Download as pdf or txt
Download as pdf or txt
You are on page 1of 15

G Model

CCR-111936; No. of Pages 15 ARTICLE IN PRESS


Coordination Chemistry Reviews xxx (2014) xxx–xxx

Contents lists available at ScienceDirect

Coordination Chemistry Reviews


journal homepage: www.elsevier.com/locate/ccr

Review

Coordination chemistry of ditopic carbanionic N-heterocyclic


carbenes
Jordan B. Waters, Jose M. Goicoechea ∗
Department of Chemistry, University of Oxford, Chemistry Research Laboratory, 12 Mansfield Road, Oxford OX1 3TA, UK

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
2. Bonding modes of N-heterocyclic carbenes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
3. “Abnormal” or mesoionic N-heterocyclic carbenes (aNHCs) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
3.1. Transition-metal aNHC complexes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
3.2. Generation of a “free” mesoionic carbene and its subsequent reactivity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
4. Ditopic carbanionic carbenes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
4.1. Metal-mediated activation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
4.2. Chemical reduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
4.3. Deprotonation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
4.3.1. Alkali-metal salts of ditopic carbanionic carbenes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
4.3.2. Group 12 complexes of ditopic carbanionic carbenes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
4.3.3. Group 13 complexes of ditopic carbanionic carbenes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
4.3.4. Group 14 complexes of ditopic carbanionic carbenes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
4.3.5. Reactivity of ditopic carbanionic carbenes towards carbon dioxide . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
4.3.6. Related compounds . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
5. Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00

a r t i c l e i n f o a b s t r a c t

Article history: The chemistry of N-heterocyclic carbenes, particularly imidazol-2-ylidenes, has been exhaustively doc-
Received 29 August 2014 umented in the literature. Amongst the many bond activation reactions available for such species (C H,
Received in revised form C C and N C), it has recently become apparent that deprotonation of the alkenic backbone can give rise
28 September 2014
to ditopic carbanionic carbenes, i.e. anionic species capable of bonding to Lewis acidic centres via two
Accepted 29 September 2014
different carbon sites (C2 and C4 or C4 and C5). In this article, we aim to provide a comprehensive review
Available online xxx
of the chemistry of such species, with a particular focus on their isolation as alkali-metal salts and their
subsequent reactivity.
Keywords:
N-heterocyclic carbenes © 2014 Elsevier B.V. All rights reserved.
Coordination chemistry
Organometallic chemistry
Acid–base chemistry

1. Introduction pioneering work of Wanzlick and Öfele [13,14], and accelerated by


Arduengo and co-workers following their seminal report describ-
The importance of N-heterocyclic carbenes (NHCs) as sup- ing the isolation of the first structurally authenticated carbene,
porting ligands in transition-metal chemistry is indisputable and 1,3-bis(adamantyl)-imidazol-2-ylidene (IAd) [15]. The extraordi-
has been extensively documented in the chemical literature nary stability of NHCs, compared to that of other acyclic all-carbon
[1–12]. This (now mature) area of chemistry was initiated by the analogues, is a result of ␲-donation by adjacent N-heteroatoms into
the empty p␲ orbital at the carbenic centre. This increased stability
has allowed for the synthesis of a wide variety of N-heterocyclic
∗ Corresponding author. carbenes. The electronic and steric parameters of such species
E-mail address: jose.goicoechea@chem.ox.ac.uk (J.M. Goicoechea). can be specifically tailored by both nitrogen and backbone-carbon

http://dx.doi.org/10.1016/j.ccr.2014.09.020
0010-8545/© 2014 Elsevier B.V. All rights reserved.

Please cite this article in press as: J.B. Waters, J.M. Goicoechea, Coord. Chem. Rev. (2014), http://dx.doi.org/10.1016/j.ccr.2014.09.020
G Model
CCR-111936; No. of Pages 15 ARTICLE IN PRESS
2 J.B. Waters, J.M. Goicoechea / Coordination Chemistry Reviews xxx (2014) xxx–xxx

Fig. 1. Bonding modes of imidazol-2-ylidenes: “classical” (I; left), and “abnormal” or


Fig. 3. Bis-C2 and bis-C4 chelating NHC ligands [34,36,37].
mesoionic (II; right). The term mesoionic is employed because such species cannot
be represented by a Lewis structure without invoking two opposing charges on the
ligand.
to the imidazole proton at the C2 position. Bearing in mind that
the acidity of the C2 proton has been determined by calculations
functionalization [16]. Accordingly, there are several reviews to be nine pKa units smaller than that of the C4 proton (pKa = 24
already published documenting the use of NHCs as ligands for and 33, respectively) [27,28], it is surprising to find that the in situ
the formation of transition-metal complexes [1–12]. Similarly, metallation of the imidazolium salt can occur exclusively at the
the strong ␴-donor ability of NHCs has also been employed in C4 position. The regioselectivity of metallation can be controlled
main-group chemistry for the isolation of low-coordinate, low- through the steric bulk of the R group, hence when R = Mes or i Pr,
oxidation state diatomic species, E2 (NHC)2 (E = Si, Ge, Sn, P, As) the sole product is 1. When R = Me, a mixture of 1 and 2 is formed.
[17–21], including, most recently a B2 (IPr)2 species (IPr = 1,3- It is important to note that 1 and 2 do not interconvert at room
bis(2,6-diisopropylphenyl)-imidazol-2-ylidene), which features a temperature or when heated to 100 ◦ C, which would suggest that 1
boron boron triple bond [22]. is the thermodynamic product of the reaction, despite the greater
acidity of the proton at the C2 position.
2. Bonding modes of N-heterocyclic carbenes The discovery of this alternative bonding mode paved the way
for an increase in the number of new complexes containing C4-
Arguably, the most ubiquitous family of N-heterocyclic car- bonded carbenes. This aspect of the coordination chemistry of
benes are those consisting of a five-membered ring core with an N-heterocyclic carbenes has been the subject of several review
unsaturated backbone, i.e. imidazol-2-ylidenes. The coordination articles in the chemical literature [29–32]. The “abnormal” bond-
chemistry of these NHCs is dominated by so-called “classical” coor- ing mode is relevant to further discussions regarding chemical
dination via the C2 position (Fig. 1, I). Numerous examples of metal bonding in ditopic carbanionic N-heterocyclic carbenes (vide infra)
complexes bearing NHCs bonded in such a manner were docu- and as such relevant examples are discussed below. For an in-
mented over the 30-year period before other coordination modes depth review of “abnormal” bonding in N-heterocyclic carbenes,
were discovered. the reader is directed to any of the excellent reviews of the area
[29–32]. We will highlight only a handful of representative exam-
3. “Abnormal” or mesoionic N-heterocyclic carbenes ples herein.
(aNHCs) C4-bonded NHC complexes of palladium are amongst the most
numerous “abnormal” complexes in the chemical literature, in
3.1. Transition-metal aNHC complexes all likelihood due to the extensive studies carried out on cat-
alytically active C2-bonded NHC palladium complexes [33–35].
The first documented example of a so-called “abnormal” (or However many other novel complexes containing “abnormally”
mesoionic) N-heterocyclic carbene (aNHC; Fig. 1, II), was reported bonded NHCs with transition metals such as ruthenium, rhodium
by Crabtree and co-workers and isolated from the reaction of and iridium are also well documented [29–32]. Rational design of
2-pyridylmethylimidazolium salts with the iridium polyhydride “abnormal” carbene complexes has been demonstrated by Albrecht
IrH5 (PPh3 )2 (Fig. 2) [23–25]. Direct metallation of imidazolium salts and co-workers who have developed a range of chelating NHC
occurs at either the C2 or C4 position, favouring one isomer or ligands which give either the bis-C2 or bis-C4-bonded NHC pal-
another as a result of either steric factors due to the nature of the ladium complexes (Fig. 3) by blocking the appropriate position on
R group, or electronic factors associated with the counter-anion of the imidazole ring with an alkyl or aryl group [34,36,37]. Likewise,
the imidazolium salt and/or the nature of the solvent [26]. “abnormally” bonded platinum(II) complexes of similar chelating
The bonding mode of 1 was unambiguously confirmed by X- NHC ligands can be synthesized by microwave-assisted double C H
ray crystallography and the presence of a low field resonance in bond activation; the corresponding platinum(IV) complexes could
the 1 H NMR spectrum in CDCl3 (8.72 ppm) which can be assigned be obtained by subsequent reaction with PhICl2 , Br2 or I2 [38].

Fig. 2. “Abnormal” (1) and “classical” (2) isomers of Crabtree’s iridium(III) complex of the 2-pyridylmethylimidazolylidene ligand [23].

