Gallium monoiodide

Last updated
Gallium monoiodide
Identifiers
3D model (JSmol)
ChemSpider
PubChem CID
  • InChI=1S/Ga.HI/h;1H/q+1;/p-1
    Key: LRPWSMQGXLANTG-UHFFFAOYSA-M
  • [Ga]I
Properties
GaI
Molar mass 196.63 g/mol
AppearancePale green solid
Hydrolysis
Except where otherwise noted, data are given for materials in their standard state (at 25 °C [77 °F], 100 kPa).

Gallium monoiodide is an inorganic gallium compound with the formula GaI or Ga4I4. It is a pale green solid and mixed valent gallium compound, which can contain gallium in the 0, +1, +2, and +3 oxidation states. It is used as a pathway for many gallium-based products. Unlike the gallium(I) halides first crystallographically characterized, [1] gallium monoiodide has a more facile synthesis allowing a synthetic route to many low-valent gallium compounds.

Contents

Synthesis

In 1990, Malcolm Green synthesized gallium monoiodide by the ultrasonication of liquid gallium metal with iodine in toluene yielding a pale green powder referred to as gallium monoiodide. [2] The chemical composition of gallium monoiodide was not determined until the early to mid-2010s despite its simple synthesis.

In 2012, the pale green gallium monoiodide was determined to be a combination of gallium metal and gallium(I,III) iodide, having the chemical composition [Ga0]2[Ga+][GaI4]. [3] However, in 2014, it was found that the incomplete reaction of gallium metal with iodine yielded gallium monoiodide with this chemical composition. Gallium monoiodide synthesized with longer reaction times for complete reaction had a different chemical composition [Ga0]2[Ga+]2[Ga2I62-]. [4]

The resultant gallium monoiodide is highly air sensitive, but stable under inert atmosphere conditions for up to a year at -35 ˚C. [4]

Characterization

When gallium monoiodide was first produced, it was proposed that gallium monoiodide is a combination of gallium metal, Ga2I3 and Ga2I4 based on the characteristic Raman spectra of these constituents. [5] This hypothesis was confirmed as two variants of gallium monoiodide were determined to have the chemical compositions [Ga0]2[Ga+][GaI4], simplified as Ga2I4·2Ga, and [Ga0]2[Ga+]2[Ga2I62-], simplified as Ga2I3·Ga. [4] [3]

When the incompletely reacted product was probed by NMR spectroscopy, it showed the presence gallium metal. [3] When probed by 127I NQR, [4] it showed the presence of Ga2I4 and further confirms the [Ga0]2[Ga+][GaI4] assignment. [6] Raman spectroscopy has also confirmed this composition assignment. [4] [7] All of the evidence from other spectroscopic methods, and power x-ray diffraction patterns, validates the assignment of [Ga0]2[Ga+][GaI4] for the incompletely reacted gallium monoiodide variant.

When the completely reacted product was probed by 127I NQR, it showed the presence of Ga2I3. [4] Raman spectroscopy has also confirmed this assignment, as it aligned with those from a Ga4I6 reference. [4] [7] Finally, power x-ray diffraction supports that this gallium monoiodide variant matches that of characteristic Ga2I3, which is different from that of GaI2. [4]

[Ga0]2[Ga+][GaI4] converts to [Ga0]2[Ga+]2[Ga2I62-] over time. [4]

Reactions and derivatives

Gallium monoiodide is used as a precursor for a variety of reactions, acting as a lewis acid and a reducing agent. Early-on, gallium monoiodide was shown to produce alkylgallium diiodides via oxidative addition by reacting liquid gallium metal and iodine in the presence of an alkyl iodide. [1] [2] [8] Since then, other organogallium complexes have been synthesized, as well as Lewis base adducts and gallium based clusters. [8]

Gallium Lewis base adducts

Reaction pathways of various gallium monoiodide Lewis base adducts. Reactions were conducted in toluene at - 78 @C. (L = phosphines, ethers, amines). GaI Lewis base adduct.png
Reaction pathways of various gallium monoiodide Lewis base adducts. Reactions were conducted in toluene at - 78 ˚C. (L = phosphines, ethers, amines).

Gallium monoiodide reacts with various monodentate Lewis bases to form Ga(II), Ga(III), or mixed valent compounds, as well as gallium-based dimers and trimers. For example, gallium monoiodide can react with primary, secondary, and tertiary amines, secondary or tertiary phosphines or ethers to form Ga(II)-Ga(II) dimers. [2] [8] [10] Gallium monoiodide can also react with triphenylphosphine (PPh3) to form Ga(III)I3PPh3. [2] It also reacts with the less sterically hindered triethylphosphine (PEt3) to form a Ga(II)-Ga(I)-Ga(II) mixed valent complex with datively coordinated PEt3 ligands. [8] [10] These reactions are believed to be a disproportionation, as gallium metal is produced in these reactions. [9]

