0% found this document useful (0 votes)
33 views29 pages

Transition-Metal-Catalyzed Suzuki-Miyaura Cross-Coupling

Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
33 views29 pages

Transition-Metal-Catalyzed Suzuki-Miyaura Cross-Coupling

Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 29

Chem Soc Rev

View Article Online


REVIEW ARTICLE
Published on 04 March 2013. Downloaded by UNIVERSIDADE FEDERAL SAO CARLOS on 27/11/2013 19:03:38.

View Journal | View Issue

Transition-metal-catalyzed Suzuki–Miyaura cross-coupling


Cite this: Chem. Soc. Rev., 2013,
reactions: a remarkable advance from palladium to
42, 5270
nickel catalysts
Fu-She Han*ab

In the transition-metal-catalyzed cross-coupling reactions, the use of the first row transition metals as catalysts
is much more appealing than the precious metals owing to the apparent advantages such as cheapness and
earth abundance. Within the last two decades, particularly the last five years, explosive interests have been
focused on the nickel-catalyzed Suzuki–Miyaura reactions. This has greatly advanced the chemistry of
transition-metal-catalyzed cross-coupling reactions. Most notably, a broad range of aryl electrophiles such as
Received 20th December 2012 phenols, aryl ethers, esters, carbonates, carbamates, sulfamates, phosphates, phosphoramides, phosphonium
DOI: 10.1039/c3cs35521g salts, and fluorides, as well as various alkyl electrophiles, which are conventionally challenging, by applying
palladium catalysts can now be coupled efficiently with boron reagents in the presence of nickel catalysts. In
www.rsc.org/csr this review, we would like to summarize the progress in this reaction.

a
Changchun Institute of Applied Chemistry, Chinese Academy of Sciences,
5625 Renmin Street, Changchun, Jilin 130022, P. R. China.
1. Introduction
E-mail: [email protected]; Tel: +86-431-85262936
b
The transition-metal-catalyzed cross-coupling reactions such as
State Key Laboratory of Fine Chemicals, Dalian University of Technology,
Dalian 116024, P. R. China
Suzuki–Miyaura,1 Heck,2 Stille,3 Kumada,4 Negishi,5 Hiyama,6
and Sonogashira7 reactions are very powerful tools for the
Fu-She Han received his BS creation of C–C bonds. These methods have been extensively
degree in chemistry from used in a rich range of academic areas including natural
Chongqing university in 1993, product synthesis, materials science, medicinal, biological,
and MS degree in organic and supramolecular chemistry, and catalysis and coordination
chemistry from Dalian University chemistry. Moreover, some of the reactions have been success-
of Technology in 1996. He then fully applied in the pharmaceutical, agrochemical, and fine
studied at Changchun Institute of chemical industrials.8 Among these powerful transformations,
Applied Chemistry (CIAC), the Suzuki–Miyaura cross-coupling, generally defined as the
Chinese Academy of Sciences transition-metal-catalyzed cross-coupling between an organo-
(CAS) and obtained his PhD boron compound and an organic (pseudo)halide, has arguably
degree in organic and polymer become the most attractive approach since its discovery
chemistry in 1999. After in 1979.9 This is attributed in a large part to the mild reac-
Fu-She Han postdoctoral research under the tion conditions and the broad functional group tolerance
supervision of Prof. Dawei Ma at of this transformation.10 Furthermore, the organoboron
Shanghai Institute of Organic Chemistry, CAS (1999–2001), Prof. reagents can be easily synthesized by various ways and exhibit
Yutaka Watanabe at Ehime University, Japan (2001–2003), Prof. high stabilities towards air and moisture.10 In addition,
Tohru Fukuyama at University of Tokyo, Japan (2003–2006), and boron compounds show only low toxicities and the boron-
Prof. Dirk G. Kurth at National Institute for Materials Science, containing by-products can be easily separated from the reac-
Japan (2006–2008), he joined CIAC as a research professor in Aug. tion mixtures.
2008. His current research interests are focused on the development Over the 30 years of research history on Suzuki–Miyaura
of novel transition-metal-catalyzed methods for the carbon–carbon cross-coupling, the palladium-based catalysts have been the
and carbon–heteroatom bond forming reactions, as well as their most frequently investigated ones. Great success has been
practical applications for the synthesis of designed or naturally achieved mainly by the means of ligand design. Nowadays,
occurring complex molecules with important biological activities. many reactions can be carried out under very mild conditions

5270 Chem. Soc. Rev., 2013, 42, 5270--5298 This journal is c The Royal Society of Chemistry 2013
View Article Online

Review Article Chem Soc Rev

even at room temperature11 and in a short time, within only


several minutes.12 Moreover, the substrate scope for either
coupling partner has been substantially expanded. Typically the
electrophiles such as sterically very congested substrates,11d,13
and the inert aryl and vinyl chlorides11,14 and sulfonate
Published on 04 March 2013. Downloaded by UNIVERSIDADE FEDERAL SAO CARLOS on 27/11/2013 19:03:38.

derivatives,15 as well as the nucleophiles such as thermally


unstable polyfluorophenyl and 2-heteroaryl boron reagents,
which are conventionally extremely challenging partners, can
now be coupled readily.16 In addition, the palladium-catalyzed
asymmetric Suzuki–Miyaura reaction has also been intensively
developed.17 Moreover, with the rapid development of nano-
technology, considerable progress has been achieved recently Fig. 1 Some typical electrophiles, which are reactive in Ni-catalyzed Suzuki–
by carrying out the reaction in a heterogeneous system by Miyaura cross-coupling reactions.
means of anchoring the catalysts onto solid supports with nano
size.18 This development would make the reaction much more
sustainable,19 and, consequently, more practical for industrial electrophiles with a sequence of aryl and vinyl halides, phenol
applications in terms of the purification of products, as well as and enol compounds, miscellaneous aryl substrates, and alkyl
the separation, recovery, and reuse of the catalysts.20 halides. The most detailed discussion will be focused on the
Owing to these important advances, the palladium-catalyzed extensive efforts devoted to the development of highly active
Suzuki–Miyaura cross-coupling has become the first choice for nickel catalysts for coupling various electrophiles, which were
chemists in a wide range of communities if a C–C bond less reactive in the presence of palladium catalysts. For a better
forming reaction is under consideration. By employing this understanding, works on palladium-catalzyed coupling reac-
reaction, almost all kinds of biaryl, aryl–vinyl, alkyl–aryl, and tions will also be introduced briefly in some parts.
alkyl–alkyl compounds can be synthesized efficiently. Needless
to elaborate any further, the importance of the palladium- 2. Ni-catalyzed Suzuki–Miyaura cross-coupling
catalyzed Suzuki–Miyaura reaction has been well recognized of aryl and vinyl electrophiles
with the 2010 Nobel Prize in chemistry together with Heck and
Negishi couplings.21 2.1 Ni-catalyzed Suzuki–Miyaura cross-coupling of aryl and
Standing on the privileged and well-established platform, vinyl halides
synthetic organic chemists have been devoted to the more 2.1.1 Coupling reactions with phosphine ligand-based
challenging issues remaining both in industry and academia. homogeneous catalysts. Before 1996, the palladium-catalyzed
To this end, much recent attention in Suzuki–Miyaura cross- Suzuki–Miyaura reaction of arene bromides and iodides was
coupling has been focused on the nickel catalysts because successfully achieved with broad generality and high effi-
nickel is much cheaper and more earth abundant than the ciency.10 However, reactions for the more economical and
palladium metal.22 Therefore, the use of nickel catalysts would widely available arene chlorides were sluggish due to their
be far more cost-effective, unless a coupling reaction is work- low reactivity (the approximate bond dissociation energies for
able with a very low level of palladium loading, or only with a Ph–X are: C–Cl = 96 kcal mol1; C–Br = 81 kcal mol1; C–I =
very high nickel catalyst loading. On the other hand, the redox 65 kcal mol1),26 although a few examples of electron-deficient
state of palladium is typically Pd(0)/Pd(II), albeit the catalysis N-heteroaryl chlorides were demonstrated to be feasible by
chemistry of high valent palladium such as Pd(III) and Pd(IV) has applying palladium catalysts.27
been increasingly investigated in recent direct C–H function- As an important solution to this issue, in 1996, Miyaura and
alization.23 In contrast, the early transition metal nickel usually co-workers reported for the first time the nickel catalyzed cross-
displays Ni(0)/Ni(II) as well as Ni(I)/Ni(III) oxidation states and is coupling of inert aryl chlorides with boronic acids (Scheme 1).28
more nucleophilic due to its smaller size.24b As such, nickel can Various aryl chlorides decorated either by an electron-withdrawing
not be simply considered to be a substitute for palladium, it group or by an electron-donating group were viable substrates in
possesses distinctive catalytic properties that palladium does the presence of 3–10 mol% of Ni(0), prepared in situ by reducing
not have. Indeed, extensive studies have clearly demonstrated the NiCl2(dppf) (dppf = 1,10 -bis(diphenylphosphino)ferrocene)
that the nickel-based catalysts were more versatile and powerful with four equivalents of BuLi or DIBAL-H. The reaction was
catalysts for the C–C,24 C–N,24b,c and C–P25 bond forming carried out under heating in the presence of a K3PO4 base. The
reactions of a diverse class of electrophiles, which are conven- ortho-substituted and heterocyclic arene chloride were also well
tionally challenging in the presence of palladium catalysts tolerated. In stark contrast, Pd(PPh3)4 was far less efficient as
(Fig. 1). shown by comparison studies. The reaction almost did not
This review will summarize the progress in nickel-catalyzed occur for aryl chloride modified by neutral or electron-donating
Suzuki–Miyaura reactions from the original development to groups. Soon after this discovery, Indolese29 demonstrated
recent advances. For clarity, we classified the substrates that, without the need of external reductants, NiCl2(dppf) could
into four major classes based on the structural feature of also catalyze the coupling reaction of various aryl chlorides and

This journal is c The Royal Society of Chemistry 2013 Chem. Soc. Rev., 2013, 42, 5270--5298 5271
View Article Online

Chem Soc Rev Review Article


Published on 04 March 2013. Downloaded by UNIVERSIDADE FEDERAL SAO CARLOS on 27/11/2013 19:03:38.

Scheme 2 Ni-catalyzed Suzuki–Miyaura cross-coupling of aryl halides and


neopentylglycolboronates. (Data from ref. 33).

ligands was a relatively general catalyst system for the cross-


coupling of aryl halides. In general, good yields were obtained
Scheme 1 Comparison of Ni- and Pd-catalyzed Suzuki–Miyaura cross-coupling for aryl chlorides, bromides, and iodides in the presence of a
of aryl chlorides. (Data from ref. 28). K3PO4 base in heated dioxane and toluene. This catalyst system
was also demonstrated to be workable when the weak nucleo-
philic neopentylglycolboronates, instead of the more reactive
boronic acids in fairly good yields. Moreover, even a 1 mol% of boronic acids, were used as nucleophiles, although an
catalyst loading was also workable in some cases. increased catalyst loading (10 mol%) and an elevated tempera-
Following these pioneering works, the Suzuki–Miyaura ture from 80 1C to 110 1C was necessary33 (Scheme 2).
cross-coupling of aryl and vinyl halides was extensively inves- In a later investigation by Bao and co-workers,34 they showed
tigated by applying the nickel–phosphine complexes. In 1999, that by changing the dppe phosphine ligand to Ph2PCH2CH2OH,
Genêt described a water-soluble nickel catalyst that was capable i.e., one of the diphenylphosphino groups in dppe was replaced
of affecting the cross-coupling of aryl or vinyl chlorides and by a hydroxy group, the derived NiCl2(Ph2PCH2CH2OH)2H2O
boronic acids using water as a cosolvent.30 The Ni(0) catalyst complex from such a P,O-hybrid ligand displayed a somewhat
was preformed in situ from 10% of NiCl2(dppe) (dppe = increased catalytic activity when used to couple aryl chlorides.
1,2-bis(diphenylphosphino)ethane), 50% of TPPTS (TPPTS = Good to excellent yields were generally obtained in the presence of
sodium triphenylphosphinotrimetasulfonate), and 50% of Zn 3 mol% of catalyst with K3PO4 as the base in isopropanol at 80 1C.
by heating. In addition, the use of readily available and low toxic isopropanol
In 2000, a more efficient and cheaper catalyst was reported as solvent was also an added advantage of Bao’s procedure.
by Miyaura and Inada.31 With the assistance of two equivalents In 2006, a very mild procedure was reported by Hu and
of PPh3 as external supporting ligands, the more affordable co-workers,35 who described that Ni(cod)2 (cod = 1,5-cyclo-
NiCl2(PPh3)2 could affect the cross-coupling of arene chlorides octadiene) with the assistance of ferrocenylmethylphosphine
and boronic acids. The catalytic efficiency of this reductant- or PPh3 ligands could catalyze the cross-coupling of aryl
free system was comparable to that of the NiCl2(dppf)–BuLi chlorides and boronic acids at room temperature. A range of
system.28 deactivated and unactivated aryl chlorides cross-coupled effi-
From these early studies, it was observed that the nickel- ciently with aryl boronic acids in high yields in the presence of
catalyzed Suzuki–Miyaura coupling was extremely sensitive to 4 mol% of Ni(cod)2 and 8–12 mol% of phosphine ligands.
the experimental conditions. Namely, a minor change of the Afterwards, Percec and co-workers36 demonstrated that by
reaction conditions gave rise to a remarkable change of the employing a catalyst system composed of 6 mol% Ni(cod)2
results. Thus, for a clear understanding of the characteristics of and 18 mol% of PCy3 (PCy3 = tricyclohexylphosphine), the
this reaction, Percec and co-workers32 carried out a compre- weakly nucleophilic neopentylglycolborate ester could also be
hensive study to inspect the effect of various parameters, such coupled with an aryl chloride at room temperature, although
as different nickel–phsophine complexes, supporting ligands, the substrate scope and limitation was not further examined.
bases, solvents, and the nature of leaving groups, on the In some recent works reported by Yang and Chen,37 the
efficacy of the coupling reactions. They found that NiCl2(dppe) pre-prepared Ni(II)-(d-aryl) complexes, [Ni(PPh3)2(1-naphthyl)X
paired with the use of an excess of dppe or PPh3 supporting (X = Cl or Br)], were proved to be also competent catalyst

5272 Chem. Soc. Rev., 2013, 42, 5270--5298 This journal is c The Royal Society of Chemistry 2013
View Article Online

Review Article Chem Soc Rev

precursors that allowed for the cross-coupling of various aryl


bromides as well as the less reactive aryl chlorides under mild
conditions. A range of deactivated, unactivated, and activated
substrates underwent effective coupling at room temperature37b in
the presence of 5 mol% of such nickel complexes and 7.5–10 mol%
Published on 04 March 2013. Downloaded by UNIVERSIDADE FEDERAL SAO CARLOS on 27/11/2013 19:03:38.

of PCy3HBF4 or PPh3 supporting ligand.


While these developments have greatly advanced the nickel-
catalyzed Suzuki–Miyaura reaction, the practicality of these
procedures is still limited. The common problems were the
requirement of a relatively high catalyst loading (typically
3–10 mol%) as well as the presence of an excess amount of
external supporting ligands (typically 1–5 equiv.). Moreover, the
use of air and moisture sensitive catalysts in some methods is
also an important concern. Very recently, Han and co-workers38
developed a much more effective and readily available nickel
catalyst - NiCl2(dppp) [dppp = 1,3-bis(diphenylphosphino)-propane].
This catalyst allowed for very general and efficient cross-
coupling of a rich variety of aryl bromides as well as the less
reactive aryl chlorides with a catalyst loading of lower than
1 mol% (Scheme 3 and 4). An array of substrates having
electron-rich and deficient substituents were well tolerated
when using K3PO4 as the base in dioxane at 100 1C. Moreover,
sterically hindered aryl and heteroaromatic halides were also
viable partners. Notably, external supporting ligands were not
needed in the transformations. In addition to the high catalytic
Scheme 4 NiCl2(dppp)-catalyzed Suzuki–Miyaura coupling of various aryl chlorides
with aryl boronic acids. (Data from ref. 38).

activity, NiCl2(dppp) is also highly stable due to the bidentate


nature of dppp, forming a stable six-membered complex with
nickel.
As a potential application, the methodology can be used for
the effective synthesis of a new fungicide Boscalids (Scheme 5),
which is currently one of the largest industrial applications of
the Suzuki–Miyaura reaction (ca. 1000 tons per year).39a In the
reported methods, the key reaction for the synthesis of the
product was the palladium-catalyzed cross-coupling of 1-chloro-
2-nitrobenzene with 4-chlorophenylboron compounds.39 Then,
reduction of NO2 in the cross-coupled biaryl gave the key
intermediate 2-amino-40 -chlorobiphenyl. Alternatively, Han and

Scheme 3 NiCl2(dppp)-catalyzed Suzuki–Miyaura coupling of various aryl bromides Scheme 5 Gram-scale synthesis of Boscalids via the Ni-catalyzed Suzuki–
with aryl boronic acids. (Data from ref. 38). Miyaura coupling. (Data from ref. 38).

