You are on page 1of 17

ChemComm

View Article Online


FEATURE ARTICLE View Journal | View Issue

The emerging chemistry of the aluminyl anion


Published on 08 December 2022. Downloaded by University of Liverpool on 12/30/2023 6:45:00 PM.

Cite this: Chem. Commun., 2023, Martyn P. Coles *a and Matthew J. Evansb
59, 503
The chemistry of low valent p-block metal complexes continues to elicit interest in the research
community, demonstrating reactivity that replicates and in some cases exceeds that of their more widely
studied d-block metal counterparts. The introduction of the first aluminyl anion, a complex containing a
formally anionic Al(I) centre charge balanced by an alkali metal (AM) cation, has established a platform
for a new area of chemical research. The chemistry displayed by aluminyl compounds is expanding
rapidly, with examples of reactivity towards a diverse range of small molecules and functional groups
now reported in the literature. Herein we present an account of the structure and reactivity of the
growing family of aluminyl compounds. In this context we examine the structural relationships between
the aluminyl anion and the AM cations, which now include examples of AM = Li, Na, K, Rb and Cs. We
Received 3rd November 2022, report on the ability of these compounds to engage in bond-breaking and bond-forming reactions,
Accepted 7th December 2022 which is leading towards their application as useful reagents in chemical synthesis. Furthermore we
DOI: 10.1039/d2cc05963k discuss the chemistry of bimetallic complexes containing direct Al–M bonds (M = Li, Na, K, Mg, Ca, Cu,
Ag, Au, Zn) and compounds with Al–E multiple bonds (E = NR, CR2, O, S, Se, Te), where both classes of
rsc.li/chemcomm compound are derived directly from aluminyl anions.

1. Introduction
1.1 Neutral, monometallic complexes of aluminium(I)
The synthesis of neutral compounds containing aluminium in
the +1 oxidation state can be traced back to the 1940’s with the
a
School of Chemical of Physical Sciences, Victoria University of Wellington, PO Box
isolation of binary compounds AlX (X = Cl, Br, I).1 It was not until
600, Wellington, 6012, New Zealand. E-mail: martyn.coles@vuw.ac.nz techniques for manipulating these meta-stable compounds were
b
School of Chemistry, Monash University, Melbourne, Victoria, Australia developed that the chemistry was able to be explored in detail,2

Martyn Coles received his BSc Matthew Evans studied at Victo-


and PhD from Durham Univer- ria University of Wellington – Te
sity in the UK, working with Herenga Waka in New Zealand,
Professor Vernon Gibson. After where he received his BSc degree
post-doctoral work at the Univer- in Chemistry in 2018. Following a
sity of Iowa (with Professor research year with Associate Pro-
Richard Jordan) and University fessor J. Robin Fulton, he com-
of California, Berkeley (with pleted his PhD degree in 2022
Professor T. Don Tilley) he under Professor Martyn Coles
started his first independent job and Dr Mathew Anker. Matthew
at the University of Sussex. He is currently works at Monash Uni-
currently a Professor of Inorganic versity in Australia as a postdoc-
Martyn P. Coles Chemistry at Victoria University Matthew J. Evans toral researcher for Prof. Came-
of Wellington – Te Herenga Waka ron Jones. His research interests
in New Zealand. His research interests focus on the synthesis and lie in the isolation of low-valent main-group complexes for small
reactivity of low-valent main group compounds, with a particular molecule activation.
passion for the group 13 elements (Al, Ga In) and the heavier group
15 elements (Sb and Bi).

This journal is © The Royal Society of Chemistry 2023 Chem. Commun., 2023, 59, 503–519 | 503
View Article Online

Feature Article ChemComm

3,5-iPr2C6H),14 and the carbazolyl-substituted example IX.15


Although the chemistry of these species is at an early stage, both
classes of complex have already demonstrated interesting
reactivity,16 including oxidative addition, cycloaddition and appli-
cations in coordination chemistry.

1.2 Low valent group 13 analogues of N-heterocyclic


carbenes
Published on 08 December 2022. Downloaded by University of Liverpool on 12/30/2023 6:45:00 PM.

Inspired by the extensive chemistry of carbenes, in which a


divalent carbon centre is supported by a dianionic ligand set to
afford a neutral complex of the general formula X2C, researchers
have explored the chemistry of isoelectronic group 13 element
(E) analogues (Fig. 2). To maintain an isoelectronic relationship
with carbenes, atom E is present in the +1 oxidation state and
the dianionic supporting ligand therefore generates an overall
negative charge, resulting in the boryl (E = B), aluminyl (E = Al),
gallyl (E = Ga) and indyl (E = In).17 Although a large number of
predominantly bidentate ligands have been applied to this area,
the N,N 0 -chelating scaffold is commonly encountered in both
carbene and the group 13 complexes, and the following discus-
sion will therefore be restricted to such systems.
The first group 13 system to be investigated within this context
was the gallyl anion [Ga(HCNtBu)2] Xa, isolated with the [K(18-c-6)-
Fig. 1 Low valent Al(I) compounds based on cyclopentadienyl (I–IV) and
b-diketiminate (BDI, V–VII) ligand systems, and aluminylenes VIII and IX.
(THF)2]+ (18-c-6 = 18-crown-6) counter-ion.18 Subsequent work
extended this chemistry to the N-Dipp substituted analogue
[Ga(HCNDipp)2] Xb, isolated as either a dimer incorporating
with the first breakthrough in the practical synthesis of isolable [K(Et2O)]+ or [K(dme)]+ cations, or as the isolated anion with a
Al(I) compounds being the reaction of AlCl with MgCp*2, affording [K2(18-c-6)3]+ cation.19 The lightest member of the series, the boryl
[AlCp*]4 (Ia, Fig. 1).3 It was subsequently shown that Ia could also anion [B(HCNDipp)2] XI was initially isolated as the [Li(dme)]2-
be accessed by the reduction of Cp*AlCl2 with potassium,4 bridged dimer, containing a direct B–Li bond.20 Examples of
circumventing the need for specialist equipment required for monomeric lithium boryls have since been reported,21 and the
handling AlX species. ‘AlCp*’ exists as the tetramer Ia in the dimeric borylpotassium complex [K{B(HCNDipp)2}]2 XII has
solid-state but undergoes a temperature dependant dissociation recently joined the family of boryls.22
into monomeric AlCp* units Ib in solution, a feature that has
been exploited in its reactivity, illustrated by its ability to act as a
ligand at transition metal centres.5 It also exists as monomer Ib in
the gas phase.6 The area of cyclopentadienyl Al(I) chemistry has
subsequently been extended to other derivatives, with the bulky
systems [C5(CH2Ph)5] (II),7 and [1,2,4-R3C5H2] (R = SiMe3 (III),8
tBu (IV)9) successfully shifting the equilibrium such that the
momomeric forms dominate.
The next major advance in Al(I) chemistry was the synthesis
of the monomeric complex Al(BDIDipp) (V, BDIDipp = [HC{C(Me)-
(NDipp)}2]), isolated from the reduction of the diiodide
precursor by potassium.10 Despite the extensive chemistry of BDI-
ligands, only two other Al-BDI compounds are known. Al(tBuBDIDipp)
VI incorporates tBu groups into the ligand framework,11 whereas
the new ‘‘superbulky’’ derivative Al(BDIDipep) (VII) replaces the
nitrogen substituents with 2,6-diisopentylphenyl (Dipep) groups.12
The rich chemistry of both Cp- and BDI-derived Al(I) systems have
been the subject of several detailed review articles.13
A noteworthy addition to the family of Al(I) compounds con-
sists of a one coordinate Al-centre in an aluminylene complex.
Fig. 2 (a) The relationship between neutral carbenes and the group 13
Two classes of aluminylene have been presented, the ter-phenyl boryl, aluminyl, gallyl and indyl anions. (b) Examples of gallyl (X), boryl
alanediyl derivative Al(AriPr8) (VIII, AriPr8 = 2,6-(2,4,6-iPr3C6H2)2- (XI, XII) and indyl (XIII, XIV) anions (dme = dimethoxyethane).

504 | Chem. Commun., 2023, 59, 503–519 This journal is © The Royal Society of Chemistry 2023
View Article Online

ChemComm Feature Article

Although the parent aluminyl and indyl anions, [Al(HCNH)2]


[In(HCNH)2] had been studied computationally,23 it was not
until 2018 that we reported the first indyl anion.24 In this study
we showed that the dianionic bis(amidodimethyl)disiloxane
ligand [O(SiMe2NDipp)2]2 (NON), which we had used in our
earlier work with low valent bismuth,25 stabilised both the
lithium indyl (NON)In–Li(THF)3 XIII that contains an In–Li
bond, and the potassium indyl [K(crypt-222)][In(NON)] (crypt-
Published on 08 December 2022. Downloaded by University of Liverpool on 12/30/2023 6:45:00 PM.