Gallium monoiodide reacts with triphenylstibine to produce an SbPh3 fragment datively bonded to a GaPhI2 fragment. [11] The difference in reactivity between PPh3 and SbPh3, a heavy atom analogue of PPh3, can be attributed to a weaker Sb-C bond, allowing for transfer of a phenyl group from antimony to gallium. This suggests that gallium monoiodide can be used as a reducing agent as well. [8] [11]

SbPh3 datively bonded to GaPhI2.png

N-heterocyclic carbenes reacts with gallium monoiodide to form a complex with a sterically hindered isopropyl ligand. [9] However, gallium monoiodide reacts with diazabutadienes and subsequent reduction by potassium metal to form Ga analogs of N-heterocyclic carbenes. [8] Other Ga-based carbenes can be produced from gallium monoiodide precursor using Li(NacNac). [8]

Gallium monoiodide reacts with multidentate Lewis bases, such as bipyridine, phenyl-terpyridine, and bis(imino)pyridine ligands to form Ga(III) complexes. [8] [12] Crystallographically, the bipyridine derivative has a distorted octahedral geometry, with a Ga–N bond length of 2.063 Å. The phenyl-terpyridine derivative adopts a distorted trigonal bipyramidal geometry where the two equatorial Ga–N bonds (as drawn) are longer than the axial Ga-N bond, with 2.104 Å and 2.007(5) Å, respectively. The average Ga-N bond length (2.071 Å) is similar to that of a neutral GaCl3(terpy) Lewis base adduct (2.086 Å). [13] The bis(imino)pyridine derivative has a distorted square-based pyramidal geometry. Like for the phenyl-terpyridine derivative, the equatorial imino Ga-N bonds (2.203 Å) are longer than the axial pyridyl Ga-N bond (2.014(7) A˚). [12] Despite these similar reactivities and bond characteristics, when gallium monoiodide was reacted with imino-substituted pyridines (RN=C(H)Py), unique reactivity was observed. Reductive coupling of the imino-substituted pyridines formed diamido-digallium(III) complexes. [12] These reactions display the ability of gallium monoiodides to form new C-C bonds.

Reaction pathways with gallium monoiodide and polydentate Lewis bases form Ga(III) salts (R = Ar, Bu ; Ar = C6H3Pr2-2,6; Py = 2-pyridyl). Reactions were conducted in toluene at - 78 @C. Only the bis(imino)pyridine derivative was reacted at 25 @C. All complexes have been crystallographically characterized. GaI polydentate Lewis base adduct.png
Reaction pathways with gallium monoiodide and polydentate Lewis bases form Ga(III) salts (R = Ar, Bu ; Ar = C6H3Pr2-2,6; Py = 2-pyridyl). Reactions were conducted in toluene at - 78 ˚C. Only the bis(imino)pyridine derivative was reacted at 25 ˚C. All complexes have been crystallographically characterized.

Gallium heterocycles

Gallium monoiodide can also be used as a precursor to form gallium-based heterocycles. Reactions with diazabutadienes, {RN=C(H)}2, forms monomers or dimers based on the substituents on the diazabutadienes. More sterically hindered substituents such as tert-butyl have resulted in the formation of gallium(II) dimers, whereas reactions with alkyl or aryl substituted diazabutadienes have formed Ga(III) monomers. [8] Gallium monoiodide can be reacted with phenyl-substituted 1,4-diazabuta-1,3-dienes to form a gallium heterocycle with a diazabutadiene monoanion. [14] EPR spectroscopy has revealed that the diazabutadiene fragment is a paramagnetic monoanionic species rather than an ene-diamido dianion or a neutral ligand. [14] Thus, gallium monoiodide undergoes a disproportionation reaction to form a gallium(III) complex with deposition of a gallium metal. [8] [14] Upon further reaction with a 1,4-dilithiated diazabutadiene, this gallium heterocycle forms a new complex with the diazabutadiene monoanion fragment datively bonded to the gallium center and an ene-diamido dianion covalently bonded to the Ga center. [14]

Ga heterocycles formed from the reaction of gallium monoiodide with 1,4-diazabuta-1,3-dienes. R = 2,6-dimethylphenyl; 2,4,6-trimethylphenyl; 2,6-diisopropylphenyl. For the 1,4-dilithiated diazabutadiene reagent, R = 2,6-dimethylphenyl. Ga heterocycles.png
Ga heterocycles formed from the reaction of gallium monoiodide with 1,4-diazabuta-1,3-dienes. R = 2,6-dimethylphenyl; 2,4,6-trimethylphenyl; 2,6-diisopropylphenyl. For the 1,4-dilithiated diazabutadiene reagent, R = 2,6-dimethylphenyl.