Please cite this article in press as: J.B. Waters, J.M. Goicoechea, Coord. Chem. Rev. (2014), http://dx.doi.org/10.1016/j.ccr.2014.09.020
G Model
CCR-111936; No. of Pages 15 ARTICLE IN PRESS
J.B. Waters, J.M. Goicoechea / Coordination Chemistry Reviews xxx (2014) xxx–xxx 3

“Abnormally” bonded NHC complexes have also been synthe-


sized fortuitously while attempting to obtain the C2-bonded NHC
by coordination of a free NHC with a metal complex [39]. In
an attempt to isolate the two-coordinate cationic nickel species
[(IPr)Ni(NO)]+ , the three-coordinate complex 5 (Fig. 4) could be
obtained from the reaction of (IPr)Ni(I)(NO) (and one equivalent
of IPr for improved yield) with TlBArF 4 (in order to abstract the
iodide anion); 5 exhibits C2 and C4 coordinated NHCs to the nickel
atom.
The Ni C2 and Ni C4 bond lengths in 5 are 1.970(3) and
1.931(3) Å, respectively, with a shorter bond length for the more
Fig. 4. Synthesis of an “abnormally” bonded nickel(II) NHC complex by direct reac-
tion of the free carbene with a cationic metal complex [39]. strongly sigma-donating abnormal NHC. However a shorter bond
to an “abnormally” bonded NHC ligand is not always the case.
For example, reaction of an iron(II) complex with two equivalents
of a dicarbene C N C pincer ligand affords a complex in which
one of the NHC units is bonded to the iron centre through the C4
carbon [40]. In the iron chelate complex, the M C bond lengths
appear to be independent of the bonding mode with very little dif-
ference between the M C2 and M C4 bond lengths (although as
mentioned previously, the steric demands of the chelating ligand
presumably play an important factor).
More recently, Layfield and co-workers reported a “classical”
to “abnormal” rearrangement occurs upon refluxing the three-
coordinate iron(II) NHC complex 6 in toluene for 3 h (Fig. 5) [41].
Fig. 5. A “classical” to “abnormal” carbene rearrangement in a three-coordinate
iron(II) complex [41].
This reaction clearly demonstrates the potential for “abnormal”
bonding in NHC complexes to be thermodynamically controlled.
It is likely that the aNHC bonding mode is favoured due to
In order to assess the electronic implications of “abnormal” the relief of steric congestion; 7 has only one adjacent N Dipp
bonding compared to the “classical” mode, one may consider the substituent adjacent to the metal bonded carbon atom. Addi-
differences in M C bond lengths, however, as a result of the tionally, as the “abnormal” carbene is more strongly ␴-donating
steric impact of the adjacent substituents, the bond length can be relative to the normal carbene, a reduction in the Fe C bond
influenced to a greater degree by these constraints rather than elec- length can be observed (c.f. 2.182(2) and 2.117(2) Å for 6 and 7,
tronic factors. This can make it difficult to disentangle competing respectively).
electronic and steric effects which influence this parameter. The C4-bonded complexes have also been synthesized by transmet-
torsion angle between the ligand and metal coordination plane in allation using silver(I) precursors. Such a strategy was employed
palladium(II) square planar complexes were typically 20–30◦ in C4- by Lassaletta and co-workers while investigating the possibil-
bonded NHCs, and increases to 35–45◦ in corresponding C2-bonded ity of “abnormal” coordination of an annulated imidazole ligand,
NHCs due to the steric repulsion from the additional adjacent N- imidazo[1,5-a]pyridine (Fig. 6) [42]. Generation of the silver com-
substituent. Albrecht demonstrated that the bond length of the plex 9 by reaction of Ag2 O with the imidazo[1,5-a]pyridinium salt
Pd X bond trans to the NHC ligand could be more indicative of 8 followed by addition of [Ir(COD)Cl]2 (COD = 1,5-cyclooctadiene)
changes in the electronic structure. The structure and geometry of resulted in the formation of the C4-bound complex 10. The “free
the NHCs in 3 and 4 were shown to be very similar, however, the carbene” 11 could be generated by deprotonation with KHMDS
average Pd Cl bond length is longer in 4 (2.404(4) Å compared to in situ and trapped by reaction with selenium to yield 12. The
2.357(2) Å in 3), implying a greater trans influence of the C4-bound presence of the fused ring in the ligand allows for effective delo-
carbene (consistent with its greater allylic character and increased calization of electron density and the C4 position can be viewed as
␴-donor ability). more electrophilic, hence with greater carbene character. This can

Fig. 6. Generation of an “abnormal” NHC complex via a transmetallation reaction [42].

Please cite this article in press as: J.B. Waters, J.M. Goicoechea, Coord. Chem. Rev. (2014), http://dx.doi.org/10.1016/j.ccr.2014.09.020
G Model
CCR-111936; No. of Pages 15 ARTICLE IN PRESS
4 J.B. Waters, J.M. Goicoechea / Coordination Chemistry Reviews xxx (2014) xxx–xxx

Fig. 7. Generation of 1,3-bis(mesityl)-4,5-dichloroimidazol-2-ylidene (13) by generation of an “abnormal” chlorinated intermediate [45].

Fig. 8. Generation of C4-functionalized NHCs (16) via the formation of mesoionic carbene intermediates (15) [46].

be seen in the longer C4 C5 bond length when compared to similar The first isolable “abnormal” carbene (Fig. 9) was synthesized
C2 bound complexes of imidazo[1,5-a]pyridines. by Bertrand and co-workers in 2009 by utilizing an imidazolium
The structurally similar imidazo[1,2-a]pyridine has also been salt with a non-labile phenyl group blocking the C2 position
used for the generation of C4-bonded complexes by Li and co- (17) [47]. Deprotonation of 17 using two equivalents of potas-
workers [43]. The C2 carbon is a part of the pyridine ring and sium bis(trimethylsilyl)amide led to the formation of the free
therefore “normal” NHC coordination is not possible. There are, aNHC 18, which could be isolated as a green powder and was
however, two different “abnormal” positions on the ligand. Met- stable in solution and as a solid for several days under an
allation can be directed by installing a blocking group at the inert atmosphere. 18 offers additional stability over unsubstituted
appropriate position on the ring. The carbonyl stretching fre- aNHCs as electron density can be delocalized into the C5-bound
quencies of iridium chloride carbonyl complexes of these ligands phenyl group. A decrease in bond angle about the C4 carbon
could be measured, and by changing the substituents, the elec- can be observed (108.0(3)◦ and 101.0(2)◦ for 17 and 18, respec-
tron donor–acceptor properties of the ligand could be evaluated. tively) which is characteristic for “abnormal” carbenes due to
Complexes which allow for electron delocalization are better ␲- the lone pair occupying an orbital of greater s character (the
acceptors granting the Ir C bond greater double bond character. HOMO of 18 was calculated to be a ␴-type lone-pair orbital). The
Similar palladium complexes could be synthesized by Ghosh and bond angle observed in NHCs for the N C2 N angle typically
co-workers for use as precatalysts in Sonogashira coupling reac- comes within the range 101–102◦ for the free carbene; whereas
tions [44]. for C2-metallated NHCs the average bond angles fall between
101◦ and 107◦ .
3.2. Generation of a “free” mesoionic carbene and its subsequent Bertrand’s isolation of a free mesoionic carbene has enabled the
reactivity direct addition of the carbene to a transition-metal or main-group
reagent in the targeted synthesis of C4-bonded complexes. Grubbs
The acidity of the backbone C H protons in imidazol-2-ylidene and co-workers reacted 18 with two equivalents of [Fe(COT)2 ]
type systems was demonstrated by Arduengo and co-workers (COT = 1,3,5,7-cyclooctatetraene) generating the highly reactive
in 1997 prior to the discovery of transition-metal complexes of species 19 (Fig. 10) [48]. Attempts to isolate a similar product using
“abnormally” bonded carbenes. Addition of two equivalents of CCl4 the saturated NHCs, SIMes and SIPr were unsuccessful affording tri-
to a tetrahydrofuran (THF) solution of 1,3-bis(mesityl)-imidazol- and tetra-metallic complexes [49].
2-ylidene (IMes) yields 1,3-bis(mesityl)-4,5-dichloroimidazol-2- Furthermore, 18 can be used as a strong Lewis base in the
ylidene 13 (Fig. 7) [45]. The reaction proceeds by generating a stabilization of heavy group 14 metallylenes [50,51]. Addition of
trichloromethyl anion in situ which is basic enough to deprotonate IPr·SiCl2 , GeCl2 ·dioxane or SnCl2 to the abnormal carbene afforded
the chlorinated imidazolium cation. The resulting mesoionic 1,3- the aNHC adduct of the group 14 element dichloride (20) (Fig. 11).
bis(mesityl)-2-chloro-imidazol-4-ylidene reacts with additional The quantitative exchange of IPr for 18 during the reaction with
CCl4 to chlorinate the C4 position. This process is then repeated IPr·SiCl2 is indicative of the greater ␴-donor strength of the abnor-
to give the dichlorinated product. The greater stability of 13 over mal carbene.
IMes is apparent since IMes is too strong a base to tolerate acidic
solvents such as chloroform, whereas 13 is formed in a reaction
that generates chloroform as a by-product.
Subsequently, the scope of this reactivity was exploited by
Bertrand and co-workers who demonstrated that addition of
IPr to various electrophilic reagents (E X where E = PhC(O),
PPh2 ; X = Cl; and E = CF3 SO2 , Me3 Si; X = CF3 SO3 ) generated the
imidazolium adducts 14 (Fig. 8) [46]. Addition of potassium
bis(trimethylsilyl)amide led to the deprotonation of the imidaz-
olium cation to generate an “abnormal” carbene as an intermediate
(15), however intramolecular rearrangement of 15 yielded the C4-
functionalized NHC 16. Fig. 9. The first isolable “abnormal” N-heterocyclic carbene [47].