222 = [2.2.2]cryptand) that has a naked indyl anion with a two-


coordinate indium. The latter anion was also isolated as the
dimeric potassium indyl, [K{In(NON)}]2 XIV.26
In this Feature Article we describe the research on the
synthesis and structures of the current family of aluminyl
anions, focussing on the characteristics of the supporting
ligand and the nature and extent of the interactions between
the anion and the counter cation. We have also discussed the
reactivity of aluminyls with a range of functional groups in the
context of bond breaking and bond forming reactions. Furthermore,
given that aluminyls provide facile entry into a class of bimetallic
complex containing unsupported Al–M bonds, we have examined
the ability of these compounds to promote a range of small Fig. 3 Potassium N,N 0 -(diamido)aluminyl compounds supported by
molecule activation processes. Finally, since the aluminyls have [xanthNON]2 (1), [NON]2 (2), [NC2NAr]2 (3Ar, Ar = Dipp, Mes) and
also provided a stable environment for the formation of Al–E [BDIAr-H]2 (4Ar, Ar = Dipp, Dipep) ligands.
multiple bonds, we have included a discussion on these new
functional groups in the context of small molecule activation. This
The potassium aluminyl [K{Al(BDIDipp-H)}]2 4Dipp was isolated
updated report compliments and extends earlier reviews on the
during an attempt to boost the reactivity of the well-known Al(I)
chemistry of the aluminyl anion.13a,27
compound Al(BDIDipp) (V)10 through the addition of potassium
bases, K[R] (R = CH2SiMe3, N{SiMe3}2).32 Instead of forming
the anticipated K[Al(BDIDipp)(R)] compounds, deprotonation
2. Synthesis and structures of of a methyl group of the ligand backbone occurred generating
aluminyls the diamido ligand, [DippNC(Me)QC(H)C(QCH2)(NDipp)]2
2.1 N,N 0 -(Diamido) aluminyl anions (BDIDipp-H) in 4Dipp. An analogous reaction of the NDipep
substituted precursor Al(BDIDipep) (VII) with K[R] was shown to
The first potassium aluminyl compound, [K{Al(xanthNON)}]2 1
be solvent dependant.12 In hexane the reaction forms exclusively
was isolated in 2018, using the 4,5-bis(2,6-diisopropylanilido)-
the potassium aluminyl [K{Al(BDIDipep-H)}]2 4Dipep, whereas in
2,7-di-tert-butyl-9,9-dimethylxanthene ([xanthNON]2) as a che-
benzene a mixture of 4Dipep and a ligand C–H activation product
lating diamido ligand scaffold (Fig. 3).28 Shortly after this report
was formed.
we described the preparation of the potassium aluminyl
[K{Al(NON)}]2 2.29 Analysis of 1 and 2 by single crystal X-ray
diffraction confirmed that the aluminium centres in 1 are 2.2 C,N-(Alkyl)(amido) and C,C 0 -(dialkyl)aluminyl anions
three-coordinate with a dative O-Al interaction stabilising It is established that the substituents adjacent to the low-valent
the formally anionic aluminium centre, whereas the alumi- C-atom directly influence the electronic structure of carbenes.33
nium centres in 2 are strictly two-coordinate. Accordingly, chelating C,N-(alkyl)(amido)-and C,C 0 -(dialkyl)
Subsequent to these reports, potassium compounds contain- ligands have been used to form potassium aluminyls with
ing seven-membered N,N 0 -(diamido) aluminyl anions with an different electronic properties. The ligands [(Me3Si)2C(CH)2NAd]2
alkyl-bridged bis(dimethylsilylamido) ligand [{CH2SiMe2NAr}2]2 ([CSiNAd]2),34 and [{H2CC(SiMe3)2}2]2 ([{CSi}2]2),35 form five-
([NC2NAr]2, Ar = Dipp, Mes), [K{Al(NC2NDipp)}]2 3Dipp,30 and membered heterocycles with C,N- and C,C 0 -coordination at
[K{Al(NC2NMes)}]2 3Mes,31 were prepared using an analogous pro- aluminium, respectively (Fig. 4). In contrast to the potassium
cedure to that employed for the synthesis of 1 and 2. The amido N,N 0 -(diamido) aluminyls 1–4, these compounds are synthe-
substituent has an effect on this chemistry, demonstrated during sized by the potassium reduction of the corresponding dialu-
the synthesis of 3Dipp and 3Mes. The major product after reacting manes (CSiNAd)Al–Al(CSiNAd) and ({CSi}2)Al–Al({CSi}2), affording
the NMes precursor (NC2NMes)Al–I with excess potassium for [K(12-c-4)2][Al(CSiNAd)] (5, 12-c-4 = 12-crown-4) and ({CSi}2)Al–
3 days was the dialumane (NC2NMes)Al–Al(NC2NMes),31 whereas K(tol)2 (6, tol = toluene) after the appropriate work-up. A DFT
under the same experimental conditions the corresponding reac- study on the electronic structures of 5 and 6 confirm that
tion with the NDipp analogue afforded the fully reduced Al(I) incorporating less electron withdrawing alkyl-groups at alumi-
aluminyl 3Dipp.30 nium raises the energy of the HOMO relative to the diamido

This journal is © The Royal Society of Chemistry 2023 Chem. Commun., 2023, 59, 503–519 | 505
View Article Online

Feature Article ChemComm

AM = Li, Na, K, Rb and Cs.42 The synthesis involved the


reduction of (NON)Al–I using graphitic RbC8 or CsC8, affording
[Rb{Al(NON)}]2 13 and [Cs{Al(NON)}]2 14, respectively. A similar
approach has since completed the series of [AM{Al(BDIDipp-H)}]2
compounds for [Rb{Al(BDIDipp-H)}]2 15 and [Cs{Al(BDIDipp-H)}]2
16.41 Although the reactions that form 15 and 16 proceed using
benzene or toluene solvent, the best yields were achieved using a
solvent-free ball-milling process involving mixtures of Al(BDIDipp)
Published on 08 December 2022. Downloaded by University of Liverpool on 12/30/2023 6:45:00 PM.

and RbC8 or CsC8.


Fig. 4 Potassium and lithium C,N-(Alkyl)(Amido) and C,C 0 -(Dialkyl) alu-
minyls supported by [CSiNAd]2 (5, Si = SiMe3, Ad = 1-adamantyl), [{CSi}2]2 2.4 Structural classification of aluminyl compounds
0
(6) and [{bo2e-CAr N}]2 (9, Ar 0 = 3,5-tBu2C6H3).
As the development of aluminyl compounds has progressed,
different structural types have been identified that can be
derivatives 1, 2 and 3Dipp, although the HOMO–LUMO energy grouped depending on the nature and extent of the cation:
gap remains similar in all five examples.27 anion interactions. Whilst this area has mainly focused on
solid-state structures determined by X-ray diffraction, the facile
conversion between different structural types (e.g. in the
2.3 Beyond potassium in aluminyl chemistry presence of a coordinating solvent) means that consideration
The use of potassium as the active reductant in the preparation of the state of the aluminyl in solution is an important and in
of 1–6 led to the exclusive formation of the corresponding some cases crucial consideration when studying the reactivity
potassium aluminyls. However, inspired by the established of these species. Indeed, as the chemistry of these systems is
phenomenon of alkali-metal mediated (AMM) reactivity,36 in explored, the role of the alkali metal in directing or even
which a group 1 metal works synergistically with another metal controlling reaction pathways is becoming evident. We have
(in this case the aluminium) to gain entry into reaction path- therefore classified the aluminyl compounds into the following
ways that may not otherwise proceed efficiently,37 we sought to structural types.
extend the family of alkali metal (AM) aluminyls to other group 2.4.1 Contacted dimeric pairs (CDPs). X-Ray crystallography
1 derivatives. has shown that the potassium aluminyls 1,28 2,29 3Ar (Ar =
2.3.1 Lithium and sodium aluminyls. Lithium and sodium Dipp,30 Mes31) and 4Ar (Ar = Dipp,32 Dipep12) exist as contacted
aluminyls have been prepared via two different synthetic dimeric pairs (CDPs) in the solid-state, in which the two alumi-
approaches: (i) the direct reaction of Al(III) iodide precursors nyl anions are linked by flanking K  p(arene) interactions
with the group 1 metal as reductants, or (ii) by a metathesis (Fig. 3). Diffusion ordered spectroscopy (DOSY) NMR experi-
reaction of the isolated potassium aluminyls with a group 1 ments on these systems are consistent with retention of these
metal salt. We have demonstrated that (NON)Al–I reacts with an associated dimeric structure in solution. In addition, the CDP
excess of Li or Na metal to afford the corresponding lithium and motif has now been structurally verified for the series of
sodium aluminyl complexes, [Li{Al(NON)}]2 7 and [Na{Al(NON)}]2 [AM{Al(NON)}]2 (AM = Li 7, Na 8, Rb 13 and Cs 14),38,42 and
8, respectively.38 Interestingly this approach is unsuccessful in the [BDIDipp-H]2 systems [AM{Al(BDIDipp-H)}]2 (AM = Li 17, Na
obtaining aluminyl anions supported by the [xanthNON]2 ligand, 18, Rb 15 and Cs 16),41 where heating samples of the initially
with the attempted lithium reduction of (xanthNON)Al–I affording isolated (BDIDipp-H)Al–Li(Et2O)2 11 and (BDIDipp-H)Al–Na-
Al–N cleavage products, and the corresponding reaction (Et2O)(TMEDA) 12 to 70 1C under vacuum afforded the corres-
with sodium giving exclusively the dialumane, (xanthNON)Al– ponding CDPs [Li{Al(BDIDipp-H)}]2 17 and [Na{Al(BDIDipp-H)}]2
0
Al(xanthNON).39 A new C,N-(alkyl)(amido) aluminyl anion 18. The same conversion was achieved when (bo2e-CAr N)Al–
has recently been reported incorporating a more robust Li(Et2O)2 9 was recrystallized from hexane, forming the CDP
0
bicyclo[2.2.2]oct-2-ene (bo2e) backbone.40 As for the synthesis [Li{Al(bo2e-CAr N)}]2 19.40
of 7, the direct reduction of the Al(III) iodide precursor with The relative orientation of the aluminyl anions within the
lithium powder afforded the lithium aluminyl, isolated from dimer varies with the alkali metal and supporting ligand, and
0
Et2O as (bo2e-CAr N)Al–Li(Et2O)2 9 (Ar 0 = 3,5-tBu2C6H3). three sub-types of CDP have emerged from structural analysis
Using the metathesis protocol, the lithium aluminyl (xanth- (Fig. 5). The lithium and sodium aluminyls 7, 8, 17, 18 and 19
NON)Al–Li(Et2O)2 10 was synthesized by reacting 1 and lithium form ‘slipped’ CDPs in which the distances from the Al to the
iodide in diethyl ether.39 A similar methodology reacting 4Dipp AM (dAl  AM) are significantly different (range DdAl  AM = 0.40–
with Li[BPh4] or Na[BPh4] gives the corresponding lithium 0.62 Å).38,40,41 DFT calculations show that a single Al–AM bond
and sodium aluminyl compounds, isolated as (BDIDipp-H)Al– exists within these ‘slipped’ CDPs, in contrast with the heavier
Li(Et2O)2 11 and (BDIDipp-H)Al–Na(Et2O)(TMEDA) 12.41 congeners (AM = K, Rb, Cs) that form ‘symmetric’ or ‘twisted’
2.3.2 Rubidium and caesium aluminyls. In collaboration dimers with the Al atoms engaging in contacts to both AM
with Mulvey and co-workers, we have now accessed aluminyl cations (vide infra).
compounds incorporating the heavier alkali metals Rb and Cs, In our initial assessment of the structures of Rb and Cs
completing the first series of aluminyls [AM{Al(NON)}]2 for CDPs 13 and 14, we noted that the aluminyl heterocycles were

506 | Chem. Commun., 2023, 59, 503–519 This journal is © The Royal Society of Chemistry 2023
View Article Online

ChemComm Feature Article

Our initial lithium reduction reactions were performed in


Et2O, affording a crude white solid after removal of the solvent.
When gently heated under vacuum during the work-up procedure,
we noted a distinct colour change to yellow and the final product
was identified as the CDP 7. Curious as to the identity of the
colourless product initially formed, we redissolved a sample of 7
in Et2O, from which we were able to isolate the lithium MIP,
(NON)Al–Li(Et2O)2 20. An analogous result was obtained with
Published on 08 December 2022. Downloaded by University of Liverpool on 12/30/2023 6:45:00 PM.