One very important reactivity of this gallium(III) heterocycle is its ability to access gallium analogues of N-heterocyclic carbenes upon reduction with potassium metal. [15] Although a gallium analogue of N-heterocyclic carbenes had been synthesized previously, [16] having access to heavier analogues of N-heterocylic carbenes from a synthetically more facile gallium monoiodide route has opened new avenues in coordination chemistry, such as access to new Ga-M bonds. [17] [18] [19]

Gallium monoiodide can also be used to access six-membered gallium(I) heterocycles that have parallels to gallium analogues of N-heterocyclic carbenes. These neutral gallium(I) heterocycles can be synthesized by reacting gallium monoiodide and Li[nacnac]. [18] [20]

Reaction of gallium monoiodide (slurry) and Li[nacnac] in a dry ice/acetone bath to access a gallium(I) heterocycle. Excess potassium metal can be added to circumvent a Ga(II) derivative of the six-member gallium(I) heterocycle. Ar = Dipp. Ga 6-member NHC new.png
Reaction of gallium monoiodide (slurry) and Li[nacnac] in a dry ice/acetone bath to access a gallium(I) heterocycle. Excess potassium metal can be added to circumvent a Ga(II) derivative of the six-member gallium(I) heterocycle. Ar = Dipp.

Cyclopentadienyl complexes

Gallium monoiodide can easily be converted to half-sandwich complexes, (pentamethylcyclopentadienyl)gallium(I) and cyclopentadienylgallium. [21] (Pentamethylcyclopentadienyl)gallium(I) can be easily produced by reacting gallium monoiodide with a potassium salt of the desired ligand under toluene to avoid side prodcuts. [22] [8]

Cyclopentadienylgallium, which is less sterically hindered than (pentamethylcyclopentadienyl)gallium(I), can also be accessed using a gallium monoiodide. This ligand can be synthesized with a metathesis reaction of NaCp with gallium monoiodide. [23] This cyclopentadienylgallium ligand has been used to access a GaCp2I complex with datively bonded cyclopentadienylgallium. This complex showcases an uncommon donor-acceptor Ga-Ga bond. Cyclopentadienylgallium can also be used to access a Lewis acid B(C6F5)3 complex with a datively bonded cyclopentadienylgallium ligand. [23] For both of these two complexes, the (pentamethylcyclopentadienyl)gallium(I) analogues have been synthesized and x-ray crystallography has supported that, as expected, (pentamethylcyclopentadienyl)gallium(I) is a slightly stronger donor than cyclopentadienylgallium.

GaCp reacts with Cr(CO)5(cyclooctene) to form a new CpGa-Cr(CO)5. The GaCp* analogue can also be accessed. GaCp Chromium complex.png
GaCp reacts with Cr(CO)5(cyclooctene) to form a new CpGa–Cr(CO)5. The GaCp* analogue can also be accessed.

Like (pentamethylcyclopentadienyl)gallium(I), cyclopentadienylgallium can also coordinate to transition metal complexes such as Cr(CO)5(cyclooctene) or Co2(CO)8 to yield CpGa–Cr(CO)5 or (thf)GaCp{Co(CO)4}2. [24] For CpGa–Cr(CO)5, the Ga-Cr bond length (239.6 pm) is similar to that for a (pentamethylcyclopentadienyl)gallium(I) analogue (240.5 pm). For this complex, the trans effect is also observed, where the Cr-CO bond trans to the cyclopentadienylgallium ligand is contracted (186 pm) relative to the cis Cr-CO bonds (189.5 pm). While cyclopentadienylgallium can act as a terminal ligand similar to (pentamethylcyclopentadienyl)gallium(I), it was determined that cyclopentadienylgallium analogues react faster than their (pentamethylcyclopentadienyl)gallium(I) counterparts. This can be attributed to the lower steric bulk of cyclopentadienylgallium. [24]

Unlike reactivity with Cr(CO)5(cyclooctene), reactivities of (pentamethylcyclopentadienyl)gallium(I) and cyclopentadienylgallium with Co2(CO)8 diverge significantly. [24] Dicobalt octacarbonyl, or Co2(CO)8, exists in various isomeric states. One such isomer contains two bridging CO ligands. When (pentamethylcyclopentadienyl)gallium(I) reacts with Co2(CO)8, two equivalents of CO gas are released, forming (CO)3Co[μ2-(η5-GaCp*)]2-Co(CO)3. This is a derivative of the dicobalt octacarbonyl complex where the bridging CO moieties are replaced by bridging (pentamethylcyclopentadienyl)gallium(I) moieties. [25] On the other hand, cyclopentadienylgallium enables oxidative addition to Co2(CO)8 to form (thf)GaCp{Co(CO)4}2, where gallium has sigma interactions to two Co(CO)4 units. The average Ga–Co bond length is 248.5 pm and gallium is in a formally +3 oxidation state in this new complex. [25] Overall, straightforward synthesis of cyclopentadienylgallium from a gallium monoiodide precursor has many merits in expanding the scope of transition metal chemistry with lower valent species.

