzky, ed. (1993), "The Concept of Aromaticity in Heterocyclic Chemistry", Advances in Heterocyclic Chemistry Volume 56, vol. 56, Academic Press, p. 375, doi:10.1016/S0065-2725(08)60196-8, ISBN   978-0-12-020756-5 , retrieved 2010-03-13
  2. Herberich, Gerhard E.; Englert, Ulli; Hostalek, Martin; Laven, Ralf (1991). "Derivate des Borols, XVI Bis(borol)nickel-Komplexe2)". Chemische Berichte. 124 (1): 17–23. doi:10.1002/cber.19911240104. ISSN   1099-0682.
  3. 1 2 Braunschweig, Holger; Kupfer, Thomas (2011-09-27). "Recent developments in the chemistry of antiaromatic boroles". Chemical Communications. 47 (39): 10903–10914. doi:10.1039/C1CC13071D. ISSN   1364-548X. PMID   21735004.
  4. 1 2 Braunschweig, Holger; Fernández, Israel; Frenking, Gernot; Kupfer, Thomas (2008). "Structural Evidence for Antiaromaticity in Free Boroles". Angewandte Chemie International Edition. 47 (10): 1951–1954. doi: 10.1002/anie.200704771 . ISSN   1521-3773. PMID   18228230.
  5. 1 2 Lo, Po-Kam; Lau, Kai-Chung (2011-02-10). "High-Level ab Initio Predictions for the Ionization Energies and Heats of Formation of Five-Membered-Ring Molecules: Thiophene, Furan, Pyrrole, 1,3-Cyclopentadiene, and Borole, C4H4X/C4H4X+ (X = S, O, NH, CH2, and BH)". The Journal of Physical Chemistry A. 115 (5): 932–939. Bibcode:2011JPCA..115..932L. doi:10.1021/jp110499c. ISSN   1089-5639. PMID   21210670.
  6. 1 2 3 4 5 6 Sindlinger, Christian P.; Ruth, Paul Niklas (2019). "A Neutral "Aluminocene" Sandwich Complex: η1- versus η5-Coordination Modes of a Pentaarylborole with ECp* (E=Al, Ga; Cp*=C5Me5)". Angewandte Chemie International Edition. 58 (42): 15051–15056. doi:10.1002/anie.201907749. ISSN   1521-3773. PMC   6856865 . PMID   31390132.
  7. 1 2 3 Herberich, G. E.; Buller, B.; Hessner, B.; Oschmann, W. (1980-08-26). "Derivative des borols: II. Pentaphenylborol: Synthese, reduktion zum dianion und komplexe von kobalt und platin". Journal of Organometallic Chemistry. 195 (3): 253–259. doi:10.1016/S0022-328X(00)93308-4. ISSN   0022-328X.
  8. So, Cheuk-Wai; Watanabe, Daisuke; Wakamiya, Atsushi; Yamaguchi, Shigehiro (2008-07-01). "Synthesis and Structural Characterization of Pentaarylboroles and Their Dianions". Organometallics. 27 (14): 3496–3501. doi:10.1021/om8002812. ISSN   0276-7333.
  9. Braunschweig, Holger; Chiu, Ching-Wen; Wahler, Johannes; Radacki, Krzysztof; Kupfer, Thomas (2010). "Chemical Reduction and Dimerization of 1-Chloro-2,3,4,5-tetraphenylborole". Chemistry – A European Journal. 16 (40): 12229–12233. doi:10.1002/chem.201001556. ISSN   1521-3765. PMID   20839188.
  10. 1 2 3 Eisch, John J.; Hota, Nalini K.; Kozima, Sinpei (1969-07-01). "Synthesis of pentaphenylborole, a potentially antiaromatic system". Journal of the American Chemical Society. 91 (16): 4575–4577. doi:10.1021/ja01044a059. ISSN   0002-7863.