This journal is c The Royal Society of Chemistry 2013 Chem. Soc. Rev., 2013, 42, 5270--5298 5273
View Article Online

Chem Soc Rev Review Article

and deficient, as well as ortho-substituted aryl halides under-


went efficient cross-coupling in excellent yields of higher than
90% in most cases. Importantly, the catalyst could be recovered
by precipitation and reused for six cycles with only negligible
loss of catalytic activity.
Scheme 6 Synthesis of 2-cyano-4 0 -methylbiphenyl via the Ni-catalyzed Suzuki–
Published on 04 March 2013. Downloaded by UNIVERSIDADE FEDERAL SAO CARLOS on 27/11/2013 19:03:38.

2.1.2 Coupling reaction with N-hetero carbene ligand-based


Miyaura coupling at a 10 gram-scale.
homogeneous catalysts. In addition to the most popularly
investigated nickel–phosphine catalyst, the nickel and
co-workers developed a more efficient pathway by applying N-heterocyclic carbene (NHC) based complexes have also been
their NiCl2(dppp) catalyzed cross-coupling method. Namely, a type of frequently investigated catalysts in the nickel-catalyzed
cross-coupling of 2-chlorobenzonitrile with a slight excess of Suzuki–Miyaura cross-coupling of arene halides.44 At the early
4-chlorophenylboronic acid (1.1 equiv.) proceeded smoothly in stage, Cavell and McGuinness45 reported on the use of mono-
the presence of only 0.4 mol% of NiCl2(dppp). This reaction dentate N-heterocarbene–Ni(II) complexes, [Ni(tmiy)2I2] or [Ni(tmiy)2-
could be carried out with gram-scale of starting material. (o-tolyl)Br] (tmiy = 1,3,4,5-tetramethylimidazol-2-ylidene) to
Hydrolysis followed by Hofmann rearrangement of the coupled catalyze the cross-coupling reaction of 4-bromoacetophenone
product gave the desired biaryl amine in 82% yield over two and phenylboronic acids. Although the activity was not high for
steps. Finally, the condensation of the biaryl amine with this individual reaction and the efficiency for other substrates
2-chloronicotinoyl chloride afforded the target Boscalid. was not examined, this result indicated that Ni–NHC complexes
The practicality of the NiCl2(dppp)-catalyzed coupling reac- are potential catalysts for the Suzuki–Miyaura reaction of aryl
tion was further demonstrated by the synthesis of 2-cyano-4 0 - halide electrophiles.
methylbiphenyl (Scheme 6). This biaryl compound is the key After this report, Lee and co-workers46 in 2005 reported a
intermediate for the synthesis of anti-hypertensive drugs such nickel complex based on a tetradentate pyridine/N-heterocyclic
as losartan, irbesartan, and valsartan,40 or AT2 antagonists, and carbene ligand (Fig. 3a). Catalytic use of this complex in
also represent one of the most important industrial applica- Suzuki–Miyaura coupling of a range of aryl iodides, bromides,
tions of Pd-catalyzed Suzuki–Miyaura reaction.41 After further and chlorides with phenylboronic acid afforded moderate to
careful optimization of their initial process,38 Han and high yields as determined by NMR analysis. However, the PPh3
co-workers showed that the coupling reaction of 2-chloro- coligand was required to enhance the activity. Based on this
benzonitrile and a slight excess of 4-methylphenyl boronic acid observation, the authors further designed and synthesized a
(1.05 equiv.) could be carried out very smoothly in 10 gram- new nickel complex whose ligand contained both NHC and
scale with 1.0 mol% of NiCl2(dppp) catalyst and a lowered phosphane functionalities (Fig. 3b).47 Such a tailored catalyst
equivalent of anhydrous K3PO4 base.42 The yield of the crude exhibited an enhanced activity and a broader applicability in
product was almost quantitative. The 1H-NMR analysis of the catalyzing the coupling reaction of various aryl halides and
crude products revealed that the purity of the desired product phenylboronic acid. The external addition of phosphine ligands
was about 95% containing only a small amount of the homo- displayed no accelerating effect on the reactive substrates and
coupled product of boronic acid as an impurity. The crude only a slightly accelerating effect on the less reactive substrates.
product can be used for the following transformation without These results indicated that the PPh2 functionality in the
further purification. carbene-P hybrid ligand can partially replace the function of
Almost at the same time, Wu and co-workers43 presented a the free phosphine ligand.
nickel catalyst stabilized by a phosphine dendrimer ligand In 2006, Inamoto and co-workers48 presented a nickel-pincer
(Fig. 2, G3DenP). This catalyst was highly active and efficient complex derived from the imidazole N-heterocyclic carbene
for the Suzuki–Miyaura reaction of aryl bromides and chlorides. ligand (Fig. 4a). This type of complex usually exhibits high
With only a 0.01–0.1 mol% of the catalyst loading, electron-rich stability toward air and moisture. Aryl bromides containing
both electron-deficient and electron-rich substituents could be
coupled with phenyl boronic acid in good yields with a catalyst
loading of 1 mol%. In addition, the less reactive aryl chlorides

Fig. 3 The structures of the Ni–NHC complexes based on the tetradentate


Fig. 2 Structure of dendrimer phosphine ligand. (Adapted from ref. 43). pyridine/N-heterocyclic carbene ligand. (Adapted from ref. 46 and 47).

5274 Chem. Soc. Rev., 2013, 42, 5270--5298 This journal is c The Royal Society of Chemistry 2013
View Article Online

Review Article Chem Soc Rev

2.1.3 Coupling reactions with miscellaneous homogeneous


catalysts. In addition to the popularly investigated nickel–
phosphine and nickel–NHC catalysts, some other nickel cata-
lysts were also investigated. Leadbeater53 and Monteiro54 have
demonstrated that the nickel complexes formed with simple
Published on 04 March 2013. Downloaded by UNIVERSIDADE FEDERAL SAO CARLOS on 27/11/2013 19:03:38.

triethylamine and 2,2 0 -bipyridine ligands, or the ligand-


free NiCl26H2O could also be used to catalyze the coupling
reactions of a certain scope of aryl bromides and iodides with
aryl boronic acids. However, the catalytic efficiency and sub-
strate scope were restricted, although these catalysts are more
affordable.
Very recently, Roglans and co-workers55 reported a class of
(0)
Ni complexes derived from polyunsaturated azamacrocyclic
Fig. 4 The structures of NHC–nickel pincer-type complexes. (Adapted from ligands (Fig. 6). To compare with the frequently used Ni(0)
ref. 48, 50, and 51). catalyst Ni(cod)2, which is formed from the olefinic cyclo-
octadiene ligand and is highly sensitive towards air, moisture,
and heating, the azamacrocyclic olefin-based Ni(0) complexes
bearing electron-deficient substituents could also be coupled in exhibit an apparently enhanced stability. Moreover, the cata-
moderate yields with a catalyst loading of 5 mol%. However, lytic use of these complexes revealed that they also displayed
a long reaction time of three to four days paired with a relatively good activity for catalyzing the Suzuki–Miyaura coupling reac-
high reaction temperature of 120 1C was essentially needed. tion of a range of aryl bromides and aryl boronic acids.
Part of these drawbacks were overcome in their later work by 2.1.4 Coupling reactions with heterogeneous catalysts.
further optimization of the reaction conditions.49 In 2009, two In the field of transition-metal-catalyzed transformations, the
types of modified pincer-type nickel complexes featuring the separation, recyclability, and reusability of the catalysts are
six-membered and five-membered metallocyclic frameworks always one of the major concerns. The most popular solution
(Fig. 4c and d) were developed by the same group.50 Catalytic to these issues is the fixation of the catalysts on a solid
use of these catalysts for the coupling of aryl halides showed an support.18 Such technology has also been routinely used in
improved activity to compare with the analogue 4a. The reac- nickel-catalyzed Suzuki–Miyaura cross-coupling by some groups.
tion time could be shortened to several hours. Moreover, both In 2000, Lipshutz and co-workers have reported a heterogeneous
aryl and alkenyl boronic acids were reactive. Later, Tu and catalyst by immobilizing the nickel on charcoal (Ni/C).56 By
co-workers51 demonstrated that the benzoimidazole-based employing four equivalents of PPh3 as the stabilizing ligand,
pincer-nickel complex (Fig. 4b) was also an effective catalyst. a variety of electron-deficient aryl and several N-heteroaryl
The coupling reaction of an array of aryl bromides and chlorides chlorides underwent smooth coupling to afford the coupled
with phenyl boronic acid proceeded in excellent yields in the products in good yields. However, the reactions for the
presence of 2 mol% catalyst with the assistance of 10 mol% of electron-rich aryl chlorides were somewhat reluctant. Later,
the PPh3 supporting ligand. study from the same group found that the catalytic efficiency
Chen and co-workers52 developed a novel type of NHC-based could be enhanced by anchoring the Ni(0) on graphite (Ni/Cg).57
dinickel complex (Fig. 5). These complexes could catalyze the The reaction could be carried out either by bath heating or
cross-coupling reactions of a broad range of aryl bromides as under microwave conditions. Moreover, the reusability of Ni/Cg
well as the less reactive aryl chlorides. Generally, with only a
0.2–0.8 mol% catalyst loading and the presence of 5.0 equiva-
lents of PPh3 ligands relative to the catalysts, the activated,
unactivated, and deactivated aryl chlorides were coupled very
effectively with aryl boronic acids in excellent yields in toluene
at 80 1C. The dinickel catalysts represent a type of the most
effective catalysts among NHC-based nickel complexes in terms
of catalytic activity and stability.

Fig. 6 Ni(0) complexes derived from polyunsaturated azamacrocyclic ligands.


Fig. 5 The structures of NHC-based dinickel complexes. (Adapted from ref. 52). (Adapted from ref. 55).

This journal is c The Royal Society of Chemistry 2013 Chem. Soc. Rev., 2013, 42, 5270--5298 5275
View Article Online

Chem Soc Rev Review Article

was also studied after the reactivation with BuLi. Further


investigation showed that the catalyst also displayed good
activity for aryl tosylates (vide infra).
Wang and co-workers58 described a nickel–metal colloid
using TBAB (tetra-n-butylammonium bromide) as stabilizer.
Published on 04 March 2013. Downloaded by UNIVERSIDADE FEDERAL SAO CARLOS on 27/11/2013 19:03:38.

Such a heterogeneous catalyst exhibited high activity for aryl


bromides and iodides without the need of phosphine ligands.
However, for the less reactive aryl chlorides, the phosphine
ligand was required for more efficient reaction.
2.1.5 Ni-catalyzed cross-coupling of aryl fluorides. Although
aryl fluorides also belong to the aryl halides, this class of
substrates is extremely inert as electrophiles in transition-
metal-catalyzed cross-couplings. In the literature, the coupling
reaction of aryl fluorides was often investigated as a subject
independent of aryl iodides, bromides, and chlorides. There-
fore, herein, we discussed the reaction of aryl fluorides as a
separate section.
Early studies on aryl fluorides were only focused on the
palladium-catalyzed couplings of the highly electron-deficient
substrates modified by strong electron-withdrawing groups
Scheme 7 Effect of directing groups on the Suzuki–Miyaura coupling of aryl
such as tricarbonylchromium(0) complexes59 and an ortho- fluorides. (Data from ref. 63).
nitro group.60 Only a very few reports investigated the
non-activated fluorobenzene.61 The first nickel-catalyzed
Suzuki–Miyaura coupling of aryl fluorides was reported by
Radius in 2006,62 who demonstrated that the perfluorinated
arenes, which are also electron-deficient substrates, underwent
effective coupling with aryl boronic acids by using a NHC-based
nickel complex. The generation of the oxidative addition inter-
mediates of aryl fluorides to nickel catalyst was confirmed by
the NMR spectroscopies and X-ray analysis. A more general
protocol was reported by Chatani and co-workers in 2011.63
Utilizing a combination of Ni(cod)2 (20 mol%), PCy3 (80 mol%),
ZrF4 (40 mol%), and CsF base, a range of deactivated aryl
fluorides could be coupled smoothly with aryl boron reagents
in toluene at 100 1C. The use of metal fluoride salts was
identified to play an important role as a cocatalyst in promoting
the reactions. ZrF4 was optimal among the various salts being
screened. Scheme 8 Synthesis of polyaryl ketone via a discriminative Suzuki–Miyaura
In-depth studies by Chatani63 and Love64 showed that the cross-coupling of aryl halide. Abbr.: dba = dibenzylideneacetone; SPhos =
proper positioning of a directing group at the ortho-position of 2-Dicyclohexylphosphino-2 0 ,6 0 -dimethoxybiphenyl. (Data from ref. 63).

fluoride was essentially important for effective cross-coupling


(Scheme 7). Typically, with a 2-N-heteroarene or imine directing
group, a Ni(cod)2 or NiCl2(PCy3)2 catalyst could catalyze the Finally, the Ni(cod)2-catalyzed coupling of aryl fluoride with an
reaction in high yields without the need of metal fluoride aryl boronate proceeded uneventfully to give the polyaryl ketone.
cocatalysts. In contrast, substrates without a directing group, 2.1.6 Reaction mechanism. Concerning the mechanism of
or with a 4-N-heteroarene or ester directing group resulted in no the nickel-catalyzed Suzuki–Miyaura coupling of aryl halides,
reaction or afforded the coupled products in a very poor yield. a Ni(0) to Ni(II) cycle is generally considered to be the most
The synthetic utility of the nickel-catalyzed protocol for aryl plausible pathway, which is similar to the palladium-catalyzed
fluorides was elaborated by the synthesis of a polyaryl ketone version.65 Namely, the oxidative addition of aryl halides to Ni(0)
via a discriminative cross-coupling of aryl iodide, chloride, generates a Ni(II) species. Transmetalation of the Ni(II) intermediate
and fluoride (Scheme 8).63 Starting from 4-chloroiodobenzene, followed by reductive elimination produced the coupled pro-
the palladium-catalyzed synthesis of benzophenone was ducts and regenerates the active Ni(0). Interestingly, a recent
carried out selectively at the iodide position. Suzuki–Miyaura work by Louie66 disclosed that Ni(I) complexes were also active
coupling of 4-fluoro-4 0 -chlorobenzophenone with 4-dimethyl- catalysts for the cross-coupling of aryl halides and boronic acids.
aminophenyl boronic acid in the presence of a Pd(OAc)2 In their studies, they found that the reaction of Ni(0)(IMes)2,
catalyst occurred exclusively at the more reactive chloride site. [IMes = 1,3-bis(2,4,6-trimethylphenyl)imidazol-2-ylidene],

5276 Chem. Soc. Rev., 2013, 42, 5270--5298 This journal is c The Royal Society of Chemistry 2013
View Article Online

Review Article Chem Soc Rev

higher stability. Thus, to best use aryl sulfonates, numerous


efforts have been devoted to the development of less expensive
and more active nickel catalysts. The first seminal reaction was
reported by Percec in 1995,69 who demonstrated that 10 mol%
of NiCl2(dppf), with a combinatorial use of 1.7 equivalents of
Published on 04 March 2013. Downloaded by UNIVERSIDADE FEDERAL SAO CARLOS on 27/11/2013 19:03:38.