Gallium clusters

Examples of Ga clusters synthesized from variants of gallium monoiodide starting materials. R = Si(SiMe3)3. For the [Ga9{Si(SiMe3)3}6] cluster, the polyhedral vertices are all Ga. Reactions were conducted in toluene at -78 @C. Ga clusters.png
Examples of Ga clusters synthesized from variants of gallium monoiodide starting materials. R = Si(SiMe3)3. For the [Ga9{Si(SiMe3)3}6] cluster, the polyhedral vertices are all Ga. Reactions were conducted in toluene at -78 ˚C.

A variety of gallium clusters have also been synthesized from gallium monoiodide. [26] These clusters have often been isolated as salts with bulky silyl or germyl anions, such as [Si(SiMe3)3]. [8] An example of an isolated gallium cluster is [Ga9{Si(SiMe3)3}6], which has a pentagonal bipyramidal polyhedral structure. It is synthesized by reacting gallium monoiodide with Li(thf)3Si(SiMe3)3 in toluene at -78 ˚C. [8] [27] This reaction has been shown to access a wide array of products, which may be attributed to the wide range of gallium monoiodide compositions that have been subsequently probed. Of these products, [Ga9{Si(SiMe3)3}6] is especially unique because Ga was found to have a very low average oxidation state (0.56) and also because this cluster has fewer R substituents than polyhedron vertices. [27] Other clusters that been isolated via similar reaction pathways include [Ga10{Si(SiMe3)3}6], which is a conjuncto-polyhedral cluster, and a closo-silatetragallane anion, which contains three 2-electron-2-center and three 2-electron-3-center bonds. [8] [28] [29] Interestingly, this latter species can only be synthesized when sub-stoichiometric quantities of I2 are utilized to access a "Ga2I3" intermediate species. [29] This is equivalent to reacting liquid gallium metal and iodine to pre-completion, which, as explained above, accesses the [Ga0]2[Ga+]2[Ga2I62-] variant of gallium monoiodide. This highlights the versatility of the gallium monoiodide precursor in accessing a wide range of gallium-based complexes.

Reaction of a diaryl Co(II) precursor with gallium monoiodide yields a nido-type Co-GaI cluster. Ellipsoids set at 50% probability. Grey = carbon, blue = cobalt, pink = gallium, and magenta = iodine. Hydrogens not depicted. Image recreated using .cif file (deposited to The Cambridge Structural Database). Co GaI.png
Reaction of a diaryl Co(II) precursor with gallium monoiodide yields a nido-type Co-GaI cluster. Ellipsoids set at 50% probability. Grey = carbon, blue = cobalt, pink = gallium, and magenta = iodine. Hydrogens not depicted. Image recreated using .cif file (deposited to The Cambridge Structural Database).

Gallium monoiodide can also form cluster-type compounds with transition metals precursors. One example is the reaction between gallium monoiodide and (2,6-Pmp2C6H3)2Co, (Pmp = C6Me5), which yields a nido-type cluster. [30] This molecule is structurally similar to cubane, where the corners are metal and bridging iodine atoms, with one corner removed. This is a particularly unique Co-GaI cluster due to its unusual geometry for transition metal compounds containing heavy group 13 atoms such as gallium. The bond critical points and bond paths, as computed with QTAIM analysis, support that while there are Co-Ga bonds, there are no Ga-Ga bonds. [30]

Bond critical points and bond paths of a Co-GaI cluster. using Multiwfn 3.8 software. Co GaI cluster.png
Bond critical points and bond paths of a Co-GaI cluster. using Multiwfn 3.8 software.

Finally, gallium monoiodide has been able to form clusters with heavy gold atoms by acting as a reducing reagent when combined with (pentamethylcyclopentadienyl)gallium(I) and triphenylphosphine-gold complexes(i.e. AuI(PPh3) or AuCl(PPh3)). [8] [32] This cluster contained the first crystallographically confirmed Ga-Au bonds, consisting of a Au3 cluster ligated by Ga ligands. In addition, NBO analysis showed that the charge on the galliums within the (pentamethylcyclopentadienyl)gallium(I) ligands were much higher than the charge on the Au atoms and the charge on the gallium atoms within the GaI2 motifs. This suggests that non-bridging Ga-Au bonds are highly polarized, whereas the µ-bridging Ga-Au bonds are more non-polar covalent in character. [32]

Ga-Au cluster formed by dropwise addition of LAuX to a mixture of GaCp*/"GaI" (excess) in dichloromethane. Ga-Au cluster.png
Ga-Au cluster formed by dropwise addition of LAuX to a mixture of GaCp*/"GaI" (excess) in dichloromethane.