the sodium analogue 8, affording (NON)Al–Na(Et2O)2 21. These


different structures exist in equilibrium in solution, and we
showed that, in addition to heating under vacuum, the CDPs
were fully regenerated when the MIPs were dissolved in a large
excess of non-coordinating solvents such as hexane or toluene.
We have investigated the interconversion between CDPs and
the corresponding MIPs for the aluminyl anions [Al(NON)] in
more detail, using a range of donor molecules (Scheme 1). The
addition of THF to the CDPs 7 and 8 affords the THF adducts,
(NON)Al–AM(THF)3 (22, AM = Li; 23, AM = Na).43 We were
unable to reform the CDPs from these MIPs, consistent with the
increased s-donor property of THF resulting in stronger AM–O
Fig. 5 Three structural sub-types noted for the contacted dimeric pairs
bonds. We were also unable to obtain the corresponding
CDPs of alkali metal aluminyls.
potassium MIP from the reaction of 2 with THF, prompting
us to examine the role of the bidentate donor ligand TMEDA. In
‘twisted’ with respect to one another. The angle y between the this case, the full series of MIPs for Li, Na and K were accessed,
least squares plane defined by Al–N–Si–O–Si–N were 66.55(4)1 in (NON)Al–AM(TMEDA)n (24, AM = Li, n = 1; 25, AM = Na, n = 2; 26,
(AM = Rb) and 66.31(5)1 (AM = Cs), which we ascribed to the AM = K, n = 2).43 A similar approach using 18-crown-6 (18-c-6) as a
steric influence of the large Rb+ and Cs+ cations. However, multidentate donor ligand forms the MIP (NC2NDipp)Al–K(18-c-6)
the relative orientations of the aluminyl heterocycles within the 27 from the CDP 3Dipp.31 This work has recently been complimen-
Rb and Cs CDPs 15 and 16 are strictly coplanar (y = 01) due to a ted by the demonstrated interconversion of 9 2 19,40 11 2 17
crystallographic inversion centre, suggesting that other ligand:AM and 12 2 18.41
interactions or packing effects also contribute to the overall 2.4.3 Separated ion pairs (SIPs). A third structure type for
structure. the aluminyls was noted in the initially synthesized potassium
2.4.2 Monomeric ion pairs (MIPs). The lack of aromatic (alkyl)(amido) aluminyl 5.34 The addition of 12-crown-4 during
substituents in the dialkyl ligand framework [{CSi}2]2 precludes
the formation of stabilizing K  p(arene) interactions that are
important in the formation of CDP structures. In this case, the
(dialkyl)-aluminyl compound is as isolated as the toluene sol-
vated species, 6.35 We refer to this structure type as a monomeric
ion pair (MIP), which is characterised by unsupported, highly
polar Al–AM bonds between the anionic and cationic compo-
nents of the aluminyl compound (Fig. 6).

Scheme 1 The conversion of Li (7), Na (8) and K (2) CDPs to MIPs using
Fig. 6 Aluminyls that have been characterised as monomeric ion pairs Et2O, THF and TMEDA. Note for simplicity, the CDP is shown as the
(MIPs) in the solid-state. symmetric variant.

This journal is © The Royal Society of Chemistry 2023 Chem. Commun., 2023, 59, 503–519 | 507
View Article Online

Feature Article ChemComm

[Li(TMEDA)2][Al(NON)] 33. Structural characterisation of both


species confirmed non interactions between the AM and the Al,
confirming these species as SIPs.

3. Reactivity of aluminyl anions


3.1 Aluminyls as nucleophiles
A defining feature of aluminyls is their nucleophilic
Published on 08 December 2022. Downloaded by University of Liverpool on 12/30/2023 6:45:00 PM.

Fig. 7 Aluminyls characterised as separated ion pairs (SIPs) in the solid- character,27 illustrated by the reactions of 1,28 3Dipp,31 and 635
state. with organic nucleophiles such as MeOTf and MeI, affording
the corresponding aluminium(III) methyl compounds. This
reactivity has been exploited in the synthesis of new bimetallic
the work-up formed the separated ion pair (SIP) [K(12-c-4)2]- complexes via a salt metathesis route with metal halide
[Al(CSiNAd)] 5, in which the potassium cation is isolated from substrates.
the aluminyl anion (Fig. 7). 3.1.1 Synthesis of bimetallic Al–M complexes. Bimetallic
The ability to form SIPs from CDPs has also been investi- compounds accessed from aluminyls generally contain
gated for the aluminyl anions [Al(xanthNON)], [Al(NON)] and unsupported Al–M bonds that are reminiscent of the Al–AM
[Al(NC2NDipp)]. For the potassium aluminyl CDPs 1, 2 and bonds in the MIPs (see Section 2.4.2), including examples of
3Dipp, the addition of [2.2.2]cryptand (crypt-222) sequesters the Al–Mg, Al–Al, Al–Ca, Al–Cu, Al–Zn, Al–Y,45 Al–Ag and Al–Au
potassium cations to form the SIPs [K(crypt-222)][Al(xanthNON)] (Fig. 9).46 The ability of these compounds to activate small
28,44 [K(crypt-222)][Al(NON)] 29,43 and [K(crypt-222)][Al(NC2NDipp)]
30.31 The addition of an excess of 18-crown-6 to 3Dipp also gives
rise to a SIP, with formula [K(18-c-6)]1.5[Al(NC2NDipp)] 31.31
Crystallographic analysis confirms that in each case there are
no contacts between the cationic and anionic components, effec-
tively releasing the ‘naked’ aluminyls for reactivity.
To date, we have isolated the only examples of SIPs for
lithium and sodium aluminyls (Fig. 8).43 As for the potassium
aluminyl, the addition of [2.2.2]cryptand to CDP 8 afforded the
SIP, [Na(crypt-222)][Al(NON)] 32. To form the lithium SIP, we
crystallised a sample of CDP 7 directly from TMEDA, affording

Fig. 9 Bimetallic complexes derived from aluminyls. NHCiPr = N,N 0 -


Fig. 8 Examples of the different structural type for the sodium aluminyl diisopropyl-4,5-dimethyl-2-ylidene; IPr = N,N 0 -bis(2,6-diisopropylphenyl)-
Na[Al(NON)]. (a) Slipped contacted dimeric pair [Na{Al(NON)}]2 (8), (b) imidazol-2-ylidene; Me2CAAC = 1-(2,6-diisopropylphenyl)-3,3,5,5-tetrame-
Monomeric ion pair (NON)Al–Na(TMEDA)2 (25), (c) Separated ion pair, thylpyrrolidin-2-ylidene; CyCAAC = 2-[2,6-bis(1-methylethyl)phenyl]-3,
[Na(crypt-222)][Al(NON)] (32). 3-dimethyl-2-azaspiro[4.5]dec-1-ylidene.

508 | Chem. Commun., 2023, 59, 503–519 This journal is © The Royal Society of Chemistry 2023
View Article Online

ChemComm Feature Article

molecules is an active area of research, with key results studied experimentally and using computational methods.53
described below. Initial reports on the reaction of the Al–Au system 42 with CO2
The monoanionic ligands [BDIAr] (Ar = Mes, Dipp) have identified the product as the dioxocarbene (xanthNON)Al(O2C)
been utilised to support the partner metal in a series of Al–M Au(PtBu3) 50 (Type-B), with crystallographic analysis confirming a
compounds, (xanthNON)Al–M(BDIMes) (34, M = Mg;28 35, M = m-1k2O,O0 :2kC-bonding mode for the CO2-unit formed by the
Zn),39 accessed via a metathesis route with the corresponding reductive insertion of CO2 into the Al–Au bond.50 This result has
(BDIMes)M–I reagents. More recently, an alternative (higher since been obtained with other Al–M compounds, affording the
yielding) route to 34 was presented,39 involving the reduction related dioxocarbenes for M = Cu: (NC2NDipp)Al(O2C)Cu(PtBu3) 51,
of (xanthNON)Al–I with Jones’ Mg(I) reagent, [(BDIMes)Mg]2.47 (NC2NDipp)Al(O2C)Cu(NHCiPr) 52, (NC2NDipp)Al(O2C)Cu(Me2CAAC)
Published on 08 December 2022. Downloaded by University of Liverpool on 12/30/2023 6:45:00 PM.

Related chemistry with potassium aluminyls 2 and 3Dipp have 53; M = Ag: (xanthNON)Al(O2C)Ag(PtBu3) 54, (NC2NDipp)Al(O2C)
afforded (NON)Al–M(BDIMes) (36, M = Mg; 37, M = Zn48) and Ag(NHCiPr) 55; M = Au: (NC2NDipp)Al(O2C)Au(NHCiPr) 56.49,51,52
(NC2NDipp)Al–M(BDIDipp) (38, M = Mg; 39, M = Ca30) When the CO2 reduction reaction was attempted with the
compounds. Al–Cu compound 40, rather than affording the dioxocarbene
The bulky phosphine ligand PtBu3 has been used to stabilise the product was identified as the carbonate (xanthNON)Al(O2-
a series of (xanthNON)Al–M(PtBu3) complexes of the coinage CO)Cu(PtBu3) 57, with a m-1k2O,O 0 :2kO00 -bonding mode for the
metals (40, M = Cu;49 41, M = Ag;49 42, M = Au50) and the alkyl carbonate identified by X-ray crystallography (Type-D).49 The
bridged derivative (NC2NDipp)Al–Cu(PtBu3) 43.51 Related com- propensity of the (xanthNON)Al–M(PtBu3) systems to form
pounds were isolated using carbenes as the neutral ligand, the carbonate complexes was investigated further, with the
affording a series of (NC2NDipp)Al–M(NHC) (44, M = Cu, NHC = results following the decrease in M–O bond strength and
NHCiPr;52 45, M = Ag, NHC = NHCiPr;51 46, M = Au, NHC = increase in M–C bond strength as the group 11 metal varies
NHCiPr;51 47, M = Cu, NHC = IPr51) and (NC2NDipp)Al–M(CAAC) from copper to gold. Hence, under 1 atmosphere of CO2, the
(48, M = Cu, CAAC = Me2CAAC;52 49, M = Ag, CAAC = CyCAAC51). Al–Cu system 40 forms the carbonate 57 at 78 1C, while the
3.1.2 Reactivity of Al–M complexes with CO2. The reported Al–Ag converted to the carbonate (xanthNON)Al(O2CO)Ag(PtBu3)
reactivity of these bimetallic complexes (Type-A, Scheme 2) has 58 over 24 h at 80 1C and the reaction of the Al–Au compound
focussed largely on the reduction of CO2. In this respect, the stopped after the initial insertion to form 50 (even when exposed
reactions of the coinage metal complexes with CO2 have been to additional CO2 under forcing conditions).49 The formation of
analogous carbonate species has since been observed when the
dioxocarbene 51 was heated to 60 1C in an atmosphere of CO2,
affording (NC2NDipp)Al(O2CO)Cu(PtBu3) 59.51
Examination of the conversion of the dioxocarbene to the
carbonate using computational analysis has identified a rate
determining step involving the extrusion of CO, to form an
intermediate m-oxo bridged species (Type-C).49,51 This was
supported experimentally for the (xanthNON)Al–M(PtBu3) system,
where the reaction of the silver aluminyl 41 with N2O formed the
spectroscopically identified (xanthNON)Al(m-O)Ag(PtBu3), which
reacted with CO2 to form the expected carbonate product 58.49
Unfortunately in this case, the proposed oxo-intermediate was
too thermally sensitive to be isolated.
We have recently examined the reactivity of Al–Zn and Al–Mg
compounds 36 and 37 with CO2.48 During this study we found
that the Al–Zn compound reacts to form the dioxocarbene
(NON)Al(O2C)Zn(BDIMes) 60, with the same Type-B regiochem-
istry as that observed in the previous examples. In contrast, the
Al–Mg compound immediately forms a carbonate complex
(Type-D), which was isolated as the trimetallic compound
[Al(NON)(O2CQO)]2[Mg(THF)4] 61, most likely the product of
a Schlenk-type ligand redistribution. To investigate whether the
Al–Mg and Al–Zn bimetallic compounds 36 and 37 allowed
access to the intermediate m-oxo species (Type-C), reactions
were performed with N2O. The products formed were shown to
contain the predicted bridging oxo-unit, (NON)Al(m-O)Zn-
Scheme 2 General reaction scheme summarizing the conversion of
(BDIMes) 62 (Fig. 10a) and (NON)Al(m-O)Mg(BDIMes) 63, verified
carbon dioxide CO2 to carbonate [CO3]2 promoted by bimetallic com-
plexes derived from aluminyls. Type-A = bimetallic complex; Type-B =
through a X-ray diffraction study. Furthermore, we showed that
dioxocarbene complex; Type-C = m-oxo complex; Type-D = carbonate both species react with additional CO2 to form the carbonates 61
complex. (M = Mg) and (NON)Al(O2CQO)Zn(BDIMes) 64 (M = Zn, Fig. 10b),

This journal is © The Royal Society of Chemistry 2023 Chem. Commun., 2023, 59, 503–519 | 509
View Article Online

Feature Article ChemComm


Published on 08 December 2022. Downloaded by University of Liverpool on 12/30/2023 6:45:00 PM.