Zn as reductant, could affect the Suzuki–Miyaura coupling of


aryl mesylates or aryl arenesulfonates with aryl boronic acids
in THF.
Although the efficiency of this cross-coupling reaction was
not sufficiently high as judged from the moderate yields, the
relatively high NiCl2(dppf) loading (10 mol%), and the use of an
excess amount of Zn as reductant, this result was a landmark
Fig. 7 Proposed mechanism for Ni(I)-catalyzed Suzuki–Miyaura cross-coupling of
aryl halides with boronic acids. (Adapted from ref. 66).
work in the history of Suzuki–Miyaura reaction. It not only
represented the first example demonstrating that nickel,
instead of palladium, could be used as a catalyst for the
with aryl halides afforded Ni(I) complexes, [NiI(IMes)2X], X = Cl Suzuki–Miyaura coupling of aryl sulfonates, but, most signifi-
or Br. The desired oxidative addition Ni(II) complexes were not cantly, also paved a new pathway for extensively investigating
observed. Catalytic use of these Ni(I) complexes in the cross- the nickel-catalyzed Suzuki–Miyaura reactions of other substrates.
coupling of the Heck and Suzuki–Miyaura reaction of aryl Indeed, on the basis of this original work, comprehensive
bromides and chlorides produced the corresponding cross- investigation has been carried out in the nickel-catalyzed
coupled products in moderate to high yields. Suzuki–Miyaura reaction of aryl halides. (See Section 2.1 for
Based on further experimental studies, they proposed a Ni(I) the nickel-catalyzed cross-coupling of aryl and vinyl halides).
to Ni(III) mechanism (Fig. 7). Namely, the oxidation of Concerning the mechanism of the nickel-catalyzed Suzuki–
Ni(0)(IMes)2 by aryl halides proceeds via the radical process to Miyaura reaction of aryl sulfonates, Percec suggested an
generate the Ni(I) species (A). Next, transmetalation between the essentially similar pathway to that of the palladium-catalzyed
Ni(I) species and metalating reagent, rather than the oxidative cross-coupling for aryl (pseudo)halides and boronic acids
addition between the Ni(I) species and aryl halide, occurs originally proposed by Suzuki and co-workers.65 Namely, the
preferentially to afford the Ni(I) species B. Then, oxidative reaction starts with the reduction of Ni(II) to Ni(0) (A) by Zn
addition of the aryl halide to B forms the Ni(III) complex C. (Fig. 8). The Ni(0) thus generated brings the cross-coupling into
Finally, reductive elimination produces the aryl–aryl coupled the catalytic cycle. First, oxidative additon of aryl sulfonates to
product and regenerates the Ni(I) catalyst. In fact, the Ni(I) to Ni(0) forms the Ni(II) species (B or B 0 ). Then, transmetalation
Ni(III) catalytic model was also proposed for the nickel-catalyzed between the boron reagent and B or B 0 produces the inter-
cross-coupling of alkyl halides (vide infra). mediate C. Finally, reductive elimination of C forms the cross-
coupled product D and regenerates the active Ni(0) catalyst.
2.2 Nickel-catalyzed Suzuki–Miyaura cross-coupling of phenol Soon after Percec’s first report, Kobayashi and co-workers70
and enol derivatives reported that lithium organoborates generated from boron
Although industrial chemistry can now provide a broad array of esters and organolithium (MeLi or n-BuLi) underwent smooth
aryl, vinyl, and alkyl halides, they are still far less available to coupling with electron-deficient aryl and alkenyl mesylates in
compare with the naturally abundant phenol and enol deriva-
tives. Thus, methods for the transition-metal-catalyzed effective
coupling of phenol and enol compounds are of considerably
importance. As one of the most attractive methods, the Suzuki–
Miyaura cross-coupling by applying a nickel catalyst has
received explosive interest, particularly in the last several years.
Progress in this area will be summarized in this section.
2.2.1 Nickel-catalyzed Suzuki–Miyaura cross-coupling of
aryl and enol sulfonates. While the rapid development of new
ligands has now substantially advanced the palladium-catalyzed
Suzuki–Miyaura coupling of aryl and vinyl sulfonates,67 earlier
works before 1995 had demonstrated only aryl triflates or
nonaflates were reactive electrophiles to couple with boronic
acids.10,68 The reaction of aryl and vinyl sulfonates such as
tosylates and mesylates were considerably challenging at that
stage. However, to compare with aryl triflates, the corre-
sponding sulfonate derivatives are far more appealing partners Fig. 8 Proposed mechanism for the Ni-catalyzed Suzuki–Miyaura cross-coupling
because of their easier preparation, lower cost, and much of aryl sulfonates. (Adapted from ref. 69).

This journal is c The Royal Society of Chemistry 2013 Chem. Soc. Rev., 2013, 42, 5270--5298 5277
View Article Online

Chem Soc Rev Review Article

the presence of 10 mol% of NiCl2(PPh3)2 at room temperature.


In comparison, aryl tosylates were somewhat less reactive than
aryl mesylates. While the mild reaction temperature was a merit
of the method, high catalyst loading and the less convenient
activation procedure for boron esters were limitations for
Published on 04 March 2013. Downloaded by UNIVERSIDADE FEDERAL SAO CARLOS on 27/11/2013 19:03:38.

extensive application.
In 1998, Miyaura and co-workers71 showed that with the
assistance of an external dppf ligand coupled with the presence
of BuLi as a reducing reagent, aryl mesylates could be coupled
with aryl boronic acids in the presence of 3–4 mol% NiCl2(dppf)
in toluene. Although this method allowed the coupling reaction
Scheme 9 Ni-catalyzed Suzuki–Miyaura cross-coupling of aryl tosylates at room
to be carried out at a lowered catalyst loading, it required the temperature. (Data from ref. 73a).
use of the relatively expensive dppf ligand.
An improved catalyst system was presented by Monteiro and
co-workers.72 The authors have utilized a combination of development of a more stable catalyst system, they found that
1.5–3.0 mol% of the NiCl2(PCy3)2 catalyst and 6–12 mol% of by replacing the PCy3 supporting ligand with a more stable
PCy3 as an external ligand for the cross-coupling of various aryl ferrocenylmethylphosphine, the cross-coupling of various aryl
tosylates and phenylboronic acid. The results showed that, in tosylates with phenylboronic acids also proceeded in high
the absence of a reducing reagent, activated, unactivated, and yields at room temperature. However, a longer reaction time
deactivated aryl tosylates could be coupled efficiently at 130 1C (2 days) and a higher amount of Ni(cod)2 (4 mol%) were
in dioxane. An active Ni(0) species in this reductant-free pro- required.35a
cedure was proposed to be generated via the transmetalation of The high catalytic activity of the Ni(cod)2 catalyst system for
boronic acid with NiCl2(PCy3)2 followed by reductive elimina- coupling aryl and alkenyl sulfonates has attracted more atten-
tion of the homocoupled biaryls. A competitive study revealed tion. Thus, Percec and co-workers36,74 demonstrated that
that the reactivity of aryl tosylate was twice as fast as aryl by applying 6 mol% of Ni(cod)2 and 12–18 mol% of PCy3,
chloride under the specified catalyst system. neopentylglycolborate esters, which are weaker nucleophiles
Then, Percec and co-workers32 carried out a detailed study to to compare with boronic acids, could also be coupled effectively
understand the effect of reaction parameters, such as different with a broad range of aryl tosylates and mesylates at room
nickel complexes with phosphines, external supporting ligands, temperature, except for the congested 2,6-disubstituted sub-
bases, solvents, and the nature of leaving groups such as strates. In addition, Molander and Beaumard75 have found that
mesylate and tosylate on the efficacy of the coupling reactions. in the presence of 10 mol% of Ni(cod)2 and 20 mol% of the
They found that NiCl2(dppe) paired with the use of an excess of PCy3HBF4 supporting ligand, both aryl mesylates and tosylates
dppe or PPh3 supporting ligands was the most general catalyst were viable electrophiles to couple with potassium organo-
system for the cross-coupling of aryl mesylates and tosylates trifluoroborates as the surrogates of boronic acids. The reaction
both in dioxane and toluene solvents using K3PO4 as the base. could be carried out in a mixed solvent of 1 : 1 t-BuOH and H2O
Although their optimized conditions allowed the reaction to be at 110 1C. Notably, the method allowed for the synthesis of
carried out at a relatively low temperature to compare with 2-heteroaromatic biaryls using potassium 2-heteroaryltrifluoro-
those reported by Monteiro72 (80 vs. 130 1C), an increased borate nucleophiles in moderate to high yields (Scheme 10).
catalyst loading of up to 5 mol% was essentially required for This is attractive since the construction of 2-heterobiaryls from
effective coupling.
In 2004, Hu and Tang73 established a very mild protocol that
allowed for the cross-coupling of aryl and alkenyl tosylates or
benzenesulfonates with boronic acids. With the presence of
3 mol% of Ni(cod)2 and 12 mol% of PCy3, cross-coupling of aryl
sulfonates proceeded very smoothly at room temperature.73a
Both electron-rich and electron-deficient aryl arenesulfonates
reacted with various boronic acids in excellent yields
(Scheme 9). In addition, the effective coupling of alkenyl
tosylates73b made their procedure useful for the synthesis of
4-substituted coumarins and 4-substituted furanones, which
are important biologically active molecules. Investigation
into the reaction mechanism by 1H NMR spectroscopic
analyses showed that the oxidative addition of aryl tosylates
to Ni(0) was a fast process. This observation indicated that
transmetalation and/or reductive elimination should be the Scheme 10 Suzuki–Miyaura cross-coupling of aryl mesylates with potassium
rate-determining step. In their later work aimed at the 2-heterotrifluoroborates. (Data from ref. 75).

5278 Chem. Soc. Rev., 2013, 42, 5270--5298 This journal is c The Royal Society of Chemistry 2013
View Article Online

Review Article Chem Soc Rev

2-heterocylic boronic acids is commonly challenging due to aryl or alkenyl boronic acids in the presence of 5 mol% catalyst
their low thermal stability. in DME at 120 1C. However, the electron-rich and ortho-
Although intensive investigations have exemplified that substituted tosylates were far less reactive. In addition, a some-
Ni(cod)2 can be a ‘‘privileged’’ catalyst for the cross-coupling what lowered reactivity was observed for aryl mesylates. In a
of aryl and alkenyl sulfonates, it’s highly sensitive nature similar work, Tu51 showed that the NHC–nickel complex derived
Published on 04 March 2013. Downloaded by UNIVERSIDADE FEDERAL SAO CARLOS on 27/11/2013 19:03:38.

towards air, moisture, and thermal effects is an important from benzoimidazole was a promising catalyst (Fig. 4b). With the
concern for wide applications. Therefore, chemists are still presence of 2 mol% of catalyst and 10 mol% of PPh3 ligand,
interested in the development of other nickel-based catalysts an array of aryl tosylates and mesylates could be coupled in
with higher stability. excellent yields. Both unactivated and activated sulfonates were
In 2008, Lipshutz and co-workers57 described a PPh3 stabi- viable substrates.
lized nickel catalyst anchored on graphite (Ni/Cg). The catalytic Despite these important progresses, most of the established
use of such a heterogeneous catalyst for the cross-coupling of procedures required a high catalyst loading (usually Z 5 mol%)
aryl tosylates showed that both activated and unactivated, as and the use of a large excess of phosphine supporting ligands.
well as ortho-substituted aryl tosylates were viable substrates In an effort towards the development of a highly active catalyst,
either by means of heating or by microwave. The reaction was Han and co-workers81 have disclosed in their recent report that
completed within 1.5 to 4 h at 180 to 200 1C. The catalyst could NiCl2(dppp) was a very active and general catalyst for the cross-
be recycled by the reactivation with BuLi. coupling of aryl sulfonates. A broad range of aryl tosylates and
In 2010, Yang and Fan76 described a Ni(II)-(s-aryl) complex, mesylates underwent very effective coupling with various aryl
NiII(PPh3)2(1-naphthyl)Cl. This type of complex displays rela- boronic acids in high to excellent yields with less than
tively good stability towards air and moisture. Study on the 1.0 mol% of NiCl2(dppp). Notably, external phosphine ligands
catalytic properties showed that the complex was a promising were not needed in the transformation. In addition to the high
catalyst for the Suzuki–Miyaura coupling of aryl sulfonates. A catalytic activity, NiCl2(dppp) is also more stable towards air
combination of 5 mol% of the complex and 5 mol% of PPh3 and moisture, and cheaper when compared with the frequently
allowed for effective cross-coupling of aryl tosylates with boronic investigated NiCl2(dppf), NiCl2(PCy3)2, and Ni(cod)2.
acids with K2CO3 as a base at 100 1C in toluene. Further 2.2.2 Nickel-catalyzed Suzuki–Miyaura cross-coupling of
optimization of the reaction conditions revealed that by repla- aryl sulfamates. Pioneered by Snieckus,82 the aryl sulfamates
cing the PPh3 ligands with PCy3HBF4, the reaction could have been studied in the nickel-catalyzed Kumada–Tamao–Corriu
proceed at room temperature and exhibited a broader substrate coupling. This type of substrate is readily available and displays
scope.37b More recently, Percec and co-workers77 found that the high stability. As a more important advantage, the sulfamate
same catalyst system was also effective for catalyzing the reac- group may be acting to direct the installation of functional groups
tion of aryl mesylates with a less reactive borate ester at room at both the ortho and para position of the aromatic ring in
temperature. In a related work, Hu and co-workers78 developed addition to the role as a leaving group. With these interesting
a similar Ni(II)-(s-aryl) complex, NiII(PCy3)2(4-MeOC6H4)OTs, features, aryl and alkenyl sulfamates have attracted much atten-
which could be also used to couple aryl and alkenyl sulfonates tion from organic chemists as potential electrophiles under the
with boronic acids at room temperature with a 5 mol% of more appealing Suzuki–Miyaura conditions.
catalyst loading. The first successful Suzuki–Miyaura coupling of aryl sulfa-
For a systematic understanding on the catalytic property of mates was reported in 2009 by Garg and co-workers.83 Under
the several frequently used catalysts such as NiCl2(PCy3)2, their optimized conditions: 5 mol% of NiCl2(PCy3)2, 4.5 equiva-
Ni(cod)2, and Ni(II)-(s-aryl) complexes, very recently, Percec lents of K3PO4 base in toluene at 110 1C, a rich range of aryl
and co-workers79 carried out comprehensive studies and sulfamates including fused and nonfused substrates, and those
compared their catalytic activity for different electrophiles modified either by electron-withdrawing or electron-donating
including sulfonates, sulfamates, carbamates, carbonates, and groups were coupled in very high yields (Scheme 11). Moreover,
esters derived from phenol derivatives, as well as different the sterically congested aryl sulfamates were also viable substrates
boron nucleophiles such as boronic acids, potassium trifluoro- although harsher conditions were needed, typically by increasing
borates, and boron esters. Although it was hard to figure out a the catalyst to 10 mol%, the boronic acids to 4.0 equivalents, the
common rule because many reactions were sensitive with a base to 7.2 equivalents, and the temperature to 130 1C. Further
minor change of the reaction conditions, aryl mesylates and studies showed that the methodology could be extensively
sulfamates are generally more reactive than other electrophiles. expanded to heteroaryl sulfamates as well as a broad range of
The reactivity order for boron nucleophiles is boronic acids, boronic acids under identical conditions.84 The utility advantages
potassium trifluoroborates, and boron esters. The details for of this protocol were elaborated by the synthesis of an anti-
the cross-coupling of sulfamates, carbamates, carbonates, and inflammatory drug flurbiprofen by the sequential use of sulfamate
esters were discussed in the following sections. group directed installation of boronate functionality at the ortho-
As an alternative study, Inamoto and Dio80 investigated the position of the sulfamate group, the orthogonal enonate coupling
catalytic activities of imidazole-based NHC–nickel complexes as at the iodide site in the presence of fluoride and sulfamate leaving
shown in Fig. 4c and d. Good efficiency was observed for the groups, and finally, the NiCl2(PCy3)2-catalyzed cross-coupling of
reaction of electron-deficient aryl or alkenyl tosylates with the the aryl sulfamate and an aryl boronic acid.

This journal is c The Royal Society of Chemistry 2013 Chem. Soc. Rev., 2013, 42, 5270--5298 5279
View Article Online

Chem Soc Rev Review Article

state enables the oxidative addition of N,N-dimethyl phenyl


sulfamate to the nickel catalyst to occur preferentially at the
Ph–O bond rather than at the O–sulfonyl bond, and, ultimately,
generate the four-centered complex IT1 (IT stands for inter-
mediates). Subsequently, the facile ligand exchange of IT1 with
Published on 04 March 2013. Downloaded by UNIVERSIDADE FEDERAL SAO CARLOS on 27/11/2013 19:03:38.

phenylboronic acid delivers the Ni(II)-borate complex IT2.


Transmetalation from IT2 to IT3 via TS4 is the rate-determining
step since a high activation energy is required. This calculated
result is also supported by the control experiments. Finally,
reductive elimination of IT3 through TS5 takes place readily to
afford the cross-coupled biaryl product.
Following these first reports, many efforts have been devoted
to improving the efficacy for cross-coupling aryl sulfamates.
In early 2011, Kappe and co-workers85 reported a procedure
with, otherwise, conditions almost identical to those employed by
Garg and Snieckus,83,84 but replacing the mode of bath heating
with microwave and the reaction time could be shortened
remarkably from 24 h to 10 min. However, the issues for the
Scheme 11 Ni-catalyzed Suzuki–Miyaura of aryl/alkenyl sulfamates with aryl requirements of relatively high catalyst loading (5 mol%) and
boronic acids; TMS = Trimethylsilyl. (Data from ref. 83). the presence of a large excess of K3PO4 base (7.0 equivalents)
remained unsolved.
Important progress for this transformation was contributed
The mechanism of this interesting transformation was by Percec’s co-workers.74,77,79 They showed in their 2011
investigated by using density functional theory (DFT).84 The report74 that in the presence of 6 mol% of Ni(cod)2 and 12 mol%
results showed that among the three possible transition states of PCy3, aryl sulfamates and a slight excess of neopentylglycolborate
TS1, TS2, and TS3 (Scheme 12, TS stands for transition state), esters (1.2 equivalents) could be coupled efficiently at room
the monoligated five-membered TS1 exhibits a much lower temperature (Scheme 13). Excellent yields were observed for
energy than the three-centered TS2 and TS3. Such a transition various combinations of a rich range of sulfamates and borate
esters including electron-rich and electron-deficient, ortho-
substituted, and heteroaromatic partners. Shortly afterwards,
the same group presented a more stable trans-Ni(II)-(1-naphthyl)-
(PPh3)2Cl complex, which also allowed for the effective coupling of
aryl sulfamates at room temperature with a loading of 5 mol%

Scheme 12 Mechanism for the NiCl2(PCy3)2-catalyzed Suzuki–Miyaura coupling Scheme 13 Ni-catalyzed Suzuki–Miyaura cross-coupling of Aryl sulfamates with
of aryl sulfamates established by computational study. (Adapted from ref. 84). borate esters at room temperature. (Data from ref. 74).