See also

Related Research Articles

A transition metal carbene complex is an organometallic compound featuring a divalent carbon ligand, itself also called a carbene. Carbene complexes have been synthesized from most transition metals and f-block metals, using many different synthetic routes such as nucleophilic addition and alpha-hydrogen abstraction. The term carbene ligand is a formalism since many are not directly derived from carbenes and most are much less reactive than lone carbenes. Described often as =CR2, carbene ligands are intermediate between alkyls (−CR3) and carbynes (≡CR). Many different carbene-based reagents such as Tebbe's reagent are used in synthesis. They also feature in catalytic reactions, especially alkene metathesis, and are of value in both industrial heterogeneous and in homogeneous catalysis for laboratory- and industrial-scale preparation of fine chemicals.

In organic chemistry and organometallic chemistry, carbon–hydrogen bond activation is a type of organic reaction in which a carbon–hydrogen bond is cleaved and replaced with a C−X bond. Some authors further restrict the term C–H activation to reactions in which a C–H bond, one that is typically considered to be "unreactive", interacts with a transition metal center M, resulting in its cleavage and the generation of an organometallic species with an M–C bond. The intermediate of this step could then undergo subsequent reactions with other reagents, either in situ or in a separate step, to produce the functionalized product.

Organogold chemistry is the study of compounds containing gold–carbon bonds. They are studied in academic research, but have not received widespread use otherwise. The dominant oxidation states for organogold compounds are I with coordination number 2 and a linear molecular geometry and III with CN = 4 and a square planar molecular geometry.

<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.

Transition metal carbyne complexes are organometallic compounds with a triple bond between carbon and the transition metal. This triple bond consists of a σ-bond and two π-bonds. The HOMO of the carbyne ligand interacts with the LUMO of the metal to create the σ-bond. The two π-bonds are formed when the two HOMO orbitals of the metal back-donate to the LUMO of the carbyne. They are also called metal alkylidynes—the carbon is a carbyne ligand. Such compounds are useful in organic synthesis of alkynes and nitriles. They have been the focus on much fundamental research.

Diimines are organic compounds containing two imine (RCH=NR') groups. Common derivatives are 1,2-diketones and 1,3-diimines. These compounds are used as ligands and as precursors to heterocycles. Diimines are prepared by condensation reactions where a dialdehyde or diketone is treated with amine and water is eliminated. Similar methods are used to prepare Schiff bases and oximes.

<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.

The phosphaethynolate anion, also referred to as PCO, is the phosphorus-containing analogue of the cyanate anion with the chemical formula [PCO] or [OCP]. The anion has a linear geometry and is commonly isolated as a salt. When used as a ligand, the phosphaethynolate anion is ambidentate in nature meaning it forms complexes by coordinating via either the phosphorus or oxygen atoms. This versatile character of the anion has allowed it to be incorporated into many transition metal and actinide complexes but now the focus of the research around phosphaethynolate has turned to utilising the anion as a synthetic building block to organophosphanes.

<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.

In chemistry, aluminium(I) refers to monovalent aluminium (+1 oxidation state) in both ionic and covalent bonds. Along with aluminium(II), it is an extremely unstable form of aluminium.

<span class="mw-page-title-main">(Pentamethylcyclopentadienyl)aluminium(I)</span> Chemical compound

(Pentamethylcyclopentadienyl)aluminium(I) is an organometallic compound with the formula Al(C5Me5) ("Me" is a methyl group; CH3). The compound is often abbreviated to AlCp* or Cp*Al, where Cp* is the pentamethylcyclopentadienide anion (C5Me5). Discovered in 1991 by Dohmeier et al., AlCp* serves as the first ever documented example of a room temperature stable monovalent aluminium compound. In its isolated form, Cp*Al exists as the tetramer [Cp*Al]4, and is a yellow crystal that decomposes at temperatures above 100 °C but also sublimes at temperatures above 140 °C.

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

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

Aluminium(I) nucleophiles are a group of inorganic and organometallic nucleophilic compounds containing at least one aluminium metal center in the +1 oxidation state with a lone pair of electrons strongly localized on the aluminium(I) center.

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

Polyfluoroalkoxyaluminates (PFAA) are weakly coordinating anions many of which are of the form [Al(ORF)4]. Most PFAA's possesses an Al(III) center coordinated by four ORF (RF = -CPh(CF3)2 (hfpp), -CH(CF3)2 (hfip), -C(CH3)(CF3)2 (hftb), -C(CF3)3 (pftb)) ligands, giving the anion an overall -1 charge. The most weakly coordinating PFAA is an aluminate dimer, [F{Al(Opftb)3}2], which possess a bridging fluoride between two Al(III) centers. The first PFAA, [Al(Ohfpp)4], was synthesized in 1996 by Steven Strauss, and several other analogs have since been synthesized, including [Al(Ohfip)4], [Al(Ohftb)4], and [Al(Opftb)4] by Ingo Krossing in 2001. These chemically inert and very weakly coordinating ions have been used to stabilize unusual cations, isolate reactive species, and synthesize strong Brønsted acids.