Please cite this article in press as: J.B. Waters, J.M. Goicoechea, Coord. Chem. Rev. (2014), http://dx.doi.org/10.1016/j.ccr.2014.09.020
G Model
CCR-111936; No. of Pages 15 ARTICLE IN PRESS
J.B. Waters, J.M. Goicoechea / Coordination Chemistry Reviews xxx (2014) xxx–xxx 5

The Whittlesey group found a similar activation occurred


between Arduengo’s carbene, IAd, and Ru3 (CO)12 [84]. The equiv-
alent reaction with Os3 (CO)12 with both IAd and It Bu gave the
abnormal carbene complex, the reaction required heating at 70 ◦ C
for 3 h due to the substitutionally inert Os CO bonds. Refluxing
a THF or pyridine solution of the abnormal osmium complexes
provided evidence by NMR spectroscopy for the formation of the
carbanionic carbene complex analogous to 22, however the reac-
tion failed to go to completion and other side products were
observed.
While exploring the reactivity of functionalized NHCs bearing
Fig. 10. Reaction of an isolable mesoionic N-heterocyclic carbene with Fe(COT)2
potential bridging functionalities, Meyer and co-workers synthe-
[48].
sized the first examples of NHCs bonded to two transition metals
through the C2 position and “abnormal” C4 position simultaneously
4. Ditopic carbanionic carbenes (Fig. 14) [85]. Reaction of an N-pyridazine functionalized imidaz-
olium salt 23 with allylpalladium chloride dimer initially produces
Ditopic carbanionic carbenes are the result of a formal depro- the anticipated mononuclear palladium complex 24.
tonation of a neutral NHC to give an anionic ligand capable of When a CD3 CN solution of 24 was left to stand, small orange
bonding through the “classical” C2 and “abnormal” C4 carbene cen- crystals of the dinuclear complex 25 grew over a period of about
tres simultaneously, hence this has also led to the description of 4 months. This is the result of two molecules of 24 reacting with
these species as anionic “dicarbenes” (NHDCs; pictured in Fig. 12). one another (presumably via initial coordination by the pyridazine
It is also possible for a ditopic carbanionic carbene to bridge two moiety), in which C H activation of the NHC backbone occurs due
metal centres via the C4 and C5 positions to give a second bonding to the close proximity of the coordinated palladium atom and sub-
mode. To date, such species are limited to NHCs with unsaturated sequent elimination of propene. In the presence of an excess of
alkenic backbones. The synthesis of such ligands can be separated [(allyl)PdCl]2 , 24 slowly generates the trimetallic complex 26 which
into three methods, namely via metal-mediated C H activation, can be viewed as two molecules of 24 bridged through the activated
chemical reduction and direct deprotonation. NHC backbone positions by a palladium(II) ion.
The capability of ditopic carbanionic carbenes to bridge two In 2012, Albrecht and co-workers synthesized a variety of
metal centres through different carbon atoms of the imidazole ring “abnormally bonded” palladium complexes by oxidative addition of
closely resembles the bonding mode of neutral 1,2,4-triazole-3,5- bis(dibenzylideneacetone)palladium(0) to 4-iodo-functionalized
diylidenes [52,53]. Both types of backbone activated NHCs involve NHCs bearing N-bound donor groups (Fig. 15). A targeted synthe-
a metal bonded to the C4 carbene centre (which has significant sis of the ditopic carbene complex 27 could be achieved with an
vinylic character), a bonding mode closely related to that observed iodo-functionalized NHC by initial metallation at the C2 position
for transition metal complexes of “abnormally bonded” carbenes by coordination with a palladium(II) species followed by oxidative-
(vide supra). addition into the C4 I bond with palladium(0) in the presence of
the bidentate ligand bipyridine [86].
4.1. Metal-mediated activation
4.2. Chemical reduction
The ability of N-heterocyclic carbenes to undergo C H, C C and
C N activation of the nitrogen-bonded functionalities has been As part of an investigation into the f-element chemistry of
well documented in the chemical literature [41,54–82]. By contrast, NHCs, Arnold and Liddle attempted the chemical reduction of sev-
metal-mediated activation of the alkenic backbone C H bonds is eral NHC complexes of the f-elements bearing amido-tethered
much rarer. The first example of the formation of a ditopic car- N-heterocyclic carbenes [87]. Reduction of an NHC complex
banionic carbene by a transition-metal complex was reported by of yttrium (28a) using potassium naphthalenide in a mixture
Whittlesey and co-workers in 2007 and features an NHC bonded of dimethoxyethane (DME) and diethyl ether (Et2 O) afforded
to two metals centres via the C4 and C5 positions simultaneously the bimetallic dimer 29a in the solid-state as characterized by
[83]. Reaction of 1,3-di-tert-butylimidazol-2-ylidene (It Bu) with X-ray crystallography (Fig. 16). 29a consists of a ditopic car-
the ruthenium carbonyl cluster Ru3 (CO)12 in THF at room tempera- banionic carbene where the C4 carbon coordinates to yttrium
ture afforded the abnormal carbene complex 21 (Fig. 13). When 21 and the C2 carbene centre is now coordinated to the potas-
was heated in THF to 50–70 ◦ C, activation of the second backbone sium cation. Close contacts between the potassium cation and
C H was observed resulting in the formation of the ␮3 –␩2 -bridging the second NHC C4 carbon can be observed in the crystal struc-
carbanionic carbene complex 22 containing a bridging hydride ture.
ligand. In order to probe the mechanism of the reduction (and sub-
sequent loss of hydrogen from the C4 position), the isostructural
samarium complex was reduced using the same conditions as
samarium(III) has a lower reduction potential than yttrium(III). The
isostructural samarium complex of 29b could be isolated in low
yield after reduction with KC10 H8 . The low yield is said to support
the formation of an intermediate one-electron reduced NHC in the
formation of 29a since the reduction of yttrium is less favourable
than the reduction of the samarium metal centre (which is rela-
tively more stable in the divalent oxidation state). 29 can react with
electrophiles such as Me3 SiCl to regenerate a C2 coordinated NHC,
functionalized with a TMS group on the backbone. Further investi-
Fig. 11. Formation of “abnormal” NHC carbene complexes of EX2 (E = Si, Ge, Sn) gation into the reduction of the parent amine–NHC allowed for the
[50,51]. characterization of a dark red toluene solution of the radical anion,

Please cite this article in press as: J.B. Waters, J.M. Goicoechea, Coord. Chem. Rev. (2014), http://dx.doi.org/10.1016/j.ccr.2014.09.020
G Model
CCR-111936; No. of Pages 15 ARTICLE IN PRESS
6 J.B. Waters, J.M. Goicoechea / Coordination Chemistry Reviews xxx (2014) xxx–xxx

Fig. 12. Bonding modes of N-heterocyclic carbanionic carbenes.

Fig. 13. Metal-mediated activation of an “abnormal” NHC to afford a complex of a ditopic carbanionic carbene bridging two metal sites via the C4 and C5 positions [83].

Fig. 14. Synthesis of the first palladium complexes of ditopic carbanionic carbenes [85].

K+ [t BuNHCH2 CH2 {:CN(CH)2 (Nt Bu)}]•− (30), by EPR spectroscopy. Having synthesized an NHC-stabilized diphosphorus fragment
These results demonstrate that the electron resides in the ␲-system by the reduction of the NHC–PCl3 adduct with KC8 , Robinson
of the NHC. Prior to this work, Arnold and co-workers had already and co-workers were able to reduce this species further using
synthesized the first stable K–NHC complex by deprotonation of an lithium metal in THF to afford the anionic carbene-stabilized parent
alcohol-functionalized imidazolium iodide [88]. phosphinidene 31 [89], where the “normal” carbene position is

Fig. 15. Synthesis of a bimetallic palladium complex of a ditopic carbanionic carbene by sequential metallation of the backbone-functionalized NHC [86].

Please cite this article in press as: J.B. Waters, J.M. Goicoechea, Coord. Chem. Rev. (2014), http://dx.doi.org/10.1016/j.ccr.2014.09.020
G Model
CCR-111936; No. of Pages 15 ARTICLE IN PRESS
J.B. Waters, J.M. Goicoechea / Coordination Chemistry Reviews xxx (2014) xxx–xxx 7

Fig. 16. Chemical reduction of an NHC complex to afford a complex of a ditopic carbanionic carbene [87].

for the carbene resonance at 180.5 ppm in the 13 C{1 H} NMR for the
protonated species, however no comparable data is available for
the lithiated species.
Our research group has also recently demonstrated the reduc-
tion of a variety of bis(mesityl) transition-metal complexes of the
widely employed IPr ligand allows for the isolation of complexes
bearing two ditopic carbanionic carbenes, both bound through the
anionic C4 position to the metal centre (Fig. 18) [92–94]. The bis-
ditopic carbanionic carbene complexes 33 (M = Mn (33a), Fe (33b),
Fig. 17. Chemical reduction of a NHC-stabilized P2 moiety to afford a phosphinidine
bonded to a ditopic carbanionic carbene [89]. Zn (33c) and Cd (33d)) can be isolated by multiple recrystallizations
from THF/hexane. This was necessary in order to remove the metal
tris(mesityl) complex, a by-product of the reaction as a result of
bound to the phosphinidene moiety and the carbanionic backbone ligand re-organization upon reduction. The balanced equation can
coordinated to a solvated lithium cation (Fig. 17). be written by invoking the formation of dihydrogen as a by-product.
31 can be viewed as the deprotonated form of the neutral Reduction of a mixture of IPr and Hg(mes)2 using KC8 in THF
NHC-stabilized phosphinidene, IPr PH (32). The first struc- (IPr does not form an adduct with Hg(mes)2 as evidenced by DOSY
turally authenticated phosphinidene was reported by Driess and NMR), the homoleptic tris-ditopic carbanionic carbene complex 34
co-workers, while IPr PH was recently crystallographically char- could be isolated from the reaction as the [K(2,2,2-crypt)]+ salt
acterized by Grützmacher and colleagues [90,91]. The 31 P NMR (Fig. 19).
spectrum for the lithiated and protonated species each display a These complexes bind to the metal centre via the backbone car-
doublet at ı = −143.0 and −134.3 ppm, with 1 JP H of 171.0 and bon while the “normal” C2 carbene centre remains uncoordinated.
165.5 Hz, respectively. These data, along with somewhat similar The availability of the free carbene to do further chemistry was
C P bond lengths of 1.763(2) and 1.752(1) Å determined by X-ray demonstrated by the reaction of the complexes with various Lewis
crystallography, indicate that bonding between the carbene and acids such as BH3 , AlMe3 and CO2 (Fig. 20), in a similar manner
phosphorus is not significantly affected by backbone deprotona- to the Lewis acid–base adducts which are already known for free
tion. It is worth noting that there is a large 1 JC P of 85.0 Hz observed NHCs [95–112].