  11. Eisch, John J.; Galle, James E.; Kozima, Sinpei. (1986-02-01). "Bora-aromatic systems. Part 8. The physical and chemical consequences of cyclic conjugation in boracyclopolyenes. The antiaromatic character of pentaarylboroles". Journal of the American Chemical Society. 108 (3): 379–385. doi:10.1021/ja00263a006. ISSN   0002-7863. PMID   22175451.
  12. Francis Gordon Albert Stone, Robert West, ed. (1996), Advances in Heterocyclic Chemistry, vol. 39, Academic Press, p. 380, ISBN   978-0-12-031139-2 , retrieved 2010-03-13
  13. Fan, Cheng; Piers, Warren E.; Parvez, Masood (2009). "Perfluoropentaphenylborole". Angewandte Chemie International Edition. 48 (16): 2955–2958. doi:10.1002/anie.200805865. ISSN   1521-3773. PMID   19145617.
  14. Braunschweig, Holger; Kupfer, Thomas (2008-09-18). "Direct functionalization at the boron center of antiaromatic chloroborole". Chemical Communications (37): 4487–4489. doi:10.1039/B808483A. ISSN   1364-548X. PMID   18802599.
  15. Fagan, Paul J.; Burns, Elizabeth G.; Calabrese, Joseph C. (1988-04-27). "Synthesis of boroles and their use in low-temperature Diels-Alder reactions with unactivated alkenes". Journal of the American Chemical Society. 110 (9): 2979–2981. doi:10.1021/ja00217a053. ISSN   0002-7863.
  16. Fagan, Paul J.; Nugent, William A.; Calabrese, Joseph C. (1994-03-01). "Metallacycle Transfer from Zirconium to Main Group Elements: A Versatile Synthesis of Heterocycles". Journal of the American Chemical Society. 116 (5): 1880–1889. doi:10.1021/ja00084a031. ISSN   0002-7863.
  17. 1 2 Rohman, Shahnaz S.; Kashyap, Chayanika; Ullah, Sabnam S.; Guha, Ankur K. (2019-09-15). "Viability of half sandwich complexes of borole with group 14 (II) ions: Structure, stability and reactivity". Polyhedron. 170: 1–6. doi:10.1016/j.poly.2019.05.023. ISSN   0277-5387. S2CID   182856373.
  18. Lee, Vladimir Ya.; Sugasawa, Haruka; Gapurenko, Olga A.; Minyaev, Ruslan M.; Minkin, Vladimir I.; Gornitzka, Heinz; Sekiguchi, Akira (2018-05-16). "From Borapyramidane to Borole Dianion". Journal of the American Chemical Society. 140 (19): 6053–6056. doi:10.1021/jacs.8b03473. ISSN   0002-7863. PMID   29669416.
  19. Herberich, Gerhard E.; Hostalek, Martin; Laven, Ralf; Boese, Roland (1990). "Borole Dianions: Metalation of 1-(Dialkylamino)-2,5-dihydro-lH-boroles and the Structure of Li2 (C4 H4 BNEt2)·TMEDA". Angewandte Chemie International Edition in English. 29 (3): 317–318. doi:10.1002/anie.199003171. ISSN   1521-3773.
  20. Herberich, Gerhard E.; Negele, Michael; Ohst, Holger (1991). "Derivate des Borols, XVII η5-[1-(Diisopropylamino)borol]metall-Komplexe: Synthesen, Protonierung, interne Rotation". Chemische Berichte. 124 (1): 25–29. doi:10.1002/cber.19911240105. ISSN   1099-0682.
  21. Braunschweig, Holger; Chiu, Ching-Wen; Radacki, Krzysztof; Brenner, Peter (2010-01-25). "Platinum substituted boroles". Chemical Communications. 46 (6): 916–918. doi:10.1039/B923652J. ISSN   1364-548X. PMID   20107649.
  22. Braunschweig, Holger; Damme, Alexander; Gamon, Daniela; Kupfer, Thomas; Radacki, Krzysztof (2011-05-16). "Synthesis and Coordination Chemistry of 1-Cymantrenyl-2,3,4,5-tetraphenylborole". Inorganic Chemistry. 50 (10): 4250–4252. doi:10.1021/ic200559d. ISSN   0020-1669. PMID   21506525.