<i>N</i>-Heterocyclic carbene boryl anion Isoelectronic structure

An N-heterocyclic carbene boryl anion is an isoelectronic structure of an N-heterocyclic carbene (NHC), where the carbene carbon is replaced with a boron atom that has a -1 charge. NHC boryl anions have a planar geometry, and the boron atom is considered to be sp2-hybridized. They serve as extremely strong bases, as they are very nucleophilic. They also have a very strong trans influence, due to the σ-donation coming from the boron atom. NHC boryl anions have stronger electron-releasing character when compared to normal NHCs. These characteristics make NHC boryl anions key ligands in many applications, such as polycyclic aromatic hydrocarbons, and more commonly low oxidation state main group element bonding.

<span class="mw-page-title-main">Organoberyllium chemistry</span> Organoberyllium Complex in Main Group Chemistry

Organoberyllium chemistry involves the synthesis and properties of organometallic compounds featuring the group 2 alkaline earth metal beryllium (Be). The area remains understudied, relative to the chemistry of other main-group elements, because although metallic beryllium is relatively unreactive, its dust causes berylliosis and compounds are toxic. Organoberyllium compounds are typically prepared by transmetallation or alkylation of beryllium chloride.

<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.

Gallylenes are a class of gallium species which are electronically neutral and in the +1-oxidation state. This broad definition may include many gallium species, such as oligomeric gallium compounds in which the gallium atoms are coordinated to each other, but these classes of compounds are often referred to as gallanes. In recent literature, the term gallylene has mostly been reserved for low valent gallium species which may have a lone pair, analogous to NHC's or terminal borylenes. They are compounds of academic interest because of their distinctive electronic properties which have been achieved for higher main group elements such as borylenes and carbenes.

<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.

Titanocene bis(trimethylsilyl)acetylene is a formally titanium(II) organometallic compound with the formula Ti(C5H5)2C2(Si(CH3)3)2. This complex and its zirconium analogue are often referred to as Rosenthal's reagent, after the first chemist to synthesize it, Uwe Rosenthal. This article will discuss its history, synthesis, structure, reactivity, and applications.