Fig. 10 (a) Crystal structure of the m-oxo complex (NON)Al(m-


O)Zn(BDIMes) 62. (b) Crystal structure of the carbonate complex, (NON)-
Al(O2CO)Zn(BDIMes) 64.

demonstrating that the m-oxo compounds are intermediates in


the conversion of CO2 to [CO3]2.
3.1.3 Reactivity of Al–M complexes towards other
substrates. In related studies, Al–M bimetallic complexes (M =
Scheme 4 Alkyne insertion and related chemistry of the bimetallic Al–Cu
Cu, Ag, Au) have been reacted with carbodiimides,
compound, (xanthNON)Al–Cu(PtBu3) 40.
RNQCQNR.39,49–52 The resulting products show two possible
bonding modes for the inserted carbodiimide (Scheme 3). For
example, the reaction of the Al–Cu compound 48 with diiso- The reactivity profile of (xanthNON)Al–Cu(PtBu3) 40 has
propylcarbodiimide (DIC) gave a m-1k2N,N 0 :2kC-bonding mode recently been extended to the insertion of internal alkynes into
(Type-E) in (NC2NDipp)Al({iPrN}2C)Cu(Me2CAAC) 65, which is an Al–Cu bonds (Scheme 4).54 The symmetrical substrate 3-hexyne
analogous bonding mode to the CO2 insertion product 53. afforded (xanthNON)Al(m-EtCQCEt)Cu(PtBu3) 67, initially formed
However, under the same conditions, the NHC supported as the syn-product that thermally isomerises to the anti-product
Al–Cu compound 44 gave the insertion product (NC2NDipp) upon gentle heating. Insertion of the non-symmetrical 1-phenyl-
Al({iPrN}2C)Cu(NHCiPr) 66 with a m-1k2N,C:2kN 0 -bonding mode 1-pentyne substrate afforded (xanthNON)Al(m-nPrCQCPh)-
(Type-F).51,52 Further studies concluded that the divergent Cu(PtBu3) 68. Analysis of the product showed a mixture of two
reactivity leading to products with Type-E and Type-F coordina- regioisomers for the syn-product, with the major product having
tion, demonstrated an ambiphilicity of the Al–Cu bond that was the alkyl substituent geminal to the aluminium. The reaction of
dependent on the identity of the carbene co-ligand in the the insertion products 67 and 68 with CO was shown to be
bimetallic starting material.52 selective for the syn-isomers affording (xanthNON)Al{m-EtCQC(Et)
CQO}Cu(PtBu3) 69 and (xanthNON)Al{m-nPrCQC(Ph)CQO}
Cu(PtBu3) 70, believed to be due to interactions with the prox-
imal aluminium centre.

3.2 Oxidative addition to aluminyls


The key to unlocking catalytic activity for many main group
metal systems is the ability to control the redox activity. Under-
standing the oxidative addition to reduced metal centres, and
onwards work towards developing reductive elimination steps
will open the opportunity to establish catalytic cycles reminiscent
of those dominant in transition metal chemistry.55 The thermo-
dynamically favoured two-electron oxidation of Al(I) to Al(III) is a
common reaction for aluminyls, and the oxidative addition of
substrates via the cleavage of s-bonds has been documented for
all systems.
3.2.1 Oxidative addition of H2. Initial work with aluminyl
compound 1 demonstrated the cleavage of H–H bonds and
oxidative addition of dihydrogen to afford [K{Al(xanthNON)
(H)2}]2 71 (Scheme 5).28 This reaction has also been studied
Scheme 3 General reaction scheme summarizing the different bonding computationally.56 Experimentally, full conversion of 1 to 71
modes for the products of carbodiimide insertion into Al–M bonds. occurred over 5 days at room temperature under 2 bar

510 | Chem. Commun., 2023, 59, 503–519 This journal is © The Royal Society of Chemistry 2023
View Article Online

ChemComm Feature Article

meric ether adduct, [K(Et2O)2][Al(NON)(H)(NHDipp)] 77,58


respectively. Compounds 2 and 5 also undergo the oxidative
addition of a hydridic E–H bond from PhSiH3, giving [K(12-c-
4)2][Al(CSiNAd)(H)(SiH2Ph)] 78,34 and [K(Et2O)][Al(NON)(H)
(SiH2Ph)] 79,58 respectively. We have also shown that 2 will
undergo oxidative addition with P–H and O–H bonds with
MesPH2 and Ar*OH (Ar* = 2,6-tBu2-4-MeC6H3) to give [K(Et2O)2]
[Al(NON)(H)(PHMes)] 80 and K[Al(NON)(H)(OAr*)] 81.58
Published on 08 December 2022. Downloaded by University of Liverpool on 12/30/2023 6:45:00 PM.

3.2.3 Oxidative addition of C–H bonds. In contrast to


Al(BDIDipp) that is known to be stable in aromatic solvents
benzene and toluene,59 the reaction of aluminyl compounds
with sp2 C–H bonds has been realised in several systems
(Scheme 7). It was initially discovered that the oxidative
addition of a C–H bond of benzene to 1 occurred over 4 days
at 57 1C to afford the (hydrido)(phenyl) aluminate product,
Scheme 5 Oxidative addition of H2 bonds to potassium (1 and 2), sodium
[K2{Al(xanthNON)(H)(C6H5)}]2 82.28 The more reactive dialkyl
(8) and lithium (20) aluminyls.
aluminyl 6 formed the corresponding oxidative addition
product [K(tol)2][Al({CSi}2)(H)(C6H5)] 83 in 2.5 hours at room
atmosphere pressure, with the isolated dihydroaluminate pro- temperature.35 In contrast, we have found that only the cae-
duct crystallizing as the dimer. Subsequently, we investigated the sium CDP 14 is able to activate aromatic substituents for
reaction of H2 with the series of lighter alkali metal aluminate the NON-aluminyl system, requiring temperatures of 80 1C for
compounds, using the MIP 20 (for solubility purposes), and the 5 days to give the corresponding product, Cs[Al(NON)(H)(C6H5)]
potassium and sodium CDPs 2 and 8, respectively.38 We found 84.42
that more forcing conditions were required to effect oxidative The reaction of 4Dipp with benzene was reportedly ‘‘very
addition of H2 for the NON-system, and performed the reactions slow’’ at 20 1C. However it was found that in this case, rather
at 100 1C under 1.5 bar of H2. Using the time taken to achieve than forming the corresponding (hydrido)(phenyl) oxidative
50% conversion (t1/2) as an approximation of the rate of reaction, addition product analogous to 82–84, when gently heated to
the relative order Li (t1/2 = 1.5 days)c Na (t1/2 = 6 days) 4 K (t1/2 = 35 1C for a week the para-phenylene product K2[1,4-{Al(BDIDipp-H)
12 days) was observed. The compounds were isolated as the (H)}2(C6H4)] 85 was formed.32 The analogous double activation
dimers [Na{Al(NON)(H)2}]2 72 and [K{Al(NON)(H)2}]2 73, product has been identified as one of the products formed during
although to obtain a pure compound for the lithium complex the decomposition of the corresponding Cs-CDP 16 in benzene,
it was necessary to crystallize the product from Et2O, affording which was crystallographically identified as the THF-adduct,
(NON)Al(m-H)2Li(Et2O)2 74. [Cs(THF)2]2[1,4-{Al(BDIDipp-H)(H)}2(C6H4)] 86.41
3.2.2 Oxidative addition of E–H bonds. The oxidative The incorporation of a methyl substituent on the C6-ring
addition of element-hydrogen (E–H) bonds has been investigated promotes competing oxidative addition pathways involving the
for acidic (N–H, O–H), hydridic (Si–H) and non-polar (P–H, C–H) aromatic or aliphatic C–H bonds (Scheme 8). A mixture of the
bonds (Scheme 6). Cleavage of N–H bonds was demonstrated
when 5 was exposed to excess ammonia, affording the amido
complex [K(12-c-4)2][Al(CSiNAd)(H)(NH2)] 75.34 The reaction of 1
and 2 with DippNH2 afforded the corresponding dimeric anilido
product [K{Al(xanthNON)(H)(NHDipp)}]2 76,57 and the mono-

Scheme 7 Oxidative addition of C–H bonds of benzene to potassium


Scheme 6 Oxidative addition of E–H bonds to aluminyls. (1, 4Dipp, 6) and caesium (14, 16) aluminyls.

This journal is © The Royal Society of Chemistry 2023 Chem. Commun., 2023, 59, 503–519 | 511
View Article Online

Feature Article ChemComm

reaction of aluminyls 1 and 2 with ethylene under ambient


conditions (1 bar, room temperature) form the corresponding
aluminacyclopropanes [{K(C6H6)}{Al(xanthNON)(Z2-C2H4)}]2
95, and [{K(C6H6)}{Al(NON)(Z2-C2H4)}]2 96,64 consistent with
44

a (1+2) cycloaddition of the aluminyl to the CQC double bond.


Analytical data show that the C2H4 protons resonate at high
field (95, dH = 0.70; 96, dH = 1.40 in C6D6 and dH = 0.81 in
THF-D8) in the 1H NMR spectrum, and the C–C bond lengths
are B20% longer than in ethene (95, 1.598(2) Å; 96, 1.587(2) Å/
Published on 08 December 2022. Downloaded by University of Liverpool on 12/30/2023 6:45:00 PM.

1.596(2) Å), suggesting an activated system.