  23. Ansorg, Kay; Braunschweig, Holger; Chiu, Ching-Wen; Engels, Bernd; Gamon, Daniela; Hügel, Markus; Kupfer, Thomas; Radacki, Krzysztof (2011). "The Pentaphenylborole–2,6-Lutidine Adduct: A System with Unusual Thermochromic and Photochromic Properties". Angewandte Chemie International Edition. 50 (12): 2833–2836. doi:10.1002/anie.201006234. ISSN   1521-3773. PMID   21387499.
  24. Wang, Zheng; Zhou, Yu; Lee, Ka-Ho; Lam, Wai Han; Dewhurst, Rian D.; Braunschweig, Holger; Marder, Todd B.; Lin, Zhenyang (2017). "DFT Studies of Dimerization Reactions of Boroles". Chemistry – A European Journal. 23 (48): 11587–11597. doi:10.1002/chem.201701737. ISSN   1521-3765. PMID   28627022.
  25. Baker, J. J.; Al Furaiji, Khadilah H. M.; Liyanage, O. Tara; Wilson, David J. D.; Dutton, Jason L.; Martin, Caleb D. (2019). "Diverse Reactivity of Dienes with Pentaphenylborole and 1-Phenyl-2,3,4,5-Tetramethylborole Dimer". Chemistry – A European Journal. 25 (6): 1581–1587. doi: 10.1002/chem.201805151 . ISSN   1521-3765. PMID   30457687.
  26. Eisch, John J.; Galle, James E.; Shafii, Babak; Rheingold, Arnold L. (1990-08-01). "Bora-aromatic systems. 12. Thermal generation and transformation of the borepin ring system: a paradigm of pericyclic processes". Organometallics. 9 (8): 2342–2349. doi:10.1021/om00158a035. ISSN   0276-7333.
  27. Fan, Cheng; Piers, Warren E.; Parvez, Masood; McDonald, Robert (2010-11-08). "Divergent Reactivity of Perfluoropentaphenylborole with Alkynes". Organometallics. 29 (21): 5132–5139. doi:10.1021/om100334r. ISSN   0276-7333.
  28. Su, Xiaojun; Baker, J. J.; Martin, Caleb D. (2020). "Dimeric boroles: effective sources of monomeric boroles for heterocycle synthesis". Chemical Science. 11 (1): 126–131. doi: 10.1039/C9SC04053F . ISSN   2041-6539. PMC   7012074 . PMID   32110363.
  29. 1 2 Lindl, Felix; Lin, Shujuan; Krummenacher, Ivo; Lenczyk, Carsten; Stoy, Andreas; Müller, Marcel; Lin, Zhenyang; Braunschweig, Holger (2019). "1,2,3-Diazaborinine: A BN Analogue of Pyridine Obtained by Ring Expansion of a Borole with an Organic Azide". Angewandte Chemie International Edition. 58 (1): 338–342. doi:10.1002/anie.201811601. ISSN   1521-3773. PMID   30394650. S2CID   205407951.
  30. Fan, Cheng; Mercier, Lauren G.; Piers, Warren E.; Tuononen, Heikki M.; Parvez, Masood (2010-07-21). "Dihydrogen Activation by Antiaromatic Pentaarylboroles". Journal of the American Chemical Society. 132 (28): 9604–9606. doi:10.1021/ja105075h. ISSN   0002-7863. PMID   20583845.
  31. Alan R. Katritzky, ed. (2001), Advances in Heterocyclic Chemistry, vol. 79, Academic Press, pp. 169–170, ISBN   978-0-12-020779-4 , retrieved 2010-03-13
  32. Tholen, Patrik; Dong, Zhaowen; Schmidtmann, Marc; Albers, Lena; Müller, Thomas (2018). "A Neutral η5-Aminoborole Complex of Germanium(II)". Angewandte Chemie International Edition. 57 (40): 13319–13324. doi:10.1002/anie.201808271. ISSN   1521-3773. PMID   30070743. S2CID   51891840.