You are on page 1of 41

Coordination Chemistry Reviews 276 (2014) 112–152

Contents lists available at ScienceDirect

Coordination Chemistry Reviews


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

Review

Ruthenium pincer complexes: Ligand design and complex synthesis


Hussein A. Younus a,b,d , Nazir Ahmad a,b , Wei Su a,b , Francis Verpoort a,b,c,e,∗
a
Laboratory of Organometallics, Catalysis and Ordered Materials, State Key Laboratory of Advanced Technology for Materials Synthesis and Processing,
Center for Chemical and Material Engineering, Wuhan University of Technology, Wuhan, China
b
Department of Applied Chemistry, Faculty of Sciences, Wuhan University of Technology, Wuhan 430070, China
c
Tomsk Polytechnic University, Lenin Avenue 30, 634050 Tomsk, Russian Federation
d
Chemistry Department, Faculty of Science, Fayoum University, Fayoum 63514, Egypt
e
Ghent University, Global Campus Songdo, 119 Songdomunhwa-Ro, Yeonsu-Gu, Incheon, South Korea

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 113
1.1. General considerations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 113
1.2. Nomenclature of pincer ligands . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 113
2. Synthesis of pincer ligands . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 114
2.1. Tuning of the pincer ligand . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 114
2.2. Pyridine-based pincer ligands . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 114
2.2.1. Phosphine-containing pincer ligands . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 114
2.2.2. NHC-containing pincer ligands . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 118
2.2.3. NNN pincer ligands . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 119
2.2.4. CNN pincer ligands . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 120
2.3. Benzene-based pincer ligands . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 123
2.3.1. Symmetrical pincer ligands . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 124
2.4. Miscellaneous pincer ligands . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 124
3. Synthesis of ruthenium pincer complexes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 127
3.1. Direct metallation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 127
3.2. C H bond activation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 133
3.3. Si H bond activation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 139
3.4. Transmetallation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 139
3.5. Trans-cyclometallation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 142
4. Effects of the pincer complex structure on the catalytic activity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 142
4.1. Hemilability and non-innocent behavior of the pincer ligand . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 142
4.2. Flexibility of the pincer framework (aliphatic versus aromatic) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 146
5. Conclusion and future prospects . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 149
Acknowledgments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 149
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 149

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

Article history: Active transition–metal complexes based on relatively inexpensive metals are considered to be a desirable
Received 6 March 2014 method for sustainable human industrial growth. Considering their cost efficiency, ruthenium complexes
Received in revised form 16 June 2014 are gaining increasing attraction, instead of palladium, rhodium, and iridium. Among the ruthenium
Accepted 18 June 2014
complexes, ruthenium pincer complexes (RPCs) have received much attention due to their outstanding
Available online 26 June 2014
performance. Various strategies have been developed for pincer ligand design and RPC synthesis, which
indicate that ligand design is a key feature of pincer chemistry. In addition, electronic and steric effects,
Keywords:
C H activation
Direct hydrogenation

∗ Corresponding author at: Ghent University Global Campus Songdo, 119 Songdomunhwa-Ro, Yeonsu-Gu, Incheon, South Korea.
E-mail addresses: Francis@whut.edu.cn, Francis.verpoort@ugent.be (F. Verpoort).

http://dx.doi.org/10.1016/j.ccr.2014.06.016
0010-8545/© 2014 Elsevier B.V. All rights reserved.
H.A. Younus et al. / Coordination Chemistry Reviews 276 (2014) 112–152 113

Direct metallation hemilability, non-innocent behavior, and the flexibility of the pincer ligand have significant effects on the
Pincer ligand catalytic performance of RPCs in hydrogenation and dehydrogenation reactions.
Ruthenium pincer complex
Transfer hydrogenation
Trans-cyclometallation
© 2014 Elsevier B.V. All rights reserved.
Trans-metallation

1. Introduction well as functional group tolerance. Excellent reviews have dis-


cussed the synthesis and catalytic applications of palladium [14]
1.1. General considerations and iridium [16] pincers, but only a few catalytic applications of
ruthenium pincers [8,12,17,34] have been reviewed, and no pre-
In the field of homogeneous catalysis, the choice of an appropri- vious reviews have addressed the design and synthesis of RPCs.
ate ligand is very important and critical for fine-tuning the catalytic Therefore, this review presents the different methods developed
activities and stereoselectivities of metal complexes. Steric and for the design and synthesis of RPCs over the last decade.
electronic properties can greatly influence the nature of the reac-
tive species and determine the course of reactions. Pincer-based
metal catalysts are exceptional in determining the balance of sta- 1.2. Nomenclature of pincer ligands
bility versus reactivity. This balance can be controlled by systematic
ligand modifications and/or variation of the metal center, thereby The nomenclature of the pincer ligand indicates the attachment
allowing enhancement of the metal complex reactivity, stability, points between the transition metal and the three atoms that coor-
and reaction selectivity. Therefore, developing a tridentate pincer dinate with the metal center, the hetero/C-atoms of the side arms,
ligand is one of the most fruitful strategies for obtaining well- and the central atom. If there are two spacers between the central
defined metal–ligand bonds. moiety (pyridine, benzene, etc.) and the donor atoms of the side
“Pincer ligands” [1] are tridentate ligands that bind tightly arms are C-atoms, or if there are no spacers, it will not be included
to three adjacent coplanar sites of a metal center, mostly in a in the symbol of the pincer ligand. However, if there are spacers
meridional fashion [2], thereby resulting in the formation of two and one or both of them are non-carbon atoms, the pincer ligand
stable cyclometallated rings. The two cyclometallated rings may be name indicates the position and type of the spacer as a superscript
five-membered [3], six-membered [4], or hybrids of five- and six- to the original attachment points. For example, pincer ligands of
membered rings [5]. Since the pioneering work reported by Shaw the type PNP with N/NH spacers between N and P are denoted as
[6] and Van Koten and Noltes [7] in the 1970s using the so-called PN NN P, and as PO NO P in the same manner. If the pincer ligand has
PCP and NCN type metal complexes with pincer ligands, pincer- a CH spacer on one side and an N/NH spacer on the other side, it
type complexes have occupied very important roles in chemistry will be symbolized as PN NC P. In addition, the alkyl or the aryl group
and chemistry-related disciplines. A wide variety of pincer struc- attached to the donor atom is included in the name as superscript
tures have been designed with various transition metals as well for the donor atom, e.g., PNPiPr and PNPPh . When a N-heterocyclic
as different ancillary ligands. Therefore, the range of pincer metal carbene is one of the donor groups, the symbol of the pincer ligand
complexes is extremely broad and increasing continuously (Fig. 1) indicates that the donor carbon atom is part of the N-heterocyclic
[3,8–23]. carbene, e.g., (NHC)CNN shows that the first C-atom is part of the
Ruthenium complexes exhibit diverse beneficial characteristics carbene. In this review, this is indicated as a superscript for the C-
including high electron transfer ability [24], high coordination abil- atom so the symbol is CNHC NN to confirm with the rules for the
ity to hetero-atoms, low redox potentials [25], Lewis acid activity nomenclature of whole pincer ligands.
[26], and unique reactivity with metal species and intermediates,
such as oxo-metals, metallacycles, carbenes, and Schiff base com-
plexes [27–32]. Consequently, ruthenium catalysis is one of the
most important catalytic tools in organic synthesis [33]. In the
last year, ruthenium pincer complexes (RPCs) represented around
13% of the total research conducted in the pincer complex area
(Fig. 2). Large numbers of novel and useful reactions have been
developed using stoichiometric and catalytic amounts of ruthe-
nium complexes. Compared with traditional ruthenium catalysts,
pincer complexes often offer higher efficiency and selectivity, as

Fig. 2. Percentage of publications in 2013 related to pincer complexes with different


Fig. 1. Publications related to pincer complexes since 2004. metal ions.
Source: Sci Finder. Source: Sci Finder.
114 H.A. Younus et al. / Coordination Chemistry Reviews 276 (2014) 112–152

methylene groups ( CH2 ), amines ( NR ), or oxygen atoms


( O ). The spacer variation can generate chiral pincers, thereby
modifying the spacers to control the enantioselectivity of the cat-
alyst. Furthermore, changing the length of the spacer has a direct
Z = C, N, S, Si, P, B, etc. effect on the coordination pocket of the pincer ligand and on its
spatial behavior during coordination. A further electronic influence
X = CH2, NH2(R), O, CO, CS, etc.
may be exerted by alkyl or aryl groups attached to the aromatic sys-
E = NR2, PR2, P(OR)2, OR, OPR2, OP(OR)2, SR, SeR, AsR, SiR, etc. tem, thus remote electronic modifications can be achieved by vary-
R = H, electron donating and electron withdrawing groups ing the R group anchored to the backbone of the ligand. In addition,
the nature of the central donor has a direct impact on the overall
Fig. 3. General pincer ligand structure. properties of the metal center. Replacements of the central carban-
ion with isoelectronic nitrogen [56,57], silicon [58–61], phosphorus
2. Synthesis of pincer ligands [62,63], or boron anion [64–66] have also been achieved in suc-
cessful syntheses and explored further in recent years. Thus, both
2.1. Tuning of the pincer ligand neutral and anionic pincers can be obtained (Fig. 3).
Various pincer structures have been reported for ruthenium
Ligand design is an increasingly important component of syn- metal ions and different synthetic strategies are available. The final
thetic chemistry because of the subtle control that ligands exert decision when selecting a convenient procedure for synthesizing
on the metal center with which they are coordinated. Pincer lig- the pincer will depend on the mean skeleton required, the spacers
ands are tridentate ligands, which usually feature a central aromatic between the central donor atom (Z) and the donor groups (E), and
ring that is ortho,ortho-disubstituted with two electron-donor sub- mainly on the arms of the pincers (X).
stituents (E). During the last 40 years, the pincer-ligand platform
has established multifunctional building blocks, which have been 2.2. Pyridine-based pincer ligands
used successfully in a number of various applications, e.g., bond
activation [8], organic synthesis [35], homogeneous and heteroge- Pyridine-based pincer transition metal complexes are of par-
neous catalysis [36], polymer chemistry [37,38], photochemistry ticular importance due to their reactivity and flexibility for
[39–41], sensors [42,43], switches [44], and as biomarkers in medic- modification, thus most of the reported RPCs are based on the
inal chemistry [45]. The wide variety of applications of pincer type pyridine skeleton. Consequently, we discuss the different synthetic
complexes is a direct indication of one of the ligands’ most attrac- methods reported for the various types of pincers, such as PNP, PNN,
tive features, i.e., the number of options for tuning their electronic PNS, CNN, CNHC NN, CNH CNCNHC , and NNN, with different structures
and steric properties without affecting the ligand’s ability to bind that serve as ligands for ruthenium complexes. Different synthetic
the metal center. strategies have been used for the direct synthesis of tridentate lig-
The most common method used to synthesize RPCs involves ands with a central pyridyl fragment.
the introduction of the ruthenium source to a predesigned pincer
ligand. The synthesis is usually based on a straightforward reaction 2.2.1. Phosphine-containing pincer ligands
but it sometimes includes functional group conversion. However, 2.2.1.1. Symmetrical ligands (PNP, PN NN P, PO NO NP). Symmetrical
sophisticated ligand design is usually involved in pincer chemistry. PNP pincer ligands are easier to produce compared with non-
For example, unsymmetrical pincer complexes with two different symmetrical ones. In general, the reaction starts with a symmetrical
donor atoms (EZE ) have been shown to possess higher catalytic ligand by applying an instant modification of the two ligand
properties compared with their symmetrical analogs [46–48]. This sides. Fitting the two side arms of the pincer with a halide
high catalytic activity could be attributed to the hemilability of the atom followed by nucleophilic substitution using a phosphine
unsymmetrical pincer ligand with respect to the metal center of the source is the desirable method for building the PNP pincers.
complex. Even a slight change in the pincer ligand structure, such as Pyridine-based PNP pincer ligands with a methylene group as
replacing one of the donor atoms in the side arms of the ligand, has a spacer have been synthesized via direct nucleophilic sub-
a significant effect on the catalytic activity as well as the reaction stitution of 2,6-bis(chloro/bromo-methyl)pyridine using sodium
product [49,50]. In addition, asymmetric catalysis can be achieved diphenylphosphide [67]. The addition of a secondary phosphine
using chiral pincer ligands [51–54]. Moreover, the steric effect of to 2,6-bis(chloromethyl)pyridine followed by selective deproto-
the substituent on the chiral moiety clearly affects the enantiose- nation using triethylamine was also described by Milstein et al.,
lectivity of the product [55]. Therefore, careful pincer ligand design where this method is more general and it avoids potential issues
depends on the target catalytic reaction, the substrates used, and with unstable phosphide intermediates [68]. Rieger et al. described
the catalytic process conditions. the synthesis of a series of PNP pincers with different aromatic
Modification of the pincer ligand backbone can be achieved substituents at phosphane via the treatment of diarylphosphanes
in a variety of ways. The most common method is changing the with 2,6-bis(chloromethyl)-pyridine in the presence of a base,
ring size (five-, six-, or seven-membered ring), changing the ring such as potassium t-butoxide [69]. Kawatsura and Hartwig synthe-
nature (aromatic versus nonaromatic), or eliminating the ring itself sized the PNP pincer ligand via selective lithiation of 2,6-lutidine
to build an acyclic pincer ligand. The nature of the side arms also followed by reaction of the lithiated species with chlorophos-
affects the design of the pincer ligand by conferring specific elec- phine reagents [70]. The yield of the latter reaction is slightly
tronic and steric properties. The donor group (E) (Fig. 3) can be lower than the two-step addition using secondary phosphines,
a heteroatom, part of a heterocyclic group, or an NHC. Moreover, but the reagents are more readily available and less expensive.
identical donor groups can be used to build a symmetrical pin- Scriban and Glueck reported the platinum-catalyzed asymmet-
cer (EZE) or different groups to construct an unsymmetrical pincer ric phosphination of 2,6-bis(bromomethyl)pyridine with PHMeIs
(EZE ). The most common differences are: (i) soft vs. hard donors, (Is = 2,4,6-tris(isopropyl)phenyl) to generate the corresponding P-
(ii) the rigidity of their binding to the metal center, (iii) the steric stereogenic diphosphine 3MeIs with up to 72% ee [71]. Recently,
constraints on their substituents, and (iv) the ligand type (e.g., Toste et al. developed a ruthenium catalyst for the enantiose-
neutral, charged, Lewis acidic, or Lewis basic). The donor groups lective alkylation of chiral racemic secondary phosphines with
(E) are connected to the central backbone by spacers (X), such as 2,6-bis(chloromethyl)pyridine (Scheme 1).
H.A. Younus et al. / Coordination Chemistry Reviews 276 (2014) 112–152 115

Scheme 1. Synthesis of PNP pincer ligands.

For the PNP family of pincer ligands, the steric, electronic, and several phosphorus-containing species were formed instead,
stereochemical parameters can be manipulated via modifications including tetraphenyldiphosphine and tetraphenyldiphosphine
of the benzylic positions and/or the phosphino alkyl/aryl groups to monoxide. However, the synthesis of these ligands could be
control the reactivity at the metal center. However, stereochemical achieved via a two-step procedure: the addition of 1 equiv. of
parameters are comparatively difficult to modify and they often Ph2 PCl to N,N-dialkyl-2,6-diaminopyridine in the presence of n-
require multistep syntheses and expensive starting materials. In BuLi, followed by the addition of another equivalent of Ph2 PCl and
addition, pincer behavior investigations have indicated that the n-BuLi to yield the desired PNP ligands 9 (Scheme 3). The former
methylene groups on the PNP ligand play significant roles in the two examples of PNP pincer ligands with O/N spacers result in a
reactivity of metal complexes. These CH2 groups are susceptible to readily accessible and tunable pincer ligand.
deprotonation by an external base or the transition metal itself,
thereby resulting in dearomatization of the ligand [17]. Conse- 2.2.1.2. Unsymmetrical ligands (PNN, PNS). Modification of the pin-
quently, replacing the reactive -CH2 - spacer on PNP with an O- cer “arm” has a profound effect on the catalytic activity of the pincer
or N- linking unit in the PO NO P and PN NN P pincer type ligands, complex. The replacement of one phosphine group by a weaker
respectively, may avoid the possibility of ligand deprotonation. The donor may facilitate ligand dissociation from the metal center and
neutral PO NO P pincer ligand 7 was prepared in a direct manner by it may increase the hemilability of this type of pincer complex,
the treatment of 2,6-dihydroxypyridine hydrochloride with dialkyl thereby increasing its catalytic activities accordingly [81]. Thus,
chlorophosphine in the presence of a base (Scheme 2). Changing the mixed or hybrid pincer systems based on pyridine have also been
base has a significant effect on the reaction yield and reaction time, designed and developed for ruthenium complexes, which increase
hence an excess of triethylamine at 65 ◦ C in tetrahydrofuran (THF) the probability of potential “hemilability” in the newly designed
yielded the product at 57% after one week [72], whereas a mixture ligand. Designing and constructing unsymmetrical pincer ligands
of NNNN-Tetramethylethylenediamine (TEMDA) and NEt3 in the is a multistep process and this is sometimes not an easy task. Start-
same conditions yielded the product at 75% within 20 h [73]. This ing with a symmetrical ligand source and modification with two
ligand can be synthesized at a higher yield (85%) and purity (99%) different donor sites in two separate steps is one possible synthetic
by combining aspects of the two methods, i.e., using a mixture of route. An easier method utilizes an unsymmetrical ligand source to
TEMDA and NEt3 in THF at 75 ◦ C for one week [74]. build the donor sites. PNN pincer ligand 12 was synthesized from
Kirchner reported the synthesis of a series of modularly the commercially available 2,6-lutidine by refluxing with 1 equiv.
designed PNP ligands based on pyridine-bearing NH-groups as NBS, thereby resulting in the formation of 2-bromomethyl-
spacers, which are easily accessible for the modification of steric, 6-methylpyridine 10, which can easily undergo nucleophilic
electronic, and stereochemical properties [75–78]. This method substitution with diethylamine to obtain 6-diethylaminomethyl-
was first reported by Haupt et al. [79,80] and then developed 6-meth-ylpyridine 11 with a yield of 66%. This intermediate
by Kirchner for the synthesis of a large collection of PNP and was deprotonated with n-BuLi at 0 ◦ C and then reacted with
PCP pincers. The ligands were prepared by the treatment of 2,6- di-t-butylchlorophosphine at −78 ◦ C to form the ligand (2-(di-tert-
diaminopyridine with 2 equiv. of the respective R2 PCl compound butyl-phosphinomethyl)-6-diethylaminomethyl)pyridine (PNN)
in the presence of a base, e.g., NEt3 and/or n-BuLi. This method 12 with a yield of 70% [82,83] (Scheme 4).
failed in the case of N,N-dialkylated diaminopyridines [17], where Furthermore, exploration of the replacement of the amine arm
in the PNN pincer ligand with a thioether group was studied
by Milstein group [84]. The PNS pincer ligand was synthesized
using a procedure similar to that reported by Szabo for the prepa-
ration of a PCS ligand [85]. Thus, reacting LiPtBu2 (BH3 ) with
2,6-bis(chloromethyl)pyridine yielded a mixture of monophos-
phine, diphosphine, and unreacted 2,6-bis(chloromethyl)pyridine.
After isolating the monophosphine 13, treatment with an excess
of sodium 2-methyl-2-propanethiolate obtained the borane-
Scheme 2. Straightforward synthesis of PO NO P pincer ligands. protected PNS ligand 14 with a yield of 99%. Deprotection of the PNS
116 H.A. Younus et al. / Coordination Chemistry Reviews 276 (2014) 112–152

Scheme 3. Synthesis of PN NN P pincer ligands.

Scheme 4. Synthesis of the hybrid PNN pincer ligands.

Scheme 5. Synthesis of the hybrid PNS pincer ligand.

ligand by refluxing in diethylamine generates the PNS pincer ligand 48 h followed by abstraction of the liberated HCl with triethylamine
15 as an air-sensitive white solid with a yield of 66% (Scheme 5). at room temperature generated the Bpy-t-Bu PNN ligand 21 [68]
Unfortunately, it is clear that the process used to prepare the air (Scheme 6). In addition, synthesis of the PNN pincer ligand based
sensitive PNS pincer ligand generated a mixture of products in more on bipyridine has been extended to obtain more flexible PNN-type
than one step. pincer ruthenium complexes [90]. Methyl bipyridinemethane was
The 2,2-bipyridine ligand has been used extensively as a prepared by reacting 2,6-lutidine with 2-fluoropyridine and n-BuLi
metal-chelating ligand due to its robust stability and ease of func- in THF, where alkylation of the bridging methylene group was per-
tionalization, thus bipyridine is the most widely used ligand [86]. formed by deprotonation using n-BuLi with the addition of slight
Milstein et al. reported highly active bipyridine-based RPCs. The excess of iodomethane in THF or 1,4-diiodobutane, and finally the
electron-rich tridentate PNN ligand based on 2,2-bipyridine, BPy- phosphination of 24a,b was achieved using a slight excess of n-
PNN, has been synthesized in many ways. In most cases, the BuLi and ClPt Bu2 to produce the PNN pincer ligand 25a,b. Ligand
synthesis starts from 6-methyl-2,2 -bipyridine (BPy-Me) 18, which 28, which bears a bridging oxygen atom, was prepared using the
has been prepared via a Stille-type cross-coupling reaction by method reported by Buchwald [91] (Scheme 7).
refluxing 2-bromo-6-picoline 17 with 2-tributylstannyl-picolines Recently [92], Gusev et al. reported the straightforward synthe-
in toluene, and in the presence of Pd(PPh3 )4 as a catalyst [87,88]. sis of a PNN pyridine-based pincer ligand via the condensation of
Treatment of 18 with 1 equiv. of LDA followed by trapping with 2-picolyl aldehyde and 2-(alkyl/aryl phosphino)ethylamine in THF
TMSCl produced 6-(trimethylsilyl)methyl-2,2 -bipyridine 19. The at ambient conditions, followed by reduction of the imine group
former reacts with hexachloroethane to yield 6-chloromethyl-2,2 - using NaBH4 (Scheme 8). Chen et al. designed a non-symmetrical
bipyridine (BPyCH2 Cl) [89]. Treatment of BPyCH2 Cl with 1.2 equiv. pyridine-based pincer ligand bearing a phosphine using NH as a
of di-tert-butyl phosphine in MeOH in a sealed tube at 50 ◦ C for spacer on one side and an oxazoline ring as a relatively weaker
H.A. Younus et al. / Coordination Chemistry Reviews 276 (2014) 112–152 117

Scheme 6. Synthesis of the PNN pincer ligand based on bipyridine.