References

  1. 1 2 Dohmeier, Carsten; Loos, Dagmar; Schnöckel, Hansgeorg (1996). "Aluminum(I) and Gallium(I) Compounds: Syntheses, Structures, and Reactions". Angewandte Chemie International Edition in English. 35 (2): 129–149. doi:10.1002/anie.199601291. ISSN   1521-3773.
  2. 1 2 3 4 Green, Malcolm L. H.; Mountford, Philip; Smout, Geoffrey J.; Speel, S. Robert (1990-01-01). "New synthetic pathways into the organometallic chemistry of gallium". Polyhedron. 9 (22): 2763–2765. doi:10.1016/S0277-5387(00)86809-6. ISSN   0277-5387.
  3. 1 2 3 Widdifield, Cory M.; Jurca, Titel; Richeson, Darrin S.; Bryce, David L. (2012-03-16). "Using 69/71Ga solid-state NMR and 127I NQR as probes to elucidate the composition of "GaI"". Polyhedron. 35 (1): 96–100. doi:10.1016/j.poly.2012.01.003. ISSN   0277-5387.
  4. 1 2 3 4 5 6 7 8 9 Malbrecht, Brian J.; Dube, Jonathan W.; Willans, Mathew J.; Ragogna, Paul J. (2014-09-15). "Addressing the Chemical Sorcery of "GaI": Benefits of Solid-State Analysis Aiding in the Synthesis of P→Ga Coordination Compounds". Inorganic Chemistry. 53 (18): 9644–9656. doi:10.1021/ic501139w. ISSN   0020-1669. PMID   25184621.
  5. Wilkinson, M.; Worrall, I. J. (1975-07-01). "Synthesis of alkylgallium diiodides". Journal of Organometallic Chemistry. 93 (1): 39–42. doi:10.1016/S0022-328X(00)94142-1. ISSN   0022-328X.
  6. Okuda, Tsutomu; Hamamoto, Hiroshi; Ishihara, Hideta; Negita, Hisao (1985-09-01). "79Br and 127I Nuclear Quadrupole Resonance of Mixed Valency Compounds Ga2X4(X=Br, I)". Bulletin of the Chemical Society of Japan. 58 (9): 2731–2732. doi: 10.1246/bcsj.58.2731 . ISSN   0009-2673.
  7. 1 2 Beamish, J. C.; Wilkinson, M.; Worrall, I. J. (1978-07-01). "Facile synthesis of the lower valent halides of gallium, Ga2X4 (X = chloride, bromide, iodide) and tetragallium hexaiodide". Inorganic Chemistry. 17 (7): 2026–2027. doi:10.1021/ic50185a069. ISSN   0020-1669.
  8. 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 Baker, Robert J.; Jones, Cameron (2005-04-11). ""GaI": A versatile reagent for the synthetic chemist". Dalton Transactions (8): 1341–1348. doi:10.1039/B501310K. hdl: 2262/69572 . ISSN   1477-9234. PMID   15824768.
  9. 1 2 3 4 Baker, Robert J.; Bettentrup, Helga; Jones, Cameron (2003). "The Reactivity of Primary and Secondary Amines, Secondary Phosphanes and N-Heterocyclic Carbenes towards Group-13 Metal(I) Halides". European Journal of Inorganic Chemistry. 2003 (13): 2446–2451. doi:10.1002/ejic.200300068. ISSN   1099-0682.
  10. 1 2 Schnepf, Andreas; Doriat, Clemens (1997-01-01). "A simple synthesis for donor-stabilized Ga2I4 and Ga3I5 species and the X-ray crystal structure of Ga3I5·3PEt3". Chemical Communications (21): 2111–2112. doi:10.1039/A703776G. ISSN   1364-548X.
  11. 1 2 Cameron Jones; Christian Schulten; Andreas Stasch (2005-04-01). "[Gal2Ph(SbPh3)]: A Rare Tertiary Stibane-Gallium Complex Formed via a Reductive Sb-C Bond Cleavage Reaction". Main Group Metal Chemistry. 28 (2): 89–92. doi: 10.1515/MGMC.2005.28.2.89 . ISSN   0792-1241. S2CID   101459028.
  12. 1 2 3 4 Baker, Robert J.; Jones, Cameron; Kloth, Marc; Mills, David P. (2004-02-03). "The reactivity of gallium(I) and indium(I) halides towards bipyridines, terpyridines, imino-substituted pyridines and bis(imino)acenaphthenes". New Journal of Chemistry. 28 (2): 207–213. doi:10.1039/B310592J. ISSN   1369-9261.
  13. Beran, G.; Carty, A. J.; Patel, H. A.; Palenik, Gus J. (1970-01-01). "A trans-effect in gallium complexes: the crystal structure of trichloro-(2,2′,2"-terpyridyl)gallium(III)". Journal of the Chemical Society D: Chemical Communications (4): 222–223. doi:10.1039/C29700000222. ISSN   0577-6171.
  14. 1 2 3 4 5 Pott, Thomas; Jutzi, Peter; Kaim, Wolfgang; Schoeller, Wolfgang W.; Neumann, Beate; Stammler, Anja; Stammler, Hans-Georg; Wanner, Matthias (July 2002). "Reactivity of "GaI" TowardN-Substituted 1,4-Diazabuta-1,3-dienes: Synthesis and Characterization of Gallium Heterocycles Containing Paramagnetic Diazabutadiene Monoanions". Organometallics. 21 (15): 3169–3172. doi:10.1021/om0200510. ISSN   0276-7333.
  15. Baker, Robert J.; Farley, Robert D.; Jones, Cameron; Kloth, Marc; Murphy, Damien M. (2002-10-03). "The reactivity of diazabutadienes toward low oxidation state Group 13 iodides and the synthesis of a new gallium(I) carbene analogue". Journal of the Chemical Society, Dalton Transactions (20): 3844–3850. doi:10.1039/B206605J. ISSN   1364-5447.
  16. Schmidt, Eva S.; Jockisch, Alexander; Schmidbaur, Hubert (1999-10-01). "A Carbene Analogue with Low-Valent Gallium as a Heteroatom in a quasi-Aromatic Imidazolate Anion". Journal of the American Chemical Society. 121 (41): 9758–9759. doi:10.1021/ja9928780. ISSN   0002-7863.
