You are on page 1of 45

CHAPTER FIVE

Metallate Complexes of the Late


Transition Metals: Organometallic
Chemistry and Catalysis
Adrián Gómez-Suárez*, David J. Nelson†, Steven P. Nolan‡,§,1
*Organic Chemistry, Bergische Universit€at Wuppertal, Wuppertal, Germany

Department of Pure and Applied Chemistry, University of Strathclyde, Glasgow, United Kingdom

Department of Chemistry and Center for Sustainable Chemistry, Ghent University, Ghent, Belgium
§
College of Science, King Saud University, Riyadh, Saudi Arabia
1
Corresponding author: e-mail address: steven.nolan@ugent.be

Contents
1. Introduction 283
2. Group: Fe, Ru, Os 284
2.1 Iron 284
2.2 Ruthenium 285
3. Group 9: Co, Rh, Ir 287
3.1 Cobalt 288
3.2 Rhodium 304
3.3 Iridium 307
4. Group 10: Ni, Pd, Pt 308
4.1 Nickel 308
4.2 Palladium 309
4.3 Platinum 314
5. Group 11: Cu, Ag, Au 315
5.1 Copper 315
5.2 Gold 316
6. Conclusions 319
Acknowledgments 319
References 320

1. INTRODUCTION
Transition metal complexes that bear a formal negative charge are
widespread in organometallic chemistry and catalysis; these are often impli-
cated as intermediates in important catalytic processes or can be catalyst

Advances in Organometallic Chemistry, Volume 69 # 2018 Elsevier Inc. 283


ISSN 0065-3055 All rights reserved.
https://doi.org/10.1016/bs.adomc.2018.02.004
284 Adrián Gómez-Suárez et al.

precursors. Often, their reactivity can be different from that of neutral ana-
logs. This chapter collects some selected examples of late transition metal ate
complexes and their use in catalysis, or in the synthesis of catalysts for impor-
tant processes. Only discrete mono- or bimetallic species are considered.
This chapter is organized by metal centre.

2. GROUP: Fe, Ru, Os


2.1 Iron
Mechanistic studies of iron catalysis are highly challenging, and the nature of
the active catalyst is often unclear.1–3 Bedford has implicated iron “ate”
complexes in cross-coupling reactions, through a series of detailed mecha-
nistic experiments (Scheme 1). Iron(II) chloride was shown to react with
excess MesMgBr to form a well-defined [Fe(Mes)3] complex, even in
the presence of chelating ligands such as tetramethylethylenediamine
(TMEDA).4 This ate species reacts far more rapidly with bromoalkanes than
[Fe(Mes)2(TMEDA)], which was previously postulated to be the key inter-
mediate arising from transmetalation. Similarly, [Fe(Bn)3] was obtained
from analogous reactions between iron(II) chloride (or iron(III) chloride)
and BnMgCl, and underwent a reversible reaction with TMEDA to form
[Fe(Bn)2(TMEDA)]; again, the ate complex was found to be much more
reactive toward organobromide-coupling partners. [Fe(Bn4)] was also
recovered during crystallization studies, represents a rare example of an
iron(III) “ate” complex, and suggests that disproportionation of [Fe
(Bn)3] can occur. An octahedral analog of [FeBn3] in which the aryl rings
bear ortho-diphenylphosphino groups has been reported and structurally
characterized.5
In contrast, it has been shown that the reduction of iron(III) salts with
PhMgBr in toluene leads to a mixture of [Fe(PhMe)2] and [FePh2(PhMe)],
with the latter being catalytically inactive in Kumada coupling reactions.6
The reactivity of iron “ate” complexes is therefore closely linked to their
oxidation state.

Scheme 1 Formation of iron “ate” complexes relevant to cross-coupling catalysis.


Metallate Complexes of the Late Transition Metals 285

Various other examples of iron “ate” complexes have been recovered—


or observed in situ—from cross-coupling reactions. F€ urstner and Neidig
have identified and characterized [FeMe4]2 (in the form of [Li
(OEt2)]2[FeMe4  MeLi]) and [FeMe3] (with a [MgCl(THF)5] counterion),
respectively.7,8 These were also isolated from reactions involving the addi-
tion of an organometal species (MeLi or MeMgCl, respectively) to an iron
halide complex. The latter complex is only stable at low temperatures, and
upon warming, forms [MgCl(THF)5][Fe8Me8] with an EPR spectra consis-
tent with that reported by Kochi9,10 in his studies of iron-catalyzed cross-
coupling reactions.11 This complex is not very reactive with organobromine
species, but becomes reactive when treated with additional MeMgCl. This
species could therefore be relevant to catalytic Kumada-type cross-coupling
reactions with iron.
Hu has reported an iron(I) ferrate complex [nBu4N]2[FePh2(η6-biphe-
nyl)], which was prepared from the reaction of excess phenyllithium with
[nBu4N]2[Fe4S4Cl4].12 Koszinowski has studied the addition of PhMgCl
to [Fe(acac)3] by mass spectrometry and has characterized a number of
iron(II) and iron(III) “ate” complexes, including [FePh3], [FePh4], and
the mixed oxidation state complex [Fe4Ph7].13 [FePh2] was not observed.
Added TMEDA did not coordinate the metal centre, but did reduce the
appearance of polynuclear complexes. NMP, SIPr, dppe, and dppp did
not coordinate the metal centre either, although dppbz led to species such
as [Fe(dppbz)], [FePh2(dppbz)], and [Fe2Ph(dppbz)2]. These feature
iron(1), iron(0), and iron(I) centres, which are rather low oxidation states.
The analysis of some catalytic reactions between PhMgCl and isopropyl
chloride showed a number of “ate” complexes to be present, including
[FePh3(iPr)]. The concentration of [FePh4] was essentially constant dur-
ing the course of the reaction, suggesting that it is not directly involved in
catalysis.
Low oxidation state iron “ate” complexes such as 1 have been shown to
be effective catalysts for alkene hydrogenation (Scheme 2).14 A series of
cobalt complexes were also tested. Reactions proceeded under quite mild
conditions. Poisoning experiments confirmed that the reaction was occur-
ring via homogeneous catalysis.

2.2 Ruthenium
Ruthenium “ate” complexes have been reported for a number of oxidation
and reduction processes. Ley has used complexes of the form [R4N][RuO4]
286 Adrián Gómez-Suárez et al.

Scheme 2 Hydrogenation of alkenes by low oxidation state iron “ate” complexes.

Scheme 3 Hydrogenation of anthracene by a ruthenium “ate” complex.

catalytically, in the presence of N-methylmorpholine N-oxide as the termi-


nal oxidant, to selectively and mildly oxidize primary alcohols to aldehydes
and secondary alcohols to ketones.15 Epoxides, esters, and many other sen-
sitive functional groups do not undergo reaction under these conditions.
Anionic ruthenium hydride complex 216 has been shown to be active for
the hydrogenation of aromatic compounds such as naphthalene (Scheme 3).17
Analogous neutral complexes were not sufficiently reactive to achieve this
transformation. Under a hydrogen atmosphere, 2 forms fac-[RuH3(PPh3)3]
which can then coordinate anthracene (with the loss of one PPh3 ligand).18,19
This then undergoes reaction with hydrogen to release tetrahydronaphthalene
Metallate Complexes of the Late Transition Metals 287

Scheme 4 A binuclear anionic ruthenium catalyst for asymmetric hydrogenation.

Scheme 5 A binuclear ruthenium “ate” catalyst for carbon dioxide hydrocondensation


with methanol.

and generate [RuH5(PPh3)2] 3, which is formally a ruthenium(IV) “ate”


complex, and also capable of mediating the hydrogenation of anthracene.
In contrast, the reduction reactions of ketones are proposed to be catalyzed
instead by neutral [RuH4(PPh3)3], which can be formed in situ from
[RuH3(PPh3)3], albeit with a clear initiation period.20
Binuclear ruthenium complex 4 has been applied in asymmetric hydro-
genation processes using alkene and ketone substrates (Scheme 4).21 It was
isolated from the reaction between [RuCl2(COD)]n and the bisphosphine
ligand, and may represent a species formed in hydrogenation reactions
where this precursor and ligand are mixed in situ.
Binuclear ruthenium “ate” complexes have also found application
in hydrocondensation reactions of carbon dioxide and methanol.22
[RuCl3(CO)3] reacts with a slight excess of KOMe to form 5, which has
a bridging methoxide and two bridging formate ligands (Scheme 5). Catalytic
reactions were carried out under 20 bar CO2 and 40 bar H2.

3. GROUP 9: Co, Rh, Ir


The most used group 9 ate complexes in catalysis are low-valent cobalt
species. However, a few examples of the application of rhodate and iridate
species are also known.
288 Adrián Gómez-Suárez et al.

3.1 Cobalt
Anionic cobalt species have been widely used or proposed as reactive inter-
mediates in catalytic transformations. Therefore, to try to organize the
extensive chemistry spanned from the use of these species, we decided to
divide it into three sections depending on the nature of the active species.

3.1.1 Reactions Catalyzed by Cobalt Carbonyl Species


In 1960, Eisenmann et al. studied the use of [Co2(CO)8] as a catalyst for the
hydroesterification of propylene oxide.23 Using ca. 3 mol% of [Co2(CO)8]
and 24 MPa of CO in MeOH at 130°C for 4 h, the authors were able to
obtain methyl 3-hydroxybutyrate in moderate yield 40%. At the time, no
mechanism was proposed for this transformation, however, later studies
showed that the active species in this process was [Co(CO)4].
Leckta reported the first example of a basic metal complex acting as a
nucleophile to catalyze acylation reactions.24 Employing 10 mol% of Na
[Co(CO)4] he was able to catalyze the acylation of alcohols, in several sol-
vents, using acetic anhydride. Moreover, he observed a change on the reac-
tion mechanism depending on the solvent: if polar solvents were used, the
main pathway involved first nucleophilic attack of the [Co(CO)4] anion
onto the anhydride to form an acylcobalt intermediate; while with nonpolar
solvents, the main pathway involved first activation of the anhydride by the
Na+ cation acting as a Lewis acid. The rationale for this mechanism switch
was that in nonpolar solvents both the alcohol and the acetic acid by-product
could strongly interact with the [Co(CO)4] via hydrogen bonds, thus
diminishing its catalytic activity and favoring the Lewis acid activation of
the anhydride. In addition, he was able to use the same catalyst (5 mol%)
to promote the Staudinger reaction between diphenylketene and an electron
deficient imine, in benzotrifluoride at room temperature, to afford the
desired β-lactam in 85% yield after 5 min.
Jacobsen reported the use of catalytic amounts of [Co2(CO)8] and
2-hydroxypyridine (6) (in a 1:2 ratio) to promote the hydroesterification
of chiral epoxides.25 By employing 5 mol% of [Co2(CO)8] and 10 mol%
of 6 under 4 MPa of CO in a 1:1 mixture of MeOH/THF at 60°C for
9 h, he was able to obtain enantiomerically pure β-hydroxy esters in excel-
lent yields (Eq. 1, Scheme 6). Inspired by this work, Denmark reported a
milder protocol for the synthesis of β-hydroxy esters from epoxides.26 By
using [Co2(CO)8] (5 mol%), in MeOH (1 M) under 0.1 MPa of CO and
at room temperature for 24 h, he was able to isolate a wide range of
Metallate Complexes of the Late Transition Metals 289

Scheme 6 Cocatalyzed ring opening of epoxides.

β-hydroxy esters in modest to high yields (11%–96%) (Eq. 2, Scheme 6).