Fig. 18. Synthesis of transition metal complexes bearing two “abnormally” bonded ditopic carbanionic carbenes by chemical reduction of NHC complexes [92–94].

Fig. 19. Synthesis of a homoleptic mercury(II) complex bearing three “abnormally” bonded ditopic carbanionic carbenes [93].

Please cite this article in press as: J.B. Waters, J.M. Goicoechea, Coord. Chem. Rev. (2014), http://dx.doi.org/10.1016/j.ccr.2014.09.020
G Model
CCR-111936; No. of Pages 15 ARTICLE IN PRESS
8 J.B. Waters, J.M. Goicoechea / Coordination Chemistry Reviews xxx (2014) xxx–xxx

Fig. 20. Reaction of complexes of ditopic carbanionic carbenes with Lewis acids demonstrating the carbenic nature of the C2 positions [93].

resulting in a ditopic trianionic carbene (39). Two low-field reso-


nances can be observed in the 13 C NMR spectrum of 39 at 195.6
and 169.5 ppm which correspond to the C2 and C4 carbons, respec-
tively.

4.3.2. Group 12 complexes of ditopic carbanionic carbenes


The reactivity of 35 towards the organometallic reagent ZnEt2
was also explored by the Robinson group in 2012 yielding a variety
of triorganozincates (Fig. 23) [115]. Stirring a slurry of 35 in hexane
Fig. 21. Deprotonation of IPr to afford a lithium salt of a ditopic carbanionic carbene
[113].
with one equivalent of ZnEt2 afforded the expected C4 functional-
ized ditopic carbanionic NHC 40 which could be crystallized from
TMEDA following complexation of the C2 bound lithium cation.
4.3. Deprotonation However, 40 was observed to slowly dissociate in THF and convert
into the mixture of products 41 and 42. The triorganozincate 41
4.3.1. Alkali-metal salts of ditopic carbanionic carbenes could be synthesized by reaction of 35 with an excess of ZnEt2 in
The generation of a ditopic carbanionic carbene requires the THF/TMEDA. 41 consists of two ditopic carbanionic NHCs bridged
activation of one of the protons of the alkenic backbone, however by a ZnEt fragment through the C4 position, and each coordinated
it was not until 2010 that Robinson and co-workers demonstrated to ZnEt2 moieties by the C2 carbene centre, a structure that bears
that a straightforward deprotonation was possible affording a a resemblance to the products reported by our group by reduction
lithium salt of a ditopic carbanionic carbene (Fig. 21). It was shown with KC8 (33c).
that IPr could be deprotonated using an equivalent of a sufficiently The novel synthesis of the ditopic carbanionic NHC complexes
strong enough base such as n BuLi [113]. Addition of n BuLi to a 43 and 44 (Fig. 24) was achieved by in situ deprotonation and
hexane slurry of the carbene at ambient temperature yields the zincation of the free NHC IPr, or the IPr·Znt Bu2 adduct, utilizing
anionic deprotonated carbene 35 ([:C{N(2,6-i Pr2 C6 H3 )}2 (CH)CLi]n ) the heteroleptic sodium zincate [(TMEDA)NaZn(TMP)(t Bu)2 ], a syn-
in near quantitative yield. This facile synthesis of an alkali metal thesis described as an alkali-metal-mediated zincation (AMMZn)
salt of a ditopic carbene makes 35 a useful starting material for the by Hevia and co-workers [116]. The scope of the alkali-metal-
potential synthesis of new complexes bearing ditopic carbanionic mediated zincation was investigated further with other NHCs,
carbenes. 35 is an amorphous, off-white powder which can readily namely It Bu, IAd and IMes, and similar reactivity could be con-
dissolve in THF or TMEDA; complexation of the lithium cation by firmed by comparison of the 2 H NMR spectrum of the in situ CD3 OD
these strongly coordinating solvents allows for the crystallization quenched reaction mixtures, however despite several attempts,
of the anionic complex. The solid state structure of 35·THF shows a the isolation of crystalline products was not possible. Unlike other
polymeric chain of anionic NHCs and Li cations. Both the C2 and C4 examples of AMMZn of aromatic and heterocyclic molecules, the
positions coordinate to lithium cations to form the chain with bond products 43 and 44 are stable to the TMPH by-product, and the
lengths of 2.175(6) and 2.216(6) Å for Li C2 which are somewhat Zn t Bu bonds remain intact. Attempts to deprotonate the NHC
longer than the Li C4 bond lengths of 2.125(6) and 2.122(6) Å (as backbone of IPr–Znt Bu2 with t BuLi were unsuccessful, instead
there are two carbene fragments in the asymmetric unit). This is as forming the IPr stabilized homoleptic zincate 45.
expected for the more strongly donating “abnormal” position. Considering the well-known ability for organozinc reagents to
Reduction of the free carbene in THF by lithium metal could sim- undergo transmetallation reactions, 43 was added to (PPh3 )AuCl
ilarly generate the ditopic carbanionic carbene as the THF adduct in THF to afford the novel dinuclear gold complex 46 (although a
(35·THF). This however gives the product in a lower yield, again, greater yield could be obtained when two equivalents of (PPh3 )AuCl
possibly with dihydrogen as the by-product. were used) (Fig. 25). This reactivity is different to that observed
Earlier this year, El-Hellani and Lavallo combined carborane for the lithiated ditopic NHC–borane reported by the Tamm group
anions in the synthesis of NHCs (Fig. 22) [114]. By synthesizing which is discussed in greater detail in Section 4.3.3 and affords a
an imidazolium precursor which has been N-functionalized with complex that is only metallated at the C2 position [117]. 46 con-
anionic carba-closo-dodecaboranes (36), subsequent deprotona- sists of a ditopic carbanionic carbene whereby a formally cationic
tion using a suitable choice of base and reaction conditions allows Au PPh3 fragment has coordinated to the anionic C4 backbone
for the isolation of either the C2 or C4 deprotonated NHC (37 and and the C2 position coordinates a formally neutral Au Cl fragment
38). Both, despite having two imidazole protons remaining, are with loss of Znt Bu2 and NaCl. When the reaction was carried out
themselves dianionic due to the carborane substituents. By using with (SMe2 )AuCl, decomposition was readily observed indicating
three equivalents of n BuLi, the C2 and C4 protons can be removed the vital role of the stabilizing PPh3 ligand.

Please cite this article in press as: J.B. Waters, J.M. Goicoechea, Coord. Chem. Rev. (2014), http://dx.doi.org/10.1016/j.ccr.2014.09.020
G Model
CCR-111936; No. of Pages 15 ARTICLE IN PRESS
J.B. Waters, J.M. Goicoechea / Coordination Chemistry Reviews xxx (2014) xxx–xxx 9

Fig. 22. Generation of ditopic carbanionic carbenes by deprotonation of carborane-functionalized NHCs [114].


Robinson’s report of the lithiated NHC [:C{N(2,6- trigonal planar coordination geometry ( bond angles = 360.0◦ ) with
i Pr C H )} (CH)CLi]
2 6 3 2 n (35) [113], encouraged us to synthesize a Zn–carbanionic carbene distance of 2.008(1) Å.
an alternative alkali metal salt of the carbanionic carbene, namely
the organo-potassium compound [:C{N(2,6-i Pr2 C6 H3 )}2 (CH)CK]n 4.3.3. Group 13 complexes of ditopic carbanionic carbenes
(47) for use as a reagent to synthesize novel complexes bearing The synthetic utility of 35 as a ditopic carbanionic carbene
ditopic carbanionic carbenes as ligands [118]. Reaction of the reagent has been illustrated by reaction with various Lewis acids
lithiated NHC with KOt Bu in Et2 O afforded the novel organo- and electrophiles. When AlMe3 or BEt3 was added to a slurry
potassium salt, KIPr, after washing with a mixture of THF and of 35 in the non-polar solvent toluene, the Lewis acid coordi-
hexane to remove the LiOt Bu by-product. Curious as to whether nates to the carbanionic backbone resulting in lithium migration
the carbanionic NHC could be isolated as the free anion, we to the C2 carbon to give Li[:C{N(2,6-i Pr2 C6 H3 )}2 (CH)C(LA)]·2THF,
attempted to sequester the potassium cation using 2,2,2-crypt, (49) after crystallization from THF (Fig. 26). However, Robinson
however this resulted in the decomposition of the cryptand and and co-workers were able to target the “normally” bound anionic
generation of the protonated NHC, a consequence of the high IPr–BEt3 complex 50 using a similar method to that employed by
Brønsted basicity of the proposed free anion. Similar to 35, 47 Roesky and co-workers when synthesizing the equivalent anionic
could be crystallized as the THF adduct forming polymeric chains IPr–BH3 complex [119,120]. Addition of n BuLi to the neutral
in the solid state of alternating anionic NHCs and potassium carbene–borane adducts in THF serves to deprotonate the backbone
cations bridging the C2 position of one NHC and the carbanionic and no reorganization of the borane moiety is observed. Never-
C4 position of another. Predictably, the molecular structure theless, heating 50 will drive the formation of the thermodynamic
of the carbanionic moiety in 47 is very similar to that in 35. product 49. This allows for functionalization at the C2 or C4 position
Reaction of 47 with Zn[N(SiMe3 )2 ]2 affords the three-coordinate as demonstrated by the reactions of 49 and 50 with GaCl3 [119].
complex [Zn{C(CH)[N(2,6-i Pr2 C6 H3 )]2 C:}{N(SiMe3 )2 }2 ]− (48) It is an interesting observation that 49 is the more thermo-
which could be isolated and crystallographically characterized dynamically stable product; on the basis of hard–soft acid–base
as a [K(2,2,2-crypt)]+ salt. This novel zinc(II) complex displays a theory, the hardest Lewis base, the carbanionic C4 position, would

Fig. 23. Synthesis of triorganozincates of ditopic carbanionic carbenes by reaction of 35 with ZnEt2 [115].