Scheme 7. Synthesis of the flexible PNN pincer ligand with different spacers.

Scheme 8. Straightforward synthesis of PNN pincer ligands.

Scheme 9. Synthesis of PN NC N pincer ligands bearing a hemilabile oxazoline ring.


118 H.A. Younus et al. / Coordination Chemistry Reviews 276 (2014) 112–152

Scheme 10. Synthesis of the symmetrical CNHC NCNHC pincer ligand.

donor on the other side, with the aim of increasing the hemilability 2,6-[(o-dialkyl)phenylimidazolium]pyridine dibromide 40a,b with
of the complex. The PN NC N pincer was prepared from commercially 2 equiv. of KHMDS produced the free carbene 41a,b with a yield
available 6-bromopicolinic acid, as shown in Scheme 9 [93–95]. of 70–80%, which is an off-white crystalline product that can
be handled as a solid under nitrogen at room temperature for
hours without exhibiting decomposition, and the structure of
2.2.2. NHC-containing pincer ligands
41a was confirmed by X-ray crystallography analysis [103]. The
NHC ligands are often considered to be phosphine mimics, but
molecule is strictly planar and it adopts a conformation where
they are generally more electron-rich than phosphines and are
the carbene and pyridine lone pairs are mutually anti, probably
more tightly bound to the metal. As a consequence, organometallic
to minimize lone pair repulsions. More conveniently, the sym-
complexes with N-heterocyclic ligands usually exhibit higher reac-
metrical CNHC NCNHC pincer ligand 45 with two methylene spacers
tivity, better stability, and a broader variety of catalytic activities
was obtained by combining 2 equiv. of substituted imidazoles with
compared with those involving phosphine ligands. Therefore, N-
the bis bromomethyl pyridine under reflux conditions in acetone
heterocyclic ligands have become more attractive in homogeneous
[104,105] (Scheme 11A). The bis triazole carbene-based CNHC NCNHC
catalysis and organic synthesis [96–99]. Incorporation of the NHC
pincer ligand is more readily accessible via ruthenium-catalyzed
moiety into the pincer backbone has allowed the synthesis of sev-
azide-alkyne cycloaddition (RuAAC) (Click reaction) [106]. Thus,
eral types of NHC-containing pincer-type ligands that are neutral
“clicking” diethynylpyridine and alkyl azide in dioxane produced
[100,101] and monoanionic [102], where the NHC moiety functions
bis(1,2,3-triazolylidene)pyridine ligand, where methylation using
as a lateral donor or as the backbone [10]. RPCs can incorporate NHC
trimethyloxonium tetrafluoroborate yielded the corresponding
ligands in one/two sides of the pincer and pincers have even been
symmetrical CNHC NCNHC pincer ligands 44 [107] (Scheme 10).
described with two different NHC ligands that coordinate to the
same Ru center.
2.2.2.2. CNHC NN pincer ligands. In general, two approaches are used
2.2.2.1. CNHC NCNHC pincer ligands. CNC ligands have a wide range to construct a CNHC NN pincer based on pyridine as a central moi-
of applications where the early transition metals support com- ety. The first approach introduces the imidazole moiety (source of
plexes with significant stability in a range of oxidation states, NHC) followed by alkylation (or N-substituted imidazole at once),
thereby providing convenient platforms for further derivatization. and then combining the ligand produced with the N-donor. In
The symmetrical CNHC NCNHC pincer ligands are easily acces- most cases, nucleophilic substitutions are used to introduce the N-
sible by the direct combination of the NHC source with the donor into the ligand structure, but the N-donor could be built in
bis-halogenated pyridine ring. Thus, heating 2,6-dibromopyridine the pincer ligand. The second approach starts with the NN donor,
with 2 equiv. of 1-alkyl imidazole in a sealed tube for one which is combined with the NHC source. Sanchez et al. reported
week at 140 ◦ C produced the pincer ligand. Treatment of the synthesis of unsymmetrical neutral N-heterocyclic carbene

Scheme 11. Synthesis of CNHC NCNHC and CNHC NN pincer ligands.


H.A. Younus et al. / Coordination Chemistry Reviews 276 (2014) 112–152 119

52 [110] (Scheme 13). The second method utilizes the commer-


cially available 2,6-dibromopyridine as a starting compound, which
has the capacity to synthesize multi-substituted pyrazole moi-
eties. The first step involves the coupling of 2,6-dibromopyridine
and pyrazole using the catalytic system Pd(OAc)2 /PPh3 [111].
Next, the heteroarylation of imidazole with 2-bromo-6-(pyrazol-
1-yl)pyridine followed by N-alkylation of the product with
1-iodoalkane generates the desired product 56 [112] (Scheme 13).

Scheme 12. Synthesis of the CNHC NN pincer ligand based on bipyridine.

2.2.3. NNN pincer ligands


pincer-type ligands CNHC NN for Ru(II) complexes [108]. The pin- The potential of tridentate NNN ligands such as Pybox
cer ligand with NHC as a hard donor and pyrrolidine as a soft [113], 2,6-bis(imino)pyridines [114], terpyridines [115], and other
donor was designed in the hope that the hemilability of the pyrro- symmetrical NNN ligands has been demonstrated in different appli-
lidine moiety might act as an internal base, possibly eliminating cations. The synthesis of this type of symmetrical pincer ligand
the need to add a base during catalysis. Starting from the sym- has been discussed previously and it is not repeated here. Sev-
metrical source, the unsymmetrical pincer was obtained first by eral 2,6-bis(pyrazol-1-yl)pyridine (BPP) derivatives of 58 have
introducing the N-substituted imidazole moiety, followed by ami- been synthesized as NNN pincer ligands by Jameson et al. [116],
nation of the alkyl halide arm. The CNHC NN ligand was prepared but the synthesis of the first neutral transition–metal complex
by treatment of 2,6-bis(bromomethyl)pyridine with a substoi- of these ligands was described only recently [117]. Symmetri-
chiometric amount of 1-aryl-1H-imidazole in refluxing acetone cal NNN pincer ligands of the bis(pyrazolyl)pyridine type can
to obtain the corresponding imidazolium salt 46. The monoalky- be prepared in a straightforward manner via the nucleophilic
lated salt undergoes a nucleophilic substitution with the secondary substitution of 2,6-dichloropyridine, 2,6-dibromopyridine, or 2,6-
amine in the presence of a base to produce the CNN pincer ligand bis(bromomethyl)pyridine with a pyrazole salt [118] (Fig. 4).
47 [104,105] (Scheme 11B). CNHC NN based on bipyridine has been Recently, Yu et al. reported the synthesis of a series of RPCs based
prepared in straightforward manner (Scheme 12) by refluxing on unsymmetrical NNN pincer ligands. The unsymmetrical pyridyl-
6-(chloromethyl)-2,2 -bipyridine 20 with 1-mesityl-1H-imidazole based bis-pyrazole pincer ligand 61 was prepared by lithiation of
38a in dry acetonitrile with a yield of 86% [109]. 2-bromo-6-(3,5-dimethylpyrazol-1-yl)pyridine 54b with n-BuLi at
Another approach has also been developed for the synthe- −78 ◦ C, followed by treatment with N,N-dimethylacetamide to pro-
sis of the CNHC NN pincer ligand with a pyrazole moiety as an duce the acetyl derivative 59a [119]. Claisen-Schmidt condensation
N-donor. A pyridyl-based (pyrazol-3-yl)-N-heterocyclic carbene of the acetyl derivative with N,N-dimethylformamide dimethy-
pincer ligand was synthesized via two different pathways. The first lacetal (DMFDMA) yielded the ␤-amino-substituted enone [120],
procedure exploits a ligand source with two different functional which after condensation with hydrazine hydrate produced the
groups in a two-step modification process. Methyl 6-bromo- unsymmetrical NNN pincer ligand 61 [121] (Scheme 14).
2-pyridinecarboxylate 49 reacts with imidazolatopotassium in The pyridyl-based pyrazolyl-imidazoly/benzoimidazolyl pincer
diglyme for 60 h, followed by Claisen condensation of the prod- ligand 62 was also prepared by formylation of 2-bromo-6-
uct with 3,3-dimethyl-2-butone, and subsequent reaction with (3,5-dimethylpyrazol-1-yl)pyridine using DMF/n-BuLi followed by
hydrazine hydrate generates 2-(imidazol-1-yl)-6-(5-tert-butyl- condensation with o-phenylenediamine [48,122]. Moreover, the
1H-pyrazol-3-yl)pyridine 51. When alkylated using 1-iodoalkane imidazolyl moiety can be converted easily into its N-methyl analog.
in refluxing acetonitrile, the former yields the CNHC NN pincer ligand It is worth mentioning that condensation of the carbonyl derivative

Scheme 13. Synthesis of CNHC NN without a spacer, which comprises pyrazole and imidazolium moieties.
120 H.A. Younus et al. / Coordination Chemistry Reviews 276 (2014) 112–152

Scheme 14. Synthesis of unsymmetrical NNN pincer ligands.

Fig. 5. CNN pincer ligand based on pyridine.

120 ◦ C [55]. In addition, the chiral mono-N-alkyl derivative and


the di-N-alkyl derivative of bis(benzimidazol-2-yl)-pyridine were
prepared by reacting bis(benzimidazol-2-yl)pyridine with 1 and
2 equiv. of 1-bromopropane in the presence of NaH [55].
The benzotriazolyl moiety has been utilized to construct
pyridyl-based benzimidazolyl–benzotriazolyl NNN pincer ligands.
Ligand synthesis begins by reacting 2,6-dibromopyridine with a
Fig. 4. Synthesis of symmetrical NNN pincer ligands. substoichiometric amount of benzotriazole in solvent-free con-
ditions to produce compound 74. Formylation of 74 followed
59 with primary amines (e.g. 4-toluidine) in the presence of p-TsOH by condensation with 1,2-phenylenediamine generates the NNN
in refluxing toluene yields the Schiff base, which after reduction ligand, which can undergo N-methylation using MeI/NaH to yield
with NaBH4 in methanol produces the NNN pincer ligand, i.e., the ligand 77, which bears a benzotriazole and a benzimidazolyl
2-(aminomethyl)-6-(3,5-dimethylpyrazol-1-yl)pyridine 64 [123] group [125] (Scheme 16).
(Scheme 14).
A chiral pyridyl-based pyrazolyl oxazolinyl pincer ligand 2.2.4. CNN pincer ligands
with a pyrazolyl NH functionality was designed and prepared Although NCN and PNP pincer ligands have been studied widely,
in a sequence of reactions, which also started with 2-bromo- the CNN motif has received much less attention. The first exam-
6-(3,5-dimethylpyrazol-1-yl), as shown in Scheme 15 [124]. ple of a CNN Ru pincer complex was reported very recently. The
Copper-catalyzed cyanation of the 2-bromopyridine derivative type A CNN tridentate pincer ligand was designed to contain the 2-
yielded compound 65, which reacted with alcohol to generate the aminomethylpyridine moiety (Fig. 5), which has a particularly high
imidate intermediate. Cyclization of the imidate with chiral amino ligand acceleration effect on the transfer hydrogenation of ketones
alcohol yielded the chiral NNN pincer ligand 66. Another chiral NNN by ruthenium(II) [126].
pincer ligand of the pyridyl-based pyrazolyl oxazolinyl pincer type The synthesis of a CNN pyridine based pincer ligand fre-
was synthesized in a similar manner by condensation of 2-acetyl- quently begins with 2-phenyl pyridine. The subsequent creation
6-bromopyridine with DMFDMA, followed by cyclization with of the N-donor at the 6-position or the modification of already
hydrazine hydrate to produce 2-(pyrazol-1-yl)-6-bromopyridine. present functional groups can produce the final pincer ligand. The
Further treatment of ligand 68 in a similar manner obtained same structure could also be assembled using the 6-substituted
ligand 69 [124]. Chiral pyridyl-based benzimidazolyl–oxazolyl NN ligand by incorporating a phenyl group. The type A CNN
NNN ligand 73 was synthesized using the same procedure, which pincer ligand can be prepared via different strategies using a
started by the esterification of compound 70, followed by reaction variety of starting compounds. The most recent method starts
with the chiral amino alcohol. Cyclization was the next step, which with ␣-cyanation of 2-arylpyridine-N-oxide with trimethylsilyl
can be performed in two ways: treatment with thionyl chloride in cyanide and dimethylcarbamyl chloride [127,128], followed by
dichloroethane at 70 ◦ C for 4 h followed by reflux with an ethanolic reduction of the cyanide group to produce 2-aminomethyl-
solution of sodium hydroxide overnight, or in a straightforward 6-aryl pyridine 81 [129]. The CNN pincer ligand with the
reaction by heating its ethereal solution with boron trifluoride at N,N-dimethylaminomethyl donor 82 can be prepared from 81 via
H.A. Younus et al. / Coordination Chemistry Reviews 276 (2014) 112–152 121

Scheme 15. Synthesis of unsymmetrical chiral NNN pincer ligands.

Scheme 16. Synthesis of pyridyl-based benzimidazolyl–benzotriazolyl NNN pincer ligands.

Scheme 17. Synthesis of achiral CNN pincer ligand based on pyridine.


122 H.A. Younus et al. / Coordination Chemistry Reviews 276 (2014) 112–152

[131] or 2,6-dibromopyridine [132] as starting materials. Chi-


ral pincers with the type A CNN structure have been prepared
by treatment of the corresponding ketones with hydroxylamine
hydrochloride and reduction of the oxime intermediate with
Zn/ammonium acetate/NH4 OH. Likewise, the optically active
ligand was obtained at a high yield via trifluoroacetic acid-mediated
hydrolysis of the (SS ,R)-N-(p-toluenesulfinyl)-2,2-dimethyl-1-(6-
phenylpyridin-2-yl)propylamine [129,133,134] (Scheme 18).
In a direct method, a CNN pincer ligand with structure B
that contained a pyrazole moiety as the N-donor was prepared
readily by palladium-catalyzed Suzuki coupling of 6-bromo-2-
(sub-pyrazol-1-yl)pyridine with p-tolylboronic acid with a yield of
86% [135] (Scheme 19).
Scheme 18. Synthesis of chiral CNN pincer ligand based on pyridine.
Baratta described the synthesis of new chiral CNN pincer lig-
ands using a chemo-enzymatic approach [136]. Candida antarctica
lipase B was used for the resolution of racemic 6-bromo-1-(2-
pyridyl)-ethanol, which was obtained by the sodium borohydride
reduction of 67. After hydrolysis of the acetyl derivative, the
homochiral alcohol was converted into the corresponding azide
with diphenylphosphoroazidate (DPPA) in the presence of DBU.
To avoid epimerization, the reaction was performed at 0 ◦ C, which
Scheme 19. Synthesis of type B CNN pincer ligand. resulted in a clean SN2 process that converted the alcohol (R)-90,
into the inverted azide (S)-91. Transformation of the azide into the
formic acid-formaldehyde methylation [130] (Scheme 17). This corresponding amine (S)-92 was achieved by treatment with Ph3 P
procedure is easier for synthesizing the type A CNN pincer and in refluxing THF/H2 O. Finally, Suzuki coupling with the appropriate
more straightforward compared with the other reported routes for substituted phenyl and naphthyl boronic acids produced the HCNN
synthesizing the same structure utilizing 6-bromopicolylaldehyde ligands (S)-93 (Scheme 20).

Scheme 20. Synthesis of chiral pincer CNN ligands using a chemo-enzymatic approach.

Scheme 21. Synthesis of chiral (A) and racemic chiral (B) CNN pincer ligands based on benzo[h]quinoline.
H.A. Younus et al. / Coordination Chemistry Reviews 276 (2014) 112–152 123

2-chlorobenzo[h]quinoline with bromotrimethylsilane to obtain 2-


bromobenzo[h]quinoline according to the Schlosser method [138].
The addition of n-butyl lithium at −78 ◦ C and DMA (R = Me), or
pivalonitrile (R = tBu), followed by reaction with NH2 OH·HCl in
ethanol resulted in the conversion of the ketones into the corre-
sponding oxime, where reduction generated the CNN pincer ligand
96b,c (Schemes 20B and 21).
Baratta et al. [139] described the application of stereo-
complementary biotransformations to the synthesis of both enan-
tiomers of 1-(6-phenylpyridin-2-yl)ethanamine 103 (Scheme 22).
The (S)-103 enantiomer was obtained via the lipase-mediated
kinetic resolution of the secondary alcohol (±)-101, followed by
Mitsunobu azidation, while baker’s yeast Saccharomyces cerevisiae
reduced the ketone derivative to give access to the enantiomeric
ligand (R)-103. This approach was then extended to the synthesis
of novel chiral 1-(benzo[h]quinolin-2-yl)ethanamine CNN pincer
ligands (R,S)-104.

2.3. Benzene-based pincer ligands

Pyridine-based pincers have been used as ligands for ruthenium


complexes, including many modifications in the mean skeleton,
donor atoms, and the spacers to generate the required properties,
Scheme 22. General strategy for the synthesis of chiral non-racemic pincer ligands.
but much less research has been performed using benzene-based
pincer ligands for ruthenium complexes. This may be related to
The synthesis of chiral and achiral CNN pincer ligands based the unusual metal–ligand interactions of the pyridine-based pin-
on benzo[h]quinoline has also been reported [137]. The reac- cer ligands. This new mode of metal–ligand cooperation based on
tion of benzo[h]quinoline N-oxide with trimethylsilylcyanide the aromatization–dearomatization of pyridine-based pincer-type
and dimethylcarbamyl chloride according to Fife’s procedure ligands has been demonstrated in the activation of many unre-
for the regioselective cyanation of pyridine 1-oxides produced active bonds. This type of metal–ligand cooperation plays a key
benzo[h]quinoline-2-carbonitrile 95. Catalytic hydrogenation at role in the recently discovered environmentally benign reactions
room temperature using a solution of benzo[h]quinoline-2- of alcohols, which are catalyzed by PNP and PNN pincer complexes
carbonitrile in acetic acid generated the 2-aminomethylbenzo[h]- of ruthenium, including hydrogenation, oxidative dehydrogena-
quinoline compound 96 with a yield of 90% (Scheme 20A). tion, dehydrogenative coupling reactions, and water splitting [15].
The substituted ligands 96b,c bearing a methyl and a tert- Unfortunately, this metal–ligand cooperation is not present in case
butyl group, respectively, were synthesized by treatment of of the benzene analog.

Scheme 23. (A) General approach for symmetrical PCP pincer ligand synthesis on the left and an alternative method for non-nucleophilic phosphorus centers on the right.
(B) General protocol for the synthesis of symmetrical NCN pincer ligands.

Scheme 24. Synthesis of NCN and CNHC CCNHC pincer ligand via the “Click” reaction.
124 H.A. Younus et al. / Coordination Chemistry Reviews 276 (2014) 112–152

Scheme 25. Synthesis of chiral PCP pincer ligand.

Scheme 26. General procedure for the synthesis of phebox.