Watanabe and Tsuji reported on the use of [Co2(CO)8] to catalyze the
ring-opening carbonylation of cyclic ethers.27 Several oxiranes were reacted
with N-silylamines in the presence of 3 mol% of [Co2(CO)8] in benzene
under 0.1 MPa of CO at room temperature for 24–50 h, to afford the desired
amides in good to excellent yields (66%–89%) (Eq. 3, Scheme 6). The authors
proposed that the active catalytic species was [(Me3Si)2R2N][Co(CO)4],
which might be generated from [Me3SiCo(CO)4] and N-silylamines.
Later, Jacobsen reported on the use of [Co2(CO)8] (2.5 mol%) to prepare
enantiopure β-hydroxy morpholine amides from chiral epoxides. This new
protocol worked in EtOAc as solvent and required low CO pressures
(0.1 MPa) to afford the desired products in good to excellent yields (56%–
85%) (Eq. 4, Scheme 6).
Alper studied the use of [Co2(CO)8] to catalyze the ring opening and
carbonylation of aziridines.28 By using 8 mol% of [Co2(CO)8] and
3.3 MPa of CO in dimethoxyethane (DME), at 100°C for 24 h, he was able
290 Adrián Gómez-Suárez et al.

to obtain β-lactams in moderate to excellent yields (42%–96%). In addition,


by changing the reaction medium from DME to THF, bicyclic β-lactams
were also accessible in up to 80% yield. The proposed mechanism for these
reactions involves either nucleophilic ring opening of the aziridine by [Co
(CO)4] or ring opening via single-electron transfer (SET) from [Co
(CO)4] for aziridines bearing electron deficient N-substituents. To support
the role of [Co(CO)4] as the catalytically active species, the reaction was
also carried out with catalytic amounts of Na[Co(CO)4], which delivered
the desired products in excellent yields, although required longer reaction
times. This observation was rationalized based on the lower solvation of
Na+ on the reaction media, thus making the [Co(CO)4] anion less reactive.
Several groups have reported on the use of cobalt carbonyl species in
combination with an activator (generally a Lewis acid) to catalyze the
ring-opening copolymerization of epoxides, aziridine, and azetidines with
CO.29–35 The proposed mechanism for this type of copolymerization reac-
tions requires the nucleophilic ring opening of the heterocycle by an anionic
cobalt carbonyl species, generally in situ generated, followed by CO insertion
into the CdCo bond. The nucleophilic O- or N-atoms of the heterocycle
then attack this acylcobalt fragment, thus promoting the polymerization
reaction.
Well-defined [Lewis acid][Co(CO)4] species have been widely employed
to catalyze the ring opening and carbonylation of epoxides. This type of com-
plex can be readily prepared and isolated from the corresponding halogenated
Lewis acid and Na[Co(CO)4]. In addition, they can be generated in situ from
the corresponding halogenated Lewis acid and [Co2(CO)8]. The general
mechanism of reactions involving [Lewis acid][Co(CO)4] generally requires
the activation of the epoxide by the Lewis acid, followed by ring opening via
nucleophilic attack of the [Co(CO)4] anion. Depending on the starting
materials and the conditions used, β-, γ-, or δ-lactones, as well as succinic
anhydrides, can be obtained from readily available epoxides. In addition, it
is possible to carry out these reactions in an enantioselective manner by
employing chiral ligands on the Lewis acid. All these advances, as well as
the use of other metal catalysts for the ring opening and carbonylation of
epoxides, has been recently reviewed by Coates,36 therefore they will not
be included in this book chapter.
Jia and Lin studied the combination of [Co2(CO)8] with LiCl to catalyze
the reaction between epoxides, benzylidenemethylamine, and CO to afford
1,3-oxazinan-4-ones.37 The reported protocol required 5 mol% of
[Co2(CO)8], 10 mol% of LiCl, a 1:1.5 ratio of imine and epoxide, 5.5 MPa
Metallate Complexes of the Late Transition Metals 291

in 1,4-dioxane at 70°C for 24 h to deliver the desired products in up to 98%


yield (Eq. 1, Scheme 7). The proposed reaction mechanism follows the path-
ways mentioned above for the ring opening and carbonylation of epoxides.
Later, Sun reported an improved catalytic system using [Ph3SiCo(CO)4] that
allowed the expansion of this protocol to the use of alkylidenemethylamines.38
Using 5 mol% of a [Ph3SiCo(CO)4]/MeOH mixture, a 1:1 ratio of epoxide
to imine, 6 MPa of CO in toluene at 50°C for 24–96 h, the desired products
were obtained in up to 94% yield (Eq. 2, Scheme 7).
Coates has also studied the application of well-defined [Lewis acid][Co
(CO)4] complexes to catalyze the Meinwal-rearrangement of epoxides to
ketones.39,40 Initial studies focused on the isomerization of monosubstituted
epoxides catalyzed by an [Al porphyrin][Co(CO)4] (7) species. Using 2 mol
% of 7 in THF at 22°C for 18 h, the desired methyl ketone derivatives were
obtained in excellent yields (76%–96%). Moreover, the reported methodol-
ogy could tolerate a wide range of functional groups, including esters,
amides, olefins, ethers, or hydroxides (Eq. 1, Scheme 8).39 Later, the same
group reported on the expansion of these studies for the regioselective isom-
erization of disubstituted trans-epoxides to methyl ketone derivatives. Using
2 mol% of [(salcy)Al(THF)2][Co(CO)4] (8) (salcy ¼ N,N0 -bis(3,5-di-tert-
butyl-salicylidene)-rac-1,2-cyclohexanediamine) in Et2O at 22°C for
18 h, the targeted products were obtained in excellent yields (78%–98%)
and selectivities (up to 50:1) (Eq. 2, Scheme 8).40 The proposed mechanism
for both transformations involves activation of the epoxide by the Lewis acid
fragment of the catalyst, followed by nucleophilic attack of the cobalt
carbonyl species and subsequent β-hydride elimination to deliver the
corresponding enolate, which will isomerize to deliver the desired ketone.

Scheme 7 Syntheses of 1,3-oxazinan-4-ones reported by Jia, Lin, and Sun.


292 Adrián Gómez-Suárez et al.

Scheme 8 Meinwal-rearrangement of epoxides to ketones reported by Coates.

Gardano reported the use of [(ECH2)Co(CO)4] (9) complexes, where


E ¼ electron withdrawing group, for the carbonylation of aryl bromides
and chlorides.41 E helps to stabilize the complex and allows its handling
under inert conditions at 0°C. Moreover, 9 could be synthesized in situ
by combining an alkyl halide and [Co(CO)4]. Although the mechanism
of the reaction was not fully elucidated, Gardano proposed the intermediacy
of [(ECH2)Co(CO)4(CO2CH3)] (10) during the catalytic cycle, which
involved a SET from 10 to the aryl halide, recombination to form a Co–aryl
bond and subsequent CO insertion followed by methanolysis of the
acylcobalt species delivered the desired ester product.
Abbayes investigated the use of [Bu4N][Co(CO)4] as catalyst for the
phase transfer carbonylation of benzylic bromides.42 This study showed that
the anionic cobalt catalyst remains in the organic phase due to rapid pairing
with the tetrabutyl ammonium cation, while the product is expelled to the
aqueous phase. Moreover, kinetic experiments showed that the rate-
limiting step of the reaction is the insertion of  OH into the acylcobalt
group. [R4N][Co(CO)4] species have also been generated in situ and used
as catalysts. For example, Alper reported the use of 5 mol% of [Co2(CO)8] in
combination with benzyltriethylammonium chloride and NaOH to gener-
ate [R4N][Co(CO)4] in situ, which was able to catalyze the debromination
of α-bromoketones in good yields (20%–69%).43
Anionic cobalt complexes have also been used in combination with
imidazolium- or pyridinium counterions as functionalized ionic liquids.
Dylon reported the synthesis, characterization, and reactivity studies of
Metallate Complexes of the Late Transition Metals 293

[Bmim][Co(CO)4] (11) (Bmim ¼ 1-butyl-3-methylimidazolium).44 This


intense blue-green ionic liquid can be readily prepared by salt metathesis
between K[Co(CO)4] and [Bmim]Cl in propanone. As mentioned before,
in situ generated [R4N][Co(CO)4] has been used in combination with
NaOH to catalyze debromination reactions.43 Since NaOH was soluble
in 11, a solution of both salts was employed to catalyze the dehalogenation
of bromoacetophenone derivatives. Extraction of the product with an
organic solvent also removed bromine from the media, thus facilitating
the recyclability of 11. [Bmim][Co(CO)4] was also employed by Lv to
catalyze the hydroesterification of ethylene oxide to afford methyl
3-hydroxypropionate, an intermediate to 1,3-propanediol, in excellent
yields.45 Using 1 mol% of [Bmim][Co(CO)4], 2 mol% of imidazole (as a
promoter to activate ethylene oxide) in MeOH under 3.7 MPa CO at
75°C for 10 h, methyl 3-hydroxypropionate was obtained in 91% yield.
Moreover, the catalyst could be readily recycled and after three cycles
afforded the product in 80% yield. Later, the same group developed a
sequential methodology where [Bmim][Co(CO)4] was used to catalyze both
the hydroesterification of ethylene oxide and the hydrogenation of the
resulting 3-hydroxypropionate to deliver 1,3-propanediol in high yields.46
For the hydroesterification step 1 mol% of [Bmim][Co(CO)4], 2 mol% of
imidazole in [Bmim]PF6 as solvent under 3.7 MPa of CO at 75°C were used
to afford 3-hydroxypropionate in excellent selectivity (92.6%), while for the
hydrogenation step 5 mol% of [Bmim][Co(CO)4], 10 mol% of imidazole in
[Bmim]PF6 as solvent under 10.5 MPa of H2 at 165°C afforded 1,3-
propanediol in 83.6% selectivity. As in the previous study, the catalyst could
be readily recovered by extraction with deionized water and reused up to
three times without losing catalytic efficiency. Following Lv work, Park
and Yoon explored the use of [Bmim][Co(CO)4] for the synthesis of
3-hydroxy butyric acid from propylene oxide. Using 1 mol% of 11,
2 mol% of imidazole (as promoter), and 6.0 MPa of CO in a 1:1 mixture
of dimethoxyethylene and water at 75°C for 56 h, they were able to convert
propylene oxide to 3-hydroxy butyric acid in moderate selectivity (49%).47
Sun and Xia described the synthesis, characterization, and reactivity
studies of a series of alkylpyridinium tetracarbonyl ionic liquids.48 The com-
pounds were synthesized by reacting K[Co(CO)4] and the corresponding
alkylpyridinium halide salt in a degassed H2O/dichloromethane mixture
(Scheme 9). The compounds were all sensitive to oxygen. However, 12,
which bears the largest alkyl chain, remained unchanged after being exposed
to air for 24 h. A crystal structure of 12 revealed that this increased stability
294 Adrián Gómez-Suárez et al.

Scheme 9 Synthesis of alkylpyridinium cobaltate species.

was due to the existence of interionic hydrogen bonds49 between the oxy-
gen atoms of the [Co(CO)4] and the hydrogen atoms of the alkyl chain of
the pyridinium moiety. These interactions increased the proximity of the
ion pair, thus affording a more stabilized [Co(CO)4] anion. Catalytic stud-
ies showed that complex 12 was effective in the alkoxycarbonylation of pro-
pylene oxide. 4 mol% of 12, 80.0 MPa of CO and propylene oxide were
reacted in MeOH at 60°C for 24 h to afford the desired product in excellent
yield (92%) and selectivity (96%). Moreover, the catalyst could be recycled
up to five times and still retain excellent selectivity (93%), albeit lower yield
was obtained (77%).

3.1.2 Reactions Catalyzed by Cobalt Alkyl Species


Anionic CoII complexes such as those of the form R4CoLi2 (R ¼ alkyl) have
been widely employed in cross-coupling reactions.50 These complexes can
be readily synthesized from alkyl lithium species and CoCl2.51 However,
most of these transformations proceed via the stoichiometric reaction of
an organic molecule with a well-defined R4CoLi2 species and, therefore,
will not be discussed in this chapter. On the catalytic side, Uchiyama studied
the ability of Me3CoLi to catalyze SET processes.52 The proposed mecha-
nism proceeded via SET from the electron rich [Me3CoII] species to a suit-
able organic molecule, followed by regeneration of the active catalyst by
reduction of the formed Me3CoIII with Mg. This catalytic system was highly
successful in a variety of organic transformations, such as reduction of ben-
zophenone, deallylation of allylic ethers, partial reduction of diketones, and
deprotection of sulfonamides. Cyclic voltammetry (CV) studies revealed
that Me3CoLi has an oxidation potential of 2.6 V vs Ag/AgCl, making
it very good reductant.
Oshima reported the use of simple cobalt salts, such as CoCl2, to catalyze
the cross-coupling between silylmethylmagnesium reagents and silyl-
substituted dribromomethane derivatives (Eq. 1, Scheme 10).53 The opti-
mized catalytic system used CoCl2 (10 mol%), and 3 equiv. of the Grignard
Metallate Complexes of the Late Transition Metals 295

Scheme 10 Cocatalyzed synthesis of di- and trisilylethenes.

reagent, in THF to afford 1,2-disilylethenes in high yields and exclusive E


conformation. Regarding the mechanism, Oshima proposed that the need
of 3 equiv. of the Grignard reagent hinted at the involvement of a cobaltate
intermediate as the active species. Due to the high efficiency and selectivity of
the catalytic system, Oshima tried to apply the same methodology for the syn-
thesis of 1,1,2-trisilylethenes. However, the reaction did not afford the desired
products. To tackle this, the authors investigated the stoichiometric reaction
between cobaltate [(R3SiCH2)4Co(MgCl)2] and dibromodisilylmethanes.
Gratifyingly, this reaction afforded the desired 1,1,2-trisilylethenes in reason-
able yields. Moreover, the use of unsymmetrical dibromodisilylmethane
afforded the corresponding 1,1,2-trisilylethenes in high E selectivity (Eq. 2,
Scheme 10). Based on these experiments, Oshima proposed a mechanism
where the key intermediate is cobaltate 13. β-Hydride elimination from
the latter species will produce the desired product. In addition, the high
E/Z selectivity could result from the more stable eclipsed configuration of
13 prior to the β-hydride elimination (Eq. 3, Scheme 10).