To investigate the onwards functionalisation of the activated
ethene, we have examined the reaction of the aluminacyclopro-
pane group in 96 with carbon monoxide, affording a series of
ethene carbonylation products (Scheme 9).64 Initial work with
dimeric 96 afforded the dialuminium product, K2[{Al(NON)
(Et)(m-OCQCHCH2)Al(NON)}] 97, containing an unprecedented
m-prop-1-ene-1,3-diyl-1k2C-1-olato-2kO ligand. To restrict the
Scheme 8 Oxidative addition of substituted arenes to potassium (1, 3Dipp,
6) aluminyls showing meta-selective and benzylic activation of C–H
reactivity to a single Al-centre, 18-crown-6 was added to 96 to
bonds. give the expected monomeric [K(18-c-6)][Al(NON)(Z2-C2H4)]
species 98. Accordingly, exposure of the aluminacyclopropane
group in 98 to CO gave the insertion product [K(18-c-6)]
meta-activated K[Al(xanthNON)(H)(3-MeC6H4)] 87 (Type-G, R 0 = [Al(NON)(CH2CH2CQO)] 99, containing the propan-1,3-diyl-
Me, R00 = H) and benzylic K[Al(xanthNON)(H)(CH2Ph)] 88 (Type-H, 1k2C-1-one ligand. This strained metallacycle underwent a
R 0 = R00 = H) products were formed in a 3 : 1 ratio when CDP 1 was thermally induced 1,2-hydrogen shift at 80 1C to afford the
heated to 80 1C in toluene for 2 days.60 In contrast, the Type-G 1-propen-1-olate compound, [K(18-c-6)(THF)2][Al(NON)(CH2-
meta-activated product K[Al({CSi}2)(H)(3-MeC6H4)] 89 formed CHQCHO)] 100.
exclusively when a sample of 6 was left at room temperature in
toluene for 12 h,61 and heating a sample of 3Dipp to 110 1C in 3.4 Aluminyl promoted oxidation chemistry
toluene for 24 h proceeded with Type-H benzylic activation to For many researchers in the area, the chemistry of aluminyls is
afford K[Al(NC2NDipp)(H)(CH2Ph)] 90.31 progressing beyond that of being an academic curiosity and
To probe the selectivity of the reaction with aluminyl com- becoming more targeted towards their application in the devel-
pound 1, the study was extended to the oxidative addition of opment of synthetically useful transformations. In this context,
xylenes and other dialkylated aromatic substrates. The reaction
with ortho-xylene (1,2-Me2C6H4) at 80 1C for 2 days exclusively
formed the benzylic activation product K[Al(xanthNON)(H)
{CH2(2-MeC6H4)}] 91 (Type-H, R 0 = Me, R00 = H), whereas
reaction with meta-xylene (1,3-Me2C6H4) under the same con-
ditions afforded a 2 : 3 mixture of meta-activated K[Al(xanthNON)
(H)(3,5-Me2C6H3)] 92 (Type-G, R 0 = R00 = Me) and benzylic
K[Al(xanthNON)(H)(CH2{3-MeC6H4})] 93 (Type-H, R 0 = H, R00 =
Me) products.60 To restrict the activation to exclusively meta-
selectively, the reaction of 1 with n-butylbenzene afforded
K[Al(xanthNON)(H)(3-nBuC6H4)] 94 (Type-G, R 0 = H, R00 = nBu).
DFT calculations performed on the reaction of 28 with n-
butylbenzene and the reaction of 6 with toluene indicated that
the activation energies required to promote meta-alumination
were lower than the competing benzylic-, ortho- and/or para-
alumination processes.60,61 Further DFT studies also revealed
the presence of Meisenheimer intermediates, consistent with a
concerted SNAr mechanism for arene C–H activation.32,44,62

3.3 Cycloaddition of aluminyls to unsaturated groups


Aluminyl compounds are known to form cycloaddition pro-
ducts with one (or more) unsaturated bonds within a substrate,
with documented examples showing addition to CRC,63 Scheme 9 Carbonylation of ethene promoted by a potassium alumina-
CQC,44,63,64 and CQN31 groups. In the simplest case, the cyclopropanes 96 and 98.

512 | Chem. Commun., 2023, 59, 503–519 This journal is © The Royal Society of Chemistry 2023
View Article Online

ChemComm Feature Article

the ability of the aluminyls to promote oxidation of small generating a linear [P4]4 ligand in K2[{Al(xanthNON)}2(m-Z2 : 2-P4)]
molecules is becoming established as an important reaction 105. The P–P bond lengths in the bridging P4-unit of 105 indicate a
type, and the following sections detail some of the areas in degree of p-bonding, confirmed using DFT calculations. It was also
which this reaction type has been recently applied. shown that the reaction of 105 with 1 atmosphere of ammonia
3.4.1 Reactivity with white phosphorus (P4). The functio- liberated PH3 from the reaction and generated a mixture of
nalisation of white phosphorus (P4) has been achieved by Al(xanthNON)(NH2)NH3 106 and [K{Al(xanthNON)(NH2)2}]2 107.
aluminyl compounds 1, 3Dipp and 5 (Scheme 10). In the Under more forcing conditions the 3Dipp/P4 system was
presence of 18-c-6, both 1 and 3Dipp form compounds containing shown to form higher P-containing oligomers. When performed
a doubly reduced [P4]2 unit in the products [K(18-c-6)][Al- at 60 1C in THF, the [P6]4 containing product, [{K(THF)2}
Published on 08 December 2022. Downloaded by University of Liverpool on 12/30/2023 6:45:00 PM.

(XanthNON)(Z1 : 1-P4)] 101,65 and [K(18-c-6)][Al(NC2NDipp)(Z1 : 1-P4)] {Al(NC2NDipp)}]2(m-P6) 108 was isolated in which the phosphorus
102,66 respectively. The corresponding ether-solvated analogue, group consisted of two cyclo-P3 subunits linked by a P–P bond.
[K(Et2O)3][Al(NC2NDipp)(Z1 : 1-P4)] 103 was obtained when the Furthermore, increasing the amount of P4 to 3 equivalents and
reaction with 3Dipp was performed in ether, which forms a 1-D heating the reaction to 60 1C for 8 days gave [{K(tol)2}
polymer with intermolecular K  P interactions. This contrasts the {Al(NC2NDipp)(m-P16)}]2 109 containing the [P16]4 cluster.
behaviour of the (alkyl)(amido) aluminyl 5 that extracted a single 3.4.2 Reactivity with ketones. A recent report described
P-atom when reacted with P4 to afford the bis-aluminyl phos- a divergent series of products obtained when the aluminyl
phide, [K(12-c-4)2][{Al(CSiNAd)}2(m-P)] 104.67 compound 3Dipp was reacted with ketones (Scheme 11).68 Based
The reaction of 101 with an additional 0.5 equivalents of the on previous work with in situ generated Al(BDIDipp),69 the
CDP 1 resulted in further reduction of the phosphorus unit, reaction was anticipated to proceed via reductive coupling of
two equivalents of ketone to afford the corresponding pinaco-
late structure, and indeed this is what is observed with aceto-
phenone to afford K[Al(NC2NDipp)({OC(H)Ph}2)] 110, which crystal-
lized as a single diastereoisomer. Although a reductive coupling
reaction is also observed when 3Dipp is reacted with 2 equivalents of
benzophenone, rather than forming the symmetrical pinacolate, a
C–C bond is formed between the ketyl carbon atom of one molecule
and an ortho-carbon atom of a phenyl group belonging to a second
equivalent, affording [K(THF)2][Al(NC2NDipp)(k2O,O0 ){OCPh2CH
(CHQCHCHQCH)CQCPhO}] 111. The reaction with one equiva-
lent of 2,4-dimethyl-3-pentanone afforded the potassium (hydrido)
(carboxyl)aluminate K[Al(NC2NDipp)(H){OC(iPr)QCMe2}] 112, which
corresponds to the formal oxidative addition of an enolisable

Scheme 11 Diverse reactivity of potassium aluminyl 3Dipp with ketones


acetophenone, benzophenone, 2,4-dimethyl-3-pentanone and 2,2,4,4-
Scheme 10 Oxidation of P4 by potassium aluminyls 1, 3Dipp and 5. tetramethyl-3-pentanone.

This journal is © The Royal Society of Chemistry 2023 Chem. Commun., 2023, 59, 503–519 | 513
View Article Online

Feature Article ChemComm

iso-propyl methine proton to aluminium. To avoid this possi- When the CDP 2 was reacted with CO under identical conditions
bility, one equivalent of the non-enolisable ketone 2,2,4,4- to that leading to the formation of 114, and the crude product
tetramethyl-3-penanone was reacted with 3Dipp. In this case, crystallized in the presence of TMEDA, the product was identified
the product K[Al(NC2N 0 ){OC(H)tBu2}] 113 (NC2N 0 = k3N,N 0 ,C- by X-ray crystallography as K5[K(TMEDA)][{Al(NONDipp)}4(C5O5)2]
[DippNSiMe2(C2H4)SiMe2NC6H3iPr{C(H)MeCH2}-2,6]3) showed 116 containing the pentameric [C5O5]5 ligand. The synthesis of
a hydroalumination pathway was in operation in which the 116 was also achieved from the reaction of the MIP 24 with CO at
ketone had been reduced to a [tBuC(H)O] alkoxy ligand, with 60 1C. Further analysis of the reaction of 2 with CO showed that it
a likely source of the hydridic hydrogen originating from an is not selective for the formation of 114, suggesting that decom-
Published on 08 December 2022. Downloaded by University of Liverpool on 12/30/2023 6:45:00 PM.

intramolecular oxidative addition of a methine C–H bond from position or higher oligomers could be forming.
an iPr substituent of a Dipp group. Shortly after this report, the reactivity of the CDP 1 and MIP
3.4.3 Reactivity with carbon monoxide. Another recent 10 with CO was described.71 Under ambient conditions, both
development in the application of aluminyl compounds for aluminyl compounds promoted tetramerization of CO in the
the reductive coupling of small molecules via C–C bond for- products to afford [{K(tol)2}{Al(xanthNON)}]2(C4O4) 117 and
mation involves their reactivity towards carbon monoxide [{Li(Et2O)}{Al(xanthNON)}]2(C4O4) 118, respectively. X-ray crystallo-
(Scheme 12). We reported that when exposed to an atmosphere graphy confirmed a trans-{syn,syn} configuration for the
of CO, the potassium aluminyl MIP 26 formed [K2{Al(NON)}2- [C4O4]4 chain, analogous to that found in 115. Interestingly,
(C4O4)]2 114, containing a m-(k2C,O-1,4-dioxido-2,3-diolato(Z)-but- when 117 was heated to 65 1C under CO at two atmosphere
2-en-1,4-diyl) tetraanion, with the cis-{anti,anti} configuration.70 pressure in THF, a further two equivalents of CO were incorpo-
The [C4O4]4 ligand is linear with bonds that approximate rated to afford K[K(THF)2][Al(xanthNON)]2(C6O6) 119 containing
to alternating C–C single and CQC double bonds, and the a novel cyclo-[C6O6]4 ligand with a C5O-pyran core. It was also
compound crystallised as the dimer with K  O interactions join- noted that a thermally induced isomerization of the trans-
ing the two monomeric units. A rearrangement of the C4-chain to {syn,syn}-C4O4 ligand in 117 to the cis-{anti,anti}-C4O4 ligand
form the corresponding (1,4-dioxido-2,3-diolato(E)-but-2-en-1,4- in the new compound [K(THF)]2[{Al(xanthNON)}2(C4O4)] 120
diyl) ligand with a trans-{syn,syn} configuration in [{K(TMEDA)} occurred when heated to 65–70 1C in the presence of K[HBEt3].
{Al(NONDipp)}]2(C4O4) 115 took place when a sample of 114 was In addition to the [C4O4]4 and [C6O6]4 ligands present in
heated under an atmosphere of CO in the presence of TMEDA. 117–120, this study contained reports of the [C4O4]6 ligand.