2.3.1. Symmetrical pincer ligands


Like PNP, symmetrical PCP pincers with dibenzylic spacers Scheme 27. Synthesis of PN CN P pincer ligand.
were synthesized via SN2 nucleophilic substitution of 1,3-
bis(halomethyl)benzene by dialkyphosphines in the presence of
triethylamine as a base, or together with direct substitution by R2 P− with O-atom as spacers (1,3-bis(dialkyphosphinito) benzene) 125
[140–142] (Scheme 23). In addition, unsymmetrical PCP pincer lig- has been prepared with a good yield from chlorodialkylphosphine
ands and chiral PCP pincer ligands with substituted CH(R) benzyl and resorcinol in THF under basic conditions [153,154]. In addition,
arms or mixed PRR groups have been described. The use of the the PO CO P pincer ligand 128 with the 4- and 6-positions blocked off
di-Grignard 1,3-C6 H4 (CH2 MgCl)2 in the synthesis of PCP ligands was prepared by reacting 2-methylresorcinol with tert-butanol in
was introduced recently by Milstein and Van Koten [143]. Simi- the presence of phosphoric acid to investigate the C H activation
larly, nucleophilic substitution of the 2,6-bis halomethyl benzene of the methyl group by ruthenium(II) salts [155] (Scheme 28).
with secondary amines generates the corresponding NCN pincer Furthermore, 1,3-di-N-imidazolyl benzene can be used as start-
ligand. The N,N,N,N-tetra methyl bis amino methyl benzene analog ing compound for preparing the bis(imidazolium) salts, which
can be prepared via methylation of ␣,␣ ,-diamino-m-xylene using are precursors of the symmetrical CNHC CCNHC pincer ligands.
a mixture of formic acid and formaldehyde [144] (Scheme 23). Akabane et al. reported the synthesis of 1,3-di-N-imidazolyl
The introduction of N- or C(NHC) donors into the pincer struc- benzene starting from 1,3-diaminobenzene, involving four steps
ture is sometimes cumbersome, but the “Click” reaction [145] [156]. Recently, Vargas described the high-yield one-step synthe-
provides a simple and rapid method for making both at the same sis of 1,3-di-N-imidazolylbenzene based on a copper-catalyzed
time. Copper catalyzed azide alkyne cycloaddition (CuAAC) using N-arylation procedure for nitrogen heterocycles. A solution of 1,3-
2,6-dialkynyl benzene and alkyl azide result in the formation of dibromobenzene and imidazole was heated in the presence of
2,6-bis(1,2,3-triazol-4-yl)benzene. The ligand created can be used copper oxide and potassium carbonate, followed by reaction with
as the NCN pincer ligand directly [146], or it can be alkylated to gen- various alkyl halides in toluene or solvent-free conditions to obtain
erate the bis imidazolium salt for use as a source of the CNHC CCNHC the bis(imidazolium) salt [157]. Subsequently, various copper pro-
pincer, as shown in Scheme 24. cedures have applied been to catalyze N-arylation of nitrogen
The chiral PCP pincer ligand 118 has been synthesized from heterocycles using aryl iodide and bromide [158–165] (Scheme 29).
the chiral diol analog 116, which was produced by the asymmet-
ric reduction of 1,3-diacetylbenzene and then converted into the
borane-protected phosphine 117 by in situ tosylation withTs2 O, 2.4. Miscellaneous pincer ligands
followed by nucleophilic attack with LiPPh2 (BH3 ). Subsequently,
the borane-protected phosphine was recrystallized from EtOAc to In contrast to the most common classes of PCsp3 P pincer
remove the meso compound. The BH3 groups were removed in an ligands, the synthesis of the dibenzobarrelene-based PCsp3 P pin-
efficient manner to produce the enantiomeric pure phosphine 118 cer complexes reported by Gelman et al. is very simple using
(Scheme 25) [147,148]. the reliable Diels–Alder cycloaddition method. Facile access to
Among the NCN pincer ligands, the bis(oxazolinyl)phenyl a readily modifiable platform also allows the tailoring of their
(abbreviated as phebox) ligand [149], which comprises two chiral steric and electronic properties. Moreover, the rigidity of the
oxazolines and a benzene backbone, is one of the most useful frame- almost full-aromatic frame and lack of labile ␣- and ␤-hydrogens
works for constructing chiral NCN pincer complexes and it has translates into robustness and conformational stability, as well
received much attention. The resulting phebox metal complexes as s-donation of the metallated bridgehead sp3 hybridized car-
can provide a C2-symmetric environment around the active metal bon. The unique topology of dibenzobarrelene molecules may be
center [51,150,151]. These phebox pincers can be readily prepared utilized to design chiral-at-frame pincer ligands, which provide
in two steps starting from isophthaloyl chloride: amide formation a more efficient chiral pocket compared with more conven-
with a chiral ␤-amino alcohol followed by oxazoline formation with tional systems [166]. Despite their structural complexity, the
methanesulfonyl chloride (Scheme 26). synthesis of these ligands can be accomplished readily using
PCP ligands 123 with NH spacers were prepared by the treat- a reliable and straightforward Diels–Alder cycloaddition strat-
ment of substituted 1,3-diaminobenzene with 2 equiv. R2 PCl in the egy. Thus, refluxing 1,8-bis-(diphenylphosphino)anthracene and
presence of NEt3 and/or n-BuLi to produce the PN CN P pincer ligand. dimethyl fumarate in xylene overnight in a nitrogen atmosphere,
All of these ligands are air stable in the solid state and in oxygen- followed by reduction of the diester adduct with LiAlH4 in THF
free solutions (Scheme 27) [152]. The phosphinito PCP pincer ligand yields the dibenzobarrelene-based pincer ligand [167].
H.A. Younus et al. / Coordination Chemistry Reviews 276 (2014) 112–152 125

Scheme 28. Synthesis of PO CO P pincer ligand.

Scheme 29. Synthesis of CCC pincer ligand with bis(imidazolium) salts.

Scheme 30. Synthesis of dibenzobarrelene-based PCsp3 P pincer and enantiopure PCS pincer ligands.

Given the general interest in the synthesis of more struc- the direct bromomethylation of acridine with bromomethyl-
turally complex chiral pincer ligands and the beneficial activities of methylether (BMME), and by reacting with the secondary
dibenzobarrelene-based ligands during chemical transformations phosphine HPiPr2 to obtain a good yield [168] (Scheme 31).
where the development of enantioselective versions is of great Silyl ligands have a strong -donating character and they have a
interest, Gelman et al. described a straightforward synthetic route stronger trans-influence than the other ligands used commonly in
for producing a new family of chiral nonracemic dibenzobarrelene- transition metal chemistry. Silyl ligands provide an electron-rich
based pincer ligands. The target molecules were synthesized via
lithiation of the racemic 1-bromo-8-diphenylphosphinotriptycene
137 and subsequent quenching with diastereomerically pure
(1R,2S,5R)-(e)-menthyl (S)-p-toluenesulfinate. The resulting mix-
ture of diastereomers was resolved using conventional column
chromatography. Hydrogenation of the diastereomerically pure
compounds led to the formation of enantiopure PCS pincer ligands
(Scheme 30).
In 2008, Gunanathan and Milstein reported an acridine-based
pincer ligand. This ligand was prepared from 4,5-bis(bromo-
methyl)acridine, which was obtained in a one-step reaction via Scheme 31. Straightforward synthesis of PNP acridine-based pincer.
126 H.A. Younus et al. / Coordination Chemistry Reviews 276 (2014) 112–152

Scheme 32. Synthesis of symmetrical (PSiP) and unsymmetrical (PSiN) silyl pincer ligand.

Scheme 33. Synthesis of phosphinodi(benzylsilane) pincer ligand.

Scheme 34. Synthesis of PCP pincer ligand based on NHC.

metal center and coordinatively unsaturated species due to their yield of 83% in a one-pot reaction by the dilithiation of PhP(o-tolyl)2
strong trans-labilizing effect. Therefore, “ancillary” silyl ligands with n-BuLi in the presence of TMEDA, followed by the addition of
may provide transition metal complexes with unique reactivi- ClMe2 SiH [172] (Scheme 33).
ties that are useful for catalysis. Simple silyl ligands usually have Transition metal complexes with NHC ligands have been a hot
high reactivity and they cannot remain on transition metals as research area. The NHC-based pincer ligand, PCNHC P, has been
“ancillary” ligands. The incorporation of a silyl group in a multi- reported as a pincer ligand for ruthenium. The PCNHC P pincer
dentate ligand framework may be a useful strategy for producing ligand, 1,3-bis(2-chloroethyl)-3H-imidazol-1-ium chloride, was
suitable “ancillary” silyl ligands. The very simple synthesis of prepared starting from imidazole via a two-step procedure because
the symmetrical PSiP-pincer ligand 144 involved the lithiation attempts to perform a one-pot reaction between imidazole and 1,2-
of (2-bromophenyl)dicyclohexylphosphine followed by coupling dichloroethane always resulted in an impure mixture. Thus, the
with dichloro(methyl)silane to obtain the PSiP pincer ligand compound 1-(2-chloroethyl)-1H-imidazole was first synthesized
[169]. The unsymmetrical analog PSiN 147 was also synthesized by a phase-transfer catalytic approach followed by reaction with
in the same manner by the lithiation of (2-bromophenyl)di- in situ generated KPPh2 in DMSO to produce 154 as an air-sensitive
tert-butylphosphine and treatment with [2-(NMe2 )C6 H4 ]SiMeHCl, white solid [173] (Scheme 34).
where the former was prepared by reacting [2-(NMe2 )C6 H4 ]MgBr A neutral PCP pincer ligand containing a central carbene was
with excess MeSiHCl2 (Scheme 32) [170]. The presence of two Si synthesized in a straightforward manner from readily available
atoms in bis(silyl) systems facilitates the generation of specific materials, which undergoes double C H activation after treatment
properties with major effects in catalysis [171], where the SiCSi with [(p-cymene)RuCl2 ]2 . First, dipyrromethane was prepared
ligand has a stronger -donor capacity than analogous PIII pincer from the Lewis acid-catalyzed condensation reaction between
species. In this context, the synthesis of silicon pincer-type ligands paraformaldehyde and an excess of pyrrole. Subsequent depro-
with Si H bonds that have “cooperating” effects may be highly sig- tonation of dipyrromethane with excess NaH in THF followed by
nificant, thus phosphinodi(benzylsilane) 150 was prepared with a the addition of 2 equiv. of iPr2 PCl generated the PCP pincer ligand
H.A. Younus et al. / Coordination Chemistry Reviews 276 (2014) 112–152 127

Fig. 6. Neutral PCP and anionic PNP and NNN pincer ligands. Scheme 36. Synthesis of [RuHCl(CO)(PNP)] pincer complex based on acridine.

155 (Fig. 6) [174]. Similarly, the anionic diarylamido phosphine


ligand [PNP] 156, which was first reported by Kaska as a pincer
ligand for rhodium [175], has also been used as a pincer ligand for
RPCs [176–178]. Phosphines are generally “soft” donors whereas
amines are “hard” donors. Thus, substituting the two phosphine
donors of PNP pincer ligands with amines resulted in the formation
of a bis(amino)amido (N2 N) ligand, which maintained the struc-
tural rigidity and stability of PNP ligands, although it exhibited
different electronic properties and stabilized metals with higher
oxidation states. Hence, the bis(amino)amido (N2 NH) pincer,
bis[(2-dimethylamino)phenyl]amine (Me N2 NH) 157, was prepared
via the Pd-catalyzed C-N coupling of N,N-dimethylaminoaniline
with o-N,N-dimethylaminobromobenzene [179,180].

3. Synthesis of ruthenium pincer complexes


Scheme 37. Direct metallation of [Ru(COD)(metallyl)2 ] to PNP and PNN in the pres-
3.1. Direct metallation ence and absence of a hydrogen atmosphere.

After constructing the pincer ligand, its incorporate into the


pincer complex is another task. Direct metallation is the simplest The electron-rich and bulky t-Bu-PNP ruthenium(II) pincer
approach, but some important points should considered before the complex 158 was prepared in a straightforward manner by the dis-
synthesis process, such as the pincer ligand structure, ruthenium placement of PPh3 from the Ru(II) complex precursor. Heating a
source, reaction solvent, and other reaction conditions. In terms of THF solution of Ru(PPh3 )3 Cl2 with a stoichiometric amount of the
the pincer ligand structure, the donor sites should have unpaired t-Bu-PNP 3t Bu2 at 65 ◦ C for 3 h in a nitrogen atmosphere yielded
electrons that are available for bonding, which are derived from the dinitrogen-bridged dimer complex 159, which appeared to
heteroatoms or generated in situ from NHCs. Moreover, the steric exist in equilibrium with its monomer 158 (Scheme 35), proba-
effect around the donor centers of the pincer ligand have a great bly due of the steric hindrance of the bulky t-butyl groups of the
impact on the probability of bonding to the metal center, so the PNP ligand and the ease of substituting the triphenylphosphine
alkyl/aryl groups around the donor sites are always a matter of con- [181]. Attempts to synthesize the same complexes by alkene dis-
cern. The ease of substituting the ancillary ligands on the ruthenium placement from the Ru(II) complex Ru(NBD)Cl2 or from the Ru(0)
center also regulates the bonding process when producing the pin- complex Ru(COD)(COT) failed to yield the desired product. The eas-
cer complex. Therefore, the most common ruthenium sources used ier reaction of the less bulky ligand iPr-PNP with RuHCl(PPh3 )3 (CO)
for direct metallation possess readily substituted triphenyl phos- in THF at room temperature overnight resulted in the forma-
phine ligands. The synthesis of RPCs via direct metallation has been tion of [RuHCl(iPr-PNP)(CO)] 160 as an off-white solid [83]. The
reported using a variety of pincer ligands: PNP, PNN, PNS, NNN, acridine-based pincer complex 162 was prepared by reacting the
CNHC NN, and CNHC N CNHC , as well as diverse ruthenium sources, electron-rich tridentate PNP ligand with [RuHCl(PPh3 )3 (CO)] in two
and in combination with different solvents. different solvent systems [182,183] (Scheme 36).

Scheme 35. Synthesis of [Ru(II)(PNP)] pincer complexes by substituting the PPh3 ligand.
128 H.A. Younus et al. / Coordination Chemistry Reviews 276 (2014) 112–152

Scheme 38. Dearomatization of PNP/PNN ruthenium pincer complexes.

Scheme 39. Synthesis of the Ru(PNS) pincer complexes and dimer formation.

Modification of the ruthenium precursor was used successfully The hemilabile PNN pincer complex 166c was prepared in a sim-
to obtain the bulky t-Bu-PNP pincer complex by olefin substitution. ilar manner to the PNP analog 166a,b by refluxing the 3i Pr pincer
Leitner reported the synthesis of a non-classical ruthenium hydride ligand with RuHCl(PPh3 )3 (CO) in THF. Deprotonation of complex
pincer complex 164, which was obtained by heating the readily 166a–c with KOtBu led to deprotonation of the benzylic phosphine
available ruthenium source [Ru(COD)(2-methylallyl)2 ] with a PNP “arm,” instead of the hydride ligand to generate the dearomatized
ligand (2,6-bis((di-t-butylphosphino)methyl)pyridine) in pentane complex 167a,b with a yield of 89% [83,188]. Similarly, the PNN
with a hydrogen atmosphere (7 atm) [184–186]. Treatment of bipyridine based pincer 18 yielded the aromatized 168 and de-
the same ruthenium precursor with either the hemi-labile PNN or aromatized pincer complexes 169 [189–191] (Scheme 38). Using
the stronger PNP pincer ligand in the absence of a hydrogen atmo- the same approach, the (PNS)RuH(Cl)CO complex 170 was prepared
sphere also led to the in situ formation of [Ru(PNN)(2-methylallyl)2 ] by refluxing the PNS ligand with RuH(Cl)CO(PPh3 )3 in THF. How-
165a and [Ru(PNP)(2-methylallyl)2 ]165b [187], probably due to the ever, attempts to synthesize the dearomatized (PNS)Ru(H)CO 171
asymmetrically bonded allyl ligands (Scheme 37). using KOtBu or KHMDS as a base with a similar protocol to that

Scheme 40. Synthesis of PNN ruthenium pincer complexes based on 2,2 -bipyridinemethane and 2,2 -oxobispyridine and attempts at their dearomatization.
H.A. Younus et al. / Coordination Chemistry Reviews 276 (2014) 112–152 129

Fig. 7. Ruthenium pincer complexes reported by Gusev et al. based on simple pincer ligands.

reported for the preparation of the dearomatized PNP or PNN com-


plexes led to the formation of the symmetrical dimeric complex
172 (Scheme 39) [84].
Using the same approach, refluxing [RuHCl(CO)(PPh3 )3 ] with a
slight excess of the electron-rich PNN ligands 25a–c in THF yielded
the PNN-Ru complexes 173a–c. The dearomatized compounds
174a,b undergo cyclometallation to yield complexes 175a,b. For
cyclometallation to proceed, the coordination sphere around the
ruthenium center has to be rearranged, which depends on the Scheme 41. Microwave synthesis of PNN ruthenium pincer complex.
flexibility of the system, thus cyclometallation is qualitatively
faster with the dimethyl derivative 174a compared with the
spyrocyclopentyl derivative 174b. Cyclometallation also occurs
diastereoselectively and yields only one diastereomer of the
cyclometallated compounds. For the 2,2-oxobispyridine complex
173c, the dearomatized complex was too unstable to be isolated
(Scheme 40).
Gusev et al. reported the synthesis of a series of RPCs 176–178
(Fig. 7) that incorporated a simple pincer ligand, which could be
prepared on a large scale from inexpensive and readily available
starting materials, where they have high reactivity in many cat-
alytic applications [92,192–194]. The monohydride complexes can Scheme 42. Direct metallation of CNHC NN pincer ligand.
be prepared easily on a large scale with high yields by reacting
[RuHCl(CO)(PPh3 )3 /AsPh3 ] with the PNP or PNN pincer ligands via produced a quantitative yield of [RuCl2 (PPh2 )(dmso)PNN] complex
conventional PPh3 or AsPh3 ligand exchange. Further treatment 180 with sufficient purity [196] (Scheme 41).
of the monohydride 176a with KOtBu in ether with a hydrogen Direct metallation has also been extended to the syn-
atmosphere generated the dihydride trans-RuH2 (CO)(PNP) 176b thesis of CNHC NN RPCs via reacting the ruthenium source
[194], but the monohydride 177a failed to yields the dihydride [RuHCl(CO)(PPh3 )3 ] and the imidazolium salt in the presence of
analog and the dimeric structure was obtained instead [192]. The a strong base. This reaction was performed in two stages: the Ru(II)
dichloride complexes were prepared by reacting the pincer lig- precursor and CNHC NN ligand were refluxed overnight in THF to
ands with RuCl2 (CO)(DMF)(PPh3 )2 or RuCl2 (PPh3 )3 in glyme and achieve NN coordination, followed by the addition of a KHMDS
toluene, respectively, for 3–4 h. In the same manner, the ruthe- solution to complete the NHC moiety bonding. The resulting solu-
nium complexes 178a–d were obtained by conventional ligand tion was refluxed for another 2 h to obtain product 181 [197]
substitution reactions using HN-(C2 H4 SEt)2 with [RuCl2 (PPh3 )3 ], (Scheme 42).
[RuHCl(PPh3 )3 ], [RuCl2 (AsPh3 )3 ], and [RuHCl(CO)(PPh3 )3 ] respec- However, reacting the same Ru precursor with CNHC NN based
tively [195]. on bipyridine following the same protocol resulted in C H acti-
Interestingly, a PNN RPC has also been prepared using vation of the bipyridine moiety instead of direct metallation
microwave radiation instead of conventional heating, which to the N-atoms of the bipyridine, thus the bipyridine moiety
sometimes yields a significant amount of side products. Accord- remained non-coordinated (Scheme 43) [109]. The X-ray single
ingly, heating [Ru(dmso)4 Cl2 ] in THF at 120 ◦ C using microwaves crystal of the complex exhibited an octahedral structure with two
130 H.A. Younus et al. / Coordination Chemistry Reviews 276 (2014) 112–152

Scheme 43. C H activation of a CNHC NN pincer based on bipyridine instead of direct metallation.
Reprinted with permission from ref. [109]. Copyright 2011 American Chemical Society.

phosphorus atoms located trans to each other, a hydride trans to


the NHC ligand, and CO trans to the C H-activated carbon of the
bipyridine ligand. A trial based on a one-pot reaction led to the
formation of an inseparable mixture of products.
The presence of bulky triphenylphosphine probably hinders the
coordination of the bipyridine ligand, thereby enforcing the C H
activation of the bipyridine moiety. Thus, changing the ruthenium
complex source into a phosphine-free precursor complex as Ru(p-
cymene)Cl2 (CO) with ligand 48 led to the formation of the desired
Fig. 8. Structure of the ruthenium(II) complex based on Me4 BPPy: correct structure
pincer complex (Scheme 44). Variable-temperature 1 H NMR mea- on the left and the originally reported structure on the right.
surements of the product in CD2 Cl2 proved that structures 184a
and 184b were in equilibrium with each other.
A CNHC NN RPC based on bipyridine-bearing triphenylphosphine
occupies three meridional sites and the two chlorides are mutu-
as an ancillary ligand was obtained in an effective manner by the
ally cis. Reaction of the Grubbs’ first-generation metathesis catalyst
coordination of [RuHCl(CO)(PPh3 )3 ] to the pre-generated free car-
analog with the free carbene 41a in THF yielded a mixture of
bene [109,198]. The free NHC ligand was generated by mixing the
products, by ligand exchange proceeded cleanly in toluene to gen-
CNHC NN pincer with LiHMDS in THF at −78 ◦ C, followed by warm-
erate a single product. Moreover, attempts to synthesize the less
ing to −30 ◦ C. Next, RuH(Cl)CO(PPh3 )3 was added and the mixture
bulky analogs of 191 by replacement of the di-isopropylphenyl sub-
was allowed to warm up slowly to room temperature. The reaction
stituents of the carbene ring with the smaller mesityl or t-Bu were
mixture was then refluxed overnight to obtain the pincer complex
unsuccessful, yielding complex mixtures [199].
185. Deprotonation of complex 185 with KHMDS in benzene or
In 2005, Wu et al. reported the first X-ray crystal structure
toluene resulted in the formation of the dearomatized complex
of a neutral ruthenium(II) complex based on 2,6-bis(3,5-
186, which was slightly unstable at room temperature (Scheme 45).
dimethylpyrazol-1-yl)pyridine (Me4 BPPy) (Fig. 8) [200]. Single
Using the same approach, Wright et al. reported the synthesis of
crystal analysis showed that the two chloride atoms were perpen-
CNHC NN based on bipyridine without a spacer via the in situ gen-
dicular to each other and one was trans to PPh3 . Based on their
eration of the free carbene 188 using KHMDS, followed by direct
results, they were able to correct the mistaken structure reported
reaction with Grubbs’ first-generation metathesis catalyst analog
previously [117]. RuCl2 (PPh3 )(Me4 BPPy) has been prepared via
to generate compound 189 (Scheme 46) [199].
direct metallation by refluxing an ethanolic solution of Me4 BPPy
Several ruthenium complexes with poly-NHC ligands have been
with RuCl3 ·3H2 O to obtain RuCl2 (Me4 BPPy), followed by reflux-
obtained starting from 2,6-pyridyl-bisimidazolium salts using dif-
ing with 1 equiv. of PPh3 in ethanol in the presence of excess Et3 N
ferent synthetic methodologies. The free carbene can be used in
[201,202].
mild and clean ligand substitution reactions in polar nonprotic
Since the first reported X-ray crystal structure of Ru(II) pin-
solvents, thereby providing high yields of pincer bis-carbene com-
cer complexes based on Me4 BPPy, the design and synthesis of
plexes. Reaction of 1 equiv. of the free carbene 41 with RuCl2 (PPh3 )3
Ru(II)NNN pincer complexes with unsymmetrical coordination
in THF yielded the ruthenium bis-carbene complex 190 as an
arms received more attention, because a metal complex that sus-
air-stable crystalline solid (Scheme 47) [103]. The X-ray single
tains a polydentate ligand with two unsymmetrical coordination
crystal analysis of compound 190 indicates that the pincer ligand
arms usually has a much higher catalytic activity due to the

Scheme 44. Direct metallation of Ru(p-cymene)Cl2 (CO) to a bipyridine-based CNHC NN pincer.