Please cite this article in press as: J.B. Waters, J.M. Goicoechea, Coord. Chem. Rev. (2014), http://dx.doi.org/10.1016/j.ccr.2014.09.020
G Model
CCR-111936; No. of Pages 15 ARTICLE IN PRESS
10 J.B. Waters, J.M. Goicoechea / Coordination Chemistry Reviews xxx (2014) xxx–xxx

Fig. 24. Alkali-metal-mediated zincation of IPr to afford bimetallic complexes of ditopic carbanionic carbenes [116].

Fig. 25. Transmetallation of a bimetallic complex of a ditopic carbanionic carbene


[116].
Fig. 27. Alkylation and silylation of “abnormally” bonded borane adducts of ditopic
carbanionic carbenes [122].
preferentially coordinate to the hardest Lewis acid, the lithium
cation. The explanation for the relatively decreased stability of 50
may be a result of steric interaction between the borane ethyl [121]. Robinson and co-workers subsequently demonstrated the
groups and the large Dipp groups. This can be seen in the bond ability to protonate the C2 carbon in 49 using HCl·NEt3 to give the
angle around the nitrogen atom between carbene and Dipp sub- abnormally bound NHC–borane which could be deprotonated with
stituents which are 131.4(4)◦ , 126.1(4)◦ and 124.9(2)◦ , 118.7(2)◦ n BuLi to regenerate the lithiated complex 49 [122]. The reaction of

for 50 and 49, respectively. Evidently, the steric bulk of the tri- 49 with electrophiles such as Me3 SiCl and MeOTf showed that func-
ethylborane increases the C2 N Dipp angle relative to the solvated tionalization would occur at the free carbene site to generate the
lithium cation. neutral, abnormally bound NHC–borane adducts 53 (Fig. 27).
Addition of two equivalents of BH3 to 35 in THF, or one equiv- Frustrated NHC–borane Lewis pairs, such as IPr or It Bu/B(C6 F5 )3 ,
alent to 49, gave the corresponding anionic NHC dinuclear borane readily activate small molecules such as dihydrogen and ammo-
species [BR3 :C{N(2,6-i Pr2 C6 H3 )}2 (CH)C:BH3 ]− (R = H (51), Et (52)) nia [123,124]. However, Tamm and co-workers discovered that an

Fig. 26. Generation of “classically” and “abnormally bonded” borane adducts of ditopic carbanionic carbenes [119,120].

Please cite this article in press as: J.B. Waters, J.M. Goicoechea, Coord. Chem. Rev. (2014), http://dx.doi.org/10.1016/j.ccr.2014.09.020
G Model
CCR-111936; No. of Pages 15 ARTICLE IN PRESS
J.B. Waters, J.M. Goicoechea / Coordination Chemistry Reviews xxx (2014) xxx–xxx 11

addition of the lithium salt to the ruthenium complex resulted only


in intractable mixtures indicating the potential importance of the
metal cation for controlling the reactivity of ditopic carbanionic
carbene reagents. The complex contains the ditopic carbanionic
carbene bound through the C4 carbon to the ruthenium centre and
the C2 position remains coordinated to the neutral Lewis acid BF3 .
Variable temperature NMR studies indicated the possibility of a
monomeric species existing in equilibrium with the dimer in solu-
tion, and the “monomer” of 57 could be isolated upon reaction with
Fig. 28. Transmetallation reactions of backbone functionalized NHCs [117]. the two-electron donor ligands PPh3 or CO. Upon abstraction of
the chloride ligand using [Ag(Et2 O)][B(C6 F5 )4 ], a rearrangement of
initially yellow toluene solution of It Bu/B(C6 F5 )3 turned colour- the carbene occurred whereby one of the isopropyl methine pro-
less over about 2 h and had lost any reactivity towards dihydrogen tons had migrated to the NHC backbone leaving the deprotonated
[124]. The “abnormal” NHC–borane adduct had formed irreversibly Dipp group coordinated to the cationic ruthenium centre, reactivity
due to the reduced steric hindrance around the C4 carbon, hence no unique to the ditopic carbanionic carbene in contrast with neutral
reactivity towards small molecules could be observed. This result NHCs.
led the group to investigate whether the NHC could be deproto- While investigating low-coordinate iron(II) hydride complexes
nated to afford the “abnormal” NHC bearing a weakly coordinating bearing chelating bis-NHC ligands, Zlatogorsky and Ingleson iso-
borane moiety. Attempts to deprotonate the zwitterionic species lated an iron hydride complex of a ditopic carbanionic carbene
were unsuccessful, however, by the same method as Robinson, [126]. Reacting two equivalents of LiBHEt3 with a bis-NHC iron(II)
addition of B(C6 F5 )3 to the in situ generated ditopic carbanionic car- iodide complex led to the formation of 58 and 59. Both products
benes 35 or the analogous lithium salt of It Bu, the desired anionic are the result of the deprotonation of the NHC backbone and sub-
abnormal NHC–borane complexes 54a and 54b could be isolated. sequent coordination of the neutral borane (Fig. 30).
Tamm then explored the reactivity of the anionic species and syn-
thesized the gold(I) complexes 55a and 55b by a transmetallation 4.3.4. Group 14 complexes of ditopic carbanionic carbenes
reaction with (Ph3 P)AuCl (Fig. 28) [117]. We have recently employed the potassium salt of deproto-
Earlier this year, Pranckevicius and Stephan isolated the lithium nated IPr, 47, as a reagent for the formation of a variety of novel
salt of the IPr–BF3 adduct and used this reagent for the synthe- three-coordinate group 14 metal complexes [118]. Reaction of
sis of a silver salt of the carbanionic NHC (56) [125]. Stirring one 47 with group 14 bis(trimethylsilyl)amides, i.e. M[N(SiMe3 )2 ]2
equivalent of [Ru(p-cymene)Cl2 ]2 with 56 gave rise to a dark red (M = Ge, Sn and Pb), lead to the formation of the three coordi-
solution containing the dimeric complex 57 (Fig. 29). Interestingly, nate complexes [M{C(CH)[N(2,6-i Pr2 C6 H3 )]2 C:}{N(SiMe3 )2 }2 ]− as

Fig. 29. Generation of a homoleptic silver(I) complex bearing two ditopic carbanionic carbenes and its subsequent reactivity [125].

Fig. 30. Reduction of an iron(II) complex of a chelating NHC to afford a backbone-functionalized complex 58 and its subsequent reactivity [126].

Fig. 31. Reactivity of a potassium salt of a ditopic carbanionic carbene towards group 14 amides [118].

Please cite this article in press as: J.B. Waters, J.M. Goicoechea, Coord. Chem. Rev. (2014), http://dx.doi.org/10.1016/j.ccr.2014.09.020
G Model
CCR-111936; No. of Pages 15 ARTICLE IN PRESS
12 J.B. Waters, J.M. Goicoechea / Coordination Chemistry Reviews xxx (2014) xxx–xxx

Dipp
Dipp
N M
N N" Dipp
THF N
2K K K[M{N(SiMe3)2}3]
M N
N
N" N" Dipp N
Dipp
Dipp
60a/c 61

Fig. 32. Schlenk equilibrium of 60 to afford 61 [118].