3.1.3 Reactions Catalyzed by Cobalt Aryloxide Complexes


Anionic cobalt(II) aryloxide complexes have been synthesized and charac-
terized by a number of research groups.54–56 However, only a couple of
reports discuss their catalytic application. Sun studied the catalytic activity
of anionic-aryloxo complexes of CuII, ZnII, NiII, and CoII for the ring-
opening polymerization of D,L-lactide.57 The aryloxo ligands were prepared
from salicylidene and aspartic acid. All the studies complexes were able to
catalyze the polymerization reaction, with the Ni-based species being the
296 Adrián Gómez-Suárez et al.

most active. Later, Xiaoping studied the use of more readily prepared [Na
(THF)6][Co(OAr)3] (14) (OAr ¼ 2,4,6-tri-tert-butylphenoxo) to catalyze
the same transformation.58 This new complex was isolated as green crystals
after reacting CoCl2 with 3 equiv. of NaOAr in THF. However, catalytic
studies revealed that 14 was not able to control the ring-opening polymer-
ization of D-lactide and mixtures of several oligomers were obtained.
Bolkon investigated the application of well-defined anionic CoIII-
aryloxo complexes for the asymmetric synthesis of cyanohydrins.59 The
catalysts were readily prepared in a one-pot reaction from salicyl aldehyde,
a chiral aminoacid, and K3[Co(CO3)3] in refluxing EtOH. Chromato-
graphic purification allowed the separation of the Λ- and Δ-bis(N-
salicylideneaminoacidato)cobaltate isomers. These new complexes were
tested on the asymmetric addition trimethylsilyl cyanide to benzaldehyde.
Although all reported complexes were able to catalyze this transformation,
only the Δ-isomer of the S-tryptophan derived complex delivered moder-
ate ee (60%), thus suggesting that the level of asymmetric induction depends
both on the ligand substituents and on the stereochemistry of the com-
plex. The mechanistic hypothesis for this reaction suggested that the
cationic counterion was responsible for the carbonyl activation, while
the cobaltate species provided a chiral environment for the asymmetric
induction. Therefore, it was reasoned that the introduction of additives
capable of increasing the Lewis acidic character of the counterion, as well
as the steric hindrance between the catalyst and the substrate, might
improve the levels of enantioselectivity. Gratifyingly, the addition of
10 mol% of triphenylphosphine increased the ee to 77%.

3.1.4 Reactions Catalyzed by Cobalt Complexes Bearing


Phosphine Ligands
Cobalt complexes of the formula [CoCl2(P)] (where P ¼ mono- or bidentate
phosphine ligand) have been extensively studied, in combination with Gri-
gnard reagents, in radical-mediated transformations.60–69 The proposed gen-
eral mechanism for these reactions involves an initial SET from an in situ
generated, anionic, Co0 species, such as [MgBr]2[CoR0 2(P)] (where R0 comes
from the corresponding Grignard reagent), and a suitable organic molecules,
such as alkyl halides. The newly formed open-shell species can then be
involved in subsequent cross-coupling or cycloisomerization reactions.
Although the general consensus is that the initial step of these transformations
requires a SET from an electron rich, low-valent Co species, the exact oxi-
dation state of this catalytically active complex has not been fully elucidated.
Metallate Complexes of the Late Transition Metals 297

Oshima has studied the application of [CoCl2(P)] as precatalyst in a wide


range of transformations.60–68 His original report involved the use of 5 mol%
of [CoCl2(dppe)] (dppe ¼ 1,2-bis(diphenylphosphino)ethane), in combina-
tion with ArMgBr, to catalyze the tandem cyclization/cross-coupling reac-
tion of bromo acetal derivatives, to obtain tetrahydrofuran and cyclopentane
derivatives in up to 84% yield (Eq. 1, Scheme 11).68 [CoCl2(dppe)] was
readily prepared by mixing CoCl2 and dppe in THF under strictly anhy-
drous conditions. Stoichiometric experiments were carried out in order
to clarify the reaction mechanism. The need for at least 4 equiv. of Grignard
reagent, relative to the catalyst, in order to obtain the product, suggested the
catalytically active species was the 17-electron ate complex [MgBr]2
[CoPh2(dppe)]. Later, the same group studied the use of these type of com-
plexes for the intramolecular Heck-type reaction of 6-halo-1-hexene deriv-
atives.60 Using 5 mol% of in situ generated [CoCl2(dppb)] (dppb ¼ 1,2-bis
(diphenylphosphino)butane), Me3SiCH2MgBr, and 6-halo-1-hexene
derivatives in refluxing THF afforded the desired cyclic products in good

Scheme 11 Selected applications of [CoCl2(P)] in catalysis.


298 Adrián Gómez-Suárez et al.

to excellent yields (58%–94%) (Eq. 2, Scheme 11). The same group also
published an intermolecular Heck-type reaction between alkyl halides
and styrene derivatives.61 In this case, the catalytic system consisted of
5 mol% of [CoCl2(dpph)] (dpph ¼ 1,6-bis(diphenylphosphino)hexane),
2.5 equiv. of Me3SiCH2MgCl, and a 1:1.5 ratio of styrene to alkyl bromide,
in refluxing ether. The desired internal olefins were obtained in up to
95% yield (Eq. 3, Scheme 11). The same catalytic system was used to pro-
mote a 3-component coupling between alkyl halides, 1,3-dienes, and
Me3SiCH2MgCl to deliver homoallylsilanes in good to excellent yields
(60%–91%) (Eq. 4, Scheme 11).63 A catalytic system based on 7 mol% of
CoCl2 and 8.5 mol% dpph was also employed for the Mizoroky–Heck-type
reaction of epoxides with styrene derivatives.62 In this case, the proposed
reaction mechanism involves first ring opening of the epoxide by MgBr2
to deliver 2-bromoethoxide derivatives that can engage in SET reactions
with a low-valent, electron rich Co species, thus following the general
mechanism described at the beginning of this section. Oshima also reported
on the cross-coupling of alkyl halides with allylic and benzylic64 Grignard
reagents.67 In this case, the catalytic system consisted of 10 mol% of
[CoCl2(dppp)] (dppp ¼ 1,3-bis(diphenylphosphino)propane), a 1:3 ratio
of alkyl halide to Grignard, in THF at 0°C, 20°C, or 40°C. The targeted
terminal products were obtained in up to 93% yield. Moreover, this protocol
could be used in a tandem cyclization/cross-coupling sequence to access
highly substituted lactones. A similar catalytic systems were employed by
Oshima to catalyze the cross-coupling of primary alkyl halides with ArMgBr
reagents.65 10 mol% of CoCl2/dppp and a 1:3 ratio of alkyl halide to Gri-
gnard were reacted in THF at 15°C for 30 min to afford the desired
alkylated aromatics in moderate yields (24%–67%).
Yoshikai has extensively studied the combination of Co salts, phosphine
ligands, and Grignard reagents for CdH activation reactions.69 The mech-
anism of these transformations remains elusive and the actual nature of the
active catalyst has not been fully elucidated, although anionic Co0 species are
often suggested as the active species. The majority of this work has recently
been reviewed by Yoshikai69 and, therefore, only a couple of examples will
be discussed to illustrate this chemistry. By using a combination of 10 mol%
of CoCl2, 20 mol% of PMePh2, and equimolar amounts of a Grignard
reagent, Yoshikai was able to catalyze the hydroarylation of internal alkynes
in good yields and excellent E/Z selectivities (Eq. 1, Scheme 12).70 The pro-
posed mechanism involves the chelation-assisted oxidative addition of the
aromatic CdH bond into an organocobalt species, followed by insertion
Metallate Complexes of the Late Transition Metals 299

Scheme 12 Selected examples of Co-catalyzed C–H activation.

of the alkyne into the CdCo bond and subsequent reductive elimination.
Due to the need of large excess of Grignard reagent to reduce the initial CoII
species, they proposed that the catalytically active species was an anionic Co0
species. The same group applied a similar catalytic system for the addition of
azoles to alkynes via CdH activation (Eq. 2, Scheme 12).71
Guerinot and Cossy have also recently investigated the use of cobalt cat-
alysts for the cross-coupling of α-bromo amides and Grignard reagents.72
The combination of 10 mol% of CoCl2/Xantphos (Xantphos ¼ 4,5-bis
(diphenylphosphino)-9,9-dimethylxanthene) in THF at 0°C promoted
the α-arylation of amides in 86% yield. Using vinyl Grignard reagents
and lower temperatures (40°C), it was possible to obtain α-alkenylated
amides in high yields. It was proposed that the reaction mechanism involved
first reduction of the CoII species with the Grignard reagent to deliver either
an anionic CoI or Co0 complex that underwent oxidative addition with the
α-bromo amide, followed by reductive elimination to deliver the desired
product.

3.1.5 Reactions Catalyzed by Cobalt Complexes Bearing Diene Ligands


Recently, the use of diene ligands to stabilize low-valent Co species has
attracted increased attention. Kambe has been particularly active in this area,
studying the combination of CoII salts and diene ligands for the cross-
coupling of alkyl Grignard reagents with alkyl halides to construct
300 Adrián Gómez-Suárez et al.

Scheme 13 Cross-coupling of alkyl Grignards and alkyl (pseudo)halides.

quaternary carbon centers (Eq. 1, Scheme 13). Initially, he investigated the


use of 2 mol% of CoCl2, 4 mol% of LiI, and 2 equiv. of 1,3-butadiene to
catalyze the cross-coupling between primary alkyl bromides and primary,
secondary, and tertiary alkyl Grignard reagents.73 The reaction proceeded
smoothly and the desired products were obtained in excellent yields
(62%–95%). Moreover, the reported protocol tolerates a wide range of func-
tional groups, including amides, amines, ethers, esters, olefins, or halides.
Regarding the mechanism, Kambe proposed an initial reduction of the CoII
salt to CoI, which undergoes transmetallation with the Grignard reagent,
followed by β-hydride elimination to deliver a Co-hydride species. Subse-
quent reaction of the latter with 3 equiv. of 1,3-butadiene delivers complex
15, which upon interaction with 1 equiv. of the Grignard reagent forms
anionic Co species 16. Nucleophilic attack of this species onto the alkyl bro-
mide would deliver the desired product and reform intermediate 15 directly,
or via complex 17 in an SN2 mechanism (Eq. 2, Scheme 13). To further
support this hypothesis, radical trap experiments were carried out to exclude
the possibility of a radical mechanism involving SET from the low-valent
Co species to the alkyl bromide. This work was later expanded to the use
Metallate Complexes of the Late Transition Metals 301

of primary pseudo halide derivatives as starting materials.74 Very recently, in


combination with Iwasaki, Kambe has reported the challenging coupling of
primary alkyl fluorides with primary, secondary, and tertiary alkyl Grignard
reagents.75 By using 1,3-pentadiene, instead of 1,3-butadiene, but otherwise
under very similar reaction conditions, it was possible to realize the desired
Csp3–Csp3 cross-coupling in up to 90% yield. The rational for the increased
reactivity in the presence of 1,3-pentadiene was the low reactivity of this
species toward Grignard reagents. 1,3-Dienes are crucial for the reaction
to proceed. Therefore, if the 1,3-diene is consumed by the Grignard, via,
e.g., a carbomagnesiation reactions, the desired cross-coupling will not take
place. Like to the previous protocol, a wide range of functional groups was
tolerated. To further demonstrate the utility of this methodology, the late
stage functionalization of an alkyl fluoride was tested, affording the desired
product in good yield.
Koszinowski has recently carried out mechanistic studies, using ESI mass
spectrometry to detect reaction intermediates that support Kambe’s catalytic
proposal.76
Jacobi von Wangelin has investigated the use of Co(acac)3 for the cross-
coupling of chlorostyrene derivatives and aryl Grignard reagents.77 Using
1 mol% of Co(acac)3 in a 10:1 THF/NMP mixture at 30°C, it was pos-
sible to obtain the desired biaryl species in moderate to excellent yields
(39%–96%). Moreover, the reaction tolerates a wide range of functional
groups and it can be performed with o-, m-, or p-chlorostyrene derivatives.
Exhaustive mechanistic studies suggested that the catalytically active species is
a CoI complex and that the reaction proceeds via the following mechanism:
initial reduction of the CoIII to CoI followed by transmetallation with the aryl
Grignard reagent to deliver an anionic CoI species. Subsequent oxidative
addition of the aryl chloride species will deliver a CoIII intermediate, which
upon reductive elimination will regenerate the active catalysts. During the
mechanistic studies, a well-defined, anionic CoI catalyst was prepared
and its catalytic activity studied (Scheme 14). CoCl2 was reacted with 3
equiv. of K(anthracene) in DME to deliver cobaltate [K(DME)2][Co(anthra-
cene)2] (18). This catalyst was also active in the aforementioned reaction, and
it was reasoned that 18 reduced the chlorostyrene starting material to gain
access to the catalytically active CoI species. Recently, Jacobi von Wangelin
has studied the application of complex 18, and its derivatives 19–23, for the
hydrogenation of olefins, ketones, and imines.14 Catalyst 19 was prepared
from in situ generated [Co(η4-naphthalene)2] and 1,5-cyclooctadiene
(COD), while catalyst 20 was prepared via the reduction of cobaltocene with
302 Adrián Gómez-Suárez et al.

Scheme 14 Cobaltate species reported by Jacobi von Wangelin.