5280 Chem. Soc. Rev., 2013, 42, 5270--5298 This journal is c The Royal Society of Chemistry 2013
View Article Online

Review Article Chem Soc Rev


Published on 04 March 2013. Downloaded by UNIVERSIDADE FEDERAL SAO CARLOS on 27/11/2013 19:03:38.

Scheme 15 Ni-catalyzed Suzuki–Miyaura reactions of aryl carbamates under


Garg’s conditions. (Data from ref. 83).
Scheme 14 NiCl2(dppp)-catalyzed Suzuki–Miyaura cross-coupling of aryl
sulfamates. (Data from ref. 86).

Ni(II) complex and 10 mol% of PCy3 as the supporting ligand.77


In addition, a comprehensive study on the effect of catalysts,
solvents, bases, and the types of different boronic compounds on
the coupling efficacy was also carried out in their recent works.79
Han and Chen,86 on the basis of their previous work on the
development of highly active nickel catalysts for the cross-
coupling of aryl halides38 and sulfonates,81 have shown that
NiCl2(dppp) is also a competent catalyst for aryl sulfamates.
Substrates featuring a fused aromatic ring, nonfused but acti-
vated phenyl, and heteroaromatic sulfamates were coupled
with only a slight excess of boronic acids (1.2 equivalents) in
very high yield in the presence of only 1–1.5 mol% of catalyst
(Scheme 14). Notably, an external supporting ligand was not
needed even with such a low catalyst loading. A substrate
substituted at the ortho-position of the sulfamate was also Scheme 16 Ni-catalyzed Suzuki–Miyaura reactions of aryl carbamates under
well-tolerated. However, the reaction conditions were less Snieckus’s conditions. (Data from ref. 89).
effective for nonactivated and deactivated substrates.
2.2.3 Nickel-catalyzed Suzuki–Miyaura cross-coupling of aryl
carbamates. Of the potential phenol derivatives being studied boronic acids and base. The same reactivity difference between
in transition-metal-catalyzed cross-coupling, aryl carbamates, aryl sulfamates and carbamates was also observed by Perece and
like aryl sulfamates, also showed important advantages in coworkers79 in their comprehensive studies on the reactivity of
directed ortho metalation (DoM) and functionalization,87 and, various aryl sulfamates and carbamates under different conditions.
therefore, are also appealing electrophiles. The application of The relatively low reactivity of aryl carbamates could be
aryl carbamates in nickel-catalyzed Kumada–Tamao–Corriu explained by the data obtained through computational studies.
coupling has been investigated since the early 1990s.88 How- A detailed mechanistic study84 by DFT indicates that the activation
ever, their coupling reactions under the Suzuki–Miyaura con- energy of the rate-determining transmetalation step for carbamates
ditions were not investigated until 2009. Garg83 and Snieckus89 (DG‡ = 30.2 kcal mol1) is much higher than that for the
reported almost simultaneously the nickel catalyzed cross- corresponding sulfamates (DG‡ = 24.7 kcal mol1). As a result,
coupling of aryl carbamates with boron compounds under aryl carbamates exhibit a lower reactivity than aryl sulfamates.
similar conditions to those they used to couple aryl sulfamates83,84 It should be mentioned that, although intensive studies
(Scheme 15 and 16). The key difference between the two have suggested that the reaction mechanism of aryl sulfamates
procedures was the utilization of different boron sources. and carbamates is essentially similar,84 some minor differences
Namely, Garg used boronic acids, whereas Snieckus employed between these two reactions were still observed. Namely, in
a mixture of boroxine and boronic acid (10 : 1 mol/mol) as the Snieckus’s study,89 they found that a certain small amount of
boron source. Their results showed that, to compare with the water would be of benefit to the coupling of carbamates with
aryl sulfamates, aryl carbamates were less reactive and required boroxines. However, excessive water is detrimental to the reac-
relatively harsher conditions by means of increasing the catalyst tion. The critical role played by a small amount of water was
loading, reaction temperature, and the molar equivalents of proposed to be responsible for generating the catalytically

This journal is c The Royal Society of Chemistry 2013 Chem. Soc. Rev., 2013, 42, 5270--5298 5281
View Article Online

Chem Soc Rev Review Article


Published on 04 March 2013. Downloaded by UNIVERSIDADE FEDERAL SAO CARLOS on 27/11/2013 19:03:38.

Scheme 17 Mechanistic study on the effect of water on the cross-coupling of


aryl carbamates. (Adapted from ref. 84).

active borate species. This result implies that the coupling model Scheme 19 Ni-catalyzed Suzuki–Miyaura reactions of aryl carbamates aided by
between aryl sulfamates with boronic acids and carbamates with microwave. (Data from ref. 85).
boroxines is slightly different. The computational study indicates
that, after the oxidative addition of phenyl carbamate to the
nickel catalyst via the five-centered transition state TS1 to and unactivated alkenyl substrates were also viable substrates.
generate the addition product IT4 (Scheme 17), the four- In addition, the reaction temperature was significantly lowered
centered IT4 tends to form a more stable six-membered water- and the reaction time was also shortened under the modified
Ni-carbamate complex IT5, which is 1.1 kcal mol1 lower in conditions. Of note, one equivalent of water was also required
energy than IT4. Consequently, coordination with water in Shi’s approach.
increases the energy barrier of transmetalation, and thus further In another effort aimed at improving the reaction efficacy of
decreases the reactivity of carbamates, although a small amount carbamates, Kappe and co-workers85 reported a microwave-
of water is important for generating the reactive species. Indeed, aided approach. By employing the same catalyst, base, and
the critical role of water has also been demonstrated by Shi and solvent applied by Garg,83 an array of aryl carbamates including
co-workers90 in their earlier work in nickel-catalyzed Suzuki– the activated, nonactivated, and deactivated substrates could be
Miyaura coupling of aryl esters with boroxines (vide infra). cross-coupled with boronic acids very rapidly in high yields
Soon after these first reports, Shi and co-workers91 described within only 10 min (Scheme 19). In addition, heteroaryl carbamates
a similar catalyst system to that of Snieckus, but employed and a sterically hindered substrate were also tolerated.
K2CO3 as base and dioxane as solvent. Under the modified 2.2.4 Nickel-catalyzed Suzuki–Miyaura cross-coupling of
conditions, not only aryl but also alkenyl carbamates could be aryl and alkenyl phosphorus derivatives. Phenol and enol
coupled efficiently with boroxines to afford the desired pro- derivatives activated by phosphorus groups are also important
ducts in good to excellent yields (Scheme 18). Deactivated aryl electrophiles in the transition-metal-catalyzed cross-coupling
reactions and have been extensively studied such as in Suzuki–
Miyaura, Stille, Negishi, and Heck reactions.92 This type of
compound offers many advantages when used as an electro-
phile, stemming from their easy preparation, greater stability,
and the generation of more benign phosphorus by-products in
synthetic applications. Particularly, their highly stable nature
would be of great value not only for handling but also for the
construction of complicated molecules through orthogonal
cross-coupling reactions. However, the high stability of aryl or
alkenyl phosphorus compounds comes at the expense of their
reactivity, making the cross-coupling reactions considerably
more difficult under common Suzuki–Miyaura conditions.
Therefore, at the early stage, only the reactions of enol phos-
phates activated by an N or O heteroatom, in other words,
derived from lactones, lactams, or acyclic amides have succeeded
when using palladium or nickel catalysts93,94 (Scheme 20a–c).
The substrates derived from common ketones were less effi-
Scheme 18 Ni-catalyzed Suzuki–Miyaura reactions of aryl carbamates under cient when applying palladium catalysts even in the presence of
Shi’s conditions. (Data from ref. 91). bulky phosphine ligands (Scheme 20d).93c,94a Effective coupling

5282 Chem. Soc. Rev., 2013, 42, 5270--5298 This journal is c The Royal Society of Chemistry 2013
View Article Online

Review Article Chem Soc Rev

inert than the CVinyl–O bond, although both of them are


sp2 C–O bonds. Thus, the design of a new strategy is essential
for achieving the Suzuki–Miyaura coupling reactions of aryl
phosphorus substrates derived from phenols.
A breakthrough to this challenging transformation was first
Published on 04 March 2013. Downloaded by UNIVERSIDADE FEDERAL SAO CARLOS on 27/11/2013 19:03:38.

achieved by Han and co-workers in 2010.95 They showed that


aryl BOPs (BOP = bis(2-oxo-3-oxazolidinyl)phosphinoxyl), a type
of aryl phosphoramide prepared from phenols and BOP-Cl
[BOP-Cl = bis(2-oxo-3-oxazolidinyl)phosphinic chloride], were
efficient electrophiles for Suzuki–Miyaura coupling reactions in
the presence of 10 mol% NiCl2(dppp) catalyst generated in situ
from 10 mol% NiCl2 and 20 mol% of dppp phosphine
(Scheme 22). The reaction was performed using K2CO3 as a
base in dioxane at 100 1C. A broad range of aryl phosphor-
amides such as fused aromatic substrates, as well as non-fused
phenyl derivatives including activated, nonactivated, and deacti-
vated, and heteroaromatic substrates were coupled effectively
with an array of aryl boronic acids. High to excellent yields were
obtained for these substrates although the yield was moderate
for an ortho-methyl substituted aryl BOP. The initial idea of
using BOP-Cl as the activation reagent of phenol was inspired
by the fact that BOP-Cl is an outstanding activator of carboxylic
acids.96
Soon after the first report, Han and co-workers97 further
reported a more efficient method for the nickel-catalyzed cross-
coupling of aryl phosphonium salts via a one-pot procedure.
Scheme 20 Pd- and Ni-catalyzed Suzuki–Miyaura cross-coupling of alkenyl A broad range of phenol compounds activated in situ by using
phosphorus compounds. (Adapted from ref. 93 and 94).
PyBroP (PyBroP = bromotripyrrolidinophosphonium hexafluoro
phosphate) exhibited a very high coupling efficiency with
various aryl boronic acids in the presence of 5 mol% of
for such substrates required the use of more active nickel catalysts
(Scheme 20e).94 Nevertheless, the reaction for sterically bulky
ketone derivatives is still difficult.
To compare with the enol phosphates, the cross-coupling
reaction of the corresponding phenol phosphorus derivatives is
considerably more challenging. As illustrated in Scheme 21,
numerous previous investigations93,94 (See Scheme 20) have
clearly demonstrated that for a phosphate molecule containing
both CVinyl–O and CAryl–O bonds, the CVinyl–O bond was cleaved
chemo-specifically under Suzuki–Miyaura conditions while the
CAryl–O remained intact. In the transformation, the diphenyl
phosphoryl moiety served as a good leaving group after
abstracting an enolic oxygen atom. These reported results
unambiguously exemplified that the CAryl–O bond is far more

Scheme 21 Comparison of the cleavability of CVinyl–O and CAryl–O bonds under Scheme 22 Ni-catalyzed Suzuki–Miyaura cross-coupling of aryl BOP. (Data
Suzuki–Miyaura conditions. from ref. 95).

This journal is c The Royal Society of Chemistry 2013 Chem. Soc. Rev., 2013, 42, 5270--5298 5283
View Article Online

Chem Soc Rev Review Article

reported approach by Han,100 when the PyBroP-activated


2,4-dihydroxyquinoline or 2,4-dihydroxypridine was employed
to couple with two different boronic acids, Ar1B(OH)2 and
Ar2B(OH)2, the first added boronic acid Ar1B(OH)2 was coupled
chemo- and regioselectively at the C-2 position in the presence
Published on 04 March 2013. Downloaded by UNIVERSIDADE FEDERAL SAO CARLOS on 27/11/2013 19:03:38.

of PdCl2(dppf), and the second boronic acid Ar2B(OH)2 was


coupled exclusively at the C-4 position in the presence of
NiCl2(dppp). Such a one-pot multi-step transformation afforded
the coupled products in good yields and displayed significant
advantages in efficiency because a six-step sequence is required
if traditional cross-coupling protocols are employed to intro-
duce the two different aryl groups, including selective protec-
tion of one of the two OHs, activation of the remaining OH,
cross-coupling, deprotection of the protected OH, reactivation
of the deprotected OH, and cross-coupling.
Scheme 23 Ni-catalyzed one-pot Suzuki–Miyaura cross-coupling of phenols via In addition to the successful coupling of aryl phosphor-
the in situ activation by PyBroP. (Data from ref. 97).
amides and phosphonium salts, the nickel-catalyzed Suzuki–
Miyaura cross-coupling of aryl phosphates were also achieved
very recently by Zhao and Cheng.101 They reported that in the
NiCl2(dppp) catalyst in dioxane at 100–110 1C (Scheme 23). The presence of 10 mol% of NiCl2(PCy3)2, aryl diethyl phosphates
use of Et3N and K3PO4 as a mixed base was crucial for effective and aryl boronic acids underwent smooth coupling with K3PO4
cross-coupling. To compare with the conventional stepwise as the base in dioxane at 110 1C (Scheme 24). Generally, both
procedures involving the preparation of activated phenol electron-donating and electron-withdrawing groups in the
electrophiles followed by cross-coupling, the one-pot operation phenyl ring periphery of either coupling partner were well-
makes the coupling reaction more appealing since such a tolerated. In addition, ortho-substituted as well as hetero-
procedure is step-economic and time-saving. Moreover, the aromatic derivatives were also viable substrates. The significance
purification of the activated intermediates is also omitted. of this methodology was also demonstrated by the synthesis of
In fact, the advantages of the one-pot coupling reaction a polyaryl molecule through the sequential Suzuki–Miyaura
mediated by PyBroP have been extensively documented pre- couplings by utilizing the different leaving abilities of bromide
viously by Kang and co-workers98 and other groups99 in the and phosphate. As shown in Scheme 25, for naphthalene
palladium-catalyzed Suzuki–Miyaura and Sonogashira-type containing a bromide and phosphate leaving group, the first
coupling of tautomerizable N-heterocyclic compounds. cross-coupling occurred exclusively at the bromide site in the
The utility advantage of the PyBroP-mediated cross-coupling presence of a palladium catalyst. Subsequent cross-coupling of
of phenols has been elaborated by the straightforward installa- the phosphate catalyzed by NiCl2(PCy3)2 proceeded also
tion of two different aryl–aryl bonds in a chemoselective smoothly to afford the desired polyaryl product.
manner through a one-pot reaction (Table 1).100 Namely, a Undoubtedly, the success in the nickel-catalyzed cross-
1 : 1 mixture of an N-heterophenolic compound such as 2-OH coupling of aryl phosphorus electrophiles and boronic acids
quinoline or pyridine and a common phenol such as 1- or has advanced remarkably the Suzuki–Miyaura reaction. Parti-
2-naphthol was first activated by PyBroP. Subsequently, boronic cularly, the ability in orthogonal cross-coupling exhibited by
acid, Ar1B(OH)2, and PdCl2(dppf) were added. After stirring the aryl phosphonium salts and phosphates are apparent advan-
reaction mixture under heating for several hours, a second tages of these approaches in addition to their easy preparation
boronic acid, Ar2B(OH)2, was added and NiCl2(dppp) was and the formation of more benign phosphorus by-products as
recharged, and the reaction vessel was further heated for mentioned above.
additional hours. Surprisingly, the results showed that in the 2.2.5 Nickel-catalyzed Suzuki–Miyaura cross-coupling of
rather complicated one-pot three-step reaction, the first added aryl and alkenyl carboxylates. For a long time, aryl carboxylates
boronic acid, Ar1B(OH)2, was coupled chemoselectively with were synthesized for the purpose of phenol protection. In
2-OH quinoline or pyridine to afford the 2-N-arylated biaryls in contrast, their use as electrophiles in transition-metal-catalyzed
very high yields. The secondly added boronic acid, Ar2B(OH)2, cross-coupling has been sparsely investigated. Indeed, the
was coupled uneventfully with naphthols, also in excellent approximate bond dissociation energy of aryl C–O is much
yields. higher than the acyl C–O bond (106 vs. 80 kcal mol1).24a These
Such a highly efficient orthogonal coupling via the one-pot data imply that the cleavage of the aryl C–O bond is a consider-
procedure was further demonstrated by the chemoselective able challenge, not only due to its high stability, but also due to
assembly of two different aryl groups in one phenol molecule the difficulties of selective cleavage from the acyl C–O bond.
having two differentiated phenolic OHs (Table 2). This is more However, given that aryl carboxylates are readily available and
challenging to compare with two intermolecular OHs owing to can be potentially used as a directing group24a for the function-
the interference between the two intramolecular OHs. In the alization of aromatic rings, it would be of great value if the

5284 Chem. Soc. Rev., 2013, 42, 5270--5298 This journal is c The Royal Society of Chemistry 2013
View Article Online

Review Article Chem Soc Rev

Table 1 One-pot construction of two different aryl–aryl bonds via the Suzuki–Miyaura cross-coupling of phenols mediated by PyBroP. (Data obtained from ref. 100)
Published on 04 March 2013. Downloaded by UNIVERSIDADE FEDERAL SAO CARLOS on 27/11/2013 19:03:38.