Fig. 33. Reactivity of 35 towards dichlorosilanes [127]. Fig. 34. Reactivity of 35 towards carbon dioxide [128].

pictured in Fig. 31 (60). The possibility of the use of a sequestering


agent such as 18-crown-6 or 2,2,2-crypt allows for the “normal” still expected to be exothermic despite the electron withdrawing
carbene site of the anionic NHC to remain free for further coordi- carboxylate group on the backbone which decreases the nucle-
nation. ophilicity of the C2 position.
Each complex bears one ditopic carbanionic carbene ligand
bound to the metal centre through the C4 position and the C2 4.3.6. Related compounds
carbene remains uncoordinated. The complexes have a trigonal 1,2,3-Triazol-5-ylidenes are isoelectronic with “abnormal”
pyramidal geometry on account of the stereogenic lone-pair on the imidazol-4-ylidenes, both are known as mesoionic carbenes (MICs)
metal centre. Only 60b is observed to be stable in solution, 60a due to the impossibility of representing them with a charge neu-
and 60c were observed to decompose via a Schlenk type equilib- tral Lewis structure. 1,2,3-Triazol-5-ylidenes have themselves been
rium to generate the bis-ditopic carbanionic carbene complexes studied as the free MIC and as ligands in metal complexes by the
[M{C(CH)[N(2,6-i Pr2 C6 H3 )]2 C:}2 {N(SiMe3 )2 }]− (61) along with Bertrand and Albrecht groups [129,130]. Recently, Bertrand and
the homoleptic metal tris-amide complexes [M{N(SiMe3 )2 }3 ]− co-workers have furthered this field of research by investigating
(Fig. 32). 61 can be synthesized in a higher yield by using two whether the remaining C4 proton could be deprotonated and sub-
equivalents of 47. sequent metal coordination to form ditopic carbanionic carbenes
There is a large degree of steric congestion around the metal (Fig. 35) which are isoelectronic with the NHC derived carbanionic
centres in 61 as evidenced by the number of NMR resonances which carbenes of type B (Fig. 1, II).
indicate restricted rotation of the N Dipp and M N(SiMe3 )2 bonds. The starting point to obtain the carbanionic complexes was ini-
The M C4 bond lengths when M = Ge are 2.067(2) and 2.038(2) Å tial coordination of the free MIC 65 with either [(allyl)PdCl]2 or CuX
(c.f. 2.059(2) Å for 60a) are shorter than those when M = Pb (2.362(5) (X = Cl or I) yielding 66 and 67 (Fig. 36) [131]. Both could be depro-
and 2.339(5) Å) in line with the difference in covalent radii of the tonated to generate the ditopic carbanionic carbene which, upon
metals. loss of the potassium halide, would result in the formation of the
C4-functionalization of an NHC by reaction of two equivalents dimeric structure 68 and the oligomeric/polymeric structure 69,
of 35 with the dichlorosilanes Ph2 SiCl2 and Me2 SiCl2 resulted in respectively. Both complexes exhibit ditopic carbanionic carbene
the formation of di-NHCs containing a bridging silane moiety (62; ligands bound to the palladium or copper atoms through both the
Fig. 33) [127]. Substitution of one chloride ligand with the carban- C4 and C5 carbons simultaneously. While 68 could be crystallo-
ionic NHC could be achieved by employing just one equivalent of graphically characterized, the oligomeric structure of 69 could be
35. The availability of the C2 carbene positions for further reactivity
was demonstrated by coordination to CuCl.

4.3.5. Reactivity of ditopic carbanionic carbenes towards carbon


dioxide
By passing a stream of dry CO2 through a solution of the anionic
lithium salt of IPr (35) or 1,3-bis(diisopropyl)-imidazol-2-ylidene
(Ii Pr; 63) allowed for the selective formation of the C4 bound car-
boxylate product (64) which could be isolated as the anionic NHC
or as the protonated imidazolium zwitterion (Fig. 34) [128]. The
formation of the dicarboxylated adduct was not observed despite Fig. 35. Isolobal analogy between anionic imidazol-4,5-diylidenes and anionic
DFT calculations indicating that carboxylation of the C2 position is 1,2,3-triazol-4,5-diylidenes.

Please cite this article in press as: J.B. Waters, J.M. Goicoechea, Coord. Chem. Rev. (2014), http://dx.doi.org/10.1016/j.ccr.2014.09.020
G Model
CCR-111936; No. of Pages 15 ARTICLE IN PRESS
J.B. Waters, J.M. Goicoechea / Coordination Chemistry Reviews xxx (2014) xxx–xxx 13

Fig. 36. Reactivity of ditopic neutral 1,2,3-triazol-5-ylidenes towards transition metals [131].

N
Mes Mes
N N
N
Mes Mes
N N
[Rh(CO)2Cl]2 [Rh(COD)Cl]2
70, M = Rh OC M M
CO
69 72
71, M = Ir OC CO or [Ir(CO)2Cl]n
Rh Rh
COD COD
N N Cl
Mes Mes
N

Fig. 37. Transmetallation reactions involving ditopic anionic 1,2,3-triazol-4,5-diylidenes [131].

determined by ESI-MS, along with NMR spectroscopy which reveals References


only one set of resonances for the mesityl groups.
69 could be treated with [(allyl)PdCl]2 to form 68 in 60% yield [1] D. Bourissou, O. Guerret, F.P. Gabbaï, G. Bertrand, Chem. Rev. 100 (2000)
and has the same boat conformation by X-ray crystallography as the 39–91.
[2] G. Bertrand (Ed.), Carbene Chemistry: From Fleeting Intermediates to Power-
analogous pyrazolate complex [Pd(␮-pz)(␩3 -allyl)]2 [132]. Addi- ful Reagents, Marcel Dekker, New York, 2002.
tion of 69 to other transition metal reagents such as [Rh(CO)2 Cl]2 , [3] W.A. Herrmann, Angew. Chem. Int. Ed. 41 (2002) 1290–1309.
[Ir(CO)2 Cl]n and [Rh(COD)Cl]2 gave the novel dinuclear complexes [4] F. Glorius (Ed.), N-Heterocyclic Carbenes in Transition Metal Catalysis, Top-
ics in Organometallic Chemistry, vol. 21, Springer-Verlag, Berlin-Heidelberg,
70, 71, 72 (Fig. 37) [131]. The lower values for the IR CO stretch- Germany, 2006.
ing frequencies in 70 compared to dinuclear rhodium complexes [5] S.P. Nolan (Ed.), N-Heterocyclic Carbenes in Synthesis, Wiley-VCH, Weinheim,
of other anionic 1,2-dihapto ligands indicate the stronger electron Germany, 2006.
[6] S. Díez-González, S.P. Nolan, Coord. Chem. Rev. 251 (2007) 874–883.
donating ability of the anionic 1,2,3-triazol-4,5-diylidene. [7] F.E. Hahn, M.C. Jahnke, Angew. Chem. Int. Ed. 47 (2008) 3122–3172.
[8] H. Jacobsen, A. Correa, A. Poater, C. Costabile, L. Cavallo, Coord. Chem. Rev.
253 (2009) 687–703.
5. Conclusions [9] S. Díez-González, N. Marion, S.P. Nolan, Chem. Rev. 109 (2009) 3612–3676.
[10] L. Mercs, M. Albrecht, Chem. Soc. Rev. 39 (2010) 1903–1912.
Recent studies have shown that N-heterocyclic carbenes such [11] S. Díez-González (Ed.), N-Heterocyclic Carbenes: From Laboratory
Curiosities to Effective Synthetic Tools, RSC Publishing, Cambridge,
as imidazol-2-ylidenes can be deprotonated using a strong base to
2010.
afford carbanionic carbenes of general formulae [:C{NR}2 (CH)C]− . [12] C.S.J. Cazin (Ed.), N-Heterocyclic Carbenes in Transition Metal Catalysis and
This development has opened a new avenue of research in NHC Organocatalysis, Catalysis by Metal Complexes, vol. 32, Springer, Dordrecht,
chemistry. Carbene deprotonation, and the concomitant gener- The Netherlands, 2011.
[13] H.-W. Wanzlick, H.-J. Schönherr, Angew. Chem. Int. Ed. 7 (1968) 141–142.
ation of an alkali-metal salt, allows for functionalization of the [14] K. Öfele, J. Organomet. Chem. 12 (1968) P42–P43.
ligand backbone with a variety of Lewis acidic substrates, thus [15] A.J. Arduengo III, R.L. Harlow, M. Kline, J. Am. Chem. Soc. 113 (1991) 361–363.
facilitating the modification of the steric and electronic properties [16] D.J. Nelson, S.P. Nolan, Chem. Soc. Rev. 42 (2013) 6723–6753.
[17] Y. Wang, Y. Xie, P. Wei, R.B. King, H.F. Schaefer III, P.v.R. Schleyer, G.H. Robin-
of the parent carbene. The ditopic nature of such species also son, Science 321 (2008) 1069–1071.
allows for the synthesis of novel multi-metallic compounds and [18] A. Sidiropoulos, C. Jones, A. Stasch, S. Klein, G. Frenking, Angew. Chem. Int. Ed.
coordination polymers. The relative ease with which such carban- 48 (2009) 9701–9704.
[19] C. Jones, A. Sidiropoulos, N. Holzmann, G. Frenking, A. Stasch, Chem. Commun.
ionic carbenes can be accessed will indubitably generate many 48 (2012) 9855–9857.
more exciting findings in the area as such systems may now be [20] Y. Wang, Y. Xie, P. Wei, R.B. King, H.F. Schaefer III, P.v.R. Schleyer, G.H. Robin-
targeted systematically and are no longer laboratory curiosities son, J. Am. Chem. Soc. 130 (2008) 14970–14971.
[21] M.Y. Abraham, Y. Wang, Y. Xie, P. Wei, H.F. Schaefer III, P.v.R. Schleyer, G.H.
isolated as undesired side-products through chemical reduction Robinson, Chem. Eur. J. 16 (2010) 432–435.
or metal-mediated C H bond activation. [22] H. Braunschweig, R.D. Dewhurst, K. Hammond, J. Mies, K. Radacki, A. Vargas,
Science 336 (2012) 1420–1422.
[23] S. Gründemann, A. Kovacevic, M. Albrecht, J.W. Faller, R.H. Crabtree, Chem.
Acknowledgements Commun. (2001) 2274–2275.
[24] S. Gründemann, A. Kovacevic, M. Albrecht, J.W. Faller, R.H. Crabtree, J. Am.
Chem. Soc. 124 (2002) 10473–10481.
We thank Christ Church College, the EPSRC and the University
[25] L.N. Appelhans, D. Zuccaccia, A. Kovacevic, A.R. Chianese, J.R. Miecznikowski,
of Oxford for financial support of this research (DTA studentship A. Macchioni, E. Clot, O. Eisenstein, R.H. Crabtree, J. Am. Chem. Soc. 127 (2005)
J.B.W.). 16299–16311.