K in the presence of COD. Complex 21 was prepared via ligand exchange of


19 or 20 with styrene. In a similar manner, 22 was prepared by reacting 22
with 1 equiv. of dibenzo[a,e]cyclooctatetraene (DCT). Finally, complex 23
was obtained by reacting 19 with 2 equiv. of DCT (Scheme 14). The cata-
lytic experiments showed that although all precatalyst except 23 were
active for the hydrogenation of olefins, 18 afforded the best performance.
The lower activity of 22 and lack of reactivity observed for 23 were attrib-
uted to the strong interactions between the DCT ligand and the CoI cen-
ter. Exhaustive mechanistic studies suggest that the reaction proceeds via an
initial π-ligand exchange to allow coordination of the olefin, followed by
H2 activation to deliver the saturated products. In the case of ketones and
imines, a SET mechanism, inducing initial radical reduction of the car-
bonyl/imine, might be in operation. This was deduced by the observation
of the corresponding pinacol product when acetophenone was reacted
with 18.

3.1.6 Reactions Catalyzed by Cobalt Complexes Bearing Amine Ligands


Some examples of cobalt-catalyzed reactions in the presence of amines as
ancillary ligands are known.78,79 The catalytic cycle for these transformations
is similar to that described in Section 3.1.4, involving first a reduction of the
CoII precatalyst to an organocobalt(0)ate complex, which can then be
Metallate Complexes of the Late Transition Metals 303

involved in SET reactions. Yorimitsu and Oshima reported the use of 5 mol
% of CoCl2 and 6 mol% of a chelating diamine ligand (24) to catalyze the
cross-coupling between primary and secondary alkyl halides and aryl Gri-
gnard reagents, to afford the corresponding cross-coupled products in excel-
lent yields (Eq. 1, Scheme 15).80 Jacobi von Wangelin studied the use of
5 mol% of CoCl2 in combination with 10 mol% of N,N,N0 ,N0 -tetra-
methyl-1,2-diaminocyclohexane (Me4-DACH) for the cross-coupling of
alkyl halides and aryl halides, in the presence of Mg (to in situ generate
the required Grignard reagent) and catalytic amounts (10 mol%) of LiCl
(Eq. 2, Scheme 15).78 Cahiez reported the use of Co salts and amine ligands
for the cross-coupling of alkyl halides with aryl and alkyl Grignard
reagents.81,82 The Csp3–Csp3 coupling required the use of 5 mol% of
CoCl22LiI in combination with 20 mol% of TMEDA in THF at 10°C

Scheme 15 Selected reactions catalyzed by Co–diamine complexes.


304 Adrián Gómez-Suárez et al.

(Eq. 3, Scheme 15),81 while the Csp3–Csp2 coupling employed 5 mol% of


Co(acac)3 and 5 mol% of TMEDA in THF at 0°C (Eq. 4, Scheme 15).82
Both protocols afforded the desired compounds in high yields. Recently,
Gerinot and Cossy reported the use of 5 mol% of CoCl2 and 6 mol% of
(R,R)-tetramethylcyclohexanediamine (TMCD) for the cross-coupling of
halogenated piperidine derivatives and aryl or vinyl Grignard reagents.79
The protocol afforded the desired products in high yields and displayed a
wide functional group tolerance (Eq. 5, Scheme 15).

3.1.7 Reactions Catalyzed by Cobalt Complexes Bearing N-Heterocyclic


Carbene Ligands
Yorimitsu and Oshima have reported a couple of studies on the use of CoCl2
in combination with N-heterocyclic carbene (NHC) ligands to catalyze the
tandem cyclization/cross-coupling of 6-halo-1-hexene derivatives with
alkynyl and trimethylsilylmethyl Grignard reagents.66,83 The proposed
mechanism for these transformations is similar to that proposed for reactions
catalyzed by Co complexes bearing phosphine ligands.

3.2 Rhodium
Although catalytic application of rhodate complexes is not very common,
some examples of the application of anionic rhodium complexes bearing
carbonyl, quinone of amine ligands are known.

3.2.1 Reactions Catalyzed by Rhodium Carbonyl Complexes


Longoni reported the hydroformylation of formaldehyde to glycolaldehyde
using a mixture of two anionic rhodium catalysts, [PPN][Rh5(CO)15] (25)
and [PPN][Rh(CO)2Cl2] (26) (PPN ¼ bis(triphenylphosphine)ammo-
nium), and triphenylphosphine. The optimal catalytic system required
1 mol% of 25, 5 mol% of 26, and 10 mol% of PPh3, under 12.6 MPa of
CO/H2 (1:1) in acetone at 110°C and afforded the desired product in up
to 95% selectivity.84 Cheng and Crotti independently reported the use of
[PPN][Rh(CO)4] (27) as catalyst for the reductive carbonylation of nitro-
benzene to carbamate derivatives.85–87 Cheng’s protocol employed 2 mol
% of 27 and 2.7 MPa of CO in tBuOH at 140°C for 12 h to deliver the
desired carbamate in 96% yield,85 while Crotti’s methodology employed
0.33 mol% of 27, 1 mol% of 2,20 -bipyridyl (bpy) as additive, 6 MPa of
CO, in a mixture of MeOH/THF (1:2.5) at 200°C for 1.5 h to afford
the desired carbamate in high selectivity.86,87 Faraone reported a bimetallic
Metallate Complexes of the Late Transition Metals 305

catalytic system, such as [Ru(Ph2PPy)3Cl][Rh(CO)2Cl2] (28), for the


hydroformylation of styrene.88 Complex 28 was synthesized by reacting
[Rh(CO)2Cl]2 with [Ru(Ph2PPy)3Cl]Cl in dichloromethane. Catalytic tests
revealed that 28 (1 mol%) was capable of catalyzing the quantitative
hydroformylation of styrene under 6 MPa of CO/H2 at 100°C in 6 h.

3.2.2 Reactions Catalyzed by Ionic Rhodium Complexes Bearing


Diamine Ligands
Alper has extensively studied the catalytic activity of ionic rhodium com-
plexes baring diamine ligands.89–96 In his original report, he described the
synthesis, characterization, and reactivity studies of this type of complexes
for the hydroformylation of olefins.89 Ionic rhodium(I) diamine complexes,
which are composed of an anionic [RhCl2(L)] and a cationic [Rh(N,N)
(L)]+ (L ¼ neutral ligand; N,N ¼ diamine), were readily prepared by reacting
[Rh(COD)Cl]2 with the corresponding diamine in toluene to afford
the desired [Rh(N,N)(COD)][RhCl2(COD)] species. Moreover, if the
reaction was performed under a CO atmosphere, [Rh(N,N)(CO)2]
[RhCl2(CO)2] species were isolated (Eq. 1, Scheme 16). The catalytic stud-
ies revealed that [Rh(TMEDA)(COD)][RhCl2(COD)] (29) was the most
active catalyst for the hydroformylation of olefins. In addition, [Rh
(TMEDA)(CO)2][RhCl2(CO)2] has been used by Alper to catalyze a wide
range of organic transformations, such as the hydroaminomethylation of
2-isopropenylanilines and 2-allylanilines (Eq. 2, Scheme 16),90,92 a three-
component synthesis of tetrahydro-1H-2-benzazepine derivatives (Eq. 3,
Scheme 16),91 the reductive allylation of conjugated aldehydes with allyl
acetates (Eq. 4, Scheme 16),93 or the reductive N-heterocyclization of
2-nitrovinylarenes and N-(2-nitroanylidene)amines (Eq. 5, Scheme 16).94,96
Moreover, he has also studied a tandem hydroformylation/reduction sequence
using in situ generated [Rh(N,N)(CO)2][RhCl2(CO)2], from [RhCl(COD)]2
and tetramethyl-1,4-diaminobutane, for the synthesis of primary alcohols from
styrene derivatives.95

3.2.3 Reactions Catalyzed by Rhodium-Quinonoid Complexes


Sweigart has investigated the use of anionic RhI quinonoid complex 30 for
the arylation of aldehydes and enones in aqueous media (Scheme 17).97–99
This species was prepared via deprotonation of the corresponding [Rh
(COD)(H2Q)]+ (H2Q ¼ hydroquinone) with either KOtBu or LiOH. 30
was active for the reaction between aryl boronic acids and aldehydes or
306 Adrián Gómez-Suárez et al.

Scheme 16 Use of ionic diamine rhodium complexes in catalysis.

Scheme 17 Synthesis of rhodium quinonoid species.

enones, without the need of external additives. The rationale for its unique
reactivity was that the basic quinonoid ligand plays a double role; it activates
the boronic acid prior to transmetallation and brings it closer to the metal
center. Moreover, later studies demonstrated that 30 could be easily recycled
and reused for upto 18 times without dramatic loss of reactivity.99
Metallate Complexes of the Late Transition Metals 307

3.2.4 Reactions Catalyzed by Diarylrhodates


Kambe has recently reported the use of [RhCl(COD)]2 as a precatalyst for
the cross-coupling of ArMgBr and vinyl ethers.100 The catalytic system
required the use of 5 mol% of [RhCl(COD)]2 and a 1:2 ratio of vinyl ether
to Grignard reagent, in THF at room temperature to afford the desired prod-
ucts in good yields (26%–92%). When β-styryl ethers were used as starting
materials, an excellent E/Z selectivity was observed. Mechanistic studies
revealed that the reaction proceeds via the intermediacy of a diarylrhodate
species, which was isolated and tested as catalyst for the model reaction
(Scheme 18).

3.3 Iridium
Similarly to rhodium, there are not many reports of the use of iridate species
in catalysis.

3.3.1 Reactions Catalyzed by Iridium Carbonyl Complexes


Faraone studied the application of a bimetallic Ru/Ir system, of the formula
[Ru(Ph2PPy)3Cl][Ir(CO)2Cl2] (31), for the hydroformylation of styrene.
This species was synthesized by reacting [Ru(Ph2PPy)3Cl]Cl with [Ir
(CO)2(p-toluidine)Cl], in dichloromethane. Unfortunately, complex 31
presented almost no reactivity in this reaction.88

3.3.2 Reactions Catalyzed by Iridium Quinonoid Complexes


Sweigart reported the synthesis and catalytic studies of [Ir(COD)(H2Q)]
[BF4] (32).101 Complex 32, which was readily synthesized from [Ir
(COD)Cl]2, AgBF4, and hydroquinone, could be converted to the anionic
iridium quinonoid complex 33 by reaction with LiOH. Catalytic tests for
the 1,4-addition of boronic acids to cyclic enones using 0.5 mol% of 31,
in combination with 2 mol% of LiOH, to generate 32 in situ, showed that
this species is a poor catalyst for this reaction (Scheme 19).

Scheme 18 Diaryl rhodate as reaction intermediate.


308 Adrián Gómez-Suárez et al.

Scheme 19 Synthesis of precatalyst 31 and conversion to 32.

Scheme 20 pH dependable equilibrium between species 33, 34, and 35.

3.3.3 Reactions Catalyzed by Iridium Complexes Bearing


a Bipyridonate Ligand
Yamaguchi has recently reported the use of anionic complex 33 to catalyze
the production of hydrogen from a MeOH/H2O mixture.102 Complex 33
was synthesized from neutral species 34 and NaOH. Interestingly, it was
observed that complexes 33, 34, and 35 were reversibly interconverted in
water depending on the pH of the medium. At basic pH the predominant
species was 33, at neutral 35, and at acidic 34. Catalytic studies revealed that
anionic complex 33, in combination with catalytic amounts of NaOH was
highly effective to promote the formation of H2 from MeOH/H2O, as well
as from a formaldehyde/H2O mixture (Scheme 20).

4. GROUP 10: Ni, Pd, Pt


4.1 Nickel
A number of nickel “ate” complexes have been reported in catalytic appli-
cations. In particular, they have been implicated in the catalytic dimeriza-
tion/alkylarylation of 1,3-dienes (Scheme 21).103 Dimerization of the
1,3-diene (e.g., butadiene) leads to complex 36, which reacts with
arylmagnesium halides to form nickel(II) “ate” complex 37. This in turn
reacts with alkyl fluorides to form 38. Reductive elimination forms the
Metallate Complexes of the Late Transition Metals 309

Scheme 21 Nickel-catalyzed dimerization/alkylation/arylation of 1,3-dienes.

product. A range of products were prepared in this way. Fluorarenes can be


used in place of the alkyl fluoride, to form 3-fluoraryl-8-aryl-1,6-octadiene
compounds.104 Ortho-substituted arylmagnesium reagents were found to
improve selectivity for these products over the alternative Kumada coupling
mechanism. Mechanistic studies were supported by the isolation and crys-
tallographic characterization of nickel “ate” complexes of the same type as
37, and DFT calculations.105

4.2 Palladium
Palladate complexes have been found to arise during the coordination of tri-
dentate ligands to PdCl2; the resulting species are often of the form [PdCl(L)]
[PdCl3(solv)].106 Similarly, attempts to prepare polymer-supported palladium
complexes have led to the spontaneous formation of [PdCl(L)]2[PdCl4]-type
species in which a palladate is the counterion.107
Anionic Pd0 complexes such as those of the form [Pd(X)(L)2] (X ¼ Cl,
OAc; L ¼ PPh3) have been implicated in a number of widely used cross-
coupling methods,108 and often form as the result of mixing a PdII source
such as PdCl2 or Pd(OAc)2 with a ligand such as PPh3 in order to form a
[Pd(L)2] complex in situ. For example, the formation and reactivity of spe-
cies such as [Pd(OAc)(PPh3)2] has been studied using electrochemical
methods, in which the current is directly proportional to the concentration
of the species under investigation (Scheme 22A).109 Oxidative addition to
this species occurs more rapidly than [Pd(PPh3)4], a widely used source of
Pd0, as a result of the presence of the acid byproduct from PPh3 oxidation
310 Adrián Gómez-Suárez et al.