H.A. Younus et al. / Coordination Chemistry Reviews 276 (2014) 112–152 131

Scheme 45. Synthesis of the CNHC NN ruthenium pincer complex based on bipyridine with triphenylphosphine as an ancillary ligand via direct metallation to free carbene.

Scheme 46. In situ generation of a free carbene-containing CNHC NN pincer and metallation with Grubbs’ first-generation catalyst analog.

hemi-labile character of one of the ligand arms. A common The synthesis of ruthenium(II) NNN complexes 202, 203 con-
method for the synthesis of a Ru(NNN) pincer complex is taining a chiral pyrazolyl–pyridyl–oxazolinyl ligand was achieved
based on the reaction of the NNN tridentate pincer ligand in a similar manner (Scheme 50). Remarkably, pincer ligands 66a
with 1.0 equiv. of RuCl2 (PPh3 )3 in refluxing toluene for 2–10 h and 66b both resulted in the formation of the same pincer com-
[55,112,121,123,124,200,203]. plex structure, whereas different structures were obtained with
Thus reaction of ligands 61 and 62 with 1.0 equiv. of pincer ligands 69a,b, depending on the substituent on the oxazo-
RuCl2 (PPh3 )3 in toluene yielded the corresponding Ru(II)pincer line moiety. The X-ray single crystal analysis of compound 203a
complexes. A pronounced advantage of complexes 194 and 196 is showed that the complex exhibits a neutral molecular structure
that the ligand can be varied by treatment with NaHCO3 to gen- with a nearly planar tridentate NNN ligand, where the ruthe-
erate the 16-electron complex, which is followed by treatment nium atom is surrounded by one PPh3 ligand, three N donor
with the ligand source in order to fine tune the activity of the pin- atoms, and two chloride atoms. The PPh3 ligand is positioned
cer complex (Scheme 48) [48,121,122]. RPCs based on the protic trans to the isopropyl group to reduce steric hindrance (Fig. 9)
2,6-di(1H-pyrazol-3-yl)pyridines ligand have also been reported, [124].
where pincer ligand 199a,b with two butylpyrazole arms reacted Recently, Zhengkun et al. reported that reacting equimolar
in a smooth manner with [RuCl2 (PPh3 )3 ] to yield the cationic amounts of ligands 76 and 77 with RuCl2 (PPh3 )3 in refluxing
bis(pyrazole) complex 200a,b (Scheme 49) [204,205]. Interest- 2-propanol led to the formation of an ionic ruthenium(II) com-
ingly, treatment of the same ruthenium source in similar reaction plex bearing a pyridyl-based benzimidazolyl–benzotriazolyl pincer
conditions with ligand 199c, but without any substituent at the 5- ligand. The X-ray crystallography of complex 204a showed that the
position of the flanking pyrazole, yielded the dinuclear complex cationic metal center is surrounded by the tridentate NNN ligand,
201, where two ruthenium atoms were bridged by two ligands two PPh3 ligands, and a chloride anion in a distorted-bipyramidal
with an unintended tautomeric form, probably to reduce steric hin- environment, with another dissociated chloride anion located in
drance around the nitrogen atoms distal to the central pyridine the vicinity (Fig. 9). Further treatment of complex 204a with K2 CO3
[205]. in DCM yielded the neutral Ru(II) complex 205 via liberation of
1 equiv. of HCl. By contrast, treatment with K2 CO3 in isopropanol at
reflux temperature generated the RuH pincer complex 206, which
could also be obtained from 205 in the same manner (Scheme 51)
[125].
Milstein et al. reported the synthesis of a series of ruthe-
nium pincers complexes based on phosphinite pincer ligands
of the PO NO P type. Different ruthenium sources have been
used to synthesize various PO NO P pincers complexes, depend-
ing on the ruthenium source used. Heating the phosphinite
pincer ligand iPr-PO NO P with HRu(Cl)(CO)(PPh3 )3 in benzene
or THF at 80 ◦ C led to the formation of pincer complex (iPr-
PO NO P)Ru(H)(Cl)(CO) 207. By contrast, treatment of the iPr-PONOP
pincer with HRu(Cl)(PPh3 )3 in the same conditions yielded (iPr-
PONOP)Ru(H)(Cl)(PPh3 )3 208 after 30 min. Treatment of complex
208 with NaHBEt3 at −35 ◦ C immediately generated the trans-
dihydride complex (iPr-PONOP)Ru(H)2 (PPh3 ) 209. It should be
noted that the trans-dihydride complex can also be prepared by
a one-step reaction with RuH2 (PPh3 )4 using ligand 7b, but the iso-
Scheme 47. Synthesis of CNHC NCNHC ruthenium pincer complexes (poly-NHC ruthe- lation of the complex from free PPh3 is a disadvantage of the latter
nium complexes) via direct coordination to free carbenes. procedure (Scheme 52) [73].
132 H.A. Younus et al. / Coordination Chemistry Reviews 276 (2014) 112–152

Scheme 48. Synthesis of [Ru(II)NNN] pincer complexes and their simple fine tuning.

Scheme 49. Synthesis of Ru(II)NNN pincer complexes with two proton-responsive pyrazole arms.

Scheme 50. Synthesis of Ru(II)NNN pincer complexes bearing a chiral pyrazolyl–pyridyl–oxazolinyl ligand.

The reaction of RuCl2 (PPh3 )3 with an equivalent amount of the Their mixtures have not been detected based on 1 H NMR spec-
corresponding PN NN P ligands 9a–f in DCM yielded the PN NN P pin- troscopy. Based on the X-ray crystallography of 210d, the geometry
cer complexes via triphenylphosphine exchange [75]. However, around ruthenium can be referred to as a distorted octahedron with
all attempts to prepare [Ru(tBu-PNP)(PPh3 )Cl2 ] were unsuccess- a meridional PNP ligand, a PPh3 ligand cis to the pyridine nitro-
ful, probably due to the presence of the sterically hindering tBu gen atom, and, remarkably, two mutually cis Cl atoms (Fig. 10).
groups and chloro ligand in the coordination sphere. The com- However, changing the ruthenium source for [RuHCl(PPh3 )3 CO]
plexes 210 are yellow to brown solids, which are stable in both successfully generated a ruthenium complex bearing bulky tert
the solid state and in oxygen-free solutions. Because of the merid- butyl groups [RuHCl(t Bu-PNP)CO] [206]. Thus, refluxing the PN NN P
ional coordination mode of the PN NN P ligands and the rigidity 9t Bu ligand with RuHCl(PPh3 )3 (CO) in THF for 12 h yielded the
of the -NHPR2 substituents, the complexes can only form two Ru(II) complex 211 as a pale yellow solid, which readily under-
mer-stereoisomers with either trans- or cis-dichloro arrangements. goes dearomatization after treatment with 1 equiv. of KOtBu in
H.A. Younus et al. / Coordination Chemistry Reviews 276 (2014) 112–152 133

Fig. 9. Solid state structures of some Ru(NNN)pincer complexes.


Copyright © 2012 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim for 203a,b. Reprinted with permission from ref [125] for 204. Copyright 2013 American Chemical Society.

Scheme 51. Synthesis of ruthenium(II) pyridyl-benzimidazolyl–benzotriazolyl pincer complexes.

Fig. 10. Molecular structure of the complexes 210d [cis-Ru(PNP-BIPOL)(PPh3 )Cl2 ] on the left and 213 [Ru(PNNNP)CO]+ on the right.
Reprinted with permission from ref. [75]. Copyright 2006 American Chemical Society for 210d. Reprinted with permission from ref. [206] for 213. Copyright 2012 Elsevier.

THF, thereby resulting in an immediate color change from orange also yielded the dearomatized form after treatment with KOtBu
to carmine. It should be noted that reacted complex 212 with (Scheme 54) [93,94].
1.5 equiv. of formic acid results in rearomatization (Scheme 53). X-
ray crystallographic analysis of complex 213 detecetd a distorted 3.2. C H bond activation
square-pyramidal geometry around the ruthenium center, with
the CO ligand trans to the pyridine nitrogen atom and the hydride The pincer PCSP 2 P undergoes C H activation using the common
located in the apical position (Fig. 10). In the same manner, reaction ruthenium precursors, e.g., RuCl3 ·3H2 O, [RuCl2 -(COD)]n , [RuCl2 (p-
of the pincer ligands 134a,b with Ru HCl(PPh3 )3 (CO) in reflux- cymene)]2 [140], RuCl2 (PPh3 )3 [207], and [RuHCl(CO)(PPh3 )3 ],
ing THF resulted in the formation of hydrido chloro-RPCs, which in the presence of a base to yield the PCP RPCs. However,
134 H.A. Younus et al. / Coordination Chemistry Reviews 276 (2014) 112–152

Scheme 55. CSp 2–H and CSp 3–H activation of PO CO P pincer ligand using
[RuHCl(CO)(PPh3 )3 ].

Scheme 52. Synthesis of dihydride ruthenium(II) and complexes based on a phos-


the general formula [RuCl(PCP)(CO)(PPh3 )]. Similarly, treatment of
phinite PONOP pincer ligand. [RuCl2 (PPh3 )3 ] with resorcinol bis(phosphinite) ligands generated
complexes of the [RuCl(PCP)(PPh3 )] type. Interestingly, a 2-methyl
[RuCl2 (COD)]n is the most fruitful source for this type of pin- C H activation occurred when C6 H-2-Me-4,6-t Bu2 -1,3-(OPPh2 )
cer construction, which involves C H activation of the phenyl 128b was treated with [RuHCl(CO)(PPh3 )3 ], thereby generating
group and NHCs. The reaction requires harsh conditions, i.e., PCP–pincer complex 217 (Scheme 55) [155].
a long reaction time and high temperature (170–190 ◦ C), so Bis(oxazolinyl)phenyl ligands (abbreviated as Phebox) provides
ethylene glycol is typically used as the reaction solvent. Reac- a C2 -symmetric and meridional configuration, as well as forming
ting RuCl2 (PPh3 )3 with the PCP pincer ligand also results in a central stronger covalent bond with the metal compared with
the C C or C H activation of Ar-CH3 , depending on the basic bis(oxazolinyl)pyridine ligands (Pybox). The synthesis of ruthe-
medium [142]. The reaction of the phosphinite PCP ligand nium complexes with the bis(oxazolinyl)phenyl moiety via C H
[C6 H4 -1,3-(OPPh2 )2 ] with [(p-cymene)RuCl2 ]2 yields the bimetal- activation was reported using commercially available ruthenium
lic species [208], but careful selection of the ruthenium source chloride in the presence of zinc powder [209], where a 4,6-
and reaction conditions can direct the reaction toward the dimethyl derivative of the chiral 1,3-bis(oxazolinyl)benzene was
desired orthometallated POCOP pincer complexes. Thus, reflux- applied so that the dimethyl substitution at 4- and 6-positions
ing [RuHCl(CO)(PPh3 )3 ] with resorcinol bis(phosphinite) pincer could prevent undesirable metallation by C H bond activation.
ligands 125a–d in toluene yielded the complexes 216a–d with Simple refluxing of a mixture of RuCl3 ·3H2 O and 121a,b in EtOH

Scheme 53. Synthesis of PN NN P ruthenium pincer complexes and their reversible dearomatization-aromatization.

Scheme 54. Synthesis of RuPN NC N pincer complexes.


H.A. Younus et al. / Coordination Chemistry Reviews 276 (2014) 112–152 135

Scheme 56. Synthesis of bis(oxazolinyl)phenyl–ruthenium(II) pincer complexes via C H bond activation with RuCl3 ·3H2 O in the presence of Zn and Mg metals.

with Et3 N resulted in the formation of non-isolable ruthenium


complexes. In addition, the use of dimeric ruthenium sources,
[RuCl2 (COD)]2 and [RuCl2 (C6 H6 )]2 , failed to generate the desired
products. However, using zinc powder in the presence of COD with
RuCl3 ·3H2 O and 121a,b successfully generated a ruthenium(II)
complex with a yield of 65%, which had the dimeric structure
of [(Phebox-ipr)Ru(CO)Cl]2 -(ZnCl2 ) 218a,b. The dimeric complex
was then treated with sodium acetate or sodium acetylaceton-
ate at room temperature for 12 h to produce (Phebox-i pr)Ru(CO)
(acetate)/(acac) 219a,b and 220a,b, respectively, with good yields
(Scheme 56) [210–213]. The X-ray analysis of complex 220a Fig. 11. General strategies for constructing a chiral pincer complex.
showed that the Ru center had a distorted octahedral geometry
with a meridionally coordinated Phebox ligand and a bidentate
acetylacetonato ligand. The CO ligand was perpendicular to the ruthenium(II) CNC-pincer bis(carbene)complexes 225 [216]. The
Phebox plane. In addition, treatment of the Phebox ligand 121a use of NEt3 provides much better results for the deprotonation of
with the same ruthenium precursor in the same conditions, where L compared with other mild (NaOAc, Na2 CO3 ) or stronger bases
zinc metal was replaced with magnesium, successfully generated (NaH, KOtBu) [216].
[RuCOCl(Phebox)(H2 O)] 221 [212]. Activating the C H of the aromatic system with ruthenium
The harsh conditions require to prepare a free NHC ligand salts to synthesize pincer complexes that incorporate the strong
may degrade the ligand, thus transmetallation has shown to be a Ru C bond is of great interest. Baratta et al. presented a new CNN
favorable procedure for obtaining NHC transition–metal complexes framework for ruthenium-based pincer complexes that comprised
[214]. However, treatment of the silver complex of imidazolium cyclometallated phenyl or benzoquinolinyl moieties, mainly via the
iodide 56d with a stoichiometric amount of the most common substitution of triphenylphosphine. Refluxing [RuCl2 (PPh3 )3 ] with
ruthenium sources, e.g., RuCl2 (PPh3 )3 or [Ru(p-cymene)Cl2 ]2 , in the CNN pincer ligand in isopropanol with a catalytic amount of tri-
DCM failed to generate the desired ruthenium(II)CNN complex. ethylamine yielded thermally stable CNN pincer complexes via C H
Likewise, trials to generate the free carbene by treatment with activation and the substitution of PPh3 as well as the elimination
1 equiv. of base (KHMDS or NaH) in THF, followed by the addition of HCl [135,137]. Similarly, complexes bearing the diphosphane
of RuCl2 (PPh3 )3 or [Ru(p-cymene)Cl2 ]2 , also failed to produce the Ph2 P(CH2 )4 PPh2 (dppb) in conjugation with the CNN pincer ligand
desired Ru(II) complex. Moreover, the one-pot reaction of CNHC NN were prepared by treating [RuCl2 (PPh3 )(dppb)] with the ligands 81,
ligand 56d and RuCl2 (PPh3 )3 in the presence of a base did not pro- 96a (Scheme 58) [137,217].
duce the desired Ru(II) complexes. Interestingly, the CNHC NN pincer There are two main strategies for constructing a chiral pin-
ligand 56d reacts with an equivalent amount of RuCl3 ·3H2 O and cer complex: the first method starts from a pincer ligand bearing
carbon monoxide in ethylene glycol at 190 ◦ C to generate the pincer chiral centers located mainly in the arms, whereas the second
complex [Ru(CNN)Cl2 CO], where further treatment with KI yielded uses a chiral auxiliary ligand, as shown in Fig. 11. A series of the
the complex 222 and, in the same manner, the pincer ligand 52 chiral RPCs that contain the CNN pincer ligand have been devel-
produced the RPC 223 (Scheme 57) [110]. Using the same protocol, oped using both strategies. The chiral complexes were prepared
Chung et al. reported the straightforward synthesis of CNHC NCNHC by reacting [RuCl2 (PPh3 )3 ] with the appropriate chiral diphos-
RPC 224 by refluxing the CNC pincer ligand with ruthenium chlo- phane using the pincer ligand (chiral/achiral) under reflux in
ride in ethylene glycol for 4 h to yield the mer-isomer as the sole 2-propanol (Scheme 59). It should be noted that the pincer com-
product [215]. plex RuCl(CNN)(dppb) was also obtained by adding HCNN and dppb
However, manipulating the ruthenium source into the dimeric to the ruthenium precursor [RuCl(␮-Cl)(␩6 -p-cymene)]2 in ethanol
[(COD)RuCl2 ]n can complete the reaction process and generate the under reflux, and in the presence of NEt3 , based on a one-pot syn-
anticipated product in milder conditions. Thus, the direct reaction thesis and without isolating the intermediate [137,218,219].
of [(COD)RuCl2 ]n and 2,6-bis(1-n-butylimidazolium-3-yl)pyridine Alternatively, Ru(CNN) pincer catalysts can be synthesized
bromide in the presence of NEt3 and reflux with ethanol yielded the by reacting the HCNN pincer ligand and the commercially
136 H.A. Younus et al. / Coordination Chemistry Reviews 276 (2014) 112–152

Scheme 57. Synthesis of [Ru(II)(CNHC NN)and CNHC NCNHC ] pincer complexes via direct C H bond activation.

Scheme 58. Synthesis of [Ru(II)(CNN)] pincer complexes via direct C H bond activation.

available [RuCl(-Cl)(6 -p-cymene)]2 to obtain both cationic and


neutral CNN cymene ruthenium complexes. Thus, treatment
of [RuCl(-Cl)(6 -p-cymene)]2 with 2 equiv. of 1-(6-arylpyridin-
2-yl)methanamine 81 in THF promptly coordinates the two
nitrogen atoms of the CNN ligand to form the [RuCl(6 -
p-cymene)(HCNN)]Cl complex 232. By contrast, the reaction
of 2-aminomethylbenzo[h]quinoline 96a with [RuCl(-Cl)(6 -p-
cymene)]2 leads to the formation of a neutral complex 234 due
to coordination to the N-atom of the amino group only. The
different behaviors of the two HCNN ligands can be ascribed
to the steric properties of the nitrogen ligands, where the
benzo[h]quinoline displays a higher rigidity and hinders the coor-
dination of the heterocyclic nitrogen to the Ru p-cymene moiety.
The cationic complex 232 reacts promptly with the monophos-
phine PPh3 in the presence of triethylamine in ethanol at reflux,
thereby yielding the pincer compound RuCl(CNNa )(PPh3 )2 233 via
displacement of p-cymene and ortho-metallation of the HCNN
ligand. In similar experimental conditions, 232 and 234 reacts
with the diphosphine Ph2 P(CH2 )4 PPh2 (dppb), thereby producing
the pincer catalysts RuCl(CNN)(dppb) 233 and 235 (Scheme 60)
Scheme 59. Synthesis of chiral Ru CNN pincer complexes. [220].
H.A. Younus et al. / Coordination Chemistry Reviews 276 (2014) 112–152 137

Fig. 12. Series of cyclometallated ruthenium complexes with CCC and CNO pincer ligands.

In addition, the regioselective metallation of the 1-pyrenyl moi- tube for 24 h. However, performing the reaction using more readily
ety toward the ruthenium center has been investigated. Reacting accessible RuCl3 ·nH2 O in ethylene glycol or EtOH under the same
[Ru(PPh3 )3 Cl2 ] with 1-pyrenaldehyde-4-R-benzoylhydrazones conditions obtained a substantially reduced yield (15%) [224]. It
(H2 pnbhR, where R = H, Me, OMe, Cl and NO2 ) in the presence was proposed that the CO source for the formation of 240a,b was
of sodium acetate yielded ortho-metallated Ru(III)(CNO) pincer derived from the decarbonylation of alcohol, which was used as
complexes with the formula trans-[Ru(pnbhR)(PPh3 )2 Cl] 236 the solvent [216]. The formation of these complexes might involve
(Fig. 12) [221]. Similarly, ruthenium(III) pincer complexes have the initial coordination of the NHC ligand to the ruthenium cen-
been synthesized and characterized that incorporate a CNO pincer ter, followed by a sequence comprising the oxidative insertion
ligand based on Schiff bases derived from 2-hydrazinopyridine of phenylene–H and reductive elimination of the hydride. More-
and 5-substituted salicylaldehyde [222]. A series of CCC-pincer over, treatment of the ruthenium complex [(CCC)Ru(CO)2 X] with
cyclo-metallated ruthenium complexes has also been synthe- AgPF6 in CH3 CN at ambient temperatures resulted in the substi-
sized, primarily by treatment of the ruthenium source RuCl3 (L) tution of the halide ligand with acetonitrile, which could also be
with AgOTf in acetone to replace the chloride atoms with replaced by NN-bidentate, as shown in Scheme 61. X-ray structure
solvent molecules, where L = 4 -di-p-anisylamino-2,2 :6 ,2 - analyses showed that these complexes had pseudo-octahedral con-
terpyridine (daatpy), bis(N-methylbenzidazolyl)pyridine (Mebip), figurations around the ruthenium center where the pincer ligand
4c 4-tolyl-2,2 :6 ,2 -terpyridine (ttpy), and trimethyl-4,4 ,4 - occupied three meridional sites.
tricarboxylate-2,2 :6 ,2 -terpyridine (Me3 tcbtpy), respectively. In general, the synthesis of [Ru(CNC/CCC)(NN)X] involves the
Subsequently, these intermediates were treated with the known construction of the pincer complex, followed by the replacement
bis-carbene precursor in the presence of KOtBu, followed by of two of the ancillary ligands with the bidentate NN ligand.
anion exchange using KPF6 to obtain the desired cyclo-metallated Alternatively, refluxing a ruthenium source that comprised NN
complexes 237 and 238 (Fig. 12) with moderate yields (21–49%). bidentate [Ru(NN)Cl4 ] 243 with pyridine-bridged bisimidazolium
X-ray analysis of the complex 237a showed that the ruthenium or bis-benzimidazolium hexafluorophosphate 244 in ethylene gly-
metal had an octahedral configuration with the two imidazole col followed by reduction using Zn granules [225] yielded the
rings and the benzene ring was essentially coplanar [223]. anticipated products. The chloride ligand can be substituted with
Benzene-based CCC-RPCs with carbonyl group as an ancil- other ligands, e.g., CH3 CN and t-BuNC, by refluxing in acetonitrile
lary ligand, 240a,b, were synthesized by heating 132a,b with or t-butylonitrile in the presence of a silver salt (silver nitrate or
[Ru(COD)Cl2 ]n in ethylene glycol/EtOH (1:2) at 165 ◦ C in a sealed silver hexafluorophosphate) (Scheme 62).