2-OH pyridines vs. common phenol Products Yield (%)

96

90

90

77

82

70

72

74

transition-metal-catalyzed cross-coupling of aryl carboxylates to be essential. The critical role played by such a certain
can be accomplished. amount of water was assumed to maintain a good balance
In a very early study, Yamamoto and co-workers found that between promoting the reactivity of boroxine and reducing the
both aryl and acyl C–O bonds could be selectively cleaved by hydrolysis of carboxylates.
Ni(0) complex under different conditions.102 Hinted by this Almost at the same time, a similar result was reported by
important observation, Shi and co-workers90 in 2008 disclosed Garg and co-workers.103 With the same nickel catalyst and base
for the first time that NiCl2(PCy3)2 was an effective catalyst for as Shi, an array of aryl pivalates could be coupled with a variety
the Suzuki–Miyaura cross-coupling of aryl carboxylates with of aryl boronic acids in toluene at 80–130 1C, depending on the
arylboroxines. With the presence of 10 mol% catalyst and using reactivity of the substrates. Moreover, the authors demon-
K3PO4 as the base in dioxane at 110 1C, a rich range of aryl strated that a one-pot cross-coupling without the isolation of
carboxylates derived from various phenols and carboxylic acids an aryl pivalate intermediate as well as the orthogonal coupling
underwent smooth coupling to afford the biaryls in good of pivalate were also feasible. In fact, these two reports received
yields (Table 3). Of note, 0.88 equivalents of water was proved considerable attention the minute they were reported,104

This journal is c The Royal Society of Chemistry 2013 Chem. Soc. Rev., 2013, 42, 5270--5298 5285
View Article Online

Chem Soc Rev Review Article

Table 2 One-pot construction of two distinct aryl–aryl bonds via the selective
cross-coupling of two intramolecular phenolic OHs. (Data obtained from ref. 100)
Published on 04 March 2013. Downloaded by UNIVERSIDADE FEDERAL SAO CARLOS on 27/11/2013 19:03:38.

Substrate Product Substrate Product

Scheme 25 Sequential coupling at bromide and phosphate site for the con-
struction of polyarenes. (Data from ref. 101).

took place selectively at the inert aryl C–O bond rather than the
much more reactive acyl C–O bond, Liu and co-workers107
carried out a detailed theoretical study through computational
calculations. The results suggested that, in consistence with the
corresponding dissociation energy of acyl and aryl C–O bonds,
the oxidative addition of acyl C–O to Ni(0) indeed takes place
more readily than that of the aryl C–O bond as seen from the
largely different energy barrier of the two processes (DG‡ between
the initial state Ni(0)L2 and the oxidative addition transition state
for acyl and aryl C–O is ca. 14.2 kcal mol1 and 22.9 kcal mol1,
respectively). However, the oxidative addition of the acyl C–O
bond to Ni(0) is readily reversible due to the relatively low energy
difference between the transition state and the oxidative addi-
tion product (DG‡ = 25.7 kcal mol1). In contrast, the energy
barrier for the reverse reaction of the addition product of aryl
C–O is considerably higher (DG‡ = 53.2 kcal mol), indicating that
the oxidative addition of the aryl C–O to Ni(0) is an irreversible
process. Based on these data, the authors proposed that the
oxidative addition product formed from the aryl C–O with Ni(0)
would transmetalate more readily than that formed from the acyl
C–O, and, consequently, result in the selective formation of the
Scheme 24 Ni-catalyzed Suzuki–Miyaura cross-coupling of aryl phosphates.
(Data from ref. 101). aryl C–O cross-coupled product.
Concerning the catalytic cycle (Scheme 26), it is suggested
that in the oxidative addition step, a mono-phosphine rather
and have stimulated explosive interest in the nickel-catalyzed than a bis-phosphine ligated Ni(0) is preferred to coordinate
cross-coupling of other phenol and enol derivatives such as the with the aryl carboxylate substrates. Oxidative addition occurs
aryl sulfamates and carbamates developed later on (vide supra). via a three-center transition state (TS6) to form the addition
On the basis of their success in the cross-coupling of aryl product (IT6). This result is somewhat different from the
carboxylates, Shi and co-workers105 further expanded this pro- proposed reaction mechanism of aryl sulfamates and carbamates,
tocol to the alkenyl carboxylates. Both acetate and pivalate whose oxidative addition has been suggested to take place
groups were good leaving groups, affording the cross-coupled through a five-centered transition state84 (vide supra). Trans-
products in good yields (Table 4). This method displayed a metalation of IT6 and boronic K[PhB(OH)3] through the inter-
broader generality than that utilizing a rhodium catalyst, as mediate IT7 and transition state TS7 affords the IT8 whose
presented by Kuwano and Yu,106 and provided an alternative facile reductive elimination via TS8 delivers the biaryl products.
pathway for the synthesis of polysubstituted olefins from the The transmetalation step is shown to be the rate-determining
readily available carbonyl compounds. step in which the K[PhB(OH)3] nucleophile in situ-generated
To elucidate the interesting results of why and how the from the boronic acid, K3PO4 base, and a suitable amount of
cross-coupling reaction of an aryl and alkenyl ester molecule water, is the real reactive species.

5286 Chem. Soc. Rev., 2013, 42, 5270--5298 This journal is c The Royal Society of Chemistry 2013
View Article Online

Review Article Chem Soc Rev

Table 3 Ni-catalyzed Suzuki–Miyaura cross-coupling of aryl carboxylates. (Data obtained from ref. 90)
Published on 04 March 2013. Downloaded by UNIVERSIDADE FEDERAL SAO CARLOS on 27/11/2013 19:03:38.

ArOC(O)R Ar 0 in boroxine Biaryl Yield (%)

73

60

51

77

75

70

67

80

75

48

This theoretical study together with those of aryl sulfamates Mechanistic study by ESI-MS spectroscopy (ESI = electrospray
and carbamates carried out by Houk and co-workers84 not only ionization) suggested that the reaction proceeded via a different
provides a better understanding on these particular cross- reaction model from that of the nickel catalyst (Fig. 9). Namely,
coupling reactions, but, more importantly, has further implica- association of palladium hydride with vinyl acetate (I), followed
tions beyond these reactions themselves for the design of more by migratory insertion (II) and b-elimination of acetate affords
powerful catalysts towards realizing more versatile and effective the species III. Subsequently, dissociation of olefin or acetate
coupling of the inert substrates. followed by transmetalation leads to the generation of the
During the study of the nickel-catalyzed Suzuki–Miyaura arylpalladium complex (V). Finally, sequential re-association
cross-coupling of aryl and alkenyl carboxylates, Larhed and of the olefin (if the olefin dissociates in III) with complex V to
co-workers108 reported a palladium-catalyzed coupling reaction afford VI, followed by migratory insertion (VII), and b-hydride
of vinyl acetates with aryl boron reagents. Somewhat different elimination (VIII) gives the styrene product and a palladium
from the nickel-catalyzed analogue reaction (vide supra), the hydride species. To compare with the nickel-catalyzed pathway,
bidentate phosphine ligands such as dppe and dppp were oxidative addition of either vinyl C–O or acyl C–O does not
much more effective in the palladium-catalyzed coupling reaction. occur in this palladium-catalyzed cycle.

This journal is c The Royal Society of Chemistry 2013 Chem. Soc. Rev., 2013, 42, 5270--5298 5287
View Article Online

Chem Soc Rev Review Article

Table 4 Ni-catalyzed Suzuki–Miyaura cross-coupling of alkenyl carboxylates.


(Data obtained from ref. 105)
Published on 04 March 2013. Downloaded by UNIVERSIDADE FEDERAL SAO CARLOS on 27/11/2013 19:03:38.

ArOC(O)R Ar 0 Product Yield (%)

88

76

95

92
Fig. 9 Pd-catalyzed cross-coupling of vinyl acetate with boron reagents.
(Adapted from ref. 108).

89

R = Ac 85 2.2.6 Nickel-catalyzed Suzuki–Miyaura cross-coupling of


R = Piv 89 aryl and alkenyl ethers. Aryl and alkenyl ethers are readily
available and highly stable chemicals both for handling and
storage. The nickel-catalyzed reactions of such substrates with
stronger nucleophilic Grignard reagents, i.e., Kumada–Tamao–
Corriu coupling, have been reported by Wenkert in 1979,109 and
have been considerably advanced later by Dankwardt,110 and
70 Shi.111 However, the transition-metal-catalyzed Suzuki–Miyaura
cross-coupling of aryl and alkenyl ethers with weak nucleo-
philic boron reagents has been largely delayed owing to the
difficulties for the selective cleavage of the inert arene C–O
bond in arene ethers.
The first Suzuki–Miyaura cross-coupling of aryl ethers
was realized by Kakiuchi and co-workers in 2004 by using a
ruthenium catalyst.112 In the transformation, a carbonyl directing
group in the substrates was required. Four years later, significant
success has been made by Chatani and coworkers113 who
achieved the cross-coupling of aryl methyl ethers with aryl
boronic esters catalyzed by a nickel catalyst (Scheme 27). By
applying 10 mol% of Ni(cod)2 and 40 mol% of the PCy3
supporting ligand, fused aromatic substrates, i.e., electron
deficient aryl methyl ethers, could be coupled in good yields
using CsF as a base in toluene at 120 1C. However, these
conditions were less effective for the non-fused phenyl methyl
ethers, i.e., the less electron deficient substrates. In addition,
alkenyl methyl ethers were also demonstrated to be compatible
to the same reaction conditions (Scheme 28).114
2.2.7 Nickel-catalyzed Suzuki–Miyaura cross-coupling of
aryl carbonates. Aryl carbonates are also an important class of
Scheme 26 Mechanism for the NiCl2(PCy3)2-catalyzed Suzuki–Miyaura
electrophiles in nickel-catalyzed Suzuki–Miyaura cross-coupling,
coupling of aryl carboxylates established by computational study. (Adapted although their use was less popular than the corresponding aryl
from ref. 107). sulfonates, sulfamates, carbamates, carboxylates, and phosphorus

5288 Chem. Soc. Rev., 2013, 42, 5270--5298 This journal is c The Royal Society of Chemistry 2013
View Article Online

Review Article Chem Soc Rev


Published on 04 March 2013. Downloaded by UNIVERSIDADE FEDERAL SAO CARLOS on 27/11/2013 19:03:38.

Scheme 27 Ni-catalyzed Suzuki–Miyaura cross-coupling of aryl methyl ethers.


(Data from ref. 113).

Scheme 30 Ni-catalyzed Suzuki–Miyaura cross-coupling of aryl methylcarbo-


nates. (Data from ref. 115).

Scheme 31 Effect of ligands on the coupling efficiency of aryl carbonate and


carboxylate. (Data from 115).

Scheme 28 Ni-catalyzed Suzuki–Miyaura cross-coupling of alkenyl methyl Recently, an improved method was reported by Kuwano and
ethers. (Data from ref. 114). Shimizu.115 By using 10 mol% of Ni(cod)2 catalyst and an equal
molar amount of the DCyPF bidentate ligand, the fused
naphthyl and activated phenyl methylcarbonates could be
compounds (vide supra). In the pioneering work of Garg,83 the
coupled with various aryl boronic acids in the presence of
coupling efficiency of three naphthyl tert-butylcarbonates was
K2CO3 base in toluene at a lowered temperature of 60 1C to
examined in the presence of a nickel catalyst (Scheme 29). Good
afford the biaryls in high yields (Scheme 30). However, the
yields were obtained for the reactions in the presence of 10 mol%
deactivated aryl carbonates were reluctant substrates. Compe-
of NiCl2(PCy3)2 catalyst and K3PO4 base in toluene at 130 1C.
titive experiments showed that the bidentate ligands such as
DCyPF were more effective than the monodentate ligand PCy3
for coupling aryl carbonates (Scheme 31). However, an opposite
outcome was obtained when the two ligands were used to
couple aryl acetate. Although there was no detailed mechanistic
study, the results indicated that the reaction model of aryl
carbonates and carboxylates should be somewhat different.
2.2.8 Nickel-catalyzed direct Suzuki–Miyaura cross-coupling
of phenol via mutual activation strategy. In numerous efforts
aimed at achieving the efficient cross-coupling of phenol and
enol compounds, the most popularly used strategy is the discovery
Scheme 29 Ni-catalyzed Suzuki–Miyaura cross-coupling of naphthyl tert-butyl of an appropriate activation reagent that is not only cheap,
carbonates. (Data from 83). readily available, and easy to react with hydroxy group of phenols,

This journal is c The Royal Society of Chemistry 2013 Chem. Soc. Rev., 2013, 42, 5270--5298 5289
View Article Online

Chem Soc Rev Review Article

but also has the ability to activate the aryl C–O bonds. Mean- intermediates, the borate moiety might be acting not only as a
while, the activated phenol intermediates should have good leaving group but also an activating group to activate the phenolic
stability for handling. Such a strategy requires multisteps, and, CAr–O bond. At the same time, the formation of the borate anion
therefore, resulting in an inevitable waste of time, energy, also activates the CAr–B bond, and, thereby, facilitating the sub-
reagents, and other related chemical materials such as solvents, sequent transmetalation reaction. This strategy was further supported
Published on 04 March 2013. Downloaded by UNIVERSIDADE FEDERAL SAO CARLOS on 27/11/2013 19:03:38.

as well as the associated environmental resources. Obviously, the by the single crystal structure of the borate anion intermediate.
direct use of phenol as an electrophile for cross-coupling is the A mechanistic study indicates that the presence of the BEt3 additive
most ideal option. However, such coupling is extremely challenging may be acting to activate the borate anion as a second Lewis base
due to the poor leaving ability of the phenolic OH group. and/or helping the oxidative addition as a radical initiator.
An important breakthrough to this challenging issue was
contributed from Shi’s group in 2011.116 Based on their precedent 2.3 Nickel-catalyzed Suzuki–Miyaura cross-coupling of other
work in the Kumada–Tamao–Corriu coupling of phenolates with aryl electrophiles
Grignard reagents,117 Shi and co-workers developed a mutual
In general, when the term ‘‘transition-metal-catalyzed cross-
activation strategy, which allowed for the direct Suzuki–Miyaura
coupling’’ such as Suzuki, Heck, Negishi, Stille, and Kumada is
cross-coupling of phenols with boroxines by applying a nickel
mentioned, the first consideration is the use of aryl halides or
catalyst. An array of naphthols could be coupled effectively with
phenol derivatives as electrophiles. Research into the other type
various boroxines under the optimized conditions: 10 mol% of
of substrates is notably deficient owing to the inconvenient
Ni(cod)2 catalyst, 40 mol% of PCy3 ligand, 1.0 equivalent of NaH
availability and/or inert nature. However, exploration of the
base, and 1.5–3.0 equivalents of BEt3 additive in mixed o-xylene
transition-metal-catalyzed coupling reactions of the ‘‘unordinary’’
and THF solvent at 110 1C (Scheme 32).
electrophiles is also greatly interesting not only due to academic
In their rational design, the authors conceived that phenolates
curiosity but also providing more options when a particular
generated from phenols and bases should react with the boronic
application is under consideration.
reagents to generate borate anions (Scheme 33). In the borate anion
As a typical example, 6-arylpurine ribonucleosides possess
interesting cytostatic activity. However, their synthesis via the
classic Suzuki–Miyaura cross-coupling of 6-(pseudo)halide sub-
strates was problematic due to the difficulties for the prepara-
tion of 6-halopurine electrophiles. As a solution, Robins and
Liu118 have developed a convenient method by applying the
nickel-catalyzed Suzuki–Miyaura coupling reaction of 6-(imid-
azol-1-yl)purine or 6-(1,2,4-triazol-4-yl)purine derivatives with
various aryl boronic acids (Scheme 34a and b). Both imidazolyl
Scheme 33 Concept of mutual activation of phenol and boronic reagent.
(Adapted from ref. 116).