Scheme 22 Formation and reactivity of [PdX(PPh3)2] complexes.

by acetate ligands. Similarly, the reduction of [PdCl2(PPh3)2] leads to [PdCl


(PPh3)2] (in equilibrium with [PdCl(PPh3)2]2 2 and [PdCl2(PPh3)2]2),
which is rendered more reactive toward oxidative addition if Li+ or Zn2+
are added to the reaction (Scheme 22B).110,111 Even if well-defined palla-
dium ate complexes are not added to the reaction deliberately, they may
well form during PdII reduction and be responsible for much of the catalytic
cycle. These processes have also been studied using density functional
theory.112
Hartwig has presented kinetic evidence that “ligandless” complexes of
the form [Pd(Ar)Br(μ-Br)]2 2 are more reactive toward alkene insertion than
the corresponding [Pd(Ar)Br(PtBu3)] complex.113 This was achieved
through detailed kinetic studies of alkene insertion into well-defined model
complexes. These results suggest that while ligand design is important for the
selectivity of Heck reactions, the “ligandless” anionic intermediates are in
fact more reactive.
Cazin and Nolan have recently prepared and characterized a series
of [NHCH][PdCl2(η3-cinnamyl)] and [NHCH][PdCl2(η3-allyl)] com-
plexes and applied them in catalysis.114 The complexes were prepared
by mixing the corresponding imidazolium salt with 0.5 equiv. of [PdCl
(η3-H2CCHCHR)]2 in refluxing acetone (Scheme 23) and were charac-
terized by methods including NMR spectroscopy and X-ray crystallogra-
phy. The X-ray data show contacts between the imidazolium C2dH and
the chloride ligands on the [PdCl2(η3-H2CCHCHR)] fragment (H … Cl
distances in the range 2.43–2.80 Å). The reaction of these complexes with
base forms known [PdCl(η3-H2CCHCHR)(NHC)] precatalysts, which
can catalyze a range of cross-coupling reactions. The new palladate com-
plexes could also catalyze a range of Suzuki reactions of aryl chlorides, and
Heck reactions of aryl bromides, if first activated by heating in the presence
of base. Such palladate complexes are also proposed to be potential inter-
mediates in the formation of PEPPSI-type precatalysts, although no evi-
dence for this has been put forward yet.115
Metallate Complexes of the Late Transition Metals 311

Scheme 23 Formation and reactivity of [NHCH][PdCl2(η3-H2CCHCHR)] complexes.

Scheme 24 Palladate complexes as intermediates in the formation of palladium–NHC


complexes.

A number of complexes of the general form [PdX3(NHC)] have been


reported in the literature. Hermann isolated complex 39 as an intermediate
in the synthesis of 40 from the corresponding diimidazolium proligand and
[Pd(OAc)2] (Scheme 24).116 Here, the acetate ligand acts as an internal
base to deprotonate the remaining imidazolium moiety, in a manner anal-
ogous to many such methods for the synthesis of NHC complexes without
preparation of the free carbene.117 Similar reactivity has been reported for
analogous mesoanionic carbenes.118 An analogous triiidopalladate is also
known, and one or two of the iodide ligands can be displaced by phosphine
nucleophiles.119
The reaction of a [PdBr(μ-Br)(NHC)] complex with NHCHBr has
been found to yield imidazolium palladate complexes. For example, when
complex 41 is heated with an imidazolium salt in CHCl3, [NHCH]
[PdBr3(NHC)] was formed (Scheme 25).120 NMR experiments suggest that
this process is reversible, and so NHCHX compounds might be able to sta-
bilize coordinatively unsaturated PdII compounds during catalytic reactions.
The formation of complex 42 was reported from the reaction of
[PdCl2(COD)] with an NHC–silver complex (Scheme 26); instead of for-
ming 43, the reaction yielded a cationic bis(NHC) complex with an anionic
palladium(II) counterion. Species of this type were found to be active in
Heck reactions of bromoarenes at catalyst loadings as low as 0.005 mol%.
312 Adrián Gómez-Suárez et al.

Scheme 25 Reaction of a [PdBr(μ-Br)(NHC)] complex with NHCHBr.

Scheme 26 Synthesis of complex 43.

Scheme 27 Navarro’s anionic Pd–NHC complex and its application in Heck reactions.

Navarro has prepared [nBu4N][PdCl3(SIPr)] from the reaction of


[PdCl2(NEt3)(SIPr)] with [nBu4N]Cl (Scheme 27A). This species was applied
in the Heck reactions of a representative palette of aryl and heteroaryl chlo-
rides and bromides with styrene. Reactions were conducted for up to 6 h
using 0.5 mol% of precatalyst.
Metallate Complexes of the Late Transition Metals 313

Strassner has tested Na[PdCl3(DMSO)] and [nBu4N][PdCl3(DMSO)]


in Suzuki–Miyaura cross-coupling reactions.121 These anionic complexes
were very poor catalysts and led to significant decomposition to palladium
black upon heating. Poisoning experiments with substoichiometric
amounts of CS2 led to no reaction at all, suggesting that these precursors
simply formed a heterogeneous catalyst in situ. However, these complexes
were active for Heck reactions with relatively low catalyst loadings
(0.05 mol%; see Scheme 28).122
A number of [NHCH]2[PdCl4], [NHCH]2[Pd2Cl6], [py.H]2[PdCl4],
and [py.H]2[Pd2Cl6] complexes have been tested in oxidative Heck coupling
reactions.123 Among these, 44 was found to be the most effective in a model
reaction when used under an atmosphere of oxygen (Scheme 29). The use of
an air atmosphere or a CuII oxidant gave lower yields.
Anionic palladium intermediates are likely to be involved in palladium-
catalyzed reactions in ionic liquids based on imidazolium scaffolds.

Scheme 28 Selected examples of Heck reactions catalyzed by anionic palladium


complexes.

Scheme 29 Oxidative Heck coupling catalyzed by complex 44.


314 Adrián Gómez-Suárez et al.

[PdCl2(COD)] reacts with these species to form [PdX4] complexes with an


imidazolium counterion, which are competent catalysts for methoxy-
carbonylation reactions.124 Complexes of this type can be converted to the
corresponding [PdX2(NHC)2] species by reaction with silver(I) oxide.125
Levason has prepared a series of [R4N][PdX4(L)] complexes in which the
anion is an octahedral PdIV species (R ¼ Me, Et, nPr, nBu; X ¼ Cl or Br;
L ¼ pyridine, PnPr3, PEt2Ph, SPPh3, AsEt3, SMe2, SeMe2).126 These were
prepared from the reaction of the corresponding PdII complex with the
appropriate halogen. Gentle heating of the complexes provoked the reduc-
tive elimination of X2.

4.3 Platinum
Anionic PtII and PtIV complexes are relatively common starting materials
when considering the organometallic chemistry, with [PtCl4] and [PtCl6]
both readily available. The coordination chemistry and spectroscopy of these
systems has been considered. Complexes of the form [PtX5(PMe3)] and
[PtX3(PMe3)] (X ¼ Cl, Br, I) are known and have been characterized by
methods including 31P NMR spectroscopy, in which the platinum(II) con-
geners were found to have δP that differed by some 30 ppm compared to the
palladium(II) analogs.127
Anionic platinum(IV) complexes are effective hydrosilylation catalysts.
Species such as [nBu4N]2[PtCl6] were found to be more efficient in the
hydrosilylation of alkynes than K2[PtCl6], even if additives were used with
the latter (such as [nBu4N]HSO4 or 18-crown-6).128
More recently, Milstein has reported an anionic platinum(0) pincer com-
plex that reacts rapidly with hexafluorobenzene via CdF bond activa-
tion.129 Complex 45 was prepared by the reduction of the platinum(II)
complex 46 by sodium and reacted with hexafluorobenzene even at
35°C (Scheme 30). This mode of reactivity could potentially form part
of a future catalytic method for CdF activation and arene functionalization
(Scheme 30).

Scheme 30 Synthesis and reactivity of an anionic platinum(0) complex.


Metallate Complexes of the Late Transition Metals 315

5. GROUP 11: Cu, Ag, Au


5.1 Copper
Lithium cuprates and related species have long been known to play a role
in various conjugate addition reactions. The area is well reviewed, so is
not discussed here.130,131 Cuprates are likely intermediates in other catalytic
reactions. For example, Yoon has reported the acceleration of the copper-
catalyzed aminohydroxylation of alkenes by ammonium chloride additives
and has suggested that cuprate formation (i.e. [R4N][CuX3]) in situ may be
responsible.132 Kambe has proposed that cuprate intermediates mediate the
allylation of fluorarenes, catalyzed by copper (Scheme 31); 133 no isolated
cuprates were presented, but the need for activation of the catalyst by
EtMgBr makes such a mechanism plausible. An analogous reaction with
alkyl fluorides was also reported.134
The role of copper “ate” complexes in amination catalysis has been
explored in some detail by Hartwig. In amination reactions catalyzed by
neutral copper(I) complexes (with, e.g., phenanthroline ligands), complexes
of the form [L2Cu][Cu(NR2)2] do not react sufficiently quickly with
haloarenes to be the active species in arylamination reactions.135 Instead,
neutral [Cu(NR2)(L)] complexes are the active species. In contrast,
when anionic ligands are used in biaryl ether synthesis, the catalytic cycle
involves rate-determining oxidative addition to a [Cu(LX)(OR)] complex

Scheme 31 Proposed mechanism for the copper-catalyzed allylation of fluoroarenes.


316 Adrián Gómez-Suárez et al.

to form [Cu(LX)(Ar)X(OR)] (LX ¼ bidentate, anionic ligand). Reduc-


tive elimination proceeds via a neutral complex formed from the dissoci-
ation of X.136.
Copper ate complexes have been implicated as intermediates in the prep-
aration of NHC–CuI complexes from imidazolium salts and copper(I)
halides. Cazin and Nolan published a preparative method in which
copper(I) chloride, bromide, or iodide was heated in acetone with an
imidazolium salt and a relatively weak base (potassium carbonate).137 This
methodology allowed the isolation of a number of [CuX(NHC)] (X ¼ Cl,
Br, I) complexes in moderate to excellent yields (49%–94%). Further
investigation showed that, upon mixing, the imidazolium salt and the
copper(I) salt formed a copper “ate” complex (Scheme 32). The addition
of base leads to the deprotonation of the imidazolium cation and formation
of the [CuX(NHC)] complex. Four examples of the intermediate
imidazolium cuprate species were characterized by single-crystal X-ray
diffraction analysis.

5.2 Gold
Recently, Nolan reported a straightforward methodology for the synthesis
of [Au(X)(NHC)] (X ¼ Cl, Br, I) from [AuCl(Me2S)] and NHCHX salts
using K2CO3 to promote the reaction.138 The methodology was very effi-
cient and could be applied for the synthesis of a wide range of [Au(X)
(NHC)], including complexes bearing (un)saturated, bulky, or alkyl
substituted NHC ligands. The key intermediate of the reaction was identi-
fied and isolated as [NHCH][AuCl2] (Scheme 33).

Scheme 32 Synthesis of [CuX(NHC)] complexes via a copper “ate” complex.

Scheme 33 Synthesis of [Au(X)(NHC)] complexes via aurate species.


Metallate Complexes of the Late Transition Metals 317

Catalytically active aurate species are not very common. Most gold-
catalyzed processes involving aurate species employ the use of NaAuCl4,
or its derivatives, as a precatalyst for the generation of the catalytically active
AuI or AuIII species. The following are some selected examples on the use of
such these aurate salts in catalytic reactions.
Utimoto reported the use of catalytic amounts of NaAuCl4 (5 mol%) in
refluxing MeCN to catalyze the intramolecular cyclization of 5-alkynylamines
to synthesize tetrahydropyridine derivatives in good to excellent yields (64%–
80%) (Eq. 1, Scheme 34).139 The same group also reported the use of 2 mol% of
NaAuCl4 in refluxing MeOH to catalyze the conversion of unactivated
alkynes into ketones or acetals. If the reaction was carried out in wet MeOH,
alkynes were obtained; while if anhydrous MeOH was used, acetals were the
main product of the reaction (Eq. 2, Scheme 34).140
Marinelli has reported the use of NaAuCl4 to catalyze the addition of
nitrogen nucleophiles into 1,3-dicarbonyl derivatives. The use of 2.5 mol
% of catalyst in EtOH at room temperature afforded the desired condensation
product in good to excellent yields (62%–98%) (Eq. 1, Scheme 35).141 In col-
laboration with Arcadi, Marinelli also reported the gold-catalyzed reaction
between 2-alkynylanilines and α,β-unsaturated enone to obtain indole deriv-
atives. The reaction proceeded using 5 mol% of NaAuCl4H2O in EtOH to
afford the desired indoles in moderate to excellent yields (28%–92%) (Eq. 2,
Scheme 35).142 The proposed mechanism involved first cycloisomerization
of the 2-alkynylphenylamine to produce the indole core, followed by 1,4-
addition of the nucleophilic indole C3 into the activated enone.
Fañanás and Rodrı́guez studied the gold-catalyzed reaction between
β-ketoesters and propargylamines to synthesize 2,5-dihydropyridines.143
β-ketoesters and propargylamines reacted in the presence of 5 mol% of

Scheme 34 Cycloisomerization and hydroxylation reactions reported by Utimoto.