Scheme 12 Summary of reactions of CDPs (1 and 2) and MIPs (10 and 26) with CO, forming complexes containing [C4O4]4 (3 isomers: {114, 120}, 115,
{117, 118}), [C5O5]5 (116) and [C6O6]4 (119) and [C4O4]6 (121) ligands.

514 | Chem. Commun., 2023, 59, 503–519 This journal is © The Royal Society of Chemistry 2023
View Article Online

ChemComm Feature Article

Thus, whilst isolated samples of 117 were resistant to further The highly reactive Al–Ooxide bond has also been used to pro-
reduction using Li[HBEt3] or K[HBEt3], the in situ reaction of mote C–C bond forming reactions (Scheme 13). The addition of
K[HBEt3] with 1/CO afforded K4[{Al(xanthNON)}2(C4O4)(BEt3)2] excess xylyl isonitrile (Ar‡NC, Ar‡ = 2,6-Me2C6H3) to a solution of
121 with the formally hexa-anionic [C4O4]6 ligand.71 To com- 130 resulted in uptake of two equivalents and formation of the C–C
plete the series of [C4O4]n ligands for n = 2, 4 and 6, the coupled product [K(THF)3][Al(xanthNON)(Ar‡NQC–C{O}QNAr‡)]
reaction of (xanthNON)Al–I with the silver salt of the pre-formed 134.57 As part of our investigations, we reacted the alumoxane 131
squarate dianion, Ag2[C4O4], was conducted, affording the with CO to form a unique ethene tetraolate ligand, [O2CQCO2]4,
expected product, {Al(xanthNON)}2(C4O4) 122. in the product [K{Al(NON)(O2C)}]2 135.81 Using computational
Published on 08 December 2022. Downloaded by University of Liverpool on 12/30/2023 6:45:00 PM.

techniques, we showed that the dimeric composition of the alumi-


3.5 Aluminium–element multiple bonds nium oxide is maintained during this reaction, which proceeds via a
Low valent aluminium species have previously been used to dioxocarbene (akin to Type-B structures) that reacts with additional
access compounds containing Al–E multiple bonds.72 For CO to form an intermediate aluminium ketene.
example, iminoalanes were synthesized from Al(BDIDipp) (V)73 3.5.3 Al–S multiple bonds. Initial evidence for the formation
or Al(tBuBDIDipp) (VI) with the N-heterocyclic carbene, 1,3-iPr2- of the Al–S multiple bond from aluminyls was obtained when
4,5-Me2-2-ylidene.74 Similarly, Al–E bonds (where E = O,75 S76 CDP 2 was reacted with CS2, affording the trithiocarbonate
and Te77) have been characterised using neutral Al(I) species as complex, [K{Al(NON)(S2CQS)}]4 136.81 As for carbonate com-
a starting reagent. The aluminyl compounds therefore offered plexes 128 and 129, the implication of this result was the
an alternative entry point into this area of chemistry and have formation of a transient ‘Al–S’ multiple bond that reacted with
been exploited in the development of a range of Al–E multiple additional CS2 in a [2+2] cycloaddition. Repeating the reaction
bonded systems. with CS2 at low temperature generated [K{Al(NON)(S2C)}]2 137
3.5.1 Al–NR multiple bonds. As a continuation of our containing the ethene tetrathiolate anion, [S2CQCS2]4.
earlier work with the indyl anion [K{In(NON)}]2,26 we examined Although this bridging ligand is analogous to the ethene tetra-
the reaction of CDP 2 with mesityl azide as a potential route to olate anion in 135, the coordination to the two aluminium
0 0
Al–NR multiple bonds.78 The reaction proceeded smoothly to centres is different. In 135, a m-(1kO1,1kO1 :2kO2,2kO2 ) mode
afford the molecular aluminium imide [K{Al(NON)(NMes)}]2 123. is adopted which generates two AlO2C heterocycles, whereas in
0 0
Examination of the Al–Nimide bond using DFT calculations 137 the ligand is m-(1kS1,1kS2 : 2kS1 ,2kS2 ) with two AlS2C2 rings.
indicated that this unit has multiple bond character but contains Attempted oxidation of aluminyl 2 directly with elemental
a significant ionic contribution. The reactivity of the Al–NR sulfur failed to generate a terminal sulfide, instead forming the
group was consistent with an unsaturated bond, demonstrated 1,2,3,4,5-tetrathiaaluminyl anion in K[Al(NON)(S4)] 138. This
by a [2+2] cycloaddition reaction with CO2 to afford the alumi- prompted a ‘top down’ approach – namely the desulfurisation
nium carbamate, [K{Al(NON)(OC{O}NMes)}]2 124. reaction of 138 with PPh3, which successfully formed the target
The analogous reactions between CDP 1 and organic sulfide, isolated as the tetrameric species [K{Al(NON)(S)}]4 139.
azides has been studied.57 Reaction with DippN3 afforded The [Al(NON)(S)] anion was also isolated as a separated ion
[K{Al(xanthNON)(NDipp)}]2 125. Reactivity was observed with H2, pair in [K(crypt-222)][Al(NON)(S)] 140.82
which adds across the Al–Nimide bond to afford [K{Al(xanthNON)- Given the bio-relevance of sulfur containing organic com-
(NHDipp)(H)}]2 126. Further demonstration of the high reac- pounds, we assessed the potential for 139 to be used as a
tivity of 125 was noted during the reaction with CO, forming
[K{Al(xanthNON)(OC{O}CQNDipp)}]2 127, which involved the
complete cleavage of the CO triple bond in CO.
3.5.2 Al–O multiple bonds. When exposed to 1 atm of CO2,
CDPs 1 and 2 afforded form the carbonate complexes,
[K{Al(xanthNON)(O2CQO)}]2 128,79 and [K{Al(NON)(O2CQO)}]4
129,80 respectively. The proposed mechanism parallels that
discussed for the bimetallic carbonates (Section 3.1.2), involving
intermediate aluminium oxide species that are generated by
extrusion of CO. Under more controlled conditions using N2O as
the oxidant, the target aluminium oxides (alumoxanes) were
isolated as the dimeric species [K{Al(xanthNON)(O)}]2 130 and
[K{Al(NON)(O)}]2 131. Reactions of 130 and 131 with additional
N2O afforded the corresponding cis-hyponitrites [K{Al(xanthNON)-
(O2N2)}]2 132 and [K(crypt-222)][Al(NON)(O2N2)] 133, whilst
exposure to CO2 formed carbonates 128 and 129. These results
support the proposal that an Al–Ooxide bond is initially generated in
the reactions with CO2, and the final products are consistent with a Scheme 13 Reaction of aluminium oxides ([K{Al(xanthNON)(O)}])2 130 and
[2+2] cycloaddition of a second molecule of CO2 to an Al–Ooxide [K{Al(NON)(O)}]2 131 with isonitrile Ar‡NC (128, Ar‡ = xylyl = 2,6-Me2C6H3)
multiple bond. and CO (129).

This journal is © The Royal Society of Chemistry 2023 Chem. Commun., 2023, 59, 503–519 | 515
View Article Online

Feature Article ChemComm

reagent for the formation of new S–C bonds (Scheme 14).


Accordingly, the reaction with CO2 afforded the thiocarbonate
compound [{K(Et2O)2}{Al(NON)(SC{O}O)}]2 141, in a [2+2]
cycloaddition reaction analogous to the formation of 124.
Furthermore, we demonstrated that reaction with benzophe-
none incorporates a single equivalent in K[Al(NON)(SC{Ph2}O)]
142 containing the k2S,O-diphenylsulfidomethanolate ligand,
while two equivalents of benzaldhyde react to give the k2O,O 0 -
Published on 08 December 2022. Downloaded by University of Liverpool on 12/30/2023 6:45:00 PM.

thio-bis(phenylmethanolate), [K{Al(NON)({OC(H)Ph}2S)}]2 143.


3.5.4 Al–Se multiple bonds. Access to the aluminium
selenide containing a terminal Al–Se bond proved more
straight forward than the corresponding sulfide, with the direct
reaction of 2 with one equivalent of elemental selenide affording
[K(THF)][Al(NON)(Se)] 144, also structurally characterized as the
Scheme 15 Synthesis of aluminium telluride 150 and its reaction with
separated ion pair, [K(crypt-222)][Al(NON)(Se)] 145.83 When two
CO2 to afford the double insertion product 151.
equivalent of selenium were used per aluminium, a diselenirane
compound was generated that was isolated as the [2.2.2]cryptand
complex, [K(crypt-222)][Al(NON)(k2-Se2)] 146. The terminal a structurally authenticated terminal Al–Te bond. When exposed
selenide bond in 144 undergoes analogous C–E bond forming to an atmosphere of CO2, compound 150 reacted to form
reactions with carbon dioxide, benzophenone and benzaldehyde [K(Et2O)2(THF/Et2O)][Al(NON)({OC(O)}2Te)] 151. Rather than
as noted for the sulfide, forming [{K(THF)3}{Al(NON)(SeC{O}- the [2+2] cycloaddition of a single equivalent of CO2 to the Al–E
O)}]2 147, [K(THF)3][Al(NON)(SeC{Ph2}O)] 148 and [K{Al(NON)- multiple bonds noted during the formation of 124 (E = NMes),
({OC(H)Ph}2Se)}]2 149, respectively (Scheme 14). 129 (E = O), 141 (E = S) and 147 (E = Se), the reaction proceeds
3.5.5 Al–Te multiple bonds. As for the preparation of the directly to give the double insertion product containing the
selenide 144, the direct reaction of 2 with elemental tellurium tellurodicarbonate ligand. An intermediate complex observed
afforded the telluride, K[Al(NON)(Te)(THF)] 150 (Scheme 15).84 spectroscopically using 13CO2 was presumed to be the initially
In contrast to 144 however, in which the THF is bound to the formed tellurocarbonate complex containing the [OC{O}Te]2
potassium, the X-ray structure of 150 shows that it coordinates ligand, which reacts rapidly with a second equivalent of CO2 to
to aluminium to generate a four-coordinate anion that contains generate 151.
3.5.6 Al–C multiple bonds. The reaction of the lithium
aluminyl CDP 19 with the bulky diaryldiazomethane derivative
5,5 0 -(diazomethylene)bis(1,3-di-tert-butylphenyl) has been used
to access a novel compound containing an Al–C multiple bond
(Scheme 16).40 The reaction proceeds with the initial formation
0
of the [1+2] cycloaddition product, [Li(Et2O)2][Al(bo2e-CAr N)(N2 =
CAr 0 2)] 152. This complex thermally decomposes at 40 1C with
0
loss of N2 to afford [Li(Et2O)][Al(bo2e-CAr N)(CAr 0 2)] 153, which
was crystallised as the 12-c-4 complex, [Li(12-c-4)2][Al(bo2e-
0 0
CAr N)(CAr 2)] 154. The crystal structure of 154 shows a short

Scheme 14 Reaction of aluminium sulfide and selenide [K{Al(NON)(S)}]4


139 and [K(THF)][Al(NON)(Se)] 144 with CO2, Ph2CO and PhHCO. Scheme 16 Formation of an Al–C multiple bond (Ar 0 = 3,5-tBu2C6H3).