Scheme 32 Ni-catalyzed direct Suzuki–Miyaura cross-coupling of phenolic com- Scheme 34 Ni-catalyzed Suzuki–Miyaura cross-coupling using imidazole and
pounds with boroxines. (Data from ref. 116). triazole as leaving groups. (Data from 118).

5290 Chem. Soc. Rev., 2013, 42, 5270--5298 This journal is c The Royal Society of Chemistry 2013
View Article Online

Review Article Chem Soc Rev


Published on 04 March 2013. Downloaded by UNIVERSIDADE FEDERAL SAO CARLOS on 27/11/2013 19:03:38.

Scheme 36 Ni-catalyzed Suzuki–Miyaura cross-coupling of aryl nitriles with


Scheme 35 Ni-catalyzed Suzuki–Miyaura cross-coupling of aryltrimethylammonium borate esters. (Data from ref. 122).
salts. (Data from ref. 119).

their equivalents also represents an important class of coupling


and triazolyl groups were good leaving groups in the presence reactions. However, the reactions of alkyl electrophiles are
of a Ni(cod)2 catalyst. considerably more challenging as compared to the corresponding
In addition, the nickel-catalyzed cross-coupling by applying aryl and alkenyl analogues due to the difficulties to undergo
aryltrimethylammonium salts, derived from N,N-dimethyl aryl- oxidative addition as well as the tendency to participate in some
amines, as electrophiles were also reported by MacMillan and side reactions such as b-H elimination and hydrodehalogenation.123
Blakey.119 Under the carefully optimized conditions: 10 mol% Despite these intrinsic difficulties, remarkable progress has been
of Ni(cod)2 and 10 mol% of the IMesHCl ligand, and CsF as a made in the Suzuki–Miyaura cross-coupling of alkyl (pseudo)halides
base in dioxane at 80 1C, the reaction displayed a broad since the palladium-catalyzed reaction of alkyl iodides with borane
generality to either coupling partners including those modified reagents was reported in 1992 by Suzuki and co-workers.124,125
by electron-rich and deficient groups as well as sterically
congested analogues (Scheme 35). This method together with 3.1 Ni-catalyzed Suzuki–Miyaura cross-coupling of alkyl
the early reported Suzuki–Miyaura coupling of aryl diazonium (pseudo)halides with aryl and alkenyl boron reagents
salts catalyzed by a palladium catalyst120 expanded significantly
the substrate scope of aryl amino compounds. The first nickel-catalyzed Suzuki–Miyaura cross-coupling of alkyl
In the efforts to expand the substrate scope of electrophiles, halides was described by Fu and Zhou in 2004.126 By applying a
several early reports have shown that nitrile groups were also catalyst system composed of 4 mol% of Ni(cod)2 and 8 mol% of
competent leaving groups in the nickel-catalyzed couplings bathophenanthroline, cyclic and acyclic secondary as well as
with Grignard and zinc reagents.121 Recently, Shi and co-workers122 primary alkyl iodides, and cyclic and acyclic secondary bromides
have achieved the nickel-catalyzed Suzuki–Miyaura coupling could be coupled with a range of aryl and alkenyl boronic acids in
reaction of aryl nitriles with boronic esters. With 10 mol% moderate to high yields in the presence of KOtBu base in s-BuOH
NiCl2(PCy3)2 and 20 mol% PCy3 coupled with the presence of at 60 1C (Scheme 37). Interestingly, the more stable exo-products
4.0 equivalents of KOtBu base and 1.5 equivalents of CuF2 were obtained predominantly either from the exo- or endo-2-
additive, various aryl nitriles and neopentylglycolborate esters bromonorbornane substrates. Later extensive mechanistic inves-
including the 2-heteroaryl substrates of either partner could be tigations suggested that this is probably attributed to the radical
coupled smoothly to afford the biaryl products in moderate to pathway for oxidative addition (see Section 3.4).
good yields (Scheme 36). Following this pioneering work, a more robust and versatile
method was contributed by the same group two years later.127
They have exemplified that a combination of 6 mol% of NiI2
3. Nickel-catalyzed Suzuki–Miyaura and 6 mol% of trans-2-aminocyclohexanol as catalyst could
cross-coupling of alkyl electrophiles expand significantly the substrate scope. Both secondary and
primary iodides or bromides with functional substituents were
Equally significant to aryl and alkenyl (pseudo)halides, the coupled with aryl boronic acids in high to excellent yields.
transition-metal-catalyzed cross-coupling of alkyl halides or The generality of the improved methodology was further

This journal is c The Royal Society of Chemistry 2013 Chem. Soc. Rev., 2013, 42, 5270--5298 5291
View Article Online

Chem Soc Rev Review Article


Published on 04 March 2013. Downloaded by UNIVERSIDADE FEDERAL SAO CARLOS on 27/11/2013 19:03:38.

Scheme 40 Ni-catalyzed Suzuki–Miyaura cross-coupling of a-pivaloxyl ketones


with aryl boronic acids. (Data from ref. 132).

Scheme 37 Ni-catalyzed Suzuki–Miyaura cross-coupling of alkyl bromides and


iodides with aryl and alkenyl boronic acids. (Data from ref. 126). coupled smoothly with a rich range of potassium aryltrifluo-
borates decorated by electron-rich and deficient substituents.
Moreover, ortho-substituted and heteroaromatic trifluoroborates
including the thermally labile 2-heterocyclic analogues could
be coupled with alkyl bromides and iodides in good yields.
In addition, alkyl chlorides were also feasible electrophiles
although moderate yields were obtained.
Scheme 38 Ni-catalyzed Suzuki–Miyaura cross-coupling for the synthesis of In a recent report by Radius and Zell,130 a NHC-stablized
(0)
aryloxetanes and arylazetidines; NaHMDS = Sodium hexamethyldisilazide. Ni complex displayed highly catalytic activity for the coupling
(Adapted from ref. 128). reaction of benzyl chloride and phenyl boronic acid. With the
presence of only 1.22 mol% of nickel complex, the reaction was
completed in two hours to afford the diphenylmethane product
demonstrated by Duncton and co-workers128 for the synthesis of
in 86% yield. However, the substrate scope and limitation of
aryloxyetanes and arylazetidines, although moderate yields were
the catalyst were not examined.
obtained for this particular transformation (Scheme 38). As also
The nickel-catalyzed Suzuki–Miyaura cross-coupling of alkyl
an important contribution from Fu and González-Bobes,127 they
halides with aryl and alkenyl boronic acids has been widely
have discovered that the use of a NiCl2–glyme complex and
investigated especially by Fu’s group. In contrast, the substrates
prolinol ligand enabled the cross-coupling of the notoriously
derived from alcohols such as alkyl tosylates were reluctant
inert secondary and primary alkyl chlorides (Scheme 39).
although their corresponding Kumada–Tamao–Corriu reaction
By employing a similar catalyst system, Molander and
with stronger Grignard reagents has been reported at the early
co-workers129 showed that various alkyl halides could also be
stage.131 Only very recently, Shi and co-workers132 disclosed that
pivaloyl esters derived from a-hydroxy ketones were viable sub-
strates for the nickel-catalyzed Suzuki–Miyaura cross-coupling.
The reaction was carried out in the presence of 10 mol% of NiCl2-
(PCy3)2 and 2.0 equivalents of NaOtBu base. Various a-pivaloxyl
ketones were coupled with a range of aryl boronic acids in a mixed
toluene and DMF solvent at 100 1C in moderate to high yields
(Scheme 40). Although the substrate scope was limited to alcohols
activated by a carbonyl group, this initial result provided an
important clue for the development of more powerful methods
for coupling unactivated alcoholic derivatives, e.g., by choosing
alternative activation strategy or designing new catalyst systems.

3.2 Ni-catalyzed Suzuki–Miyaura cross-coupling of alkyl


halides with alkyl borane reagents
Scheme 39 Ni-catalyzed Suzuki–Miyaura cross-coupling of alkyl chlorides with
aryl boronic acids; KHMDS = Potassium hexamethyldisilazide; TBS = tert-butyl- With their great success in nickel-catalyzed Suzuki–Miyaura
dimethylsilyl. (Data from ref. 127). coupling reactions of alkyl halides with aryl or alkenyl boronic acids,

5292 Chem. Soc. Rev., 2013, 42, 5270--5298 This journal is c The Royal Society of Chemistry 2013
View Article Online

Review Article Chem Soc Rev


Published on 04 March 2013. Downloaded by UNIVERSIDADE FEDERAL SAO CARLOS on 27/11/2013 19:03:38.

Scheme 41 Ni-catalyzed Suzuki–Miyaura cross-coupling of alkyl bromides and


iodides with alkyl borane reagents; Cbz = Benzyloxylcarbonyl. (Data from 133).

Fu and Saito133 achieved the more challenging alkyl–alkyl


Suzuki–Miyaura cross-coupling with a nickel catalyst in 2007.
By employing an optimized catalyst system composed of 6 mol%
of NiCl2 glyme and 8 mol% of the trans-N,N 0 -dimethylcyclo-
hexanediamine ligand, the cross-coupling of a variety of
secondary as well as primary bromides and iodides with alkyl Scheme 42 Ni-catalyzed asymmetric Suzuki–Miyaura cross-coupling of secondary
9-BBN (BBN = 9-Borabicyclo[3,3,1]nonyl) borane reagents pro- homobenzylic bromides with alkyl boranes. (Data from ref. 135).

ceeded smoothly even at room temperature in dioxane solvent


(Scheme 41). Interestingly, two equivalents of i-BuOH coupled
with the use of 1.2 equivalents of KOtBu base was essentially The established reaction conditions were Ni(cod)2 (10 mol%),
required, which was proposed to activate the alkylborane for chiral diamine ligand 1 (12 mol%), KOtBu base (1.2 equivalents)
transmetalation with nickel metal. and i-BuOH additive (2.0 equivalents) in di-iso-propyl ether
For the less reactive alkyl chlorides, they found that just by solvent at 5 1C. Under these conditions, acyclic secondary alkyl
replacing the NiCl2 glyme, trans-N,N 0 -dimethylcyclohexane bromides having an aryl substitution at the a-position of
diamine ligand, and dioxane solvent, which was an effective bromide were coupled effectively with various alkyl boranes.
combination for the coupling alkyl iodides and bromides, with For such substrates, high enantioselectivities as well as good
the corresponding congeners of NiBr2 glyme, trans-N,N 0 - yields were observed (Scheme 42a). However, for an acyclic
dimethyl-1,2-diphenylethylenediamine, and iso-propyl ether, substrate whose aryl group was located at the b-position and a
respectively, a broad range of secondary alkyl chlorides including cyclic substrate, only poor enantioselectivity was obtained
cyclic and acyclic, as well as functionalized substrates could (Scheme 42b). The authors suggested that such a strong
also be coupled with 9-BBN boranes in high yields at room dependence of enantioselectivity on the position of the aryl
temperature.134 In addition, a primary alkyl chloride was also a substituents may be attributed to the existence of an appro-
viable coupling partner although the substrate scope was not priate secondary interaction between the CH2Ar and the chiral
extensively inspected. catalyst.
On the basis of this fundamental discovery, Fu and
3.3 Ni-catalyzed asymmetric Suzuki–Miyaura cross-coupling co-workers carried out comprehensive studies and established
of alkyl halides with borane reagents several sets of specified conditions to couple various types of
Having extensively investigated the nickel-catalyzed racemic alkyl halide electrophiles. As summarized in Scheme 43, by
alkyl–aryl and alkyl–alkyl couplings, chemists shifted their atten- employing the identical catalyst NiBr2 diglyme and base KOtBu
tions to the reactions of an asymmetric version. The first successful but through finely tuning the chiral ligands, additives, and
protocol was developed by Fu and Saito in 2008 (Scheme 42).135 solvents, asymmetric Suzuki–Miyaura cross-coupling of various

This journal is c The Royal Society of Chemistry 2013 Chem. Soc. Rev., 2013, 42, 5270--5298 5293
View Article Online

Chem Soc Rev Review Article


Published on 04 March 2013. Downloaded by UNIVERSIDADE FEDERAL SAO CARLOS on 27/11/2013 19:03:38.

Fig. 10 Proposed mechanism for the Ni-catalyzed cross-coupling of alkyl


halides. (Adapted from ref. 138–140).

for the preparation of synthetically useful chiral intermediates


and building blocks with versatile functionalities.

3.4 Reaction mechanism of Ni-catalyzed Suzuki–Miyaura


cross-coupling of alkyl halides
Based on the extensive experimental and theoretical
studies,136b,138–140 it is proposed that a Ni(I) to Ni(III) catalytic
cycle is the most favourable way for nickel-catalyzed Suzuki–
Miyaura cross-coupling of alkyl halides (Fig. 10). In the initial
step, Ni(I)(L)X (L = ligands, X = halides), generated presumably
through the comproportionation of Ni(II) and Ni(0) complexes or
the disproportionation of the Ni(II) complex, was first trans-
metalated with boron reagents to form a Ni(I)(L)(R1) complex.
Then, oxidative addition of alkyl R2X to Ni(I)(L)(R1) produces the
Ni(III)(L)(R1)(R2)X through a radical pathway. Finally, reductive
elimination of the Ni(III) complex delivers the cross-coupled
products R1–R2 and regenerates the active catalyst species
Ni(I)(L)X. Of note, this catalytic cycle somewhat deviates from
the case of aryl (pseudo)halides whose coupling is generally
considered to undergo a Ni(0) to Ni(II) model although the Ni(I) to
Ni(III) process has been proposed in very few reports (vide supra).
In addition, unlike the aryl halides, the transmetalation of the
borane reagent to Ni(I) takes place prior to the oxidative addition
step in the catalytic cycle of alkyl halides.

4. Conclusions
This review summarized the progress in nickel-catalyzed
Scheme 43 Ni-catalyzed asymmetric Suzuki–Miyaura cross-coupling of various
alkyl halide electrophiles. (Adapted from ref. 136–139). Suzuki–Miyaura cross-coupling reactions since 1995. In the
not-too-long research history, especially the most recent
five years, we have observed an explosive development of this
types of alkyl halides was achieved with satisfactory yields and chemistry. Of particular importance is that the coupling reac-
high enantioselectivities. So far, the investigated substrates included tions of a broad range of aryl electrophiles such as sulfamates,
a-, g-, and d-haloamides136 (Scheme 43a), the unactivated secondary carbamates, carboxylates, ethers, carbonates, phosphoramides,
alkyl bromohydrin dreivatives137 (Scheme 43b), the unactivated phosphonium salts, phosphates, and even the common phenols,
secondary a-halogen tertiary amines138 (Scheme 43c), the as well as a broad range of alkyl substrates including both
a-halo-N-carbamates and sulfonamides139 (Scheme 43d and e), secondary and primary alkyl iodides, bromides, and chlorides
and the b-sulfones139 (Scheme 43f). In addition to the signifi- have been achieved by using purpose-designed nickel catalysts.
cantly broadened substrate scope, the N-based functionalities However, these substrates are usually unreactive in the
(Scheme 43c–e) and the sulfone group (Scheme 43f) were found presence of palladium-based catalyst systems. In addition,
to play a crucial role in directing the asymmetric couplings. some types of substrates such as aryl halides and sulfonates,
Consequently, these comprehensive studies not only estab- which were coupled conventionally by applying palladium or a
lished some fundamental principles for the redesign of new high amount of nickel catalyst could now be achieved with very
catalyst systems, but more importantly, provided powerful tools low level of nickel catalysts. Needless to say, these achievements

5294 Chem. Soc. Rev., 2013, 42, 5270--5298 This journal is c The Royal Society of Chemistry 2013
View Article Online

Review Article Chem Soc Rev

have advanced greatly the chemistry of transition-metal-catalyzed 67, 5553; (d) G. Altenhoff, R. Goddard, C. W. Lehmann
cross-coupling reactions. Despite these remarkable progresses, and F. Glorius, Angew. Chem., Int. Ed., 2003, 42, 3690;
many more challenging works still await further investigation. (e) S. Wan, S. R. Wang and W. Lu, J. Org. Chem., 2006,
Many of the newly developed reactions still require high catalyst 71, 4349; ( f ) T. Hama, D. A. Culkin and J. F. Hartwig, J. Am.
loading paired with the use of a large excess of phosphine or Chem. Soc., 2006, 128, 4976; ( g) M. R. Biscoe, B. P. Fors and
Published on 04 March 2013. Downloaded by UNIVERSIDADE FEDERAL SAO CARLOS on 27/11/2013 19:03:38.