Please cite this article in press as: J.B. Waters, J.M. Goicoechea, Coord. Chem. Rev. (2014), http://dx.doi.org/10.1016/j.ccr.2014.09.020
G Model
CCR-111936; No. of Pages 15 ARTICLE IN PRESS
14 J.B. Waters, J.M. Goicoechea / Coordination Chemistry Reviews xxx (2014) xxx–xxx

[26] A. Kovacevic, K.R. Meadows, M. Counts, D.J. Arthur, Inorg. Chim. Acta 373 [75] C.Y. Tang, A.L. Thompson, S. Aldridge, J. Am. Chem. Soc. 132 (2010)
(2011) 259–261. 10578–10591.
[27] R.W. Alder, P.R. Allen, S.J. Williams, J. Chem. Soc. Chem. Commun. (1995) [76] J. Navarro, O. Torres, M. Martín, E. Sola, J. Am. Chem. Soc. 133 (2011)
1267–1268. 9738–9740.
[28] A.M. Magill, B.F. Yates, Aust. J. Chem. 57 (2004) 1205–1210. [77] B.R. Galan, M. Gembicky, P.M. Dominiak, J.B. Keister, S.T. Diver, J. Am. Chem.
[29] P.L. Arnold, S. Pearson, Coord. Chem. Rev. 251 (2007) 596–609. Soc. 127 (2005) 15702–15703.
[30] M. Albrecht, Chem. Commun. (2008) 3601–3610. [78] R.A. Diggle, S.A. Macgregor, M.K. Whittlesey, Organometallics 27 (2008)
[31] O. Schuster, L. Yang, H.G. Raubenheimer, M. Albrecht, Chem. Rev. 109 (2009) 617–625.
3445–3478. [79] T. Bolaño, M.L. Buil, M.A. Esteruelas, S. Izquierdo, R. Lalrempuia, M. Oliván, E.
[32] R.H. Crabtree, Coord. Chem. Rev. 257 (2013) 755–766. Oñate, Organometallics 29 (2010) 4517–4523.
[33] N. Hadei, E.A.B. Kantchev, C.J. O’Brien, M.G. Organ, Org. Lett. 7 (2005) [80] S. Caddick, F.G.N. Cloke, P.B. Hitchcock, A.K. de K. Lewis, Angew. Chem. Int.
3805–3807. Ed. 43 (2004) 5824–5827.
[34] M. Heckenroth, E. Kluser, A. Neels, M. Albrecht, Angew. Chem. Int. Ed. 46 [81] S. Burling, M.F. Mahon, R.E. Powell, M.K. Whittlesey, J.M.J. Williams, J. Am.
(2007) 6293–6296. Chem. Soc. 128 (2006) 13702–13703.
[35] H. Lebel, M.K. Janes, A.B. Charette, S.P. Nolan, J. Am. Chem. Soc. 126 (2004) [82] L.J.L. Häller, M.J. Page, S. Erhardt, S.A. Macgregor, M.F. Mahon, M.A. Naser, A.
5046–5047. Vélez, M.K. Whittlesey, J. Am. Chem. Soc. 132 (2010) 18408–18416.
[36] M. Heckenroth, E. Kluser, A. Neels, M. Albrecht, Dalton Trans. (2008) [83] C.E. Ellul, M.F. Mahon, O. Saker, M.K. Whittlesey, Angew. Chem. Int. Ed. 46
6242–6249. (2007) 6343–6345.
[37] M. Heckenroth, A. Neels, M.G. Garnier, P. Aebi, A.W. Ehlers, M. Albrecht, Chem. [84] M.R. Crittall, C.E. Ellul, M.F. Mahon, O. Saker, M.K. Whittlesey, Dalton Trans.
Eur. J. 15 (2009) 9375–9386. (2008) 4209–4211.
[38] V. Khlebnikov, M. Heckenroth, H. Müller-Bunz, M. Albrecht, Dalton Trans. 42 [85] U.J. Scheele, S. Dechert, F. Meyer, Chem. Eur. J. 14 (2008) 5112–5115.
(2013) 4197–4207. [86] A. Krüger, E. Kluser, H. Müller-Bunz, A. Neels, M. Albrecht, Eur. J. Inorg. Chem.
[39] M.S. Varonka, T.H. Warren, Organometallics 29 (2010) 717–720. (2012) 1394–1402.
[40] A.A. Danopoulos, N. Tsoureas, J.A. Wright, M.E. Light, Organometallics 23 [87] P.L. Arnold, S.T. Liddle, Organometallics 25 (2006) 1485–1491.
(2004) 166–168. [88] P.L. Arnold, M. Rodden, C. Wilson, Chem. Commun. (2005) 1743–1745.
[41] B.M. Day, T. Pugh, D. Hendriks, C.F. Guerra, D.J. Evans, F.M. Bickelhaupt, R.A. [89] Y. Wang, Y. Xie, M.Y. Abraham, R.J. Gilliard Jr., P. Wei, H.F. Schaefer III, P.v.R.
Layfield, J. Am. Chem. Soc. 135 (2013) 13338–13341. Schleyer, G.H. Robinson, Organometallics 29 (2010) 4778–4780.
[42] M. Alcarazo, S.J. Roseblade, A.R. Cowley, R. Fernández, J.M. Brown, J.M. Las- [90] K. Hansen, T. Szilvási, B. Blom, S. Inoue, J. Epping, M. Driess, J. Am. Chem. Soc.
saletta, J. Am. Chem. Soc. 127 (2005) 3290–3291. 135 (2013) 11795–11798.
[43] G. Song, Y. Zhang, X. Li, Organometallics 27 (2008) 1936–1943. [91] A.M. Tondreau, Z. Benkő, J.R. Harmer, H. Grützmacher, Chem. Sci. 5 (2014)
[44] A. John, M.M. Shaikh, P. Ghosh, Dalton Trans. (2009) 10581–10591. 1545–1554.
[45] A.J. Arduengo III, F. Davidson, H.V.R. Dias, J.R. Goerlich, D. Khasnis, W.J. Mar- [92] R.A. Musgrave, R.S.P. Turbervill, M. Irwin, J.M. Goicoechea, Angew. Chem. Int.
shall, T.K. Prakasha, J. Am. Chem. Soc. 119 (1997) 12742–12749. Ed. 51 (2012) 10832–10835.
[46] D. Mendoza-Espinosa, B. Donnadieu, G. Bertrand, J. Am. Chem. Soc. 132 (2010) [93] J.B. Waters, R.S.P. Turbervill, J.M. Goicoechea, Organometallics 32 (2013)
7264–7265. 5190–5200.
[47] E. Aldeco-Perez, A.J. Rosenthal, B. Donnadieu, P. Parameswaran, G. Frenking, [94] R.A. Musgrave, R.S.P. Turbervill, M. Irwin, R. Herchel, J.M. Goicoechea, Dalton
G. Bertrand, Science 326 (2009) 556–559. Trans. 43 (2014) 4335–4344.
[48] V. Lavallo, A. El-Batta, G. Bertrand, R.H. Grubbs, Angew. Chem. Int. Ed. 50 [95] D.P. Curran, A. Solovyev, M. Makhlouf Brahmi, L. Fensterbank, M. Malacria, E.
(2011) 268–271. Lacôte, Angew. Chem. Int. Ed. 50 (2011) 10294–10317.
[49] V. Lavallo, R.H. Grubbs, Science 326 (2009) 559–562. [96] Y. Wang, B. Quillian, P. Wei, C.S. Wannere, Y. Xie, R.B. King, H.F. Schaefer III,
[50] A.P. Singh, R.S. Ghadwal, H.W. Roesky, J.J. Holstein, B. Dittrich, J.-P. Demers, P.v.R. Schleyer, G.H. Robinson, J. Am. Chem. Soc. 129 (2007) 12412–12413.
V. Chevelkov, A. Lange, Chem. Commun. 48 (2012) 7574–7576. [97] P. Bissinger, H. Braunschweig, T. Kupfer, K. Radacki, Organometallics 29
[51] A.P. Singh, P.P. Samuel, K.C. Mondal, H.W. Roesky, N.S. Sidhu, B. Dittrich, (2010) 3987–3990.
Organometallics 32 (2013) 354–357. [98] A. Solovyev, S.-H. Ueng, J. Monot, L. Fensterbank, M. Malacria, E. Lacôte, D.P.
[52] O. Guerret, S. Solé, H. Gornitzka, M. Teichert, G. Trinquier, G. Bertrand, J. Am. Curran, Org. Lett. 12 (2010) 2998–3001.
Chem. Soc. 119 (1997) 6668–6669. [99] A. Adolf, U. Vogel, M. Zabel, A.Y. Timoshkin, M. Scheer, Eur. J. Inorg. Chem.
[53] O. Guerret, S. Solé, H. Gornitzka, G. Trinquier, G. Bertrand, J. Organomet. Chem. (2008) 3482–3492.
600 (2000) 112–117. [100] N. Kuhn, G. Henkel, T. Kratz, J. Kreutzberg, R. Boese, A.H. Maulitz, Chem. Ber.
[54] J. Huang, E.D. Stevens, S.P. Nolan, Organometallics 19 (2000) 1194–1197. 126 (1993) 2041–2045.
[55] R.F.R. Jazzar, S.A. Macgregor, M.F. Mahon, S.P. Richards, M.K. Whittlesey, J. [101] T. Ramnial, H. Jong, I.D. McKenzie, M. Jennings, J.A.C. Clyburne, Chem. Com-
Am. Chem. Soc. 124 (2002) 4944–4945. mun. (2003) 1722–1723.
[56] S. Burling, M.F. Mahon, B.M. Paine, M.K. Whittlesey, J.M.J. Williams, [102] M. Makhlouf Brahmi, J. Monot, M. Desage-El Murr, D.P. Curran, L. Fensterbank,
Organometallics 23 (2004) 4537–4539. E. Lacôte, M. Malacria, J. Org. Chem. 75 (2011) 6983–6985.
[57] K. Abdur-Rashid, T. Fedorkiw, A.J. Lough, R.H. Morris, Organometallics 23 [103] W.-C. Shih, C.-H. Wang, Y.-T. Chang, G.P.A. Yap, T.-G. Ong, Organometallics 28
(2004) 86–94. (2009) 1060–1067.
[58] N.M. Scott, R. Dorta, E.D. Stevens, A. Correa, L. Cavallo, S.P. Nolan, J. Am. Chem. [104] A.R. Kennedy, R.E. Mulvey, S.D. Robertson, Dalton Trans. 39 (2010)
Soc. 127 (2005) 3516–3526. 9091–9099.
[59] R. Cariou, C. Fischmeister, L. Toupet, P.H. Dixneuf, Organometallics 25 (2006) [105] X.-W. Li, J. Su, G.H. Robinson, Chem. Commun. (1996) 2683–2684.
2126–2128. [106] A.-L. Schmitt, G. Schnee, R. Welter, S. Dagorne, Chem. Commun. 46 (2010)
[60] R. Corberán, M. Sanaú, E. Peris, J. Am. Chem. Soc. 128 (2006) 3974–3979. 2480–2482.
[61] S.H. Hong, A. Chlenov, M.W. Day, R.H. Grubbs, Angew. Chem. Int. Ed. 46 (2007) [107] H.A. Duong, T.N. Tekavec, A.M. Arif, J. Louie, Chem. Commun. (2004) 112–113.
5148–5151. [108] N. Kuhn, M. Steimann, G. Weyers, Z. Naturforsch. 54b (1999) 427–433.
[62] K. Vehlow, S. Gessler, S. Blechert, Angew. Chem. Int. Ed. 46 (2007) 8082–8085. [109] J.D. Holbrey, W.M. Reichert, I. Tkatchenko, E. Bouajila, O. Walter, I. Tommasi,
[63] S. Burling, B.M. Paine, D. Nama, V.S. Brown, M.F. Mahon, T.J. Prior, P.S. Pregosin, R.D. Rogers, Chem. Commun. (2003) 28–29.
M.K. Whittlesey, J.M.J. Williams, J. Am. Chem. Soc. 129 (2007) 1987–1995. [110] B.R. Van Ausdall, N.F. Poth, V.A. Kincaid, A.M. Arif, J. Louie, J. Org. Chem. 76
[64] C.E. Cooke, M.C. Jennings, R.K. Pomeroy, J.A.C. Clyburne, Organometallics 26 (2011) 8413–8420.
(2007) 6059–6062. [111] B.R. Van Ausdall, J.L. Glass, K.M. Wiggins, A.M. Aarif, J. Louie, J. Org. Chem. 74
[65] A.A. Danopoulos, D. Pugh, J.A. Wright, Angew. Chem. Int. Ed. 47 (2008) (2009) 7935–7942.
9765–9767. [112] B. Bantu, G.M. Pawar, K. Wurst, U. Decker, A.M. Schmidt, M.R. Buchmeiser,
[66] Y. Ohki, T. Hatanaka, K. Tatsumi, J. Am. Chem. Soc. 130 (2008) 17174–17186. Eur. J. Inorg. Chem. (2009) 1970–1976.
[67] E.M. Leitao, S.R. Dubberley, W.E. Piers, Q. Wu, R. McDonald, Chem. Eur. J. 14 [113] Y. Wang, Y. Xie, M.Y. Abraham, P. Wei, H.F. Schaefer III, P.v.R. Schleyer, G.H.
(2008) 11565–11572. Robinson, J. Am. Chem. Soc. 132 (2010) 14370–14372.
[68] S. Burling, E. Mas-Marzá, J.E.V. Valpuesta, M.F. Mahon, M.K. Whittlesey, [114] A. El-Hellani, V. Lavallo, Angew. Chem. Int. Ed. 53 (2014) 4489–4493.
Organometallics 28 (2009) 6676–6686. [115] Y. Wang, Y. Xie, M.Y. Abraham, R.J. Gilliard Jr., P. Wei, C.F. Campana, H.F.
[69] O. Torres, M. Martín, E. Sola, Organometallics 28 (2009) 863–870. Schaefer III, P.v.R. Schleyer, G.H. Robinson, Angew. Chem. Int. Ed. 51 (2012)
[70] C. Zhang, Y. Zhao, B. Li, H. Song, S. Xu, B. Wang, Dalton Trans. (2009) 10173–10176.
5182–5189. [116] D.R. Armstrong, S.E. Baillie, V.L. Blair, N.G. Chabloz, J. Diez, J. Garcia-Alvarez,
[71] L. Vieille-Petit, X. Luan, M. Gatti, S. Blumentritt, A. Linden, H. Clavier, S.P. Nolan, A.R. Kennedy, S.D. Robertson, E. Hevia, Chem. Sci. 4 (2013) 4259–4266.
R. Dorta, Chem. Commun. (2009) 3783–3785. [117] S. Kronig, E. Theuergarten, C.G. Daniliuc, P.G. Jones, M. Tamm, Angew. Chem.
[72] A.E.W. Ledger, M.F. Mahon, M.K. Whittlesey, J.M.J. Williams, Dalton Trans. Int. Ed. 51 (2012) 3240–3244.
(2009) 6941–6947. [118] J.B. Waters, J.M. Goicoechea, Dalton Trans. 43 (2014) 14239–14248.
[73] C. Zhang, F. Luo, B. Cheng, B. Li, H. Song, S. Xu, B. Wang, Dalton Trans. (2009) [119] M. Chen, Y. Wang, R.J. Gilliard Jr., P. Wei, N.A. Schwartz, G.H. Robinson, Dalton
7230–7235. Trans. 43 (2014) 14211–14214.
[74] C.Y. Tang, W. Smith, A.L. Thompson, D. Vidovic, S. Aldridge, Angew. Chem. Int. [120] A. Jana, R. Azhakar, G. Tavčar, H.W. Roesky, I. Objartel, D. Stalke, Eur. J. Inorg.
Ed. 50 (2011) 1359–1362. Chem. (2011) 3686–3689.