Scheme 60. Synthesis of Ru CNN pincer complexes and the effects of the steric properties of the nitrogen ligands on coordination.
138 H.A. Younus et al. / Coordination Chemistry Reviews 276 (2014) 112–152

Scheme 61. Synthesis of ruthenium(II) carbonyl complexes with benzene-based CCC-pincer bis-carbene ligands.

Scheme 62. Synthesis of ruthenium(II) complexes bearing bipyridine and the N-heterocyclic carbene-based CNC pincer ligand.

Scheme 63. Synthesis of ruthenium pincer complexes via CSP 3 -H double activation.

Bulky pincer complexes of ruthenium based on a propane skele- 16-e monohydride, whereas the reaction with 1,3-bis(di-tert-
ton are capable of achieving C H activation and H-elimination butylphosphinomethyl)cyclohexane 246 generated two olefin and
from the pincer ligand backbone to produce mixtures of olefin two alkylidene isomers of 16-e RuHCl[2,6-(CH2 PBut2 )], which all
and carbene products. Thus, reacting [RuCl2 (p-cymene)]2 with N,N- resulted from dehydrogenation of the ligand backbone. However,
bis(di-tert-butylphosphino)-1,3-diaminopropane yielded a mix- when the reaction was performed under hydrogen, the dihydride
ture of alkylidenes, Fischer carbenes, and olefin isomers of the products RuH2 Cl[2,6-(CH2 PBut2 )2 C6 H9 ] 247 were obtained. The
H.A. Younus et al. / Coordination Chemistry Reviews 276 (2014) 112–152 139

key intermediates associated with the C H activation reactions 1,5-cyclooctadiene, respectively. Finally, the structure of the prod-
were identified as 14-electron paramagnetic species RuCl(PCP), uct was confirmed by X-ray crystallographic analysis (Scheme 64)
which undergo hydrogenation to generate the 16-electron dihy- [170].
drides, RuH2 Cl(PCP) [226]. In addition, the PCSP 3 P pincer ligand
that encompasses a central carbene undergoes double C H 3.4. Transmetallation
activation. Heating a mixture of the PCP pincer ligand 155, [Ru(p-
cymene)Cl2 ]2 , and triethylamine for 12 h in toluene resulted in the Silver NHC complexes readily undergo exchange reactions
formation of the Ru(H)(Cl)(PCP) pincer complex as a reddish solid with binuclear halide-bridged metal complexes of rhodium and
with a yield of 75% (Scheme 63). Interestingly, exposure of this type ruthenium [229]. Transmetallation of the (NHC)CNN Ag(I) pincer
of carbene complex to strongly coordinating ligands such as CO complex with a Ru(II) source in ambient conditions, the high yield,
induced a quantitative 1,2-H shift [174]. and lack of a requirement for a strong base makes transmetallation
Recently [227], Musa et al. reported the synthesis of an air-stable an attractive choice for the synthesis of RPCs. However, few RPCs
ruthenium complex bearing the dibenzobarrelene-based cooper- have been prepared via transmetallation from silver and lithium
ating ligand. This complex was prepared via CSP 3 -H activation intermediates due to the restriction of transmetallation to pincer
by reacting the bifunctional ligand with the ruthenium precur- complexes that bear NHCs. During transmetallation, the silver or
sor Ru2 Cl4 (CO)6 in chloroform at room temperature. Analysis of lithium intermediates were treated with a halogenated ruthenium
the reaction mixture by 31 P NMR indicated the presence of three precursor in DCM. The Ru(II) CNN 254 and CNC 256 pincer com-
isomers of the target compound with a ratio of 1:0.3:0.14, which plexes were prepared by in situ transmetallation using the silver
appear to differ at the chloride ligand position. Moreover, X-ray carbene complexes. In light-free conditions and at room temper-
crystallographic analysis confirmed that the ruthenium center was ature, a solution of the pincer ligand solution in DCM was treated
slightly distorted from the octahedral geometry because the com- with Ag2 O or the soluble precursor AgOC(CF3 )3 [230], which work
plex contained the dibenzobarrelene-based ligand with a high as both base and halide scavengers [108], followed by the addition
degree of bend (Fig. 13). of [RuHCl(CO)(PPh3 )3 ] to obtain the product. The attempted de-
aromatization of the CNN pincer complex 254 with KOtBu yielded
3.3. Si H bond activation the desired de-aromatized Ru(II)pincer complex 2255. By con-
trast, reacting the CNC pincer complex 256 with KOtBu in the
Silyl ligands have a strong ␴-donating character and a stronger same conditions failed to yield the de-aromatized pincer com-
trans influence than the ligands that are used commonly in tran- plex and the Ru(II) hydride complex 257 was formed instead
sition metal chemistry [228]. Thus, silyl ligands may produce an (Scheme 65). The deactivation and recovery of homogeneous cat-
electron-rich metal center and coordinatively unsaturated species alysts, as well as metal separation from the organic substrates, are
due to their strong trans-labilizing effect. However, simple silyl lig- difficult challenges in homogenous catalysis. Ruthenium pincers
ands usually have high reactivity and cannot remain on transition with a triethoxysilyl group have been immobilized on MCM-41
metals as “ancillary” ligands. The incorporation of a silyl group in a [108,231], MCM-41/Al, or MCM-41/Sn [232]. Supported complexes
multidentate ligand framework may be a useful strategy for making have been prepared by refluxing a mixture of the pincer com-
“ancillary” silyl ligands. Two methods can be used to incorporate plexes with a pendant silyloxy group and the support in toluene
silyl ligands into the ruthenium pincer structure: (1) incorpora- for 16 h. The catalyst prepared in this manner had a metal loading
tion of one silyl group at the center of a multidentate framework of 0.06 mmol of metal/g in the support, according to atomic absorp-
[169], and (2) attachment of two silyl groups in a rigid multidentate tion analysis. The activity and recyclability of the immobilized
framework [172]. catalyst were explored, which showed that the solid complexes
RPCs have been prepared via Si–H activation using a variety remained active and recyclable. Moreover, no deactivation was
of ruthenium sources and pincer ligands. For example, ter- observed after repeated recycling (Fig. 14).
tiary silane, [PSiP]H, was treated with an equimolar amount Similarly, Suárez et al. reported the synthesis of Ru(II)CNC pincer
of [RuCl2 (COD)] in dry THF with an ambient temperature in complexes 260a–d via a separable silver intermediate by treatment
the presence of NEt3 and oxidative addition of the Si H bond of bis-imidazolium salts 258 with 1 equiv. of Ag2 O in CH2 Cl2 at
to the ruthenium(II) center occurred, which was followed by room temperature, thereby yielding the bimetallic silver complexes
reductive-elimination to obtain Ru(PSiP) complex 250 with high 259a–d, where further treatment with RuHCl(CO)(PPh3 )3 obtained
yields [169]. The phosphinodibenzylsilane compound also acts as the product in the fac-coordination mode, probably because of
a pincer-type ligand. “PSi2 Hx ” can adopt different coordination the steric hindrance of the bulky group on the imidazole ring
modes with ruthenium via diverse types of Si H bond activa- thus it cannot be considered as a pincer complex (Scheme 66).
tion. At room temperature, compound 150 reacts in deuterated Single-crystal X-ray diffraction of 260a (X = BF4 ) showed that
benzene/toluene, or THF, with bis(dihydrogen) RuH2 (H2 )2 (PCy3 )2 this complex comprised a distorted octahedral structure that
to yield compound 251 due to the formal substitution of two contained the CNC pincer coordinated in a fac configuration
dihydrogen and one tricyclohexylphosphine ligands in the ruthe- (C2(NHC)–Ru–C2(NHC) = 101.3(8)◦ ), whereas the CO was located
nium source [RuH2 (H2 )2 (PCy3 )2 ] [172]. Attempts to synthesize trans to the pyridine nitrogen atom of the pincer system. Ini-
the (t Bu-PSiN-Me)Ru complex by reacting the PSiN pincer ligand tial trials to synthesize the complex via direct metallation to
147 with common Ru(II) precursors, such as (PPh3 )3 RuCl2 and the free carbenes by reacting the imidazolium salt 258a (X = Br)
[(p-cymene)RuCl2 ]2 , in the presence of Et3 N failed to yield the with different ruthenium precursors (RuHCl(PPh3 )3 , RuCl2 (PPh3 )3 ,
anticipated products, and they typically led to the formation of mul- RuHCl(CO)(PPh3 )3 , and RuH2 (CO)(PPh3 )3 ) in the presence of a base
tiple intractable products. By contrast, heating the pincer ligand produced an inseparable mixture of products [233].
PSiN with 1 equiv. of (COD)Ru(2-methylallyl)2 for 18 h at 85 ◦ C Transmetallation from silver intermediates was also extended
led to the formation of a single unpredicted product 253. How- to the synthesis of phosphinated RPCs based on imidazole moiety.
ever, the NMR features of product 253 were not consistent with The PCNHC P RPC 263 and 264 were prepared via transmetallation
the formation of the (t Bu-PSiN-Me)Ru(2-methylallyl) complex. of the silver complex 261. Treatment of PCNHC P·HCl with silver
Instead, the 1 H and 13 C NMR spectra indicated the formation of oxide resulted in the formation of a trisilver complex containing
a cyclooctenyl (PSiC)Ru carbene complex due to double C H bond 2 equiv., of PCNHC P ligand. Reacting the silver complex 261 with
activation of a NMe group in the t Bu-PSiN-Me ligand and in the RuCl2 (PPh3 )3 produced a bimetallic ruthenium(II)NHC complex
140 H.A. Younus et al. / Coordination Chemistry Reviews 276 (2014) 112–152

Fig. 13. Synthesis of the dibenzobarrelene-based ruthenium pincer complex via CSP 3 -H activation.
Reprinted with permission from ref. [227]. Copyright 2013 American Chemical Society.

Scheme 64. Syntheses of ruthenium pincer complexes via Si H activation.

Scheme 65. Synthesis of ruthenium CNN and CNC pincer complexes via transmetallation from in situ generated silver complexes.
H.A. Younus et al. / Coordination Chemistry Reviews 276 (2014) 112–152 141

Fig. 14. Heterogenization of ruthenium CNN pincer complexes MCM-41, MCM-41/Al, or MCM-41/Sn.
Reprinted with permission from ref. [232]. Copyright 2012 Elsevier.

Scheme 66. Synthesis of Ru(CNC) pincer complexes via transmetallation from isolated silver complexes.

Scheme 67. Synthesis of ruthenium PCP pincer complex based on imidazole via silver transmetallation.

Scheme 68. Synthesis of ruthenium(N2 N) pincer complexes via lithium transmetallation.


142 H.A. Younus et al. / Coordination Chemistry Reviews 276 (2014) 112–152

in the facial coordination and, interestingly, this fac-complex can ligand is a sterically and electronically tunable ligand that sup-
also be produced from a direct reaction between PCNHC P·HCl and ports catalysis when coordinated with the usual catalytic active
RuCl2 (PPh3 )3 without the need for a base, although this synthesis metals. New pincer systems may even allow the emergence of
is more complicated and less clean than the silver carbene trans- useful concepts that can gain general acceptance and aid the
fer reaction. After refluxing a solution of 262 in 1,2-dichloroethane design of molecular structures orientated toward a given prop-
in a CO atmosphere, RuCl2 (CO)(PCNHC P) 263 was formed with an erty. RPCs have various catalytic applications, e.g., hydrogenation
octahedral geometry where the PCNHC P ligand was chelated in a reactions [243] (direct and indirect), oxidative dehydrogenation
meridional fashion. Reduction of the ruthenium dihalide pincer reactions [34] (acceptor and acceptorless), amination of alcohols
complex with excess NaBH4 in ethanol led to the isolation of the [168,244], hydrolysis of amines [245,246], hydroboration of termi-
ruthenium hydride RuHCl(CO)(PCNHC P) 267 (Scheme 67) [234]. nal alkynes [247], cyclopropanation of olefins [212], water splitting
Transmetallation from lithiated reagents has also been inves- [81], alkynylation of carbonyl compounds with terminal alkynes
tigated in the preparation of RPCs. Lithiation of Me N2 NH 157 [211,213], transesterification [248], olefin metathesis [199], and
produced the dimeric lithium complex [(Me NN2 )Li]2 265, where the alcohol racemization [249]. The fundamental differences between
structure was confirmed by X-ray crystallographic analysis. Treat- various pincer classes result in dramatic variation in their reac-
ment of 265 with Ru(PPh3 )3 Cl2 at room temperature yielded the tivities and catalytic properties. RPCs that catalyze hydrogenation,
[(Me N2 N)Ru(PPh3 )Cl] complex 266 (Scheme 68). X-ray crystallo- oxidative dehydrogenation, and other catalytic applications are
graphic analysis showed that the Ru center was five-coordinated controlled by many factors, in addition to the pincer ligand struc-
and surrounded by the Me N2 N ligand, Cl anion, and PPh3 ligand in a ture. Thus, knowledge of the main factors that affect these activities
square-pyramidal environment, where the N2 N pincer ligand and and how to implement them in pincer ligands is essential for tuning
Cl anion comprised the plane, and the PPh3 ligand occupied an axial the pincer ligand to improve the performance of a catalyst.
position [235].
4.1. Hemilability and non-innocent behavior of the pincer ligand
3.5. Trans-cyclometallation
Hemilabile ligands are polydentate ligands that contain at least
two different types of chemical functionalities and they are capable
Trans-cyclometallation refers to the exchange of cyclomet-
of binding to a metal center [250]. These functionalities are often
allated ligands at the metal center without the formation of
selected so that they are very different from each other to increase
significant and detectable amounts of purely inorganic com-
the differentiation between their resulting interactions with the
pounds, as described by Van Koten in 2000 [236]. During
metal center, and thus their reactivity and chemoselectivity. These
trans-cyclometallation, the relatively weakly coordinated ligand is
functionalities will influence the bonding/reactivity of the other
substituted by an incoming stronger bonding ligand, which also
ligands that bind to the metal, particularly those in the trans posi-
involves aryl C H oxidative-addition/reductive-elimination. Thus,
tion. Combining hard and soft donors in the same ligand, which
the addition of a PCP ligand to the cyclometallated ruthenium pre-
is often called a hybrid or heteroditopic ligand, is a major goal
cursor 268 is expected to result in the replacement of the hard
because it is hoped that different and contrasting chemistries may
amine ligands by phosphine ligands, which are known to be much
be associated within the same molecule, thereby leading to novel
softer ␦-donors. The development of the trans-cyclometallation
and unprecedented properties in the metal complexes obtained.
procedure addressed problems such as the presence of free phos-
An essential feature of hemilabile ligands is the possession of at
phine or HCl, and the need for using chlorinated solvents, which
least one substitutionable labile donor function, whereas the other
may give rise to secondary products. The pincer ligand 267a
donor group(s) remains firmly bound to the metal center. The pres-
was prepared in a trial of the functionalization of aryl ligands as
ence of a hemilabile ligand in a pincer complex may significantly
templates for incorporating catalytic active TM fragments onto
affect the reactivity of incoming substrates and promote transfor-
polymeric supports or well-defined carbosilane dendritic supports,
mations that would not occur otherwise (Tables 1 and 2). Examples
and treatment with the Ru(II) precursor RuCl2 (PPh3 )3 led to the
include the totally different behaviors of Milstein PNP and PNN
isolation of para-H complex 272 due to protodesilylation. The anal-
pincer complexes during the hydrogenation of esters and in
ogous complex RuCl[2,6-(Me2 NCH2 )2 C6 H3 ](PPh3 ) 268 reacts with
oxidative dehydrogenation reactions. Remarkably, the symmetri-
the ligand precursor 267a and generates the desired complex 270a
cal bis(di(tert-butyl)phosphinomethyl)-pyridine ligand (tBu-PNP)
with an overall yield of 70% (Scheme 69) [237].
exhibits very low activity in the hydrogenation of esters (Table 1,
Refluxing the PCP pincer ligand with an NCN RPC resulted in
entries 2 and 5), whereas the PNN analog has a much higher activity
ligand exchange to form a PCP RPC, thereby releasing the free NCN
with a major ligand effect that is attributable to the ligand hemil-
ligand [238,239]. The same method has been applied successfully
ability of the PNN pincer (Table 1, entries 3 and 6). The key features
to ruthenation of “cartwheel” type ligand systems that contain six
that make this reaction successful are as follows. The ligand is hemi-
potential metal-binding sites. Thus, refluxing 273 with 6 equiv. of
labile and it possesses a cooperative basic site in the ␤-position
[RuCl(NCN)(PPh3 )] 268 in benzene for 20 h resulted in the forma-
relative to the metal (the deprotonated benzylic carbon). The basic
tion of an air-sensitive hexakisruthenium complex where the color
site allows the addition of H2 to the de-aromatized PNN complex,
changed from dark purple to dark green (Fig. 15) [240–242].
thereby leading to the formation of the trans-dihydride complex,
while decoordination of the hemilabile amine arm facilitates the
4. Effects of the pincer complex structure on the catalytic binding of the ester substrate. Hydride transfer to the ester carbonyl
activity and the subsequent amine arm coordination and de-aromatization
of the pyridine core regenerate the complex, as well as eliminat-
The primary role of the ligand in a catalyst complex is to sta- ing a hemiacetal, which exist in equilibrium with the aldehyde in
bilize the transition metal in different oxidation states, thereby solution. The aldehyde is then hydrogenated to the corresponding
accommodating the metal during the catalytic reaction. Pincer lig- alcohol via a similar cycle (Fig. 16) [188].
ands provide a good opportunity to blend three different or similar Similarly, during the acceptorless dehydrogenation of alco-
coordination sites with the central metal, where each can have hols to carbonyl compounds with the PNP pincer complex 167b,
a specific role in stabilizing different oxidation states. In addi- alcohols are converted into the corresponding ketones with the
tion to its strong coordination with the metal center, a pincer liberation of dihydrogen. A significant improvement was achieved
H.A. Younus et al. / Coordination Chemistry Reviews 276 (2014) 112–152 143

Scheme 69. Synthesis of ruthenium PCP pincer complexes via trans-cyclometallation from Ru(II) NCN pincer complexes.

Fig. 15. Ruthenation of a “cartwheel” type ligand system via trans-cyclometallation.

when a ligand was constructed that combined hemilability and


cooperativity. Replacing one arm in the tridentate PNP ligand
with an aminomethyl group generated a PNN pincer ligand that
introduced hemilability due to the weaker binding of the amino

Scheme 70. Catalytic hydrogenation of carbonyl compounds with ruthenium pincer


complexes.

function. Modulating the donor sites and thus creating hemilabil-


ity in the pincer structure affected the catalytic activity as well as
controlling the reaction product. During the dehydrogenative cou-
pling of alcohols (or ester) and primary amines, the PNN pincer
complexes 165a [187,251] and 167c [49] promoted the formation
of amides, whereas the PNP analog 167b [50] led to imine for-
mation with the liberation of molecular hydrogen and water [50]
(Scheme 70).
The striking difference in the catalytic activity of the PNN and
PNP complexes can be rationalized as follows. For the PNN com-
plex with the hemilabile amine arm, the coordinated aldehyde
is attacked by the primary amine, thereby forming a quaternary
ammonium intermediate, which is followed by intramolecular pro-
ton transfer to the dearomatized phosphine arm. In the case of
the PNP complex, however, the attack on the aldehyde by the
Fig. 16. Proposed simplified catalytic cycle for the synthesis of esters, amides, and
imines from alcohols catalyzed by PNP and PNN RPCs (mechanisms that so not amine occurs after it is released into solution, thereby generat-
involve amine arm dissociation are also possible). ing a free hemiaminal intermediate that liberates water. Using the
144 H.A. Younus et al. / Coordination Chemistry Reviews 276 (2014) 112–152

Table 1
Catalytic activities of different pincer complexes in direct hydrogenation reactions.