318 Adrián Gómez-Suárez et al.

Scheme 35 Condensation and cycloisomerization reactions reported by Marinelli.

Scheme 36 Synthesis of 2,5-dihydropyridines reported by Fañanás and Rodríguez.

NaAuCl4H2O and anhydrous MeOH at 40°C to afford the desired prod-


ucts in moderate to excellent yields (Scheme 36). The proposed reaction
mechanism involves first condensation of the amine with the ketone,
followed by a 6-endo-dig cyclization from the enaminic carbon to the acti-
vated alkyne, to deliver the desired product.
Arcadi and Michelet reported a one-pot gold-catalyzed methodology for
the synthesis of mono- or difluorinated indole derivatives from
2-alkynylanilines.144 The protocol involved first the gold-catalyzed (5 mol
% of NaAuCl4H2O) cycloisomerization of the 2-alkynylaniline to generate
and indole derivative, followed by the addition of 1 or 3 equiv. of Selectfluor©
to afford the desired mono- or difluorinated products (Eq. 1, Scheme 37). The
same group also reported a gold-catalyzed synthesis of 2-arylbenzoxazinones
via the sequential oxidative cyclization of 2-alkynylaniline derivatives in
the presence of oxone as the oxidant (Eq. 2, Scheme 37).145 Interestingly,
if AgNO3 was used as the catalyst for this transformation, instead of
NaAuCl4H2O, the reaction afforded a different product.
Very recently, Cao and You have reported the chemoselective
α-methylenation of aromatic ketones catalyzed by NaAuCl4H2O.146 The
optimized reaction conditions required the use of 5 mol% of NaAuCl4H2O
and 1.1 equiv. of Selectfluor©, in DMSO at 100°C to deliver the desired
products in excellent yields (Scheme 38).
Metallate Complexes of the Late Transition Metals 319

Scheme 37 Sequential cycloisomerization/fluorination and cycloisomerization/oxida-


tion reactions.

Scheme 38 Gold-catalyzed α-methylenation.

6. CONCLUSIONS
The applications of anionic late transition metal complexes in catalysis,
catalysis synthesis, and fundamental processes relevant to catalysis have been
discussed. Much of this work is underpinned by the isolation and character-
ization of relevant anionic complexes, but in other areas the involvement of
such species is not certain. The reactivity of anionic species can often be
quite different from that of neutral analogs, and so this is a potentially very
fruitful area for future study.

ACKNOWLEDGMENTS
A.G.S. thanks the Fonds der Chemischen Industrie (FCI) for a Liebig Fellowship. D.J.N.
thanks the University of Strathclyde for a Chancellor’s Fellowship.
320 Adrián Gómez-Suárez et al.

REFERENCES
1. Bedford RB. How low does iron go? Chasing the active species in Fe-catalyzed cross-
coupling reactions. Acc Chem Res. 2015;48(5):1485–1493.
2. Cassani C, Bergonzini G, Wallentin C-J. Active species and mechanistic pathways in
iron-catalyzed C–C bond-forming cross-coupling reactions. ACS Catal. 2016;6(3):
1640–1648.
3. Mako TL, Byers JA. Recent advances in iron-catalysed cross coupling reactions and
their mechanistic underpinning. Inorg Chem Front. 2016;3(6):766–790.
4. Bedford RB, Brenner PB, Carter E, et al. TMEDA in iron-catalyzed Kumada coupling:
amine adduct versus Homoleptic “ate” complex formation. Angew Chem Int Ed.
2014;53(7):1804–1808.
5. Jacobs BP, Wolczanski PT, MacMillan SN. High- and low-spin chelate complexes of
iron featuring κ-C,X-CH2C6H4X (X ¼ NMe2, PMe2, PPh2) and κ-C,P-CH2PMe2
ligands. J Organomet Chem. 2017;847:132–139.
6. Clemancey M, Cantat T, Blondin G, Latour J-M, Dorlet P, Lefèvre G. Structural
insights into the nature of Fe0 and FeI low-valent species obtained upon the reduction
of iron salts by aryl Grignard reagents. Inorg Chem. 2017;56(7):3834–3848.
7. Al-Afyouni MH, Fillman KL, Brennessel WW, Neidig ML. Isolation and characteri-
zation of a Tetramethyliron(III) ferrate: an intermediate in the reduction pathway of
ferric salts with MeMgBr. J Am Chem Soc. 2014;136(44):15457–15460.
8. F€urstner A, Krause H, Lehmann CW. Unusual structure and reactivity of a Homoleptic
“super-ate” complex of iron: implications for Grignard additions, cross-coupling reac-
tions, and the Kharasch deconjugation. Angew Chem Int Ed. 2006;45(3):440–444.
9. Smith RS, Kochi JK. Mechanistic studies of iron catalysis in the cross coupling of
alkenyl halides and Grignard reagents. J Org Chem. 1976;41(3):502–509.
10. Kwan CL, Kochi JK. Electron spin resonance studies of the reduction of transition
metal complexes with Grignard reagents. 1. Dianion radicals of β-diketonates. J Am
Chem Soc. 1976;98(16):4903–4912.
11. Muñoz III SB, Daifuku SL, Brennessel WW, Neidig ML. Isolation, characterization,
and reactivity of Fe8Me 12: Kochi’s S ¼ 1/2 species in iron-catalyzed cross-couplings
with MeMgBr and ferric salts. J Am Chem Soc. 2016;138(24):7492–7495.
12. Zhurkin FE, Wodrich MD, Hu X. A monometallic iron(I) organoferrate.
Organometallics. 2017;36(3):499–501.
13. Parchomyk T, Koszinowski K. Ate complexes in iron-catalyzed cross-coupling reac-
tions. Chem A Eur J. 2016;22(44):15609–15613.
14. B€uschelberger P, G€artner D, Reyes-Rodriguez E, et al. Alkene metalates as hydroge-
nation catalysts. Chem A Eur J. 2017;23(13):3139–3151.
15. Griffith WP, Ley SV, Whitcombe GP, White AD. Preparation and use of tetra-n-
butylammonium per-ruthenate (TBAP reagent) and tetra-n-propylammonium per-
ruthenate (TPAP reagent) as new catalytic oxidants for alcohols. J Chem Soc Chem
Commun. 1987;(21):1625–1627.
16. Pez GP, Grey RA, Corsi J. Anionic metal hydride catalysts. 1. Synthesis of potassium
hydrido(phosphine)ruthenate complexes. J Am Chem Soc. 1981;103(25):7528–7535.
17. Grey RA, Pez GP, Wallo A. Selective homogeneous catalytic hydrogenation of poly-
nuclear aromatics. J Am Chem Soc. 1980;102(18):5948–5949.
18. Wilczynski R, Fordyce WA, Halpern J. Coordination chemistry and catalytic proper-
ties of hydrido(phosphine)ruthenate complexes. J Am Chem Soc. 1983;105(7):
2066–2068.
19. Fordyce WA, Wilczynski R, Halpern J. Hydrido(phosphine)ruthenate complexes and
their role in the catalytic hydrogenation of arenes. J Organomet Chem. 1985;296(1):
115–125.
Metallate Complexes of the Late Transition Metals 321

20. Linn DE, Halpern J. Roles of neutral and anionic ruthenium polyhydrides in the cat-
alytic hydrogenation of ketones and arenes. J Am Chem Soc. 1987;109(10):2969–2974.
21. Ohta T, Tonomura Y, Nozaki K, Takaya H, Mashima K. An anionic Dinuclear
BINAP ruthenium(II) complex: crystal structure of [NH2Et2][{RuCl((R)-
p-MeO-BINAP)}2(μ-cl)3] and its use in asymmetric hydrogenation. Organometallics.
1996;15(6):1521–1523.
22. S€uss-Fink G, Soulie J-M, Rheinwald G, Stoeckli-Evans H, Sasaki Y. Hydrocondensation
of carbon dioxide with methanol catalyzed by anionic ruthenium complexes: isolation,
structural characterization, and catalytic implications of the Dinuclear anion
[Ru2(CO)4(μ2-η2-CO2CH3)2(μ2-OCH3)Cl2. Organometallics. 1996;15(15):3416–3422.
23. Eisenmann J, Yamartino R, Howard JJ. Notes- preparation of methyl
β-hydroxybutyrate from propylene oxide, carbon monoxide, methanol, and dicobalt
octacarbonyl. J Org Chem. 1961;26(6):2102–2104.
24. Wack H, Drury WJ, Taggi AE, Ferraris D, Lectka T. Nucleophilic metal complexes as
acylation catalysts: solvent-dependent “switch” mechanisms leading to the first cata-
lyzed Staudinger reaction. Org Lett. 1999;1(12):1985–1988.
25. Hinterding K, Jacobsen EN. Regioselective carbomethoxylation of chiral epoxides: a
new route to enantiomerically pure β-hydroxy esters. J Org Chem. 1999;64(7):
2164–2165.
26. Denmark SE, Ahmad M. Carbonylative ring opening of terminal epoxides at atmo-
spheric pressure. J Org Chem. 2007;72(25):9630–9634.
27. Yoshihisa W, Kazuhiro N, Kunsan Z, Fumio O, Teruyuki K, Yasushi T. Co2(CO)8-
catalyzed ring-opening carbonylation of cyclic ethers using N-silylamines. Bull Chem
Soc Jpn. 1994;67(3):879–882.
28. Piotti ME, Alper H. Inversion of stereochemistry in the Co2(CO)8-catalyzed carbon-
ylation of aziridines to β-lactams. The first synthesis of highly strained trans-bicyclic
β-lactams. J Am Chem Soc. 1996;118(1):111–116.
29. Jia L, Sun H, Shay JT, Allgeier AM, Hanton SD. Living alternating copolymerization of
N-alkylaziridines and carbon monoxide as a route for synthesis of poly-β-peptoids.
J Am Chem Soc. 2002;124(25):7282–7283.
30. Takeuchi D, Sakaguchi Y, Osakada K. Alternating copolymerization of propylene
oxide with carbon monoxide catalyzed by Co complex and Co/Ru complexes.
J Polym Sci A Polym Chem. 2002;40(24):4530–4537.
31. Darensbourg DJ, Phelps AL, Gall NL, Jia L. Mechanistic studies of the copolymeriza-
tion reaction of Aziridines and carbon monoxide to produce poly-β-peptoids. J Am
Chem Soc. 2004;126(42):13808–13815.
32. Lee JT, Alper H. Alternating copolymerization of propylene oxide and carbon mon-
oxide to form aliphatic polyesters. Macromolecules. 2004;37(7):2417–2421.
33. Liu G, Jia L. Cobalt-catalyzed Carbonylative copolymerization of N-alkylazetidines
and tetrahydrofuran. Angew Chem Int Ed. 2006;45(1):129–131.
34. Chai J, Liu G, Chaicharoen K, Wesdemiotis C, Jia L. Cobalt-catalyzed Carbonylative
polymerization of Azetidines. Macromolecules. 2008;41(23):8980–8985.
35. Lin S, Yu X, Tu Y, Xu H, Cheng SZD, Jia L. Poly([small beta]-alanoid-block-[small
beta]-alanine)s: synthesis via cobalt-catalyzed carbonylative polymerization and self-
assembly. Chem Commun. 2010;46(24):4273–4275.
36. Kramer JW, Rowley JM, Coates GW. Ring-Expanding Carbonylation of Epoxides.
In: Organic Reactions. John Wiley & Sons, Inc.; 2004.
37. Zhang Y, Ji J, Zhang X, Lin S, Pan Q, Jia L. Cobalt-catalyzed cyclization of carbon
monoxide, imine, and epoxide. Org Lett. 2014;16(8):2130–2133.
38. Liu L, Sun H. [HCo(CO)4]-catalyzed three-component cycloaddition of epoxides,
imines, and carbon monoxide: facile construction of 1,3-oxazinan-4-ones. Angew
Chem Int Ed. 2014;53(37):9865–9869.
322 Adrián Gómez-Suárez et al.