516 | Chem. Commun., 2023, 59, 503–519 This journal is © The Royal Society of Chemistry 2023
View Article Online

ChemComm Feature Article

Al–CAr 0 2 bond and DFT calculations identify key MOs containing mechanistic investigation using DFT concluded that an ionic
a p-component in the Al–C bond. pathway involving the Al–Sc bimetallic complex was operating.

Latest developments 4. Conclusions


There are two notable publications in this area that have been In the relatively brief time since their discovery, the chemistry
published since this original manuscript was submitted. We of aluminyls has expanded rapidly to encompass a diverse
have continued to explore the oxidative addition of E–H bonds range of reactions. Whilst aspects of aluminyl reactivity mimic
Published on 08 December 2022. Downloaded by University of Liverpool on 12/30/2023 6:45:00 PM.

to the [AM{Al(NON)}]2 (Section 3.2.2), confirming that the that documented with known Al(I) compounds, their enhanced
full series of alkali metal aluminyls form the (silyl)(hydrido) activity (compared with their neutral counterparts) and ability
product, [AM(L)n][Al(NON)(H)(SiH2Ph)] (AM(L)n = Li(Et2O) 155, to react with new substrates through the activation of otherwise
Na(Et2O) 156, K(Et2O) 79, Rb(Et2O) 157 and Cs(THF)2 158).85 stable chemical bonds is becoming increasingly evident.
Structural analysis of the isomorphous Na, K and Rb crystal Furthermore, it is apparent that the chemistry of the aluminyl
structures confirmed a tendency to form intramolecular anions cannot be considered in isolation. There is mounting
AM  (arene) interactions for the heavier alkali metals, noted evidence that the group 1 metal cations play a fundamental role
from a higher hapticity to the SiPh and NDipp groups. in directing the reactivity of these systems through synergistic
A new addition to the family of N,N 0 -(diamido) aluminyl interactions with many substrates. This is being developed by
anions has been reported and the chemistry of its bimetallic expanding the family of potassium aluminyls to encompass all
complex with scandium investigated.86 The relatively simple stable members of group 1 metals, and through an enhanced
1,3-propanediamido ligand [DippN(CH2)3NDipp]2 ([NC3N]2) control of the cation and anion interactions that are present.
supports the potassium aluminyl that has been isolated as the With these new tools in hand, the current goals of researchers
CDP 159 and the [K(crypt-222)]+ SIP, 160 (Scheme 17). Metathesis in this field involve the development of synthetically useful
reactions with Sc(NR2)2Cl(THF)n (R = SiMe3, n = 1; R = iPr, n = 2) transformations that will allow aluminyls to take a position at
affords the bimetallic Al–Sc complexes, (NC3N)Al–Sc(NR2)2(THF) the forefront of p-block metal chemistry.
161 (R = SiMe3) and 162 (R = iPr). Compound 161 undergoes a
thermal (room temperature) activation of a methyl group from
SiMe3 to afford a methylene-/hydrido-bridged species 163. When Conflicts of interest
the NiPr2 derivative 162 was reacted with Ph2CHBr in benzene, a
There are no conflicts to declare.
dialuminated 1,4-cyclohexadiene product 164 was isolated that
derives from the dearomatisation of C6H6. The side product
of the reaction was identified as 1,1,2,2-tetraphenylethane and Acknowledgements
This works was supported by Government funding from the
Marsden Fund Council, managed by the Royal Society Te
Apārangi (Grant Number: MFP-VUW2020).

References
1 W. Klemm, E. Voss and K. Geiersberger, Z. Anorg. Allg. Chem., 1948,
256, 15–24.
2 (a) M. Tacke and H. Schnoeckel, Inorg. Chem., 1989, 28, 2895–2896;
(b) M. Mocker, C. Robl and H. Schnöckel, Angew. Chem., Int. Ed.
Engl., 1994, 33, 1754–1755; (c) A. Ecker and H. Schnöckel, Z. Anorg.
Allg. Chem., 1996, 622, 149–152.
3 C. Dohmeier, C. Robl, M. Tacke and H. Schnöckel, Angew. Chem.,
Int. Ed. Engl., 1991, 30, 564–565.
4 S. Schulz, H. W. Roesky, H. J. Koch, G. M. Sheldrick, D. Stalke and
A. Kuhn, Angew. Chem., Int. Ed. Engl., 1993, 32, 1729–1731.
5 (a) S. González-Gallardo, T. Bollermann, R. A. Fischer and
R. Murugavel, Chem. Rev., 2012, 112, 3136–3170; (b) C. Gemel,
T. Steinke, M. Cokoja, A. Kempter and R. A. Fischer, Eur. J. Inorg.
Chem., 2004, 4161–4176; (c) G. Linti and H. Schnöckel, Coord. Chem.
Rev., 2000, 206–207, 285–319; (d) R. A. Fischer and J. Weiß, Angew.
Chem., Int. Ed., 1999, 38, 2830–2850.
6 A. Haaland, K.-G. Martinsen, S. A. Shlykov, H. V. Volden,
C. Dohmeier and H. Schnoeckel, Organometallics, 1995, 14,
3116–3119.
7 C. Dohmeier, E. Baum, A. Ecker, R. Köppe and H. Schnöckel,
Organometallics, 1996, 15, 4702–4706.
Scheme 17 Reactions of a new potassium N,N 0 -(diamido)aluminyl 8 H. Sitzmann, M. F. Lappert, C. Dohmeier, C. Üffing and
compound. H. Schnöckel, J. Organomet. Chem., 1998, 561, 203–208.

This journal is © The Royal Society of Chemistry 2023 Chem. Commun., 2023, 59, 503–519 | 517
View Article Online

Feature Article ChemComm

9 A. Hofmann, T. Tröster, T. Kupfer and H. Braunschweig, Chem. Sci., 38 M. J. Evans, M. D. Anker, C. L. McMullin, S. E. Neale and M. P. Coles,
2019, 10, 3421–3428. Angew. Chem., Int. Ed., 2021, 60, 22289–22292.
10 C. Cui, H. W. Roesky, H.-G. Schmidt, M. Noltemeyer, H. Hao and 39 M. M. D. Roy, J. Hicks, P. Vasko, A. Heilmann, A.-M. Baston,
F. Cimpoesu, Angew. Chem., Int. Ed., 2000, 39, 4274–4276. J. M. Goicoechea and S. Aldridge, Angew. Chem., Int. Ed., 2021, 60,
11 X. Li, X. Cheng, H. Song and C. Cui, Organometallics, 2007, 26, 22301–22306.
1039–1043. 40 C. Yan and R. Kinjo, Angew. Chem., Int. Ed., 2022, 61, e202211800.
12 S. Grams, J. Mai, J. Langer and S. Harder, Organometallics, 2022, 41, 41 S. Grams, J. Mai, J. Langer and S. Harder, Dalton Trans., 2022, 51,
2862–2867. 12476–12483.
13 (a) K. Hobson, C. J. Carmalt and C. Bakewell, Chem. Sci., 2020, 11, 42 T. X. Gentner, M. J. Evans, A. R. Kennedy, S. E. Neale,
6942–6956; (b) M. Zhong, S. Sinhababu and H. W. Roesky, Dalton C. L. McMullin, M. P. Coles and R. E. Mulvey, Chem. Commun.,
Trans., 2020, 49, 1351–1364; (c) Y. Liu, J. Li, X. Ma, Z. Yang and 2022, 58, 1390–1393.
Published on 08 December 2022. Downloaded by University of Liverpool on 12/30/2023 6:45:00 PM.