  17. Baker, Robert J.; Jones, Cameron; Mills, David P.; Murphy, Damien M.; Hey-Hawkins, Evamarie; Wolf, Robert (2006-12-14). "The reactivity of gallium-(I), -(II) and -(III) heterocycles towards Group 15 substrates: attempts to prepare gallium–terminal pnictinidene complexes". Dalton Transactions (1): 64–72. doi:10.1039/B511451A. ISSN   1477-9234. PMID   16357962.
  18. 1 2 3 Baker, Robert J.; Jones, Cameron (2005-09-01). "The coordination chemistry and reactivity of group 13 metal(I) heterocycles". Coordination Chemistry Reviews. 36th International Conference on Coordination Chemistry, Merida, Mexico, July 2004. 249 (17): 1857–1869. doi:10.1016/j.ccr.2004.12.016. ISSN   0010-8545.
  19. Jones, Cameron; Rose, Richard P.; Stasch, Andreas (2007-07-10). "Synthesis and characterisation of zinc gallyl complexes: First structural elucidations of Zn–Ga bonds". Dalton Transactions (28): 2997–2999. doi:10.1039/B706402K. ISSN   1477-9234. PMID   17622416.
  20. 1 2 Hardman, Ned J.; Eichler, Barrett E.; Power, Philip P. (2000-01-01). "Synthesis and characterization of the monomer Ga{(NDippCMe)2CH} (Dipp = C6H3Pri2-2,6): a low valent gallium(I) carbene analogue". Chemical Communications (20): 1991–1992. doi:10.1039/B005686N. ISSN   1364-548X.
  21. Loos, Dagmar; Baum, Elke; Ecker, Achim; Schnöckel, Hansgeorg; Downs, Anthony J. (1997). "Hexameric Aggregates in Crystalline (Pentamethylcyclopentadienyl)gallium(I) at 200 K". Angewandte Chemie International Edition in English. 36 (8): 860–862. doi:10.1002/anie.199708601. ISSN   1521-3773.
  22. Jutzi, Peter; Schebaum, Lars Oliver (2002-07-15). "A novel synthetic route to pentaalkylcyclopentadienylgallium(I) compounds". Journal of Organometallic Chemistry. 654 (1): 176–179. doi:10.1016/S0022-328X(02)01429-8. ISSN   0022-328X.
  23. 1 2 Schenk, Christian; Köppe, Ralf; Schnöckel, Hansgeorg; Schnepf, Andreas (2011). "A Convenient Synthesis of Cyclopentadienylgallium – The Awakening of a Sleeping Beauty in Organometallic Chemistry". European Journal of Inorganic Chemistry. 2011 (25): 3681–3685. doi:10.1002/ejic.201100672. ISSN   1099-0682.
  24. 1 2 3 4 Naglav, Dominik; Tobey, Briac; Schnepf, Andreas (2013). "Application of GaCp as a Ligand in Coordination Chemistry: Similarities and Differences to GaCp*". European Journal of Inorganic Chemistry. 2013 (24): 4146–4149. doi:10.1002/ejic.201300401. ISSN   1099-0682.
  25. 1 2 Jutzi, Peter; Neumann, Beate; Reumann, Guido; Stammler, Hans-Georg (1998-03-01). "Pentamethylcyclopentadienylgallium (Cp*Ga): Alternative Synthesis and Application as a Terminal and Bridging Ligand in the Chemistry of Chromium, Iron, Cobalt, and Nickel". Organometallics. 17 (7): 1305–1314. doi:10.1021/om970913a. ISSN   0276-7333.
  26. Schnepf, Andreas; Schnöckel, Hansgeorg (2002). "Metalloid Aluminum and Gallium Clusters: Element Modifications on the Molecular Scale?". Angewandte Chemie International Edition. 41 (19): 3532–3554. doi:10.1002/1521-3773(20021004)41:19<3532::AID-ANIE3532>3.0.CO;2-4. ISSN   1521-3773. PMID   12370894.
  27. 1 2 Köstler, Wolfgang; Linti, Gerald (1997). "Synthesis and Structure of a Tetragallane [R4Ga4I3]− and a Polyhedral Nonagallane [R6Ga9]". Angewandte Chemie International Edition in English. 36 (23): 2644–2646. doi:10.1002/anie.199726441. ISSN   1521-3773.
  28. Kehrwald, Michael; Köstler, Wolfgang; Rodig, Alexander; Linti, Gerald; Blank, Thomas; Wiberg, Nils (2001-03-01). "Ga10[Si(SiMe3)3]6, [Ga10(SitBu3)6]-, and [Ga13(SitBu3)6]-Syntheses and Structural Characterization of Novel Gallium Cluster Compounds". Organometallics. 20 (5): 860–867. doi:10.1021/om000703p. ISSN   0276-7333.
  29. 1 2 Linti, Gerald; Köstler, Wolfgang; Piotrowski, Holger; Rodig, Alexander (1998). "A Silatetragallane—Classical Heterobicyclopentane or closo-Polyhedron?". Angewandte Chemie International Edition. 37 (16): 2209–2211. doi:10.1002/(SICI)1521-3773(19980904)37:16<2209::AID-ANIE2209>3.0.CO;2-3. ISSN   1521-3773. PMID   29711437.
  30. 1 2 3 4 Blundell, Toby J.; Taylor, Laurence J.; Valentine, Andrew J.; Lewis, William; Blake, Alexander J.; McMaster, Jonathan; Kays, Deborah L. (2020-07-21). "A transition metal–gallium cluster formed via insertion of "GaI"". Chemical Communications. 56 (58): 8139–8142. doi: 10.1039/D0CC03559A . ISSN   1364-548X. PMID   32691803.
  31. Lu, Tian; Chen, Feiwu (2012). "Multiwfn: A multifunctional wavefunction analyzer". Journal of Computational Chemistry. 33 (5): 580–592. doi:10.1002/jcc.22885. ISSN   1096-987X. PMID   22162017. S2CID   13508697.
  32. 1 2 3 Anandhi, U.; Sharp, Paul R. (2004). "A Gallium-Coated Gold Cluster". Angewandte Chemie International Edition. 43 (45): 6128–6131. doi:10.1002/anie.200461295. ISSN   1521-3773. PMID   15549757.