39. Lamb JR, Jung Y, Coates GW. Meinwald-type rearrangement of monosubstituted


epoxides to methyl ketones using an [Al porphyrin]+[Co(CO)4] catalyst. Org Chem
Front. 2015;2(4):346–349.
40. Lamb JR, Mulzer M, LaPointe AM, Coates GW. Regioselective isomerization of 2,3-
disubstituted epoxides to ketones: an alternative to the Wacker oxidation of internal
alkenes. J Am Chem Soc. 2015;137(47):15049–15054.
41. Foà M, Francalanci F, Bencini E, Gardano A. Cobalt-catalysed carbonylation of aryl
halides. J Organomet Chem. 1985;285(1):293–303.
42. Des Abbayes H, Buloup A, Tanguy G. Kinetics and mechanism of the phase-transfer-
catalyzed carbonylation of benzyl bromide by cobalt tetracarbonyl anion.
Organometallics. 1983;2(12):1730–1736.
43. Alper H, Logbo KD, des Abbayes H. Cobalt carbonyl catalyzed dehalogenation and
coupling reactions by phase transfer catalysis. Tetrahedron Lett. 1977;18(33):2861–2864.
44. Brown RJC, Dyson PJ, Ellis DJ, Welton T. 1-Butyl-3-methylimidazolium cobalt
tetracarbonyl [bmim][Co(CO)]: a catalytically active organometallic ionic liquid. Chem
Commun. 2001;(18):1862–1863.
45. Lv Z, Wang H, Li J, Guo Z. Hydroesterification of ethylene oxide catalyzed by
1-butyl-3-methylimidazolium cobalt tetracarbonyl ionic liquid. Res Chem Intermed.
2010;36(9):1027–1035.
46. Guo Z, Wang H, Lv Z, Wang Z, Nie T, Zhang W. Catalytic performance of [bmim]
[Co(CO)4] functional ionic liquids for preparation of 1,3-propanediol by coupling of
hydroesterification-hydrogenation from ethylene oxide. J Organomet Chem.
2011;696(23):3668–3672.
47. Rajendiran S, Park G, Yoon S. Direct conversion of propylene oxide to 3-hydroxy
butyric acid using a cobalt carbonyl ionic liquid catalyst. Catalysts. 2017;7(8):228.
48. Deng F-G, Hu B, Sun W, Chen J, Xia C-G. Novel pyridinium based cobalt carbonyl
ionic liquids: synthesis, full characterization, crystal structure and application in catal-
ysis. Dalton Trans. 2007;(38):4262–4267.
49. K€ olle P, Dronskowski R. Hydrogen bonding in the crystal structures of the ionic liquid
compounds butyldimethylimidazolium hydrogen sulfate, chloride, and chloroferrate
(II,III). Inorg Chem. 2004;43(9):2803–2809.
50. Kauffmann T. Nonstabilized alkyl complexes and alkyl-cyano-ate complexes of iron(II)
and cobalt(II) as new reagents in organic synthesis. Angew Chem Int Ed Engl.
1996;35(4):386–403.
51. Kauffmann T, Hopp G, Laarmann B, Stegemann D, Wingberm€ uhle D. Selektive
kohlenstoff-kohlenstoff-verkn€ upfungen mit nucleophilen thermolabilen methylcobalt-
reagenzien. Tetrahedron Lett. 1990;31(4):511–514.
52. Uchiyama M, Matsumoto Y, Nakamura S, et al. Development of a catalytic electron
transfer system mediated by transition metal ate complexes: applicability and tunability
of electron-releasing potential for organic transformations. J Am Chem Soc.
2004;126(28):8755–8759.
53. Ohmiya H, Yorimitsu H, Oshima K. Regio- and stereoselective approach to 1,2-di- and
1,1,2-trisilylethenes by cobalt-mediated reaction of silyl-substituted dibromomethanes
with silylmethylmagnesium reagents. Angew Chem Int Ed. 2005;44(22):3488–3490.
54. Yuan C-X, Zhou K-D, Xu X-P, Zhang Y, Ji S-J. Syntheses and structural character-
ization of σ-/π-aryloxo cobalt(II) complexes and radical aryloxo sodium complex.
Polyhedron. 2015;87:245–250.
55. Buzzeo MC, Iqbal AH, Long CM, et al. Homoleptic cobalt and copper phenolate A2
[M(OAr)4] compounds: the effect of phenoxide fluorination. Inorg Chem. 2004;43(24):
7709–7725.
56. Boyle TJ, Rodriguez MA, Ingersoll D, et al. A novel family of structurally characterized
lithium cobalt double aryloxides and the nanoparticles and thin films generated there-
from. Chem Mater. 2003;15(20):3903–3912.
Metallate Complexes of the Late Transition Metals 323

57. Sun J, Shi W, Chen D, Liang C. The ring-opening polymerization of D,L-lactide cat-
alyzed by new complexes of Cu, Zn, Co, and Ni Schiff base derived from salicylidene
and L-aspartic acid. J Appl Polym Sci. 2002;86(13):3312–3315.
58. Yuan C, Xu X, Zhang Y, Ji S. Ring-opening polymerization of L-Lactide by an anionic
cobalt(II) aryloxide. Chin J Chem. 2012;30(7):1474–1478.
59. Belokon’ YN, Maleev VI, Mal’fanov IL, et al. Anionic chiral cobalt(III) complexes as
catalysts of asymmetric synthesis of cyanohydrins. Russ Chem Bull. 2006;55(5):821–827.
60. Fujioka T, Nakamura T, Yorimitsu H, Oshima K. Cobalt-catalyzed intramolecular
Heck-type reaction of 6-halo-1-hexene derivatives. Org Lett. 2002;4(13):2257–2259.
61. Ikeda Y, Nakamura T, Yorimitsu H, Oshima K. Cobalt-catalyzed Heck-type reaction
of alkyl halides with styrenes. J Am Chem Soc. 2002;124(23):6514–6515.
62. Ikeda Y, Yorimitsu H, Shinokubo H, Oshima K. Cobalt-mediated Mizoroki–Heck-
type reaction of epoxide with styrene. Adv Synth Catal. 2004;346(13–15):1631–1634.
63. Mizutani K, Shinokubo H, Oshima K. Cobalt-catalyzed three-component coupling
reaction of alkyl halides, 1,3-dienes, and trimethylsilylmethylmagnesium chloride.
Org Lett. 2003;5(21):3959–3961.
64. Ohmiya H, Tsuji T, Yorimitsu H, Oshima K. Cobalt-catalyzed cross-coupling reac-
tions of alkyl halides with allylic and benzylic Grignard reagents and their application
to tandem radical cyclization/cross-coupling reactions. Chem A Eur J.
2004;10(22):5640–5648.
65. Ohmiya H, Wakabayashi K, Yorimitsu H, Oshima K. Cobalt-catalyzed cross-coupling
reactions of alkyl halides with aryl Grignard reagents and their application to sequential
radical cyclization/cross-coupling reactions. Tetrahedron. 2006;62(10):2207–2213.
66. Someya H, Ohmiya H, Yorimitsu H, Oshima K. Cobalt-catalyzed sequential
cyclization/cross-coupling reactions of 6-halo-1-hexene derivatives with Grignard
reagents and their application to the synthesis of 1,3-diols. Tetrahedron.
2007;63(35):8609–8618.
67. Tsuji T, Yorimitsu H, Oshima K. Cobalt-catalyzed coupling reaction of alkyl halides
with allylic Grignard reagents. Angew Chem Int Ed. 2002;41(21):4137–4139.
68. Wakabayashi K, Yorimitsu H, Oshima K. Cobalt-catalyzed tandem radical cyclization
and cross-coupling reaction: its application to benzyl-substituted heterocycles. J Am
Chem Soc. 2001;123(22):5374–5375.
69. Gao K, Yoshikai N. Low-valent cobalt catalysis: new opportunities for C–H func-
tionalization. Acc Chem Res. 2014;47(4):1208–1219.
70. Gao K, Lee P-S, Fujita T, Yoshikai N. Cobalt-catalyzed hydroarylation of alkynes
through chelation-assisted C H bond activation. J Am Chem Soc. 2010;132(35):
12249–12251.
71. Ding Z, Yoshikai N. Cobalt-catalyzed addition of azoles to alkynes. Org Lett.
2010;12(18):4180–4183.
72. Barde E, Guerinot A, Cossy J. Cobalt-catalyzed cross-coupling of α-Bromo amides
with Grignard reagents. Org Lett. 2017;19(22):6068–6071.
73. Iwasaki T, Takagawa H, Singh SP, Kuniyasu H, Kambe N. Co-catalyzed cross-
coupling of alkyl halides with tertiary alkyl Grignard reagents using a 1,3-butadiene
additive. J Am Chem Soc. 2013;135(26):9604–9607.
74. Iwasaki T, Takagawa H, Okamoto K, Singh SP, Kuniyasu H, Kambe N. The cobalt-
catalyzed cross-coupling reaction of alkyl halides with alkyl Grignard reagents: a new
route to constructing quaternary carbon centers-. Synthesis. 2014;46(12):1583–1592.
75. Iwasaki T, Yamashita K, Kuniyasu H, Kambe N. Co-catalyzed cross-coupling reaction
of alkyl fluorides with alkyl Grignard reagents. Org Lett. 2017;19(14):3691–3694.
76. Kreyenschmidt F, Koszinowski K. Low-valent ate complexes formed in cobalt-catalyzed
cross-coupling reactions with 1,3-dienes as additives. Chemistry. 2018;24(5):1168–1177.
77. Gulak S, Stepanek O, Malberg J, et al. Highly chemoselective cobalt-catalyzed biaryl
coupling reactions. Chem Sci. 2013;4(2):776–784.
324 Adrián Gómez-Suárez et al.

78. Czaplik WM, Mayer M, Jacobi von Wangelin A. Direct cobalt-catalyzed cross-
coupling between aryl and alkyl halides. Synlett. 2009;2009(18):2931–2934.
79. Gonnard L, Guerinot A, Cossy J. Cobalt-catalyzed cross-coupling of 3- and
4-iodopiperidines with Grignard reagents. Chem A Eur J. 2015;21(36):12797–12803.
80. Ohmiya H, Yorimitsu H, Oshima K. Cobalt(diamine)-catalyzed cross-coupling reac-
tion of alkyl halides with arylmagnesium reagents: stereoselective constructions of
arylated asymmetric carbons and application to total synthesis of AH13205. J Am Chem
Soc. 2006;128(6):1886–1889.
81. Cahiez G, Chaboche C, Duplais C, Giulliani A, Moyeux A. Cobalt-catalyzed cross-
coupling reaction between functionalized primary and secondary alkyl halides and ali-
phatic Grignard reagents. Adv Synth Catal. 2008;350(10):1484–1488.
82. Cahiez G, Chaboche C, Duplais C, Moyeux A. A new efficient catalytic system for the
chemoselective cobalt-catalyzed cross-coupling of aryl Grignard reagents with primary
and secondary alkyl bromides. Org Lett. 2009;11(2):277–280.
83. Someya H, Ohmiya H, Yorimitsu H, Oshima K. N-Heterocyclic carbene ligands in
cobalt-catalyzed sequential cyclization/cross-coupling reactions of 6-halo-1-hexene
derivatives with Grignard reagents. Org Lett. 2007;9(8):1565–1567.
84. Marchionna M, Longoni G. Hydroformylation of formaldehyde with the [Rh(CO)
2Cl2]- and [Rh5(CO)15-x(PPh3)x]- system: a case of synergetic catalysis with two
combined rhodium carbonyl species in different oxidation states. Organometallics.
1987;6(3):606–610.
85. Liu C-H, Cheng C-H. The role of alcohol in the catalytic reductive carbonylation of
nitrobenzenes to carbamates in the presence of Rh(CO)4 or Ru3(CO)12. J Organomet
Chem. 1991;420(1):119–123.
86. Ragaini F, Cenini S, Demartin F. Mechanistic studies of the carbonylation of nitroben-
zene catalysed by the [Rh(CO)4]/bipy system. X-ray structure of [PPN][Rh(CO)2-
ON[C6H3Cl2)C(O)O]; [PPN +¼(PPh3)2N +; bipy ¼ 2,2[prime or minute]-
bipyridyl]. J Chem Soc Chem Commun. 1992;(19):1467–1468.
87. Ragaini F, Cenini S, Fumagalli A, Crotti C. [Rh(CO)4], [Rh5(CO)15], and bime-
tallic clusters as catalysts for the carbonylation of nitrobenzene to methyl phe-
nylcarbamate. J Organomet Chem. 1992;428(3):401–408.
88. Drommi D, Nicolò F, Arena CG, Bruno G, Faraone F, Gobetto R. Synthesis and
structural characterization of a new series of ruthenium(II) complexes containing the
short bite Ph2PPy ligand. Cooperative effect between the anionic rhodium and cat-
ionic ruthenium species in the catalytic hydroformylation of styrene by [Ru
(Ph2PPy)3Cl] [Rh(CO)2Cl2]. Inorg Chim Acta. 1994;221(1):109–116.
89. Kim JJ, Alper H. Ionic diamine rhodium(i) complexes-highly active catalysts for the
hydroformylation of olefins. Chem Commun. 2005;41(24):3059–3061.
90. Vieira TO, Alper H. Rhodium(I)-catalyzed hydroaminomethylation of 2-
isopropenylanilines as a novel route to 1,2,3,4-tetrahydroquinolines. Chem Commun.
2007;43(26):2710–2711.
91. Vieira TO, Alper H. An efficient three-component one-pot approach to the synthesis
of 2,3,4,5-tetrahydro-1H-2-benzazepines by means of rhodium-catalyzed
hydroaminomethylation. Org Lett. 2008;10(3):485–487.
92. Okuro K, Alper H. Ionic diamine rhodium complex catalyzed hydroaminomethylation
of 2-allylanilines. Tetrahedron Lett. 2010;51(38):4959–4961.
93. Vasylyev M, Alper H. Rhodium-catalyzed reductive allylation of conjugated aldehydes
with allyl acetate. J Org Chem. 2010;75(8):2710–2713.
94. Okuro K, Gurnham J, Alper H. Ionic diamine rhodium complex catalyzed reductive
N-heterocyclization of 2-Nitrovinylarenes. J Org Chem. 2011;76(11):4715–4720.
95. Cheung LLW, Vasapollo G, Alper H. Synthesis of alcohols via a rhodium-catalyzed
hydroformylation—reduction sequence using tertiary bidentate amine ligands. Adv
Synth Catal. 2012;354(10):2019–2022.
Metallate Complexes of the Late Transition Metals 325