E Catalyst Substrate Reaction conditions Conv. (%) Ref

Cat. mol% PH2 (bar) T (◦ C) t (h) Base (mol%)

Cinnamaldehyde 0.5 40 30 20 B1 (1.0) 87

1 [196]
Cl
N PPh2
Ru 1,1 ,1 -trimethylacetophenone 0.5 50 50 20 B1 (1.0) 99
DMSO Isobutyrophenone 0.5 50 50 20 B1 (1.0) 99
N
H2 Cl Methyl heptafluorobutanoate 0.5 50 140 24 B2 (1.0) 100
180 Dimethyl-o-phthalate 0.5 60 150 60 B2 (1.5) 100

H Ethyl benzoate 1.0 5.3 115 16 – 7.5


2 Pi Pr 2 [188]

N Ru CO

PiPr 2
167a Ethyl benzoate 1.0 3.5 160 16 – 12

H Ethyl benzoate 1.0 5.3 115 4 – 99.2


Pi Pr 2 Hexyl hexanoate 1.0 5.3 115 5 – 82.2
Benzyl benzoate 1.0 5.3 115 7 – 98.5
N Ru CO Methyl benzoate 1.0 5.3 115 4 – 100
3 [188]
Ethyl butyrate 1.0 5.3 115 4 – 100
NEt 2 Ethyl acetate 1.0 5.3 115 12 – 86
Tert-butyl acetate 1.0 5.3 115 24 – 10.5
167d Dimethyl terephthalate 1.0 5.3 115 5 – 100

Ethyl benzoate 1.0 5.3 105 2 B1 (1.0) 99.6

N
4 H [197]
N Phenyl ethylacetate 1.0 5.3 105 2 B1 (8.0) 99.7
Ru N Methyl acetate 1.0 5.3 105 2 B1 (8.0) 96
N Ethyl acetate 1.0 5.3 105 2 B1 (8.0) 99
Br Benzyl acetate 1.0 5.3 105 3 B1 (8.0) 90
dipp CO Tert-butyl acetate 1.0 5.3 105 2 B1 (8.0) 93
275 Diethyl succinate 2.0 5.3 105 2 B1 (.08) 99

H Butyl butyrate 0.5 10 110 12 – 10


5 [263]
PtBu2
N Ru H
t H B H
P Bu2
H
276 Benzyl benzoate 0.5 10 110 12 – 17

H Hexyl hexanoate 0.5 10 110 12 – 94

6 PtBu2 [263]

N Ru H
NEt2 H B H
Butyl butyrate. 10 110 12 – 98
277 H Benzyl benzoate 10 110 12 – 99
Methyl benzoate 10 110 12 – 97

Ethyl benzoate 1.0 5.4 135 2 B1 (1.0) 97


N
H
N N [109]

7 Ru Mes Benzyl benzoate 5.4 135 2 B1 (1.0) 91


CO
N PPh Ethyl butyrate 5.4 135 2 B1 (1.0) 98
3
Pentyl pentanoate 5.4 135 2 B1 (1.0) 96
186 Glycolide 1.0 10 110 48 B1 (1.0) 58
Glycolide 10 110 48 – 0 [198]
l-lactide 50 110 12 B1 (1.0) 67
H.A. Younus et al. / Coordination Chemistry Reviews 276 (2014) 112–152 145

Table 1 (Continued)

E Catalyst Substrate Reaction conditions Conv. (%) Ref



Cat. mol% PH2 (bar) T ( C) t (h) Base (mol%)

Glycolide 1.0 10 110 48 – 93

[198]

Glycolide 50 110 12 – 85
l-lactide 10 110 48 – 82
l-lactide 50 110 12 – 91
N-benzyl-2-methoxyacetamide 1.0 10 110 48 – 90
N-hexyl-2-methoxyacetamide 1.0 10 110 48 – 91
N-hexylfuran-2-carboxamide 1.0 10 110 48 – 69
N H P N-benzylbenzamide 1.0 10 110 48 – 57
N-ethyl acetamide 1.0 10 110 48 – 71
8 Ru N-methylpropionamide 1.0 10 110 48 – 68
N N-phenylacetamide 1.0 10 110 48 – 95
[189]

CO N-phenylhexanamide 1.0 10 110 48 – 92


N-phenylbenzamide 1.0 10 110 48 – 92
169
N-acetylmorphiline 1.0 10 110 48 – 98
N-formylmorpholine 1.0 10 110 48 – 98
2-methoxy-1-(pyrrolidin-1-yl)ethanone 1.0 10 110 48 – 97
Dimethyl carbonate 0.1 10 110 48 – 96
Dimethyl carbonate 0.2 50 110 14 – 89
Dimethyl carbonate 1.0 10 100 8 – >99
Methyl benzylcarbamate 1.0 10 110 48 – 97 [190]
Methyl pentylcarbamate 1.0 10 110 48 – 94
Methyl formate 0.06 10 110 48 – 96
Methyl formate 0.02 50 110 14 – 94

Dimethyl carbonate 0.04 40 110 3.5 – >99


9 [190]
N H P
Ru
Et2 N
CO
Dimethyl carbonate 0.04 60 110 1.0 – >99
167c Methyl formate 0.1 9 145 36 – >99

H Methyl benzoate 0.1 50 100 16 B3 (10) 98

H
N PPh 2 Methyl undecanoate 50 100 16 B3 (10) 98
10 [258]
Ru Dimethyl succinate 50 100 16 B3 (10) 100
Methyl 2-(benzyloxy)acetate 50 100 16 B3 (10) 100
P CO Methyl 2-(piperidin-1-yl)acetate 50 100 16 B3 (10) 100
Ph2
Methyl 3-methoxypropanoate 50 100 16 B3 (10) 11
Cl
Methyl 3-(N,N-Me2 )propanoate 50 100 16 B3 (10) <10
176c Isopropyl benzoate 50 100 16 B3 (10) 98
Tert-butyl benzoate 50 100 16 B3 (10) 96

H c Ethylene carbonate 0.1 50 140 0.5 B1 (1.0) 99

H
11 N PR 2 [264]
a Ethylene carbonate 0.1 50 140 0.5 B1 (1.0) 74
Ru
f Ethylene carbonate 0.1 50 140 0.5 B1 (1.0) 16
P CO g Ethylene carbonate 0.1 50 140 0.5 B1 (1.0) 76
R2
h Ethylene carbonate 0.1 50 140 0.5 B1 (1.0) 24
Cl 176
O
a; R=i Pr, c; R=Ph,
f; R=tBu, g; R=Cy, O O
R1 R3
h;R=1-Ad
R2 n R4
c n=0, 1 0.1 50 140 4–10 B1 (1.0) >99

H3C
N PPh2
Ru
P CO
Ph2
Cl
12 278 Ethylene carbonate 0.1 50 140 10 B1 (1.0) – [264]
146 H.A. Younus et al. / Coordination Chemistry Reviews 276 (2014) 112–152

Table 1 (Continued)

E Catalyst Substrate Reaction conditions Conv. (%) Ref



Cat. mol% PH2 (bar) T ( C) t (h) Base (mol%)

x a Methyl 2,2-difluoroacetate 0.02 10 40 6 B1 (25) >99


H PPh 2
N [265]
13
Ru
P CO
Ph2 x b Methyl 2,2-difluoroacetate 0.02 10 40 6 B1 (25) >99
a; X= H, Cl c Methyl 2,2-difluoroacetate 0.02 10 40 21 B1 (25) 65
b; X= H,H a Methyl 2,2,2-trifluoroacetate 0.05 25 40 24 B1 (25) 98
279a-c
c; X= Cl, Cl a Methyl acetate 0.02 10 40 18 B1 (25) 17
a Methyl 3,3,3-trifluoropropanoate 0.05 30 40 22 B1 (25) 5

i Pr
2H
P Co
Ru
N N
H
Cl
14 177a Methyl benzoate 0.05 50 100 1.7 B1 (1.0) 99 [192]

Ph 2 Cl Methyl benzoate 0.025 50 40 16 B4 (5.0) 98


P PPh 3 Ethyl acetate 0.005 50 40 16 B5 (1.0) 100
Ru Methyl hexanoate 0.005 50 40 18 B4 (5.0) 94
15 Methyl 2-methoxyacetate 0.025 50 40 16 B1 (1.0) 100 [92]
N N Methyl 2-hydroxypropanoate 0.05 50 40 16 B4 (10) 98
H N-benzylideneaniline 0.002 50 40 16 B1 (1.0) 100
Cl N-benzylidenepropan-1-amine 0.05 50 40 16 B1 (1.0) 100
177d
Methyl benzoate 0.025 50 40 6 B1 (1.0) 95
Ethyl benzoate 0.005 50 40 16 B5 (1.0) 85
Dimethyl phthalate 0.1 50 100 1.2 B4 (1.0) 96
Dimethyl isophthalate 0.05 50 40 16 B4 (1.0) 93
Dimethyl isophthalate 0.1 50 100 1 B4 (1.0) 100
Ethyl acetate 0.0025 50 40 14 B5 (1.0) 95
Ethyl acetate 0.0012 50 40 21 B5 (1.0) 73
Methyl hexanoate 0.005 50 40 24 B4 (1.0) 81
Et Cl Methyl 2-methoxyacetate 0.01 50 60 16 B4 (1.0) 100
S PPh3 Methyl 3-methoxypropanoate 0.05 50 23 21 B4 (1.0) 97
Tert-butyl acetate 0.025 50 100 1 B1 (1.0) 100
Ru Methyl non-3-enoate 0.05 50 40 8 B1 (1.0) 100
16 N SEt Methyl 2-hydroxypropanoate 0.05 50 100 1 B4 (1.0) 93 [195]
H Oxepan-2-one 0.01 50 100 2 B4 (1.0) 99
Cl N-benzylideneaniline 0.005 50 23 1.5 B1 (1.0) 100
N-(1-phenylethylidene)aniline 0.002 50 40 1 B4 (1.0) 63
178b N-(1-phenylethylidene)propan-1-amine 0.05 50 40 6 B1 (1.0) 100
Acetophenone 0.0025 50 40 24 B1 (1.0) 100
Cyclohexanone 0.005 50 23 1 B1 (1.0) 100
4-(tert-butyl)cyclohexanone 0.005 50 23 2 B1 (1.0) 100
2-pentanone 0.005 50 40 1 B1 (1.0) 100
Benzaldehyde 0.01 50 100 1.5 B4 (1.0) 94
Styrene 0.05 50 40 20 B1 (1.0) 100
1-pentene 0.05 50 40 4848 B4 (1.0) 75
2-pentene 0.05 50 40 B4 (1.0) 13

B1 : KOt Bu, B2 : LiHBEt3 , B3 : NaOMe, B4 : KOMe, B5 : NaOEt.

same approach, the dehydrogenative coupling of ␤-amino alcohols of up to 7.2 × 105 h−1 that increase more than 100-fold in a few
generates pyrazines and peptide formation, which depend on the seconds (Table 2, entries 3 and 4). The exceptionally high catalytic
presence or absence of the hemilabile arm. Thus, for the bulky PNP activity of these complexes is also attributed to the hemilability of
complex 167b that lacks a hemilabile amine “arm,” dissociation of the ligand and the ease of conversion from the saturated complex to
the aldehyde and its attack by the amino alcohol occur in solution, the coordinately unsaturated precatalyst in the reaction conditions.
thereby yielding an imine (by water liberation from a hemiami- Moreover, complexes that lack the NH functionality (Table 1, entry
nal), eventually producing a pyrazine after the aromatization of 12) have very low catalytic activities because the pincer ligand is
a presumed 1,4-dihydropyrazine intermediate. For the hemilabile incapable of unsaturated precatalyst formation, thereby supporting
PNN complex 167c, a sequence that involves nucleophilic attack the principle of non-innocent ligand cooperation.
by the amine group of the second amino alcohol molecule on the
aldehyde, which is coordinated with the PNN complex, eventually 4.2. Flexibility of the pincer framework (aliphatic versus
yields a peptide (Schemes 71 and 72) [252]. aromatic)
Ruthenium(II) complexes that incorporate the unsymmetrical
NNN pincer ligand exhibit an exceptionally high catalytic activity RPCs exhibit unprecedented catalytic activity in both acceptor
for the transfer hydrogenation of ketones in refluxing isopropyl and acceptorless dehydrogenation reactions [34]. In acceptorless
alcohol compared with the symmetrical ligand, with final TOFs dehydrogenation reactions, the catalytic process has been proposed
H.A. Younus et al. / Coordination Chemistry Reviews 276 (2014) 112–152 147

Table 2
Catalytic activities of different pincer complexes in the transfer hydrogenation of carbonyl compounds.
O OH i OH
+ PrOH, Base
R1 R2 + CH3 COCH 3
R1 R2 .

E Catalyst Substrate Reaction conditions Conv. (%) Final TOF (×104 h−1 ) Ref.

H-donor Base T (◦ C) t (h)


i
NMe 2 Cyclohexanone PrOH KOH 82 3.3 98 0.11 [266]
PPh3
1
Ru Cl

i
NMe2 Acetophenone PrOH KOH 82 44 70 0.003
i
280 Benzophenone PrOH KOH 82 90 90 0.008
i
2-pentanone PrOH KOH 82 4.5 0.1
i
PPh 2 Cyclohexanone PrOH KOH 82 1.3 98 2.7 [266]
PPh 3
2 Ru OTf

PPh2 i
Acetophenone PrOH KOH 82 0.5 90 0.9
i
281 Benzophenone PrOH KOH 82 108 98 0.01
i
2-pentanone PrOH KOH 82 1.5 90 0.2
i i
Cyclohexanone PrOH PrOK 82 1/12 98 0.58 [200]

N N N
3 Cl
N N
Ru
i i
Ph3 P Acetophenone PrOH PrOK 82 1/12 96 0.57
i i
Cl Benzophenone PrOH PrOK 82 3 89 0.01
192 1-phenyl propanone i
PrOH i
PrOK 82 3 95 0.06
i i
1-chloroacetophenone PrOH PrOK 82 1/12 100 0.6
i i
2-chloroacetophenone PrOH PrOK 82 1/12 98 0.58

Cyclohexanone i
PrOH i
PrOK 82 2.7 × 10−3 98 70.5 [121]

4 N N
Cl
N N NH Acetophenone i
PrOH i
PrOK 82 16 × 10−3 95 11.4
Ru
Benzophenone i
PrOH i
PrOK 82 8.3 × 10−3 98 23.5
Ph 3P 1-phenyl propanone i
PrOH i
PrOK 82 16 × 10−3 98 11.7
Cl
1-chloroacetophenone i
PrOH i
PrOK 82 2.7 × 10−3 96 69.1
194 2-chloroacetophenone i
PrOH i
PrOK 82 2.7 × 10−3 95 22.8

H Cyclohexanone i
PrOH i
PrOK 82 2.7 × 10−3 100 72.5 [48,122]
N
N N
5 Cl
N N
Ru
Cyclohexanone i
PrOH i
PrOK 25 33 × 10−3 99 2.97
Ph3 P
Cl Acetophenone i
PrOH i
PrOK 82 2.7 × 10−3 98 70.5
Acetophenone i
PrOH i
PrOK 25 500 × 10−3 96 0.19
196a
Benzophenone i
PrOH i
PrOK 82 2.7 × 10−3 98 70.5
Benzophenone i
PrOH i
PrOK 25 250 × 10−3 97 0.38
i i
Cyclohexanone PrOH PrOK 82 1.5 100 0.03 [110]

t Bu N N
6 I
N N Ru N
H nBu i i
I CO Acetophenone PrOH PrOK 82 2.0 98 0.02
i i
223
1-phenyl propanone PrOH PrOK 82 2.0 98 0.02
i i
1-chloroacetophenone PrOH PrOK 82 12 98 0.004
i i
2-chloroacetophenone PrOH PrOK 82 0.5 98 0.09

Cyclohexanone i
PrOH i
PrOK 82 16 × 10−3 97 6.0 [135]

7 N N
PPh 3
N Acetophenone i
PrOH i
PrOK 82 16 × 10−3 98 5.88
Ru
Ph3 P 1-phenyl propanone i
PrOH i
PrOK 82 16 × 10−3 99 5.94
1-chloroacetophenone i
PrOH i
PrOK 82 16 × 10−3 99 5.94
Cl
2-Acetonaphthone i
PrOH i
PrOK 82 16 × 10−3 97 5.82
227
Benzophenone i
PrOH i
PrOK 33.3 × 10−3 98 2.94
i
Cyclohexanone PrOH – 40 16 99 – [93]

8 O
N N H
tBu
2P Ru N
Acetophenone i
PrOH – 40 16 74 –
i
CO 4-Bromoacetophenone PrOH – 40 10 97 –
i
215 1-Acetonaphthone PrOH – 40 5 93 –
i
3-Hexanone PrOH – 40 16 99 –
148 H.A. Younus et al. / Coordination Chemistry Reviews 276 (2014) 112–152

Table 2 (Continued)

E Catalyst Substrate Reaction conditions Conv. (%) Final TOF (×104 h−1 ) Ref.

H-donor Base T ( C) t (h)
i
Cyclohexanone PrOH – 82 16 100 – [206]

9
N N H NH
t Bu P Ru PtBu
2 Acetophenone i
PrOH – 82 16 83a –
CO 4-Bromoacetophenone i
PrOH – 82 16 94a –
i
212 2-Heptanone PrOH – 82 16 93 –

Acetophenone i
PrOH i
PrONa 82 83.3 × 10−3 98 93 [267]
Ph 2
10 P

N Ru
2-chlorooacetophenone i
PrOH i
PrONa 82 33.3 × 10−3 99 68
Cl P
Ph2 2-MeO-acetophenone i
PrOH i
PrONa 82 33.3 × 10−3 96 54
NH 2
t
Acetophenone b i
PrOH i
PrONa 82 66.6 × 10−3 98 66
Bu Acetophenone i
PrOH i
PrONa 82 83.3 × 10−3 98 110 [137,219,268]
282a,b
Cl Acetophenone i
PrOH i
PrONa 60 166 × 10−3 98 18 [136,218]
N
11 Ru
Ar 2P N
Cy 2P H
H
Fe 2,4-OMe2 -acetophenone i
PrOH i
PrONa 82 33.3 × 10−3 98 130
2-chlorooacetophenone i
PrOH i
PrONa 60 33.3 × 10−3 97 15
231b
2-acetyl pyridine i
PrOH i
PrONa 60 500 × 10−3 99 12
Ar= 4-MeO-3,5-Me 2C 6H 2
2-Acetonaphthone i
PrOH i
PrONa 60 83.3 × 10−3 97 20

Acetophenone, P = PPh3 i
PrOH NaOH 82 83.3 × 10−3 95 0.57 [269]

12 O O
N
Cl
N N
Ph Ru
Ph
P
Cl Acetophenone, P = Pi Pr3 i
PrOH NaOH 82 50 × 10−3 97 0.97
283 1-phenyl propanone, P = Pi Pr3 i
PrOH NaOH 82 83.3 × 10−3 97 0.58
P= PPh3, Pi Pr3 3-bromooacetophenone, P = Pi Pr3 i
PrOH NaOH 82 16.6 × 10−3 91 2.7c

Acetophenone, L = PMe3 i
PrOH NaOH 82 250 × 10−3 97 – [270]

13 O O
N
Cl
N N
iPr iPr Ru
Cl L
Acetophenone, L = CNBn i
PrOH NaOH 82 250 × 10−3 92 –
284 Acetophenone, L = CNCy i
PrOH NaOH 82 500 × 10−3 86 –
L= PMe3 , CNBn, CNCy, MeCN Acetophenone, L = MeCN i
PrOH NaOH 82 250 × 10−3 47 –
a
A mixture of alcohol and substituted styrene was obtained.
b
Catalyst prepared in situ from the pincer ligand and RuCl2 (PPh3 )(dppb).
c
TOF at t = 1 min.

to proceed via cooperative interactions among substrates with the


basic site of the pincer ligand and with the electrophilic metal.
A hydrogen acceptor is not necessary because the ligands play
an active role in the hydrogen abstraction and liberation process.

Scheme 72. Synthesis of peptides and pyrazines from ␤-amino alcohols via pincer
ligand-controlled selectivity.