H. W. Roesky, Coord. Chem. Rev., 2018, 374, 387–415; (d) Y.-C. Tsai, 43 M. J. Evans, M. D. Anker, M. G. Gardiner, C. L. McMullin and
Coord. Chem. Rev., 2012, 256, 722–758; (e) M. Asay, C. Jones and M. P. Coles, Inorg. Chem., 2021, 60, 18423–18431.
M. Driess, Chem. Rev., 2011, 111, 354–396; ( f ) S. Nagendran and 44 J. Hicks, P. Vasko, J. M. Goicoechea and S. Aldridge, J. Am. Chem.
H. W. Roesky, Organometallics, 2008, 27, 457–492; ( g) H. W. Roesky Soc., 2019, 141, 11000–11003.
and S. S. Kumar, Chem. Commun., 2005, 4027–4038; 45 K. Sugita and M. Yamashita, Chem. – Eur. J., 2020, 26, 4520–4523.
(h) M. N. Sudheendra Rao, H. W. Roesky and G. Anantharaman, 46 The aluminium-titanium complex ({CSi}2)Al(m-OiPr)2Ti(OiPr)2 has
J. Organomet. Chem., 2002, 646, 4–14; (i) C. Dohmeier, D. Loos and also been isolated from an aluminyl starting reagent. In this
H. Schnöckel, Angew. Chem., Int. Ed. Engl., 1996, 35, 129–149. instance the metals do not form a direct bond but are linked by
14 J. D. Queen, A. Lehmann, J. C. Fettinger, H. M. Tuononen and bridging isopropoxide ligands:K. Sugita and M. Yamashita, Organo-
P. P. Power, J. Am. Chem. Soc., 2020, 142, 20554–20559. metallics, 2020, 39, 2125–2129.
15 (a) X. Zhang and L. L. Liu, Angew. Chem., Int. Ed., 2021, 60, 47 S. P. Green, C. Jones and A. Stasch, Science, 2007, 318, 1754–1757.
27062–27069; (b) A. Hinz and M. P. Müller, Chem. Commun., 2021, 48 M. J. Evans, G. H. Iliffe, S. E. Neale, C. L. McMullin, J. R. Fulton,
57, 12532–12535. M. D. Anker and M. P. Coles, Chem. Commun., 2022, 58,
16 X. Zhang, Y. Mei and L. L. Liu, Chem. – Eur. J., 2022, e202202102. 10091–10094.
17 As noted previously (ref. 27), the recommended nomenclature for 49 C. McManus, J. Hicks, X. Cui, L. Zhao, G. Frenking, J. M. Goicoechea
these anions is boranyl, alumanyl, gallanyl and indinanyl (based on and S. Aldridge, Chem. Sci., 2021, 12, 13458–13468.
uncharged EH2. See R. M. Hartshorn and A. T. Hutton, in Nomen- 50 J. Hicks, A. Mansikkamäki, P. Vasko, J. M. Goicoechea and
clature of Inorganic Chemistry, IUPAC Recommendations 2005, IUPAC S. Aldridge, Nat. Chem., 2019, 11, 237–241.
Red Book, ed. N. G. Connelly and T. Damhus, RSC Publishing, 51 H.-Y. Liu, S. E. Neale, M. S. Hill, M. F. Mahon and C. L. McMullin,
London, 2005, However, for consistency with the existing literature Dalton Trans., 2022, 51, 3913–3924.
and in agreement with most researchers in the area, we have 52 H.-Y. Liu, R. J. Schwamm, M. S. Hill, M. F. Mahon, C. L. McMullin
adopted the terms boryl, aluminyl, gallyl and indyl to describe and N. A. Rajabi, Angew. Chem., Int. Ed., 2021, 60, 14390–14393.
generic EX2 systems and their corresponding anions. 53 (a) D. Sorbelli, L. Belpassi and P. Belanzoni, J. Am. Chem. Soc., 2021,
18 E. S. Schmidt, A. Jockisch and H. Schmidbaur, J. Am. Chem. Soc., 143, 14433–14437; (b) D. Sorbelli, L. Belpassi and P. Belanzoni, Inorg.
1999, 121, 9758–9759. Chem., 2022, 61, 1704–1716; (c) D. Sorbelli, L. Belpassi and
19 R. J. Baker, R. D. Farley, C. Jones, M. Kloth and D. M. Murphy, P. Belanzoni, Chem. Sci., 2022, 13, 4623–4634; (d) X. Guo, T. Yang,
J. Chem. Soc., Dalton Trans., 2002, 3844–3850. Y. Zhang, F. K. Sheong and Z. Lin, Inorg. Chem., 2022, 61, 10255–10262.
20 Y. Segawa, M. Yamashita and K. Nozaki, Science, 2006, 314, 113–115. 54 C. McManus, A. E. Crumpton and S. Aldridge, Chem. Commun.,
21 Y. Segawa, Y. Suzuki, M. Yamashita and K. Nozaki, J. Am. Chem. Soc., 2022, 58, 8274–8277.
2008, 130, 16069–16079. 55 T. Chu and G. I. Nikonov, Chem. Rev., 2018, 118, 3608–3680.
22 A. V. Protchenko, P. Vasko, M. Á. Fuentes, J. Hicks, D. Vidovic and 56 N. Villegas-Escobar, A. Toro-Labbé and H. F. Schaefer Iii, Chem. –
S. Aldridge, Angew. Chem., Int. Ed., 2021, 60, 2064–2068. Eur. J., 2021, 27, 17369–17378.
23 (a) N. Metzler-Nolte, New J. Chem., 1998, 22, 793–795; 57 A. Heilmann, J. Hicks, P. Vasko, J. M. Goicoechea and S. Aldridge,
(b) A. Sundermann, M. Reiher and W. W. Schoeller, Eur. J. Inorg. Angew. Chem., Int. Ed., 2020, 59, 4897–4901.
Chem., 1998, 305–310; (c) H. M. Tuononen, R. Roesler, J. L. Dutton 58 M. J. Evans, M. D. Anker and M. P. Coles, Inorg. Chem., 2021, 60,
and P. J. Ragogna, Inorg. Chem., 2007, 46, 10693–10706. 4772–4778.
24 R. J. Schwamm, M. D. Anker, M. Lein, M. P. Coles and C. M. Fitchett, 59 T. Chu, I. Korobkov and G. I. Nikonov, J. Am. Chem. Soc., 2014, 136,
Angew. Chem., Int. Ed., 2018, 57, 5885–5887. 9195–9202.
25 R. J. Schwamm, J. R. Harmer, M. Lein, C. M. Fitchett, S. Granville 60 J. Hicks, P. Vasko, A. Heilmann, J. M. Goicoechea and S. Aldridge,
and M. P. Coles, Angew. Chem., Int. Ed., 2015, 54, 10630–10633. Angew. Chem., Int. Ed., 2020, 59, 20376–20380.
26 M. D. Anker, M. Lein and M. P. Coles, Chem. Sci., 2019, 10, 61 S. Kurumada, K. Sugita, R. Nakano and M. Yamashita, Angew.
1212–1218. Chem., Int. Ed., 2020, 59, 20381–20384.
27 J. Hicks, P. Vasko, J. M. Goicoechea and S. Aldridge, Angew. Chem., 62 J. J. Cabrera-Trujillo and I. Fernández, Chem. – Eur. J., 2020, 26,
Int. Ed., 2021, 60, 1702–1713. 11806–11813.
28 J. Hicks, P. Vasko, J. M. Goicoechea and S. Aldridge, Nature, 2018, 63 K. Sugita, R. Nakano and M. Yamashita, Chem. – Eur. J., 2020, 26,
557, 92–95. 2174–2177.
29 R. J. Schwamm, M. D. Anker, M. Lein and M. P. Coles, Angew. Chem., 64 M. J. Evans, S. E. Neale, M. D. Anker, C. L. McMullin and M. P. Coles,
Int. Ed., 2019, 58, 1489–1493. Angew. Chem., Int. Ed., 2022, 61, e202117396.
30 R. J. Schwamm, M. P. Coles, M. S. Hill, M. F. Mahon, C. L. McMullin, 65 M. M. D. Roy, A. Heilmann, M. A. Ellwanger and S. Aldridge, Angew.
N. A. Rajabi and A. S. S. Wilson, Angew. Chem., Int. Ed., 2020, 59, 3928–3932. Chem., Int. Ed., 2021, 60, 26550–26554.
31 R. J. Schwamm, M. S. Hill, H.-Y. Liu, M. F. Mahon, C. L. McMullin 66 M. S. Hill, M. F. Mahon, C. L. McMullin, S. E. Neale, K. G. Pearce and
and N. A. Rajabi, Chem. – Eur. J., 2021, 27, 14971–14980. R. J. Schwamm, Z. Anorg. Allg. Chem., 2022, 648, e202200224.
32 S. Grams, J. Eyselein, J. Langer, C. Färber and S. Harder, Angew. 67 K. Koshino and R. Kinjo, Organometallics, 2020, 39, 4183–4186.
Chem., Int. Ed., 2020, 59, 15982–15986. 68 H.-Y. Liu, M. S. Hill and M. F. Mahon, Chem. Commun., 2022, 58,
33 D. Bourissou, O. Guerret, F. P. Gabbaı̈ and G. Bertrand, Chem. Rev., 6938–6941.
2000, 100, 39–92. 69 C. Cui, S. Köpke, R. Herbst-Irmer, H. W. Roesky, M. Noltemeyer,
34 K. Koshino and R. Kinjo, J. Am. Chem. Soc., 2020, 142, 9057–9062. H.-G. Schmidt and B. Wrackmeyer, J. Am. Chem. Soc., 2001, 123,
35 S. Kurumada, S. Takamori and M. Yamashita, Nat. Chem., 2020, 12, 36–39. 9091–9098.
36 T. X. Gentner and R. E. Mulvey, Angew. Chem., Int. Ed., 2021, 60, 70 M. J. Evans, M. G. Gardiner, M. D. Anker and M. P. Coles, Chem.
9247–9262. Commun., 2022, 58, 5833–5836.
37 (a) J. M. Gil-Negrete and E. Hevia, Chem. Sci., 2021, 12, 1982–1992; 71 A. Heilmann, M. M. D. Roy, A. E. Crumpton, L. P. Griffin, J. Hicks,
(b) S. D. Robertson, M. Uzelac and R. E. Mulvey, Chem. Rev., 2019, J. M. Goicoechea and S. Aldridge, J. Am. Chem. Soc., 2022, 144,
119, 8332–8405. 12942–12953.

518 | Chem. Commun., 2023, 59, 503–519 This journal is © The Royal Society of Chemistry 2023
View Article Online

ChemComm Feature Article

72 (a) D. Franz and S. Inoue, Dalton Trans., 2016, 45, 9385–9397; 79 J. Hicks, A. Heilmann, P. Vasko, J. M. Goicoechea and S. Aldridge,
(b) P. Bag, C. Weetman and S. Inoue, Angew. Chem., Int. Ed., 2018, Angew. Chem., Int. Ed., 2019, 58, 17265–17268.
57, 14394–14413; (c) C. Weetman, Chem. – Eur. J., 2021, 27, 80 M. D. Anker and M. P. Coles, Angew. Chem., Int. Ed., 2019, 58,
1941–1954; (d) F. Dankert and C. Hering-Junghans, Chem. Commun., 18261–18265.
2022, 58, 1242–1262. 81 M. D. Anker, C. L. McMullin, N. A. Rajabi and M. P. Coles, Angew.
73 N. J. Hardman, C. Cui, H. W. Roesky, W. H. Fink and P. P. Power, Chem., Int. Ed., 2020, 59, 12806–12810.
Angew. Chem., Int. Ed., 2001, 40, 2172–2174. 82 M. J. Evans, M. D. Anker, C. L. McMullin, S. E. Neale, N. A. Rajabi
74 J. Li, X. Li, W. Huang, H. Hu, J. Zhang and C. Cui, Chem. – Eur. J., and M. P. Coles, Chem. Sci., 2022, 13, 4635–4646.
2012, 18, 15263–15266. 83 M. D. Anker and M. P. Coles, Angew. Chem., Int. Ed., 2019, 58,
75 D. Neculai, H. W. Roesky, A. M. Neculai, J. Magull, B. Walfort and 13452–13455.
D. Stalke, Angew. Chem., Int. Ed., 2002, 41, 4294–4296. 84 M. J. Evans, M. D. Anker, C. L. McMullin, N. A. Rajabi and
Published on 08 December 2022. Downloaded by University of Liverpool on 12/30/2023 6:45:00 PM.

76 T. Chu, S. F. Vyboishchikov, B. Gabidullin and G. I. Nikonov, Angew. M. P. Coles, Chem. Commun., 2021, 57, 2673–2676.
Chem., Int. Ed., 2016, 55, 13306–13311. 85 G. M. Ballmann, M. J. Evans, T. X. Gentner, A. R. Kennedy,
77 D. Franz, T. Szilvási, E. Irran and S. Inoue, Nat. Commun., 2015, J. R. Fulton, M. P. Coles and R. E. Mulvey, Inorg. Chem., 2022, 61,
6, 10037. 19838–19846.
78 M. D. Anker, R. J. Schwamm and M. P. Coles, Chem. Commun., 2020, 86 G. Feng, K. L. Chan, Z. Lin and M. Yamashita, J. Am. Chem. Soc.,
56, 2288–2291. 2022, 144, 22662–22668.

This journal is © The Royal Society of Chemistry 2023 Chem. Commun., 2023, 59, 503–519 | 519

You might also like