96. Okuro K, Gurnham J, Alper H. Ionic diamine rhodium complex catalyzed reductive
N-heterocyclization of N-(2-nitroarylidene)amines. Tetrahedron Lett. 2012;53(6):
620–622.
97. Son SU, Kim SB, Reingold JA, Carpenter GB, Sweigart DA. An anionic rhodium η4-
Quinonoid complex as a multifunctional catalyst for the arylation of aldehydes with
arylboronic acids. J Am Chem Soc. 2005;127(35):12238–12239.
98. Trenkle WC, Barkin JL, Son SU, Sweigart DA. Highly efficient 1,4-additions of
electron-deficient aryl boronic acids with a novel rhodium(I) quinonoid catalyst.
Organometallics. 2006;25(15):3548–3551.
99. Kim SB, Cai C, Faust MD, Trenkle WC, Sweigart DA. A water-stable organometallic
rhodium quinone catalyst and its recyclability. Organometallics. 2009;28(8):2625–2628.
100. Iwasaki T, Miyata Y, Akimoto R, Fujii Y, Kuniyasu H, Kambe N. Diarylrhodates as
promising active catalysts for the arylation of vinyl ethers with Grignard reagents. J Am
Chem Soc. 2014;136(26):9260–9263.
101. Kim SB, Cai C, Faust MD, Trenkle WC, Sweigart DA. The synthesis and catalytic
activity of the iridium(I) hydroquinone complex [(H2Q)Ir(COD)]+. J Organomet
Chem. 2009;694(1):52–56.
102. Fujita K-I, Kawahara R, Aikawa T, Yamaguchi R. Hydrogen production from a
methanol–water solution catalyzed by an anionic iridium complex bearing a functional
bipyridonate ligand under weakly basic conditions. Angew Chem Int Ed. 2015;54(31):
9057–9060.
103. Iwasaki T, Min X, Fukuoka A, Kuniyasu H, Kambe N. Nickel-catalyzed dimerization
and alkylarylation of 1,3-dienes with alkyl fluorides and aryl Grignard reagents. Angew
Chem Int Ed. 2016;55(18):5550–5554.
104. Iwasaki T, Fukuoka A, Min X, Yokoyama W, Kuniyasu H, Kambe N. Mul-
ticomponent coupling reaction of perfluoroarenes with 1,3-butadiene and aryl Gri-
gnard reagents promoted by an anionic Ni(II) complex. Org Lett. 2016;18(19):
4868–4871.
105. Iwasaki T, Fukuoka A, Yokoyama W, et al. Nickel-catalyzed coupling reaction of alkyl
halides with aryl Grignard reagents in the presence of 1,3-butadiene: mechanistic stud-
ies of four-component coupling and competing cross-coupling reactions. Chem Sci.
2018;.
106. Hazell A, McKenzie CJ, Preuss Nielsen L. Mono-, di- and poly-nuclear transition-
metal complexes of a bis(tridentate) ligand: towards p-phenylenediamine-bridged
co-ordination polymers. J Chem Soc, Dalton Trans. 1998;(11):1751–1756.
107. Lang C, Pahnke K, Kiefer C, Goldmann AS, Roesky PW, Barner-Kowollik C. Con-
secutive modular ligation as an access route to palladium containing polymers. Polym
Chem. 2013;4(21):5456–5462.
108. Amatore C, Jutand A. Anionic Pd(0) and Pd(II) intermediates in palladium-catalyzed
Heck and cross-coupling reactions. Acc Chem Res. 2000;33(5):314–321.
109. Amatore C, Carre E, Jutand A, M’Barki MA. Rates and mechanism of the formation
of zerovalent palladium complexes from mixtures of Pd(OAc)2 and tertiary phos-
phines and their reactivity in oxidative additions. Organometallics. 1995;14(4):
1818–1826.
110. Amatore C, Azzabi M, Jutand A. Stabilization of bis(triphenylphosphine)palladium(0)
by chloride ions. Electrochemical generation of highly reactive zerovalent palladium
complexes. J Organomet Chem. 1989;363(3):C41–C45.
111. Amatore C, Azzabi M, Jutand A. Role and effects of halide ions on the rates and mech-
anisms of oxidative addition of iodobenzene to low-ligated zerovalent palladium com-
plexes Pd0(PPh3)2. J Am Chem Soc. 1991;113(22):8375–8384.
112. Kozuch S, Shaik S, Jutand A, Amatore C. Active anionic zero-valent palladium cata-
lysts: characterization by density functional calculations. Chem A Eur J. 2004;10(12):
3072–3080.
326 Adrián Gómez-Suárez et al.

113. Carrow BP, Hartwig JF. Ligandless, anionic, arylpalladium halide intermediates in the
Heck reaction. J Am Chem Soc. 2010;132(1):79–81.
114. Zinser CM, Nahra F, Brill M, et al. A simple synthetic entryway into palladium cross-
coupling catalysis. Chem Commun. 2017;53(57):7990–7993.
115. Organ MG, Chass GA, Fang D-C, Hopkinson AC, Valente C. Pd-NHC (PEPPSI)
complexes: synthetic utility and computational studies into their reactivity. Synthesis.
2008;2008(17):2776–2797.
116. Herrmann WA, Schwarz J, Gardiner MG. High-yield syntheses of sterically demanding
bis(N-heterocyclic carbene) complexes of palladium. Organometallics. 1999;18(20):
4082–4089.
117. Nelson DJ. Accessible syntheses of late transition metal (pre)catalysts bearing
N-heterocyclic Carbene ligands. Eur J Inorg Chem. 2015;2015(12):2012–2027.
118. Heckenroth M, Kluser E, Neels A, Albrecht M. Palladation of diimidazolium salts at the
C4 position: access to remarkably electron-rich palladium(ii) centers. Dalton Trans.
2008;(44):6242–6249.
119. Zamora MT, Ferguson MJ, McDonald R, Cowie M. Carbene-anchored/pendent-
imidazolium species as precursors to di-N-heterocyclic carbene-bridged mixed-metal
complexes. Dalton Trans. 2009;(35):7269–7287.
120. Huynh HV, Han Y, Ho JHH, Tan GK. Palladium(II) complexes of a sterically bulky,
benzannulated N-heterocyclic carbene with unusual intramolecular C  HPd and
CcarbeneBr interactions and their catalytic activities. Organometallics. 2006;25(13):
3267–3274.
121. Schroeter F, Strassner T. Cationic versus anionic palladium species in the Suzuki–
Miyaura cross-coupling. Eur J Inorg Chem. 2017;2017(36):4231–4236.
122. Schroeter F, Soellner J, Strassner T. Cross-coupling catalysis by an anionic palladium
complex. ACS Catal. 2017;7(4):3004–3009.
123. Silarska E, Trzeciak AM. Oxygen-promoted coupling of arylboronic acids with olefins
catalyzed by [CA]2[PdX4] complexes without a base. J Mol Cat A, Chem. 2015;408:
1–11.
124. Zawartka W, Trzeciak AM, Ziółkowski JJ, Lis T, Ciunik Z, Pernak J.
Methoxycarbonylation of Iodobenzene in ionic liquids. A case of inhibiting effect of
imidazolium halides. Adv Synth Catal. 2006;348(1213):1689–1698.
125. Yang X, Fei Z, Geldbach TJ, et al. Suzuki coupling reactions in ether-functionalized ionic
liquids: the importance of weakly interacting cations. Organometallics. 2008;27(15):
3971–3977.
126. Gulliver DJ, Levason W. Co-ordination chemistry of higher oxidation states. Part 3.
Palladium(IV) complexes with neutral unidentate ligands. J Chem Soc Dalton Trans.
1982;(10):1895–1898.
127. Goggin PL, Goodfellow RJ, Haddock SR, Knight JR, Reed FJS, Taylor BF.
Vibrational spectra and nuclear magnetic resonance parameters of some platinum(IV)
complexes [PtX4L2] and [PtX5L]-(X ¼ Cl, Br, or I; L ¼ NMe3, PMe3, AsMe3, or
SMe2). J Chem Soc Dalton Trans. 1974;(5):523–533.
128. Iovel IG, Goldberg YS, Shymanska MV, Lukevics E. Quaternary onium
hexachloroplatinates: novel hydrosilylation catalysts. Organometallics. 1987;6(7):1410–1413.
129. Schwartsburd L, Cohen R, Konstantinovski L, Milstein D. A pincer-type anionic plat-
inum(0) complex. Angew Chem Int Ed. 2008;47(19):3603–3606.
130. Alexakis A, B€ackvall JE, Krause N, Pàmies O, Dieguez M. Enantioselective copper-
catalyzed conjugate addition and allylic substitution reactions. Chem Rev. 2008;108(8):
2796–2823.
131. Erdik E. Tetrahedron Report Number 159: copper(I) catalyzed reactions of
organolithiums and Grignard reagents. Tetrahedron. 1984;40(4):641–657.
Metallate Complexes of the Late Transition Metals 327

132. Benkovics T, Du J, Guzei IA, Yoon TP. Anionic Halocuprate(II) complexes as catalysts
for the oxaziridine-mediated aminohydroxylation of olefins. J Org Chem.
2009;74(15):5545–5552.
133. Iwasaki T, Okamoto K, Kuniyasu H, Kambe N. Cu-catalyzed reductive coupling of
perfluoroarenes with 1,3-dienes. Chem Lett. 2017;46(10):1504–1507.
134. Iwasaki T, Shimizu R, Imanishi R, Kuniyasu H, Kambe N. Copper-catalyzed
Regioselective Hydroalkylation of 1,3-dienes with alkyl fluorides and Grignard
reagents. Angew Chem Int Ed. 2015;54(32):9347–9350.
135. Tye JW, Weng Z, Johns AM, Incarvito CD, Hartwig JF. Copper complexes of anionic
nitrogen ligands in the Amidation and Imidation of aryl halides. J Am Chem Soc.
2008;130(30):9971–9983.
136. Giri R, Brusoe A, Troshin K, Wang JY, Font M, Hartwig JF. Mechanism of the
Ullmann Biaryl ether synthesis catalyzed by complexes of anionic ligands: evidence
for the reaction of Iodoarenes with ligated anionic CuI intermediates. J Am Chem
Soc. 2018;140(2):793–806.
137. Santoro O, Collado A, Slawin AMZ, Nolan SP, Cazin CSJ. A general synthetic route
to [Cu(X)(NHC)] (NHC ¼ N-heterocyclic carbene, X ¼ Cl, Br, I) complexes. Chem
Commun. 2013;49(89):10483–10485.
138. Collado A, Gómez-Suárez A, Martin AR, Slawin AM, Nolan SP. Straightforward syn-
thesis of [Au(NHC)X] (NHC ¼ N-heterocyclic carbene, X ¼ Cl, Br, I) complexes.
Chem Commun. 2013;49(49):5541–5543.
139. Fukuda Y, Utimoto K. Preparation of 2,3,4,5-Tetrahydropyridines from
5-alkynylamines under the catalytic action of gold(III) salts. Synthesis. 1991;1991(11):
975–978.
140. Fukuda Y, Utimoto K. Effective transformation of unactivated alkynes into ketones or
acetals with a gold(III) catalyst. J Org Chem. 1991;56(11):3729–3731.
141. Arcadi A, Bianchi G, Giuseppe SD, Marinelli F. Gold catalysis in the reactions of 1,3-
dicarbonyls with nucleophiles. Green Chem. 2003;5(1):64–67.
142. Alfonsi M, Arcadi A, Aschi M, Bianchi G, Marinelli F. Gold-catalyzed reactions of
2-alkynyl-phenylamines with α,β-enones. J Org Chem. 2005;70(6):2265–2273.
143. Fañanás FJ, Arto T, Mendoza A, Rodrı́guez F. Synthesis of 2,5-dihydropyridine deriv-
atives by gold-catalyzed reactions of β-ketoesters and propargylamines. Org Lett.
2011;13(16):4184–4187.
144. Arcadi A, Pietropaolo E, Alvino A, Michelet V. One-pot gold-catalyzed
aminofluorination of unprotected 2-alkynylanilines. Org Lett. 2013;15(11):2766–2769.
145. Arcadi A, Chiarini M, Vecchio LD, Marinelli F, Michelet V. Silver- versus gold-
catalyzed sequential oxidative cyclization of unprotected 2-alkynylanilines with oxone.
Chem Commun. 2016;52(7):1458–1461.
146. Zhu H, Meng X, Zhang Y, et al. Chemoselective α-methylenation of aromatic ketones
using the NaAuCl4/Selectfluor/DMSO system. J Org Chem. 2017;82(23):
12059–12065.

You might also like