Please cite this article in press as: J.B. Waters, J.M. Goicoechea, Coord. Chem. Rev. (2014), http://dx.doi.org/10.1016/j.ccr.2014.09.020
G Model
CCR-111936; No. of Pages 15 ARTICLE IN PRESS
J.B. Waters, J.M. Goicoechea / Coordination Chemistry Reviews xxx (2014) xxx–xxx 15

[121] Y. Wang, Y. Xie, M.Y. Abraham, P. Wei, H.F. Schaefer III, P.v.R. Schleyer, G.H. [128] M. Vogt, C. Wu, A.G. Oliver, C.J. Meyer, W.F. Schneider, B.L. Ashfeld, Chem.
Robinson, Organometallics 30 (2011) 1303–1306. Commun. 49 (2013) 11527–11529.
[122] Y. Wang, M.Y. Abraham, R.J. Gilliard Jr., P. Wei, J.C. Smith, G.H. Robinson, [129] G. Guisado-Barrios, J. Bouffard, B. Donnadieu, G. Bertrand, Angew. Chem. Int.
Organometallics 31 (2012) 791–793. Ed. 49 (2010) 4759–4762.
[123] P.A. Chase, D.W. Stephan, Angew. Chem. Int. Ed. 47 (2008) 7433–7437. [130] P. Mathew, A. Neels, M. Albrecht, J. Am. Chem. Soc. 130 (2008)
[124] D. Holschumacher, T. Bannenberg, C.G. Hrib, P.G. Jones, M. Tamm, Angew. 13534–13535.
Chem. Int. Ed. 47 (2008) 7428–7432. [131] X. Yan, J. Bouffard, G. Guisado-Barrios, B. Donnadieu, G. Bertrand, Chem. Eur.
[125] C. Pranckevicius, D.W. Stephan, Chem. Eur. J. 20 (2014) 6597–6602. J. 18 (2012) 14627–14631.
[126] S. Zlatogorsky, M.J. Ingleson, Dalton Trans. 41 (2012) 2685–2693. [132] A. Singhal, R. Mishra, S.K. Kulshreshtha, P.V. Bernhardt, E.R.T. Tiekink, J.
[127] R.S. Ghadwal, S.O. Reichmann, E. Carl, R. Herbst-Irmer, Dalton Trans. 43 (2014) Organomet. Chem. 691 (2006) 1402–1410.
13704–13710.

Please cite this article in press as: J.B. Waters, J.M. Goicoechea, Coord. Chem. Rev. (2014), http://dx.doi.org/10.1016/j.ccr.2014.09.020

You might also like