However, the bulky ligands in these catalysts may hinder the abil-
ity of the substrate to interact with both sites in the complex,
which may explain why tertiary amides are difficult to synthe-
size via acceptorless dehydrogenation coupling of alcohol and
amines [49,253–255]. Various rigid RPCs have been produced based
Scheme 71. Dehydrogenative coupling of alcohols and amines into amides and on 2,6-disubstituted benzene and pyridine, but new members
imines, depending on the hemilability of the pincer ligand. of this family, including the complexes 176, 178, 278, and 279,
H.A. Younus et al. / Coordination Chemistry Reviews 276 (2014) 112–152 149

have started to receive greater attention recently (Tables 1 and 2) ligands used by these RPCs could be tested with other earth-
[92,192,194,256–259]. The flexibility of the pincer framework has abundant metals such as iron, copper, nickel, and cobalt. Given
a significant impact on catalytic hydrogenation and acceptorless the ease of pincer complex immobilization and the strong bind-
dehydrogenation reactions may be attributed to the following: ing of pincer ligands to a metal center, heterogenization of RPCs
(i) highly flexible pincer ligands support different coordination may prove to be particularly valuable in the development of tan-
geometries, thereby facilitating the mer-fac interconversion, which dem systems. In these heterogeneous systems, metal leaching is
might be involved in the catalytic cycle; (ii) the loss of the steric not likely to be a problem, which is promising for the development
effect in the central aromatic pyridine/benzene ring allows the of ruthenium pincer catalysts in industrial applications and other
desired cooperative interactions between the substrates and the pincer complexes.
metal/ligand framework [260,261]. Transition metal complexes
with sulfur ligands are active catalysts in a considerable number
of homogeneous reactions [262]. Therefore, replacing N- and P- Acknowledgments
donors with softer S-donors, as well as the non-innocent behavior
and flexibility of the designed pincer ligand, can yield state-of- The authors would like to express their deep appreciation of
the-art catalysts for use in hydrogenation and dehydrogenation the State Key Lab of Advanced Technology for Materials Synthesis
reactions (Table 1, entry 16) [195]. and Processing for financial support (Wuhan University of Tech-
nology). F.V. acknowledges the Chinese Central Government for an
“Expert of the State” position in the program of “Thousand talents,”
5. Conclusion and future prospects as well as the support of the Natural Science Foundation of China
(No. 21172027) and the Research Fund - Flanders (FWO) (project
Significant achievements have been made in the chemistry of No. 3G022912). H.A.Y. and N. A. express their deep appreciation of
pincer-type complexes over the last decade. The wide variety of the Chinese Scholarship Council (CSC) for financial support via their
applications of pincer type molecular structures is a direct indi- PhD study grants 2012GXZ639 and 2012GXZ641, respectively.
cation of one of the ligands’ most attractive features, i.e., various
options for tuning their electronic and steric properties without References
affecting the ligand’s ability to bind to the metal center. RPCs have
received much attention as some of the most important ruthenium [1] G. van Koten, Pure Appl. Chem. 61 (1989) 1681.
complexes. Various strategies have been presented for constructing [2] D.W. Lee, C.M. Jensen, D. Morales-Morales, Organometallics 22 (2003)
4744–4749.
RPC skeletons but improving ligand design is almost an art. Sym- [3] M. Albrecht, G. van Koten, Angew. Chem. Int. Ed. 40 (2001) 3750–3781.
metrical ligands are easy to produce, but unsymmetrical ligands are [4] G. Alesso, M.A. Cinellu, S. Stoccoro, A. Zucca, G. Minghetti, C. Manassero, S.
receiving much attention for use as RPCs due to their hemilability Rizzato, O. Swang, M.K. Ghosh, Dalton Trans. 39 (2010) 10293–10304.
[5] V.A. Kozlov, D.V. Aleksanyan, Y.V. Nelyubina, K.A. Lyssenko, A.A. Vasil’ev, P.V.
and outstanding catalytic applications. Pyridine-based RPCs have Petrovskii, I.L. Odinets, Organometallics 29 (2010) 2054–2062.
been explored in depth for synthesis and many catalytic applica- [6] C.J. Moulton, B.L. Shaw, J. Chem. Soc., Dalton Trans. (1976) 1020–1024.
tions, but RPCs with other motifs remain relatively or completely [7] G. Van Koten, K. Timmer, J.G. Noltes, A.L. Spek, J. Chem. Soc., Chem. Commun.
(1978) 250–252.
underdeveloped.
[8] M.E. Van der Boom, D. Milstein, Chem. Rev. 103 (2003) 1759–1792.
Various procedures have been reported for the synthesis of [9] J.T. Singleton, Tetrahedron 59 (2003) 1837–1857.
RPCs, e.g., direct metallation, C/Si H bond activation, transmet- [10] D. Pugh, A.A. Danopoulos, Coord. Chem. Rev. 251 (2007) 610–641.
[11] D. Morales-Morales, C.M. Jensen, The Chemistry of Pincer Compounds, Else-
allation, and trans-cyclometallation. Regardless of whether the
vier, Amsterdam, 2007.
reaction occurs or not, many factors control the reaction prod- [12] D. Benito-Garagorri, K. Kirchner, Acc. Chem. Res. 41 (2008) 201–213.
uct, e.g., donor atoms, steric hindrance in the pincer ligand around [13] J.I. van der Vlugt, J.N. Reek, Angew. Chem. Int. Ed. 48 (2009) 8832–8846.
the donor atoms, the ability to displace the ligand in the ruthe- [14] N. Selander, K.l.n.J. Szabó, Chem. Rev. 111 (2010) 2048–2076.
[15] D. Milstein, Top. Catal. 53 (2010) 915–923.
nium source, and the reaction conditions. Direct metallation is used [16] J. Choi, A. MacArthur, M. Brookhart, A.S. Goldman, Chem. Rev. 111 (2011)
widely because of its applicability to most pincer ligands based on 1761.
pyridine-bearing heteroatoms on both side arms, e.g., PNP, PNN, [17] C. Gunanathan, D. Milstein, Acc. Chem. Res. 44 (2011) 588–602.
[18] M. Albrecht, M.M. Lindner, Dalton Trans. 40 (2011) 8733–8744.
PNS, PO NO P, PN NN P, PN NC N, and NNN. Numerous RPCs have been [19] J.-L. Niu, X.-Q. Hao, J.-F. Gong, M.-P. Song, Dalton Trans. 40 (2011) 5135–5150.
synthesized via CSP 2 H bond activation, but only a few examples [20] S. Schneider, J. Meiners, B. Askevold, Eur. J. Inorg. Chem. 2012 (2012) 412–429.
have been reported based on CSP 3 H activation. RPCs have also [21] G. van Koten, D. Milstein, A. Castonguay, Organometallic Pincer Chemistry,
Springer, 2013.
been prepared via Si H activation using a variety of ruthenium [22] W. Leis, H.A. Mayer, W.C. Kaska, Coord. Chem. Rev. 252 (2008) 1787–1797.
sources and pincer ligands. Bulky groups around the donor atoms [23] C.S. Creaser, W.C. Kaska, Inorg. Chim. Acta 30 (1978) L325–L326.
have a great impact on the ability to form a complex and Si H acti- [24] M. Kitamura, R. Noyori, Hydrogenation and transfer hydrogenation, in: Ruthe-
nium in Organic Synthesis, Wiley-VCH Verlag GmbH & Co. KGaA, 2005, pp.
vation might be facilitated by CSP 3 H activation. Transmetallation
3–52.
from NHC-ligated Ag(I) complexes and lithiated reagents, as well [25] S.-I. Murahashi, N. Komiya, Oxidation reactions, in: Ruthenium in Organic
as trans-cyclometallation, have been utilized successfully for the Synthesis, Wiley-VCH Verlag GmbH & Co. KGaA, 2005, pp. 53–93.
[26] R.F.R. Jazzar, E.P. Kündig, Ruthenium Lewis acid-catalyzed reactions, in:
preparation of RPCs.
Ruthenium in Organic Synthesis, Wiley-VCH Verlag GmbH & Co. KGaA, 2005,
RPCs have various catalytic applications, e.g., direct/indirect pp. 257–276.
hydrogenation reactions, acceptor/acceptorless oxidative dehydro- [27] Y. Yamamoto, K. Itoh, Carbon–Carbon bond formations via ruthenacycle inter-
genation reactions, amination of alcohols, hydrolysis of amines, mediates, in: Ruthenium in Organic Synthesis, Wiley-VCH Verlag GmbH & Co.
KGaA, 2005, pp. 95–128.
hydroboration of terminal alkynes, cyclopropanation of olefins, [28] E. Colacino, J. Martinez, F. Lamaty, Coord. Chem. Rev. 251 (2007) 726–764.
water splitting, alkynylation of carbonyl compounds with terminal [29] V. Dragutan, I. Dragutan, L. Delaude, A. Demonceau, Coord. Chem. Rev. 251
alkynes, transesterification, olefin metathesis, and alcohol racem- (2007) 765–794.
[30] R. Drozdzak, B. Allaert, N. Ledoux, I. Dragutan, V. Dragutan, F. Verpoort, Coord.
ization. In addition, electronic and steric effects are two main Chem. Rev. 249 (2005) 3055–3074.
factors that have significant influences on the catalytic activities [31] F.B. Hamad, T. Sun, S. Xiao, F. Verpoort, Coord. Chem. Rev. 257 (2013)
of these superior types of pincer complexes during hydrogena- 2274–2292.
[32] T. Naota, H. Takaya, S.-I. Murahashi, Chem. Rev. 98 (1998) 2599–2660.
tion and dehydrogenation reactions, due to: (i) the hemilability [33] S.-I. Murahashi, Ruthenium in Organic Synthesis, Wiley, 2006.
and non-innocent behavior the pincer ligand, and (ii) the flexibil- [34] C. Gunanathan, D. Milstein, Science (2013) 341.
ity of the pincer framework (aliphatic versus aromatic). The pincer [35] J.M. Serrano-Becerra, D. Morales-Morales, Curr. Org. Synth. 6 (2009) 169–192.
150 H.A. Younus et al. / Coordination Chemistry Reviews 276 (2014) 112–152

[36] K. Szabó, Pincer complexes as catalysts in organic chemistry, in: G. [87] U.S. Schubert, C. Eschbaumer, M. Heller, Org. Lett. 2 (2000) 3373–3376.
van Koten, D. Milstein (Eds.), Organometallic Pincer Chemistry, Springer, [88] M. Heller, U.S. Schubert, J. Org. Chem. 67 (2002) 8269–8272.
Berlin/Heidelberg, 2013, pp. 203–241. [89] S.A. Savage, A.P. Smith, C.L. Fraser, J. Org. Chem. 63 (1998) 10048–10051.
[37] S. Sarkar, K.P. McGowan, S. Kuppuswamy, I. Ghiviriga, K.A. Abboud, A.S. Veige, [90] R. Barrios-Francisco, E. Balaraman, Y. Diskin-Posner, G. Leitus, L.J. Shimon, D.
J. Am. Chem. Soc. 134 (2012) 4509–4512. Milstein, Organometallics 32 (2013) 2973–2982.
[38] L. Luconi, J. Klosin, A.J. Smith, S. Germain, E. Schulz, J. Hannedouche, G. [91] R.A. Altman, S.L. Buchwald, Org. Lett. 9 (2007) 643–646.
Giambastiani, Dalton Trans. 42 (2013) 16056–16065. [92] D. Spasyuk, D.G. Gusev, Organometallics 31 (2012) 5239–5242.
[39] G.R. Freeman, J.G. Williams, Metal complexes of pincer ligands: excited states, [93] T. Chen, L.-P. He, D. Gong, L. Yang, X. Miao, J. Eppinger, K.-W. Huang, Tetrahe-
photochemistry, and luminescence, in: Organometallic Pincer Chemistry, dron Lett. 53 (2012) 4409–4412.
Springer, 2013, pp. 89–129. [94] G. Zeng, T. Chen, L. He, I. Pinnau, Z. Lai, K.W. Huang, Chem. Eur. J. 18 (2012)
[40] C.K. Prier, D.A. Rankic, D.W. MacMillan, Chem. Rev. 113 (2013) 5322–5363. 15940–15943.
[41] M. Albrecht, R.A. Gossage, M. Lutz, A.L. Spek, G. van Koten, Chem. Eur. J. 6 [95] L.-P. He, T. Chen, D. Gong, Z. Lai, K.-W. Huang, Organometallics 31 (2012)
(2000) 1431–1445. 5208–5211.
[42] M. Albrecht, M. Schlupp, J. Bargon, G.v. Koten, Chem. Commun. (2001) [96] M. Beller, C. Bolm, Transition Metals for Organic Synthesis: Building Blocks
1874–1875. and Fine Chemicals, 2 Volume Set, Wiley-VCH, New York, 2004.
[43] M. Schlupp, T. Weil, A.J. Berresheim, U.M. Wiesler, J. Bargon, K. Müllen, Angew. [97] J.C.C. Chen, I.J.B. Lin, J. Chem. Soc., Dalton Trans. (2000) 839–840.
Chem. Int. Ed. 40 (2001) 4011–4015. [98] A.A. Danopoulos, D. Pugh, H. Smith, J. Saßmannshausen, Chem. Eur. J. 15
[44] M.-C. Lagunas, R.A. Gossage, A.L. Spek, G. van Koten, Organometallics 17 (2009) 5491–5502.
(1998) 731–741. [99] A.A. Danopoulos, D. Pugh, J.A. Wright, Angew. Chem. Int. Ed. 47 (2008)
[45] M. Albrecht, G. Rodríguez, J. Schoenmaker, G. van Koten, Org. Lett. 2 (2000) 9765–9767.
3461–3464. [100] E. Peris, J.A. Loch, J. Mata, R.H. Crabtree, Chem. Commun. (2001) 201–202.
[46] I. Moreno, R. SanMartin, B. Ines, M.T. Herrero, E. Domínguez, Curr. Org. Chem. [101] S. Gründemann, M. Albrecht, J.A. Loch, J.W. Faller, R.H. Crabtree,
13 (2009) 878–895. Organometallics 20 (2001) 5485–5488.
[47] M.R. Eberhard, S. Matsukawa, Y. Yamamoto, C.M. Jensen, J. Organomet. Chem. [102] F.E. Hahn, M.C. Jahnke, T. Pape, Organometallics 26 (2007) 150–154.
687 (2003) 185–189. [103] A.A. Danopoulos, S. Winston, W.B. Motherwell, Chem. Commun. (2002)
[48] F. Zeng, Z. Yu, Organometallics 27 (2008) 2898–2901. 1376–1377.
[49] C. Gunanathan, Y. Ben-David, D. Milstein, Science 317 (2007) 790–792. [104] M.r. Boronat, A. Corma, C. González-Arellano, M. Iglesias, F. Sánchez,
[50] B. Gnanaprakasam, J. Zhang, D. Milstein, Angew. Chem. Int. Ed. 49 (2010) Organometallics 29 (2009) 134–141.
1468–1471. [105] C. del Pozo, A. Corma, M. Iglesias, F.l. Sánchez, Organometallics 29 (2010)
[51] H. Nishiyama, Chem. Soc. Rev. 36 (2007) 1133–1141. 4491–4498.
[52] H. Nishiyama, J.-i. Ito, Chem. Commun. 46 (2010) 203–212. [106] L. Zhang, X. Chen, P. Xue, H.H. Sun, I.D. Williams, K.B. Sharpless, V.V. Fokin, G.
[53] O.A. Wallner, V.J. Olsson, L. Eriksson, K.J. Szabó, Inorg. Chim. Acta 359 (2006) Jia, J. Am. Chem. Soc. 127 (2005) 15998–15999.
1767–1772. [107] S. Sinn, B. Schulze, C. Friebe, D.G. Brown, M. Jäger, E. Altuntaş, J. Kübel, O.
[54] D. Morales-Morales, R.E. Cramer, C.M. Jensen, J. Organomet. Chem. 654 (2002) Guntner, C.P. Berlinguette, B. Dietzek, U.S. Schubert, Inorg. Chem. 53 (2014)
44–50. 2083–2095.
[55] W. Ye, M. Zhao, W. Du, Q. Jiang, K. Wu, P. Wu, Z. Yu, Chem. Eur. J. 17 (2011) [108] C. del Pozo, M. Iglesias, F.l. Sánchez, Organometallics 30 (2011) 2180–2188.
4737–4741. [109] E.E. Fogler, Y. Balaraman, G. Ben-David, L. Leitus, J.W. Shimon, D. Milstein,
[56] C. Mazet, L.H. Gade, Chem. Eur. J. 9 (2003) 1759–1767. Organometallics 30 (2011) 3826–3833.
[57] O. Vechorkin, V. Proust, X. Hu, Am. Chem. Soc. J. 131 (2009) 9756–9766. [110] F. Zeng, Z. Yu, Organometallics 27 (2008) 6025–6028.
[58] M. Stradiotto, K.L. Fujdala, T.D. Tilley, Chem. Commun. (2001) 1200–1201. [111] Sun X.-J., Yu Z.-K., Wu S.-Z., W.-J. Xiao, Organometallics 24 (2005) 2959–2963.
[59] J. Takaya, N. Iwasawa, J. Am. Chem. Soc. 130 (2008) 15254–15255. [112] F. Zeng, Z. Yu, J. Org. Chem. 71 (2006) 5274–5281.
[60] E. Morgan, D.F. MacLean, R. McDonald, L. Turculet, J. Am. Chem. Soc. 131 [113] J.M. Fraile, J.I. García, J.A. Mayoral, Coord. Chem. Rev. 252 (2008) 624–646.
(2009) 14234–14236. [114] V.C. Gibson, C. Redshaw, G.A. Solan, Chem. Rev. 107 (2007) 1745–1776.
[61] E.E. Korshin, G. Leitus, L.J. Shimon, L. Konstantinovski, D. Milstein, Inorg. Chem. [115] S.D. Cummings, Coord. Chem. Rev. 253 (2009) 449–478.
47 (2008) 7177–7189. [116] D.L. Jameson, K.A. Goldsby, J. Org. Chem. 55 (1990) 4992–4994.
[62] M. Blug, M. Doux, X. Le Goff, P. Maître, F.o. Ribot, P. Le Floch, N. Mézailles, [117] N.J. Beach, G.J. Spivak, Inorg. Chim. Acta 343 (2003) 244–252.
Organometallics 28 (2009) 2020–2027. [118] A.R. Karam, E.L. Catarí, F. López-Linares, G. Agrifoglio, C.L. Albano, A. Díaz-
[63] E.J. Derrah, C. Martin, S. Mallet-Ladeira, K. Miqueu, G. Bouhadir, D. Bourissou, Barrios, T.E. Lehmann, S.V. Pekerar, L.A. Albornoz, R. Atencio, Appl. Catal. A:
Organometallics 32 (2013) 1121–1128. Gen. 280 (2005) 165–173.
[64] Y. Segawa, M. Yamashita, K. Nozaki, J. Am. Chem. Soc. 131 (2009) 9201–9203. [119] Z.R. Reeves, K.L.V. Mann, J.C. Jeffery, J.A. McCleverty, M.D. Ward, F. Barigelletti,
[65] H. Ogawa, M. Yamashita, Dalton Trans. 42 (2013) 625–629. N. Armaroli, J. Chem. Soc., Dalton Trans. (1999) 349–356.
[66] J. Borau-Garcia, Pincer Ligands with Strong ␴-Donors at the Central Posi- [120] J.S. Fleming, E. Psillakis, S.M. Couchman, J.C. Jeffery, J.A. McCleverty, M.D.
tion.PhD, University of Calgary, 4-5. (2013), http://hdl.handle.net/11023/423 Ward, J. Chem. Soc., Dalton Trans. (1998) 537–544.
[67] W.V. Dahlhoff, S.M. Nelson, J. Chem. Soc. A (1971) 2184–2190. [121] W. Jin, L. Wang, Z. Yu, Organometallics 31 (2012) 5664–5667.
[68] D. Hermann, M. Gandelman, H. Rozenberg, L.J. Shimon, D. Milstein, [122] F. Zeng, Z. Yu, Organometallics 28 (2009) 1855–1862.
Organometallics 21 (2002) 812–818. [123] M. Zhao, Z. Yu, S. Yan, Y. Li, J. Organomet. Chem. 694 (2009) 3068–3075.
[69] G. Müller, M. Klinga, M. Leskelä, B. Rieger, Z. Anorg, Allg. Chem. 628 (2002) [124] W. Ye, M. Zhao, Z. Yu, Chem. Eur. J. 18 (2012) 10843–10846.
2839–2846. [125] W. Du, P. Wu, Q. Wang, Z. Yu, Organometallics 32 (2013) 3083–3090.
[70] M. Kawatsura, J.F. Hartwig, Organometallics 20 (2001) 1960–1964. [126] W. Baratta, P. Da Ros, A. Del Zotto, A. Sechi, E. Zangrando, P. Rigo, Angew.
[71] C. Scriban, D.S. Glueck, J. Am. Chem. Soc. 128 (2006) 2788–2789. Chem. Int. Ed. 43 (2004) 3584–3588.
[72] W.H. Bernskoetter, S.K. Hanson, S.K. Buzak, Z. Davis, P.S. White, R. Swartz, K.I. [127] Z. Huo, T. Kosugi, Y. Yamamoto, Acta. Chim. Slov. 56 (2009) 659–663.
Goldberg, M. Brookhart, J. Am. Chem. Soc. 131 (2009) 8603–8613. [128] C. Liu, J. Luo, L. Xu, Z. Huo, ARKIVOC 1 (2013) 154–174.
[73] H. Salem, L.J. Shimon, Y. Diskin-Posner, G. Leitus, Y. Ben-David, D. Milstein, [129] W. Baratta, M. Bosco, G. Chelucci, A. Del Zotto, K. Siega, M. Toniutti, E. Zan-
Organometallics 28 (2009) 4791–4806. grando, P. Rigo, Organometallics 25 (2006) 4611–4620.
[74] S. Kundu, W.W. Brennessel, W.D. Jones, Inorg. Chem. 50 (2011) 9443–9453. [130] S.H. Pine, B.L. Sanchez, J. Org. Chem. 36 (1971) 829–832.
[75] D. Benito-Garagorri, E. Becker, J. Wiedermann, W. Lackner, M. Pollak, K. Mere- [131] C.-l. Chuang, K. Lim, Q. Chen, J. Zubieta, J.W. Canary, Inorg. Chem. 34 (1995)
iter, J. Kisala, K. Kirchner, Organometallics 25 (2006) 1900–1913. 2562–2568.
[76] D. Benito-Garagorri, V. Bocokić, K. Mereiter, K. Kirchner, Organometallics 25 [132] D. Song, R.H. Morris, Organometallics 23 (2004) 4406–4413.
(2006) 3817–3823. [133] G. Chelucci, S. Baldino, R. Solinas, W. Baratta, Tetrahedron Lett. 46 (2005)
[77] D. Benito-Garagorri, J. Wiedermann, M. Pollak, K. Mereiter, K. Kirchner, 5555–5558.
Organometallics 26 (2007) 217–222. [134] G. Chelucci, S. Baldino, S. Chessa, G.A. Pinna, F. Soccolini, Tetrahedron: Asym-
[78] B. Bichler, C. Holzhacker, B. Stöger, M. Puchberger, L.F. Veiros, K. Kirchner, metry 17 (2006) 3163–3169.
Organometallics 32 (2013) 4114–4121. [135] W. Du, L. Wang, P. Wu, Z. Yu, Chem. Eur. J. 18 (2012) 11550–11554.
[79] W. Schirmer, U. Flörke, H.J. Haupt, Z. Anorg, Allg. Chem. 545 (1987) 83–97. [136] W. Baratta, F. Benedetti, A. Del Zotto, L. Fanfoni, F. Felluga, S. Magnolia, E.
[80] W. Schirmer, U. Flörke, H.J. Haupt, Z. Anorg, Allg. Chem. 574 (1989) 239–255. Putignano, P. Rigo, Organometallics 29 (2010) 3563–3570.
[81] S.W. Kohl, L. Weiner, L. Schwartsburd, L. Konstantinovski, L.J. Shimon, Y. Ben- [137] W. Baratta, M. Ballico, S. Baldino, G. Chelucci, E. Herdtweck, K. Siega, S. Mag-
David, M.A. Iron, D. Milstein, Science 324 (2009) 74–77. nolia, P. Rigo, Chem. Eur. J. 14 (2008) 9148–9160.
[82] J. Zhang, M. Gandelman, D. Herrman, G. Leitus, L.J. Shimon, Y. Ben-David, D. [138] M. Schlosser, F. Cottet, Eur. J. Org. Chem. 2002 (2002) 4181–4184.
Milstein, Inorg. Chim. Acta 359 (2006) 1955–1960. [139] F. Felluga, W. Baratta, L. Fanfoni, G. Pitacco, P. Rigo, F. Benedetti, J. Org. Chem.
[83] J. Zhang, G. Leitus, Y. Ben-David, D. Milstein, J. Am. Chem. Soc. 127 (2005) 74 (2009) 3547–3550.
10840–10841. [140] D.G. Gusev, M. Madott, F.M. Dolgushin, K.A. Lyssenko, M.Y. Antipin,
[84] M. Gargir, Y. Ben-David, G. Leitus, Y. Diskin-Posner, L.J. Shimon, D. Milstein, Organometallics 19 (2000) 1734–1739.
Organometallics 31 (2012) 6207–6214. [141] D. Amoroso, A. Jabri, G.P.A. Yap, D.G. Gusev, E.N. dos Santos, D.E. Fogg,
[85] M. Gagliardo, N. Selander, N.C. Mehendale, G. van Koten, R.J. Klein Gebbink, Organometallics 23 (2004) 4047–4054.
K.J. Szabo, Chem. Eur. J. 14 (2008) 4800–4809. [142] M.E. van der Boom, H.-B. Kraatz, L. Hassner, Y. Ben-David, D. Milstein,
[86] C. Kaes, A. Katz, M.W. Hosseini, Chem. Rev. 100 (2000) 3553–3590. Organometallics 18 (1999) 3873–3884.
H.A. Younus et al. / Coordination Chemistry Reviews 276 (2014) 112–152 151