NHC ligands. Moreover, some of the catalysts are air and S. L. Buchwald, J. Am. Chem. Soc., 2008, 130, 6686;
moisture sensitive. These shortcomings limit the practicality of (h) T. Ogata and J. F. Hartwig, J. Am. Chem. Soc., 2008,
the methods in potential industrial processes. In fact, although 130, 13848; (i) C. V. Reddy, J. V. Kingston and J. G. Verkade,
tremendous lab research has succeeded in nickel-catalyzed J. Org. Chem., 2008, 73, 3047; ( j ) B. Gabriele, R. Mancuso,
Suzuki–Miyaura cross-coupling, their industrial applications are G. Salerno and P. Plastina, J. Org. Chem., 2008, 73, 756.
far less successful than the corresponding palladium catalysts. In 12 For a representative example, see: J. L. Bolliger and
addition, the cross-couplings of sterically congested aryl and alkyl C. M. Frech, Chem.–Eur. J., 2010, 16, 4075.
substrates by applying nickel catalysts have been largely unex- 13 For recent reviews, see: (a) R. Jana, T. P. Pathak and
plored. Finally, alcohols are one of the most abundant chemicals M. S. Sigman, Chem. Rev., 2011, 111, 1417; (b) C. Valente,
from nature, however, the utilization of these natural resources in S. Çalimsiz, K. H. Hoi, D. Mallik, M. Sayah and M. G.
Suzuki–Miyaura cross-coupling is almost unknown. Chemists Organ, Angew. Chem., Int. Ed., 2012, 51, 3314. For selected
have created great achievements, but they are facing more examples, see: (c) J. Yin, M. P. Rainka, X.-X. Zhang and S. L.
challenges. Let’s await surprises! Buchwald, J. Am. Chem. Soc., 2002, 124, 1162; (d) B. Bhayana,
B. P. Fors and S. L. Buchwald, Org. Lett., 2009, 11, 3954.
Acknowledgements 14 For some reviews, see: (a) A. F. Littke and G. C. Fu, Angew.
Chem., Int. Ed., 2002, 41, 4176; (b) S. Würtz and F. Glorius,
The author would like to acknowledge the ‘‘Hundred Talent Acc. Chem. Res., 2008, 41, 1523.
Program’’ of Chinese Academy of Sciences (CAS), State Key 15 For selected examples, see: (a) M. Huffman and N. Yasuda,
Laboratory of Fine Chemicals (KF1201), and the Innovation Synlett, 1999, 471; (b) H. Nguyen, X. Huang and S. L.
Program of Jilin Province (20111802) for financial support. Buchwald, J. Am. Chem. Soc., 2003, 125, 11818; (c) J. Wu,
Q. Zhu, L. Wang, R. Fathi and Z. Yang, J. Org. Chem., 2003,
Notes and references 68, 670; (d) D. Steinhuebel, J. M. Baxter, M. Palucki and
I. W. Davies, J. Org. Chem., 2005, 70, 10124; (e) J. M. Baxter,
1 A. Suzuki, Pure Appl. Chem., 1985, 57, 1749. D. Steinhuebel, M. Palucki and I. W. Davies, Org. Lett.,
2 R. F. Heck, Org. React., 1982, 27, 345. 2005, 7, 215; ( f ) L. Zhang, T. Meng and J. Wu, J. Org.
3 (a) D. Milstein and J. K. Stille, J. Am. Chem. Soc., 1979, Chem., 2007, 72, 9346; ( g) B. Bhayana, B. P. Fors and
101, 4981; (b) D. Milstein and J. K. Stille, J. Am. Chem. Soc., S. L. Buchwald, Org. Lett., 2009, 11, 3954.
1979, 101, 4992; (c) D. Milstein and J. K. Stille, J. Org. 16 T. Kinzel, Y. Zhang and S. L. Buchwald, J. Am. Chem. Soc.,
Chem., 1979, 44, 1613. 2010, 132, 14073.
4 E. I. Negishi and F. Liu, in Metal-Catalyzed Cross-Coupling 17 (a) A. N. Cammidge and K. V. L. Crépy, Chem. Commun.,
Reactions, ed. F. Diederich and P. J. Stang, Wiley-VCH, 2000, 1723; (b) J. Yin and S. L. Buchwald, J. Am. Chem. Soc.,
Weinheim, Germany, 1998, pp. 1–47. 2000, 122, 12051; (c) A.-S. Castanet, F. Colobert,
5 (a) E. Negishi, Acc. Chem. Res., 1982, 15, 340; (b) E. Erdik, P.-E. Broutin and M. Obringer, Tetrahedron: Asymmetry,
Tetrahedron, 1992, 48, 9577. 2002, 13, 659; (d) A. Herrbach, A. Marinetti, O. Baudoin,
6 Y. Nakao and T. Hiyama, Chem. Soc. Rev., 2011, 40, 4893; D. Guénard and F. Guéritte, J. Org. Chem., 2003, 68, 4897;
and references therein. (e) J. F. Jensen and M. Johannsen, Org. Lett., 2003, 5, 3025;
7 K. Sonogashira, in Comp. Org. Synth., ed. B. M. Trost and ( f ) K. Mikami, T. Miyamoto and M. Hatano, Chem. Com-
I. Fleming, Pergamon, Oxford, 1991, vol. 3, p. 521. mun., 2004, 2082; (g) A. N. Cammidge and K. V. L. Crépy,
8 For a recent review, see: C. Torborg and M. Beller, Tetrahedron, 2004, 60, 4377; (h) O. Baudoin, Eur. J. Org.
Adv. Synth. Catal., 2009, 351, 3027; and related references Chem., 2005, 4223; (i) M. Genov, A. Almorı́n and P. Espinet,
therein. Chem.–Eur. J., 2006, 12, 9346; ( j) A. Bermejo, A. Ros,
9 (a) N. Miyaura and A. Suzuki, J. Chem. Soc., Chem. Commun., R. Fernández and J. M. Lassaletta, J. Am. Chem. Soc.,
1979, 866; (b) N. Miyaura, K. Yamada and A. Suzuki, 2008, 130, 15798; (k) K. Sawai, R. Tatumi, T. Nakahodo
Tetrahedron Lett., 1979, 20, 3437; (c) N. Miyaura, T. Yanagi and H. Fujihara, Angew. Chem., Int. Ed., 2008, 47, 6917;
and A. Suzuki, Synth. Commun., 1981, 11, 513. (l ) Y. Uozumi, Y. Matsuura, T. Arakawa and Y. M. A.
10 N. Miyaura and A. Suzuki, Chem. Rev., 1995, 95, 2457. Yamada, Angew. Chem., Int. Ed., 2009, 48, 2708; (m) X. Shen,
11 For selected examples, see: (a) J. P. Wolfe, R. A. Singer, G. O. Jones, D. A. Watson, B. Bhayana and S. L. Buchwald,
B. H. Yang and S. L. Buchwald, J. Am. Chem. Soc., 1999, J. Am. Chem. Soc., 2010, 132, 11278; (n) S.-S. Zhang,
121, 9550; (b) A. F. Littke, C. Dai and G. C. Fu, J. Am. Chem. Z.-Q. Wang, M.-H. Xu and G.-Q. Lin, Org. Lett., 2010,
Soc., 2000, 122, 4020; (c) N. Kataoka, Q. Shelby, 12, 5546; (o) T. Yamamoto, Y. Akai, Y. Nagata and
J. P. Stambuli and J. F. Hartwig, J. Org. Chem., 2002, M. Suginome, Angew. Chem., Int. Ed., 2011, 50, 8844;

This journal is c The Royal Society of Chemistry 2013 Chem. Soc. Rev., 2013, 42, 5270--5298 5295
View Article Online

Chem Soc Rev Review Article

(p) L. Sun and W.-M. Dai, Tetrahedron, 2011, 67, 9072; 32 V. Percec, G. M. Golding, J. Smidrkal and O. Weichold,
(q) W. Tang, N. D. Patel, G. Xu, X. Xu, J. Savoie, S. Ma, J. Org. Chem., 2004, 69, 3447.
M.-H. Hao, S. Keshipeddy, A. G. Capacci, X. Wei, Y. Zhang, 33 B. M. Rosen, C. Huang and V. Percec, Org. Lett., 2008,
J. J. Gao, W. Li, S. Rodriguez, B. Z. Lu, N. K. Yee and 10, 2597.
C. H. Senanayake, Org. Lett., 2012, 14, 2258. 34 L. Zhou, Q. Miao, R. He, X. Feng and M. Bao, Tetrahedron
Published on 04 March 2013. Downloaded by UNIVERSIDADE FEDERAL SAO CARLOS on 27/11/2013 19:03:38.

18 For a recent review on nanocatalysts for Suzki-Miyaura cross- Lett., 2007, 48, 7899.
coupling, see: A. Fihri, M. Bouhrara, B. Nekoueishahraki, 35 (a) Z.-Y. Tang, S. Spinella and Q.-S. Hu, Tetrahedron Lett.,
J.-M. Basset and V. Polshettiwar, Chem. Soc. Rev., 2011, 2006, 47, 2427; (b) Z.-Y. Tang and Q.-S. Hu, J. Org. Chem.,
40, 5181. 2006, 71, 2167.
19 (a) B. M. Trost, Science, 1991, 254, 1471; (b) B. M. Trost, 36 D. A. Wilson, C. J. Wilson, B. M. Rosen and V. Percec, Org.
Science, 1983, 219, 245; (c) B. M. Trost and G. Dong, Nature, Lett., 2008, 10, 4879.
2008, 456, 485. 37 (a) C. Chen and L.-M. Yang, Tetrahedron Lett., 2007,
20 Á. Molnár, Chem. Rev., 2011, 111, 2251. 48, 2427; (b) X.-H. Fan and L.-M. Yang, Eur. J. Org. Chem.,
21 X.-F. Wu, P. Anbarasan, H. Neumann and M. Beller, Angew. 2011, 1467.
Chem., Int. Ed., 2010, 49, 9047. 38 Y.-L. Zhao, Y. Li, S.-M. Li, Y.-G. Zhou, F.-Y. Sun, L.-X. Gao
22 The approximate cost for some commonly used palladium and F.-S. Han, Adv. Synth. Catal., 2011, 353, 1543.
and nickel catalysts by Aldrich Chemical Co., Inc. are: 39 (a) A. M. Rouhi, Chem. Eng. News, 2004, 82, 49; (b) J.
PdCl2 (99%) = RMB 8.8  104 per mol; PdCl2(dppp) = Ehrenfreund, C. Lamberth, H. Tobler and H. Walter, WO
RMB 3.98  105 per mol; PdCl2(PCy3)2 (95%) = RMB 9.1  Patent, 2004058723; (c) F.-X. Felpin, E. Fouquet and C. Zakri,
105 per mol; NiCl2 (98%) = RMB 1.3  103 per mol; Adv. Synth. Catal., 2009, 351, 649.
NiCl2(dppp) = RMB 8.3  104 per mol; NiCl2(PCy3)2 40 (a) C. A. Bernhart, P. M. Perreaut, B. P. Ferrari, Y. A.
(97%) = RMB 3.1  105 per mol. Muneaux, J.-L. A. Assens, J. Clément, F. Haudricourt,
23 For some recent reviews, see: (a) X. Chen, K. M. Engle, C. F. Muneaux, J. E. Taillades, M.-A. Vignal, J. Gougat,
D.-H. Wang and J.-Q. Yu, Angew. Chem., Int. Ed., 2009, P. R. Guiraudou, C. A. Lacour, A. Roccon, C. F. Cazaubon,
48, 5094; (b) L.-M. Xu, B.-J. Li, Z. Yang and Z.-J. Shi, Chem. J.-C. Brelière, G. L. Fur and D. Nisato, J. Med. Chem., 1993,
Soc. Rev., 2010, 39, 712; (c) C. C. C. Johansson and 36, 3371; (b) P. Bühlmayer, P. Furet, L. Criscione,
T. J. Colacot, Angew. Chem., Int. Ed., 2010, 49, 676; M. de Gasparo, S. Whitebread, T. Schmidlin, R. Lattmann
(d) A. J. Hickman and M. S. Sanford, Nature, 2012, 484, 177. and J. Wood, Bioorg. Med. Chem. Lett., 1994, 4, 29.
24 For recent reviews containing the Ni-catalyzed cross- 41 (a) S. Cacchi and G. Fabrizi, Chem. Rev., 2005, 105, 2873;
couplings involving C–O bond, see: (a) D.-G. Yu, B.-J. Li (b) P. Herbert, F. Christoph, K. Renat, A. Juan, M. Axel and
and Z.-J. Shi, Acc. Chem. Res., 2010, 43, 1486; (b) B. M. R. Thomas, WO Patent, 2008025673.
Rosen, K. W. Quasdorf, D. A. Wilson, N. Zhang, 42 Unpublished results.
A.-M. Resmerita, N. K. Garg and V. Percec, Chem. Rev., 43 L. Wu, J. Ling and Z.-Q. Wu, Adv. Synth. Catal., 2011,
2011, 111, 1346; (c) B.-J. Li, D.-G. Yu, C.-L. Sun and Z.-J. Shi, 353, 1452.
Chem.–Eur. J., 2011, 17, 1728; (d) T. Mesganaw and 44 For a recent review for nickel-catalyzed cross-coupling with
N. K. Garg, Org. Process Res. Dev., 2013, 17, 29. pincer ligands, see: Z.-X. Wang and N. Liu, Eur. J. Inorg.
25 (a) Y.-L. Zhao, G.-J. Wu and F.-S. Han, Chem. Commun., Chem., 2012, 901, wherein, the nickel-pincer type catalyst
2012, 48, 5868, and references therein; (b) Y.-L. Zhao, for the Suzuki–Miyaura cross-coupling was also summarized.
G.-J. Wu, Y. Li, L.-X. Gao and F.-S. Han, Chem.–Eur. J., 45 D. S. McGuinness and K. J. Cavell, Organometallics, 1999,
2012, 18, 9622, and references therein. 18, 1596.
26 (a) V. V. Grushin and H. Alper, Chem. Rev., 1994, 94, 1047; 46 P. L. Chiu, C.-L. Lai, C.-F. Chang, C.-H. Hu and H. M. Lee,
(b) V. V. Grushin and H. Alper, in Activation of Unreactive Organometallics, 2005, 24, 6169.
Bonds and Organic Synthesis, ed. S. Murai, Springer, Berlin, 47 C.-C. Lee, W.-C. Ke, K.-T. Chan, C.-L. Lai, C.-H. Hu and
1999. H. M. Lee, Chem.–Eur. J., 2007, 13, 582.
27 (a) M. B. Mitchell and P. J. Wallbank, Tetrahedron Lett., 48 K. Inamoto, J. Kuroda, K. Hiroya, Y. Noda, M. Watanabe
1991, 32, 2273; (b) N. M. Ali, A. McKillop, M. B. Mitchell, and T. Sakamoto, Organometallics, 2006, 25, 3095.
R. A. Rebelo and P. J. Wallbank, Tetrahedron, 1992, 48, 8117; 49 K. Inamoto, J. Kuroda, T. Sakamoto and K. Hiroya,
(c) S. Achab, M. Guyot and P. Potier, Tetrahedron Lett., 1993, Synthesis, 2007, 2853.
34, 2127; (d) D. Janietz and M. Bauer, Synthesis, 1993, 33. 50 K. Inamoto, J. Kuroda, E. Kwon, K. Hiroya and T. Doi,
28 (a) S. Saito, M. Sakai and N. Miyaura, Tetrahedron Lett., J. Organomet. Chem., 2009, 694, 389.
1996, 37, 2993; (b) S. Saito, S. Oh-tani and N. Miyaura, 51 T. Tu, H. Mao, C. Herbert, M. Xu and K. H. Dötz, Chem.
J. Org. Chem., 1997, 62, 8024. Commun., 2010, 46, 7796.
29 A. F. Indolese, Tetrahedron Lett., 1997, 38, 3513. 52 Y. Zhou, Z. Xi, W. Chen and D. Wang, Organometallics,
30 J.-C. Galland, M. Savignac and J.-P. Genét, Tetrahedron 2008, 27, 5911.
Lett., 1999, 40, 2323. 53 N. E. Leadbeater and S. M. Resouly, Tetrahedron, 1999,
31 K. Inada and N. Miyaura, Tetrahedron, 2000, 56, 8657. 55, 11889.

5296 Chem. Soc. Rev., 2013, 42, 5270--5298 This journal is c The Royal Society of Chemistry 2013
View Article Online

Review Article Chem Soc Rev

54 D. Zim and A. L. Monteiro, Tetrahedron Lett., 2002, 43, 4009. 75 G. A. Molander and F. Beaumard, Org. Lett., 2010, 12, 4022.
55 S. Brun, A. Pla-Quintana and A. Roglans, Organometallics, 76 X.-H. Fan and L.-M. Yang, Eur. J. Org. Chem., 2010, 2457.
2012, 31, 1983. 77 P. Leowanawat, N. Zhang, M. Safi, D. J. Hoffman,
56 B. H. Lipshutz, J. A. Sclafani and P. A. Blomgren, M. C. Fryberger, A. George and V. Percec, J. Org. Chem.,
Tetrahedron, 2000, 56, 2139. 2012, 77, 2885.
Published on 04 March 2013. Downloaded by UNIVERSIDADE FEDERAL SAO CARLOS on 27/11/2013 19:03:38.

57 B. H. Lipshutz, T. Butler and E. Swift, Org. Lett., 2008, 78 C.-H. Xing, J.-R. Lee, Z.-Y. Tang, J. R. Zheng and Q.-S. Hu,
10, 697. Adv. Synth. Catal., 2011, 353, 2051.
58 E. You, P. Li and L. Wang, Synthesis, 2006, 1465. 79 (a) P. Leowanawat, N. Zhang and V. Percec, J. Org. Chem.,
59 (a) D. A. Widdowson and R. Wilhelm, Chem. Commun., 2012, 77, 1018; (b) N. Zhang, D. J. Hoffman, N. Gutsche,
1999, 2211; (b) R. Wilhelm and D. A. Widdowson, J. Chem. J. Gupta and V. Percec, J. Org. Chem., 2012, 77, 5956.
Soc., Perkin Trans. 1, 2000, 3808. 80 J. Kuroda, K. Inamoto, K. Hiroya and T. Doi, Eur. J. Org.
60 (a) Y. M. Kim and S. Yu, J. Am. Chem. Soc., 2003, 125, 1696; Chem., 2009, 2251.
(b) D. A. Widdowson and R. Wilhelm, Chem. Commun., 81 H. Gao, Y. Li, Y.-G. Zhou, F.-S. Han and Y.-J. Lin, Adv.
2003, 578; (c) K. Mikami, T. Miyamoto and M. Hatano, Synth. Catal., 2011, 353, 309.
Chem. Commun., 2004, 2082; (d) S. Bahmanyar, B. C. Borer, 82 T. K. Macklin and V. Snieckus, Org. Lett., 2005, 7, 2519.
Y. M. Kim, D. M. Kurtz and S. Yu, Org. Lett., 2005, 7, 1011; 83 K. W. Quasdorf, M. Riener, K. V. Petrova and N. K. Garg,
(e) M. R. Cargill, G. Sandford, A. J. Tadeusiak, D. S. Yufit, J. Am. Chem. Soc., 2009, 131, 17748.
J. A. K. Howard, P. Kilickiran and G. Nelles, J. Org. Chem., 84 K. W. Quasdorf, A. Antoft-Finch, P. Liu, A. L. Silberstein,
2010, 75, 5860. A. Komaromi, T. Blackburn, S. D. Ramgren, K. N. Houk,
61 (a) J. R. Ruiz, C. Jiménez-Sanchidrián and M. Mora, V. Snieckus and N. K. Garg, J. Am. Chem. Soc., 2011,
J. Fluorine Chem., 2006, 127, 443; (b) B. J. Gallon, 133, 6352.
R. W. Kojima, R. B. Kaner and P. L. Diaconescu, Angew. 85 M. Baghbanzadeh, C. Pilger and C. O. Kappe, J. Org. Chem.,
Chem., Int. Ed., 2007, 46, 7251. 2011, 76, 1507.
62 T. Schaub, M. Backes and U. Radius, J. Am. Chem. Soc., 86 G.-J. Chen and F.-S. Han, Eur. J. Org. Chem., 2012, 3575.
2006, 128, 15964. 87 V. Snieckus, Chem. Rev., 1990, 90, 879.
63 M. Tobisu, T. Xu, T. Shimasaki and N. Chatani, J. Am. 88 For selected examples, see: (a) S. Sengupta, M. Leite,
Chem. Soc., 2011, 133, 19505. D. S. Raslan, C. Quesnelle and V. Snieckus, J. Org. Chem.,
64 A. D. Sun and J. A. Love, Org. Lett., 2011, 13, 2750. 1992, 57, 4066; (b) P. M. Wehn and J. Du Bois, Org. Lett.,
65 N. Miyaura, K. Yamada, H. Suginome and A. Suzuki, J. Am. 2005, 7, 4685; (c) N. Yoshikai, H. Matsuda and
Chem. Soc., 1985, 107, 972. E. Nakamura, J. Am. Chem. Soc., 2009, 131, 9590.
66 K. Zhang, M. Conda-Sheridan, S. R. Cooke and J. Louie, 89 A. Antoft-Finch, T. Blackburn and V. Snieckus, J. Am. Chem.
Organometallics, 2011, 30, 2546. Soc., 2009, 131, 17750.
67 See ref. 11–17, and some very recent examples: 90 B.-T. Guan, Y. Wang, B.-J. Li, D.-G. Yu and Z.-J. Shi, J. Am.
(a) Z.-Y. Wang, G.-Q. Chen and L.-X. Shao, J. Org. Chem., Chem. Soc., 2008, 130, 14468.
2012, 77, 6608; (b) K. H. Chung, C. M. So, S. M. Wong, 91 L. Xu, B.-J. Li, Z.-H. Wu, X.-Y. Lu, B.-T. Guan, B.-Q. Wang,
C. H. Luk, Z. Zhou, C. P. Lau and F. Y. Kwong, Chem. K.-Q. Zhao and Z.-J. Shi, Org. Lett., 2010, 12, 884.
Commun., 2012, 48, 1967; (c) M. Lüthy and R. J. K. Taylor, 92 For a recent review, see: J. D. Sellars and P. G. Steel, Chem.
Tetrahedron Lett., 2012, 53, 3444; (d) J. F. Cı́vicos, Soc. Rev., 2011, 40, 5170.
D. A. Alonso and C. Nájera, Eur. J. Org. Chem., 2012, 3670. 93 For selected examples of palladium-catalyzed cross-coupling of
68 For reviews, see: (a) J. S. Blakeney, R. C. Reid, G. T. Le and enol phosphorus derivatives, see: (a) F. Lepifre, S. Clavier,
D. P. Fairlie, Chem. Rev., 2007, 107, 2960; (b) J.-P. Corbet P. Bouyssou and G. Coudert, Tetrahedron, 2001, 57, 6969;
and G. Mignani, Chem. Rev., 2006, 106, 2651. (b) I. B. Campbell, J. Guo, E. Jones and P. G. Steel, Org. Biomol.
69 V. Percec, J.-Y. Bae and D. H. Hill, J. Org. Chem., 1995, Chem., 2004, 2, 2725; (c) U. S. Larsen, L. Martiny and
60, 1060. M. Begtrup, Tetrahedron Lett., 2005, 46, 4261; (d) B. Cottineau,
70 (a) Y. Kobayashi and R. Mizojiri, Tetrahedron Lett., 1996, I. Gillaizeau, J. Farard, M.-L. Auclair and G. Coudert, Synlett,
37, 8531; (b) Y. Kobayashi, A. D. William and R. Mizojiri, 2007, 1925; (e) H. Fuwa and M. Sasaki, Org. Lett., 2007, 9, 3347;
J. Organomet. Chem., 2002, 653, 91. ( f ) J. Guo, J. D. Harling, P. G. Steel and T. M. Woods, Org.
71 M. Ueda, A. Saitoh, S. Oh-tani and N. Miyaura, Tetrahedron, Biomol. Chem., 2008, 6, 4053; (g) L. Pedzisa, I. W. Vaughn
1998, 54, 13079. and R. Pongdee, Tetrahedron Lett., 2008, 49, 4142;
72 D. Zim, V. R. Lando, J. Dupont and A. L. Monteiro, (h) P. Steel and T. Woods, Synthesis, 2009, 3897.
Org. Lett., 2001, 3, 3049. 94 For nickel-catalyzed cross-coupling of enol phosphorus
73 (a) Z.-Y. Tang and Q.-S. Hu, J. Am. Chem. Soc., 2004, derivatives, see: (a) Y. Nan and Z. Yang, Tetrahedron Lett.,
126, 3058; (b) Z.-Y. Tang and Q.-S. Hu, Adv. Synth. Catal., 1999, 40, 3321; (b) A. L. Hansen, J.-P. Ebran, T. M. Gøgsig
2004, 346, 1635. and T. Skrydstrup, Chem. Commun., 2006, 4137; (c) A. L.
74 P. Leowanawat, N. Zhang, A.-M. Resmerita, B. M. Rosen Hansen, J.-P. Ebran, T. M. Gøgsig and T. Skrydstrup, J. Org.
and V. Percec, J. Org. Chem., 2011, 76, 9946. Chem., 2007, 72, 6464.

This journal is c The Royal Society of Chemistry 2013 Chem. Soc. Rev., 2013, 42, 5270--5298 5297
View Article Online

Chem Soc Rev Review Article

95 Y.-L. Zhao, Y. Li, Y. Li, L.-X. Gao and F.-S. Han, Chem.–Eur. 119 S. B. Blakey and D. W. C. MacMillan, J. Am. Chem. Soc.,
J., 2010, 16, 4991. 2003, 125, 6046.
96 J. Diago-Meseguer, A. L. Palomo-Coll, J. R. Fernández- 120 S. Darses, J. T. Jeffrey, J. P. Genet, J. L. Brayer and
Lizarbe and A. Zugaza-Bilbao, Synthesis, 1980, 547. J. P. Demoute, Tetrahedron Lett., 1996, 37, 3857.
97 G.-J. Chen, J. Huang, L.-X. Gao and F.-S. Han, Chem.–Eur. J., 121 (a) J. A. Miller, Tetrahedron Lett., 2001, 42, 6991;
Published on 04 March 2013. Downloaded by UNIVERSIDADE FEDERAL SAO CARLOS on 27/11/2013 19:03:38.

2011, 17, 4038. (b) J. A. Miller and J. W. Dankwardt, Tetrahedron Lett.,


98 For a review, see: (a) F.-A. Kang, Z. Sui and W. V. Murray, 2003, 44, 1907; (c) J. A. Miller, J. W. Dankwardt and
Eur. J. Org. Chem., 2009, 461. For research articles, see: J. M. Penney, Synthesis, 2003, 1643; (d) J. M. Penney and
(b) F.-A. Kang, Z. Sui and W. V. Murray, J. Am. Chem. Soc., J. A. Miller, Tetrahedron Lett., 2004, 45, 4989.
2008, 130, 11300; (c) F.-A. Kang, J. C. Lanter, C. Cai, Z. Sui 122 D.-G. Yu, M. Yu, B.-T. Guan, B.-J. Li, Y. Zheng, Z.-H. Wu
and W. V. Murray, Chem. Commun., 2010, 46, 1347. and Z.-J. Shi, Org. Lett., 2009, 11, 3374.
99 (a) C. Shi and C. C. Aldrich, Org. Lett., 2010, 12, 2286; 123 T.-Y. Luh, M.-K. Leung and K.-T. Wong, Chem. Rev., 2000,
(b) V. P. Mehta, S. G. Modha and E. Van der Eycken, J. Org. 100, 3187.
Chem., 2010, 75, 976; (c) A. Sharma, D. Vachhani and 124 T. Ishiyama, S. Abe, N. Miyaura and A. Suzuki, Chem. Lett.,
E. Van. der Eycken, Org. Lett., 2012, 14, 1854. 1992, 691.
100 S.-M. Li, J. Huang, G.-J. Chen and F.-S. Han, Chem. Commun., 125 For some representative reviews, see: (a) M. R. Netherton
2011, 47, 12840. and G. C. Fu, Adv. Synth. Catal., 2004, 346, 1525;
101 H. Chen, Z. Huang, X. Hu, G. Tang, P. Xu, Y. Zhao and (b) A. Frisch and M. Beller, Angew. Chem., Int. Ed., 2005,
C.-H. Cheng, J. Org. Chem., 2011, 76, 2338. 44, 674; (c) B. Liégault, J.-L. Renaud and C. Bruneau, Chem.
102 J. Ishizu, T. Yamamoto and A. Yamamoto, Chem. Lett., Soc. Rev., 2008, 37, 290; (d) B. D. Sherry and A. Fürstner,
1976, 1091. Acc. Chem. Res., 2008, 41, 1500; (e) A. Rudolph and
103 K. W. Quasdorf, X. Tian and N. K. Garg, J. Am. Chem. Soc., M. Lautens, Angew. Chem., Int. Ed., 2009, 48, 2656.
2008, 130, 14422. 126 J. Zhou and G. C. Fu, J. Am. Chem. Soc., 2004, 126, 1340.
104 L. J. Gooben, K. Gooben and C. Stanciu, Angew. Chem., 127 F. González-Bobes and G. C. Fu, J. Am. Chem. Soc., 2006,
Int. Ed., 2009, 48, 3569. 128, 5360.
105 C.-L. Sun, Y. Wang, X. Zhou, Z.-H. Wu, B.-J. Li, B.-T. Guan 128 M. A. J. Duncton, M. A. Estiarte, D. Tan, C. Kaub, D. J. R.
and Z.-J. Shi, Chem.–Eur. J., 2010, 16, 5844. O’Mahony, R. J. Johnson, M. Cox, W. T. Edwards, M. Wan,
106 J.-Y. Yu and R. Kuwano, Angew. Chem., Int. Ed., 2009, J. Kincaid and M. G. Kelly, Org. Lett., 2008, 10, 3259.
48, 7217. 129 G. A. Molander, O. A. Argintaru, I. Aron and S. D. Dreher,
107 Z. Li, S.-L. Zhang, Y. Fu, Q.-X. Guo and L. Liu, J. Am. Chem. Org. Lett., 2010, 12, 5783.
Soc., 2009, 131, 8815. 130 T. Zell and U. Radius, Z. Anorg. Allg. Chem., 2011,
108 J. Lindh, J. Sävmarker, P. Nilsson, P. J. R. Sjöberg and 637, 1858.
M. Larhed, Chem.–Eur. J., 2009, 15, 4630. 131 (a) J. Terao, H. Watanabe, A. Ikumi, H. Kuniyasu and
109 (a) E. Wenkert, E. L. Michelotti and C. S. Swindell, J. Am. N. Kambe, J. Am. Chem. Soc., 2002, 124, 4222; (b) B. L. H.
Chem. Soc., 1979, 101, 2246; (b) E. Wenkert, E. L. Taylor, E. C. Swift, J. D. Waetzig and E. R. Jarvo, J. Am.
Michelotti, C. S. Swindell and M. Tingoli, J. Org. Chem., Chem. Soc., 2011, 133, 389.
1984, 49, 4894. 132 K. Huang, G. Li, W.-P. Huang, D.-G. Yu and Z.-J. Shi, Chem.
110 J. W. Dankwardt, Angew. Chem., Int. Ed., 2004, 43, 2428. Commun., 2011, 47, 7224.
111 (a) B.-T. Guan, S. Xiang, T. Wu, Z. Sun, B. Wang, K. Zhao 133 B. Saito and G. C. Fu, J. Am. Chem. Soc., 2007, 129, 9602.
and Z.-J. Shi, Chem. Commun., 2008, 1437; (b) B.-T. Guan, 134 Z. Lu and G. C. Fu, Angew. Chem., Int. Ed., 2010, 49,
S. Xiang, B. Wang, Z. Sun, Y. Wang, K. Zhao and Z.-J. Shi, 6676.
J. Am. Chem. Soc., 2008, 130, 3268. 135 B. Saito and G. C. Fu, J. Am. Chem. Soc., 2008, 130, 6694.
112 F. Kakiuchi, M. Usui, S. Ueno, N. Chatani and S. Murai, 136 (a) P. M. Lundin and G. C. Fu, J. Am. Chem. Soc., 2010,
J. Am. Chem. Soc., 2004, 126, 2706. 132, 11027; (b) S. L. Zultanski and G. C. Fu, J. Am. Chem.
113 M. Tobisu, T. Shimasaki and N. Chatani, Angew. Chem., Soc., 2011, 133, 15362.
Int. Ed., 2008, 47, 4866. 137 N. A. Owston and G. C. Fu, J. Am. Chem. Soc., 2010,
114 T. Shimasaki, Y. Konno, M. Tobisu and N. Chatani, 132, 11908.
Org. Lett., 2009, 11, 4890. 138 Z. Lu, A. Wilsily and G. C. Fu, J. Am. Chem. Soc., 2011,
115 R. Kuwano and R. Shimizu, Chem. Lett., 2011, 40, 913. 133, 8154.
116 D.-G. Yu and Z.-J. Shi, Angew. Chem., Int. Ed., 2011, 139 A. Wilsily, F. Tramutola, N. A. Owston and G. C. Fu, J. Am.
50, 7097. Chem. Soc., 2012, 134, 5794.
117 D.-G. Yu, B.-J. Li, S.-F. Zheng, B.-T. Guan, B.-Q. Wang and 140 (a) B. L. H. Taylor and E. R. Jarvo, J. Org. Chem., 2011,
Z.-J. Shi, Angew. Chem., Int. Ed., 2010, 49, 4566. 76, 7573; (b) Z. Li, Y.-Y. Jiang and Y. Fu, Chem.–Eur. J.,
118 J. Liu and M. J. Robins, Org. Lett., 2004, 6, 3421. 2012, 18, 4345.

5298 Chem. Soc. Rev., 2013, 42, 5270--5298 This journal is c The Royal Society of Chemistry 2013

You might also like