[143] D.M. Roddick, Tuning of PCP pincer ligand electronic and steric properties, in: [196] M.L. Clarke, M.B. Díaz-Valenzuela, A.M. Slawin, Organometallics 26 (2007)
Organometallic Pincer Chemistry, Springer, 2013, pp. 49–88. 16–19.
[144] P. Steenwinkel, J.T.B.H. Jastrzebski, B.-J. Deelman, D.M. Grove, H. Kooijman, [197] Y. Sun, C. Koehler, R. Tan, V.T. Annibale, D. Song, Chem. Commun. 47 (2011)
N. Veldman, W.J.J. Smeets, A.L. Spek, G. van Koten, Organometallics 16 (1997) 8349–8351.
5486–5498. [198] E. Balaraman, E. Fogler, D. Milstein, Chem. Commun. 48 (2012) 1111–1113.
[145] H.C. Kolb, M. Finn, K.B. Sharpless, Angew. Chem. Int. Ed. 40 (2001) 2004–2021. [199] J.A. Wright, A.A. Danopoulos, W.B. Motherwell, R.J. Carroll, S. Ellwood, J.
[146] W.-W. Yang, L. Wang, Y.-W. Zhong, J. Yao, Organometallics 30 (2011) Organomet. Chem. 691 (2006) 5204–5210.
2236–2240. [200] H. Deng, Z. Yu, J. Dong, S. Wu, Organometallics 24 (2005) 4110–4112.
[147] L. McKinstry, T. Livinghouse, Tetrahedron 51 (1995) 7655–7666. [201] D.-H. Jo, H.-J. Yeo, Bull. Korean Chem. Soc. 14 (1993) 682.
[148] J.M. Longmire, X. Zhang, M. Shang, Organometallics 17 (1998) 4374–4379. [202] C.A. Bessel, R.F. See, D.L. Jameson, M.R. Churchill, K.J. Takeuchi, J. Chem. Soc.,
[149] H. Nishiyama, H. Sakaguchi, T. Nakamura, M. Horihata, M. Kondo, K. Itoh, Dalton Trans. (1993) 1563–1576.
Organometallics 8 (1989) 846–848. [203] M. Zhao, Z. Yu, S. Yan, Y. Li, Tetrahedron Lett. 50 (2009) 4624–4628.
[150] J.-i. Ito, H. Nishiyama, Synlett (2012) 509–523. [204] T. Jozak, D. Zabel, A. Schubert, Y. Sun, W.R. Thiel, Eur. J. Inorg. Chem. 2010
[151] H. Nishiyama, J.-i. Ito, T. Shiomi, T. Hashimoto, T. Miyakawa, M. Kitase, Pure (2010) 5135–5145.
Appl. Chem. 80 (2008) 743–749. [205] A. Yoshinari, A. Tazawa, S. Kuwata, T. Ikariya, Chem. Asian J. 7 (2012)
[152] D. Benito-Garagorri, V. Bocokic, K. Mereiter, K. Kirchner, Organometallics 25 1417–1425.
(2006) 3817–3823. [206] L.-P. He, T. Chen, D.-X. Xue, M. Eddaoudi, K.-W. Huang, Organomet. Chem. J.
[153] D. Morales-Morales, C. Grause, K. Kasaoka, R.o. Redón, R.E. Cramer, C.M. 700 (2012) 202–206.
Jensen, Inorg. Chim. Acta 300 (2000) 958–963. [207] D. Amoroso, A. Jabri, G.P. Yap, D.G. Gusev, E.N. dos Santos, D.E. Fogg,
[154] D. Morales-Morales, Mini-Rev. Org. Chem. 5 (2008) 141–152. Organometallics 23 (2004) 4047–4054.
[155] R.B. Bedford, M. Betham, M.E. Blake, S.J. Coles, S.M. Draper, M.B. Hursthouse, [208] R. Cerón-Camacho, V. Gómez-Benítez, R. Le Lagadec, D. Morales-Morales, R.A.
P.N. Scully, Inorg. Chim. Acta 359 (2006) 1870–1878. Toscano, J. Mol. Catal. A: Chem. 247 (2006) 124–129.
[156] R. Akabane, M. Tanaka, K. Matsuo, N. Koga, K. Matsuda, H. Iwamura, J. Org. [209] J.-i. Ito, S. Ujiie, H. Nishiyama, Organometallics 28 (2008) 630–638.
Chem. 62 (1997) 8854–8861. [210] J.-i. Ito, S. Ujiie, H. Nishiyama, Chem. Commun. (2008) 1923–1925.
[157] V.C. Vargas, R.J. Rubio, T.K. Hollis, M.E. Salcido, Org. Lett. 5 (2003) 4847–4849. [211] J.-i. Ito, R. Asai, H. Nishiyama, Org. Lett. 12 (2010) 3860–3862.
[158] H.-J. Cristau, P.P. Cellier, J.-F. Spindler, M. Taillefer, Chem. Eur. J. 10 (2004) [212] J.i. Ito, S. Ujiie, H. Nishiyama, Chem. Eur. J. 16 (2010) 4986–4990.
5607–5622. [213] J.i. Ito, K. Fujii, H. Nishiyama, Chem. Eur. J. 19 (2013) 601–605.
[159] R.A. Altman, E.D. Koval, S.L. Buchwald, J. Org. Chem. 72 (2007) 6190–6199. [214] J.C. Garrison, W.J. Youngs, Chem. Rev. 105 (2005) 3978–4008.
[160] L. Liang, Z. Li, X. Zhou, Org. Lett. 11 (2009) 3294–3297. [215] S.U. Son, K.H. Park, Y.-S. Lee, B.Y. Kim, C.H. Choi, M.S. Lah, Y.H. Jang, D.-J. Jang,
[161] D. Wang, F. Zhang, D. Kuang, J. Yu, J. Li, Green Chem. 14 (2012) 1268–1271. Y.K. Chung, Inorg. Chem. 43 (2004) 6896–6898.
[162] S. Ganesh Babu, R. Karvembu, Ind. Eng. Chem. Res. 50 (2011) 9594–9600. [216] M. Poyatos, J.A. Mata, E. Falomir, R.H. Crabtree, E. Peris, Organometallics 22
[163] Y. Zhu, Y. Shi, Y. Wei, Monatsh. Chem./Chem. Mon. 141 (2010) 1009–1013. (2003) 1110–1114.
[164] A. Rit, T. Pape, A. Hepp, F.E. Hahn, Organometallics 30 (2010) 334–347. [217] W. Baratta, G. Chelucci, S. Gladiali, K. Siega, M. Toniutti, M. Zanette, E. Zan-
[165] T.O. Howell, A.J. Huckaba, T.K. Hollis, Org. Lett. 16 (2014) 2570–2572. grando, P. Rigo, Angew. Chem. Int. Ed. 44 (2005) 6214–6219.
[166] D. Gelman, R. Romm, PC (sp 3) P transition metal pincer complexes: proper- [218] W. Baratta, G. Chelucci, S. Magnolia, K. Siega, P. Rigo, Chem. Eur. J. 15 (2009)
ties and catalytic applications, in: Organometallic Pincer Chemistry, Springer, 726–732.
2013, pp. 289–317. [219] W. Baratta, L. Fanfoni, S. Magnolia, K. Siega, P. Rigo, Eur. J. Inorg. Chem. 2010
[167] S. Musa, I. Shaposhnikov, S. Cohen, D. Gelman, Angew. Chem. Int. Ed. 50 (2011) (2010) 1419–1423.
3533–3537. [220] S. Zhang, W. Baratta, Organometallics 32 (2013) 3339–3342.
[168] C. Gunanathan, D. Milstein, Angew. Chem. Int. Ed. 47 (2008) 8661–8664. [221] K. Nagaraju, S. Pal, J. Organomet. Chem. 737 (2013) 7–11.
[169] Y.-H. Li, X.-H. Ding, Y. Zhang, W.-R. He, W. Huang, Inorg. Chem. Commun. 15 [222] K. Nagaraju, R. Raveendran, S. Pal, S. Pal, Polyhedron 33 (2012) 52–59.
(2012) 194–197. [223] Y.-M. Zhang, J.-Y. Shao, C.-J. Yao, Y.-W. Zhong, Dalton Trans. 41 (2012)
[170] A.J. Ruddy, S.J. Mitton, R. McDonald, L. Turculet, Chem. Commun. 48 (2012) 9280–9282.
1159–1161. [224] A.R. Naziruddin, Z.-J. Huang, W.-C. Lai, W.-J. Lin, W.-S. Hwang, Dalton Trans.
[171] S. Lachaize, L. Vendier, S. Sabo-Etienne, Dalton Trans. 39 (2010) 8492–8500. 42 (2013) 13161–13172.
[172] V. Montiel-Palma, M.A. Muñoz-Hernández, C.A. Cuevas-Chávez, L. Vendier, [225] L.-H. Chung, K.-S. Cho, J. England, S.-C. Chan, K. Wieghardt, C.-Y. Wong, Inorg.
M. Grellier, S. Sabo-Etienne, Inorg. Chem. 52 (2013) 9798–9806. Chem. 52 (2013) 9885–9896.
[173] H.M. Lee, J.Y. Zeng, C.-H. Hu, M.-T. Lee, Inorg. Chem. 43 (2004) 6822–6829. [226] V.F. Kuznetsov, K. Abdur-Rashid, A.J. Lough, D.G. Gusev, J. Am. Chem. Soc. 128
[174] W. Weng, S. Parkin, O.V. Ozerov, Organometallics 25 (2006) 5345–5354. (2006) 14388–14396.
[175] A.M. Winter, K. Eichele, H.-G. Mack, S. Potuznik, H.A. Mayer, W.C. Kaska, J. [227] S. Musa, S. Fronton, L. Vaccaro, D. Gelman, Organometallics 32 (2013)
Organomet. Chem. 682 (2003) 149–154. 3069–3073.
[176] R. Çelenligil-Çetin, L.A. Watson, C. Guo, B.M. Foxman, O.V. Ozerov, [228] J. Zhu, Z. Lin, T.B. Marder, Inorg. Chem. 44 (2005) 9384–9390.
Organometallics 24 (2004) 186–189. [229] H.M. Wang, I.J. Lin, Organometallics 17 (1998) 972–975.
[177] L.A. Watson, J.N. Coalter Iii, O. Ozerov, M. Pink, J.C. Huffman, K.G. Caulton, New [230] T.K. Maishal, J.-M. Basset, M. Boualleg, C. Coperet, L. Veyre, C. Thieuleux, Dalton
J. Chem. 27 (2003) 263–273. Trans. (2009) 6956–6959.
[178] L.-C. Liang, Coord. Chem. Rev. 250 (2006) 1152–1177. [231] C. del Pozo, A. Corma, M. Iglesias, F. Sánchez, Green Chem. 13 (2011)
[179] Z. Csok, O. Vechorkin, S.B. Harkins, R. Scopelliti, X. Hu, J. Am. Chem. Soc. 130 2471–2481.
(2008) 8156–8157. [232] C. del Pozo, A. Corma, M. Iglesias, F. Sánchez, J. Catal. 291 (2012) 110–116.
[180] O. Vechorkin, V. Proust, X. Hu, Angew. Chem. Int. Ed. 49 (2010) 3061–3064. [233] M. Hernández-Juárez, M. Vaquero, E. Álvarez, V. Salazar, A. Suárez, Dalton
[181] J. Zhang, M. Gandelman, L.J. Shimon, H. Rozenberg, D. Milstein, Trans. 42 (2013) 351–354.
Organometallics 23 (2004) 4026–4033. [234] P.L. Chiu, H.M. Lee, Organometallics 24 (2005) 1692–1702.
[182] C. Gunanathan, B. Gnanaprakasam, M.A. Iron, L.J. Shimon, D. Milstein, J. Am. [235] P. Ren, O. Vechorkin, Z. Csok, I. Salihu, R. Scopelliti, X. Hu, Dalton Trans. 40
Chem. Soc. 132 (2010) 14763–14765. (2011) 8906–8911.
[183] J.R. Khusnutdinova, Y. Ben-David, D. Milstein, J. Am. Chem. Soc. 136 (2014) [236] M. Albrecht, P. Dani, M. Lutz, A.L. Spek, G. van Koten, J. Am. Chem. Soc. 122
2998–3001. (2000) 11822–11833.
[184] M.H.G. Prechtl, Y. Ben-David, D. Giunta, S. Busch, Y. Taniguchi, W. Wisniewski, [237] P. Dani, T. Karlen, R.A. Gossage, W.J. Smeets, A.L. Spek, G. van Koten, J. Am.
H. Görls, R.J. Mynott, N. Theyssen, D. Milstein, W. Leitner, Chem. Eur. J. 13 Chem. Soc. 119 (1997) 11317–11318.
(2007) 1539–1546. [238] P. Dani, M. Albrecht, G.P. van Klink, G. van Koten, Organometallics 19 (2000)
[185] S. Busch, W. Leitner, Chem. Commun. (1999) 2305–2306. 4468–4476.
[186] M.H. Prechtl, M. Hölscher, Y. Ben-David, N. Theyssen, R. Loschen, D. Milstein, [239] M. Gagliardo, P.A. Chase, S. Brouwer, G.P. van Klink, G. van Koten,
W. Leitner, Angew. Chem. Int. Ed. 46 (2007) 2269–2272. Organometallics 26 (2007) 2219–2227.
[187] M.H. Prechtl, K. Wobser, N. Theyssen, Y. Ben-David, D. Milstein, W. Leitner, [240] H.P. Dijkstra, M. Albrecht, G.v. Koten, Chem. Commun. (2002) 126–127.
Catal. Sci. Technol. 2 (2012) 2039–2042. [241] H.P. Dijkstra, M. Albrecht, S. Medici, G.P.M. van Klink, G. van Koten, Adv. Synth.
[188] J. Zhang, G. Leitus, Y. Ben-David, D. Milstein, Angew. Chem. Int. Ed. 45 (2006) Catal. 344 (2002) 1135–1141.
1113–1115. [242] M. Gagliardo, P.A. Chase, M. Lutz, A.L. Spek, F. Hartl, R.W. Havenith, G.P. van
[189] E. Balaraman, B. Gnanaprakasam, L.J. Shimon, D. Milstein, J. Am. Chem. Soc. Klink, G. van Koten, Organometallics 24 (2005) 4553–4557.
132 (2010) 16756–16758. [243] P.A. Dub, T. Ikariya, ACS Catal. 2 (2012) 1718–1741.
[190] E. Balaraman, C. Gunanathan, J. Zhang, L.J. Shimon, D. Milstein, Nat. Chem. 3 [244] F. Klasovsky, J.C., Pfeffer, T., Tacke, T., Haas, A., Martin, J., Deutsch, A. Köckritz,
(2011) 609–614. WO Patent 2,012,031,884 (2012).
[191] E. Balaraman, Y. Ben-David, D. Milstein, Angew. Chem. 123 (2011) [245] J.R. Khusnutdinova, Y. Ben-David, D. Milstein, Angew. Chem. 125 (2013)
11906–11909. 6389–6392.
[192] D. Spasyuk, S. Smith, D.G. Gusev, Angew. Chem. Int. Ed. 51 (2012) 2772–2775. [246] E. Milczek, N. Boudet, S. Blakey, Angew. Chem. Int. Ed. 47 (2008) 6825–6828.
[193] D. Spasyuk, S. Smith, D.G. Gusev, Angew. Chem. 125 (2013) 2598–2602. [247] C. Gunanathan, M. Hölscher, F. Pan, W. Leitner, J. Am. Chem. Soc. 134 (2012)
[194] M. Bertoli, A. Choualeb, A.J. Lough, B. Moore, D. Spasyuk, D.G. Gusev, 14349–14352.
Organometallics 30 (2011) 3479–3482. [248] B. Gnanaprakasam, Y. Ben-David, D. Milstein, Adv. Synth. Catal. 352 (2010)
[195] D. Spasyuk, S. Smith, D.G. Gusev, Angew. Chem. Int. Ed. 52 (2013) 2538–2542. 3169–3173.
152 H.A. Younus et al. / Coordination Chemistry Reviews 276 (2014) 112–152

[249] G. Bossi, E. Putignano, P. Rigo, W. Baratta, Dalton Trans. 40 (2011) 8986–8995. [259] N.J. Oldenhuis, V.M. Dong, Z. Guan, Tetrahedron 70 (2014) 4213–4218.
[250] P. Braunstein, F. Naud, Angew. Chem. Int. Ed. 40 (2001) 680–699. [260] X. Yang, ACS Catal. 3 (2013) 2684–2688.
[251] B. Gnanaprakasam, D. Milstein, J. Am. Chem. Soc. 133 (2011) 1682–1685. [261] X. Yang, ACS Catal. 4 (2014) 1129–1133.
[252] B. Gnanaprakasam, E. Balaraman, Y. Ben-David, D. Milstein, Angew. Chem. Int. [262] J.C. Bayón, C. Claver, A.M. Masdeu-Bultó, Coord. Chem. Rev. 193–195 (1999)
Ed. 50 (2011) 12240–12244. 73–145.
[253] L.U. Nordstrøm, H. Vogt, R. Madsen, J. Am. Chem. Soc. 130 (2008) [263] J. Zhang, E. Balaraman, G. Leitus, D. Milstein, Organometallics 30 (2011)
17672–17673. 5716–5724.
[254] N.D. Schley, G.E. Dobereiner, R.H. Crabtree, Organometallics 30 (2011) [264] Z. Han, L. Rong, J. Wu, L. Zhang, Z. Wang, K. Ding, Angew. Chem. Int. Ed. 51
4174–4179. (2012) 13041–13045.
[255] D. Srimani, E. Balaraman, B. Gnanaprakasam, Y. Ben-David, D. Milstein, Adv. [265] T. Otsuka, A. Ishii, P.A. Dub, T. Ikariya, J. Am. Chem. Soc. 135 (2013) 9600–9603.
Synth. Catal. 354 (2012) 2403–2406. [266] P. Dani, T. Karlen, R.A. Gossage, S. Gladiali, G. van Koten, Angew. Chem. 112
[256] M. Nielsen, A. Kammer, D. Cozzula, H. Junge, S. Gladiali, M. Beller, Angew. (2000) 759–761.
Chem. Int. Ed. 50 (2011) 9593–9597. [267] W. Baratta, M. Bosco, G. Chelucci, A. Del Zotto, K. Siega, M. Toniutti, E. Zan-
[257] M. Nielsen, E. Alberico, W. Baumann, H.-J. Drexler, H. Junge, S. Gladiali, M. grando, P. Rigo, Organometallics 25 (2006) 4611–4620.
Beller, Nature 495 (2013) 85–89. [268] S. Zhang, W. Baratta, Organometallics 32 (2013) 3339–3342.
[258] W. Kuriyama, T. Matsumoto, O. Ogata, Y. Ino, K. Aoki, S. Tanaka, K. [269] D. Cuervo, M.P. Gamasa, J. Gimeno, Chem. Eur. J. 10 (2004) 425–432.
Ishida, T. Kobayashi, N. Sayo, T. Saito, Org. Process Res. Dev. 16 (2011) [270] D. Cuervo, E. Menéndez-Pedregal, J. Díez, M.P. Gamasa, J. Organomet. Chem.
166–171. 696 (2011) 1861–1867.

You might also like