nath2019

Download as pdf or txt
Download as pdf or txt
You are on page 1of 24

Volume 10

Number 1

Catalysis
7 January 2020
Pages 1–292

Science &
Technology
rsc.li/catalysis

ISSN 2044-4761

PERSPECTIVE
Tadashi Ema et al.
Macrocyclic multinuclear metal complexes acting as catalysts
for organic synthesis
Catalysis
Science &
Technology
PERSPECTIVE View Article Online
View Journal | View Issue

Macrocyclic multinuclear metal complexes acting


Cite this: Catal. Sci. Technol., 2020,
as catalysts for organic synthesis
10, 12
Bikash Dev Nath, Kazuto Takaishi and Tadashi Ema *

Metal clusters in nature and artificial systems are known to exhibit excellent catalytic activities. Among
Published on 21 November 2019. Downloaded on 1/2/2020 10:29:18 PM.

various coordination systems, macrocyclic multinuclear metal complexes have unique structural features
and great potential in organic synthesis. The multiple metal centers in a macrocyclic framework act as
specific sites for the binding and activation of substrates, showing high catalytic activity and selectivity as a
result of the cooperative effect of the multiple metal centers and robustness originating from macrocyclic
skeletons. Most of them are precisely designed and synthesized, while occasionally, nice complexes also
Received 19th September 2019, emerge from a fortuitous combination of metal ions and ligands. Here we overview the recent
Accepted 21st November 2019
achievements of (i) multinuclear metal complexes with a covalently-linked macrocyclic ligand (category A),
(ii) multinuclear metal complexes forming a macrocyclic skeleton (category B), and (iii) self-assembled
DOI: 10.1039/c9cy01894h
supramolecular coordination complexes (category C). Various organic reactions catalyzed by these
rsc.li/catalysis coordination systems are summarized here.

1. Introduction functions, and some of them act as catalysts for organic


synthesis and enzyme model reactions.1–5 Metal clusters can
1.1 Classification of macrocyclic multinuclear metal be created in the cavity of a covalently-bonded macrocyclic
complexes polydentate ligand,6 or they can be synthesized by the self-
Multinuclear metal complexes have attracted much attention assembly of metal ions and non-macrocyclic polydentate
of chemists because of their fascinating structures and ligands. Simultaneous coordinations of multiple metal ions
can lead to the effective recognition and activation of
substrates.7,8
Multinuclear metal complexes can be classified into two
Division of Applied Chemistry, Graduate School of Natural Science and Technology,
Okayama University, Tsushima, Okayama 700-8530, Japan. types: macrocyclic and non-macrocyclic ones. The former
E-mail: ema@cc.okayama-u.ac.jp possesses a macrocyclic framework useful for the spatial

Bikash Dev Nath was born in Kazuto Takaishi was born in


Barisal, Bangladesh in 1984. He Ehime, Japan in 1980. He
received his bachelor's degree in received his bachelor's degree in
2010 and master's degree in 2003 and master's degree in
2011 from Jahangirnagar 2005 from Okayama University,
University. He worked as a and he obtained his doctor's
synthesis engineer at Pilarquim degree in 2008 from Kyoto
(Shanghai) Co., Ltd. in China University under the supervision
from 2011 to 2016. He of Professor Takeo Kawabata. He
completed his doctor's degree in then moved to RIKEN as a
2019 at Okayama University postdoctoral fellow. He was
under the supervision of appointed as an Assistant
Bikash Dev Nath Professor Tadashi Ema. His Kazuto Takaishi Professor at Seikei University in
subject during doctoral study 2012. He moved to Okayama
was “Novel Macrocyclic Multinuclear Ni(II) and Zn(II) Complexes University to join the Tadashi Ema research group as a Senior
for Catalytic CO2 Conversions”. Assistant Professor in 2015. His main research topic is the
synthesis of chiral functional compounds.

12 | Catal. Sci. Technol., 2020, 10, 12–34 This journal is © The Royal Society of Chemistry 2020
View Article Online

Catalysis Science & Technology Perspective

disposition of not only metal ions but also catalytic


functional groups. The precise preorganization of metal ions
and catalytic functional groups is a great advantage of the
macrocyclic ones over the non-macrocyclic ones. Macrocyclic
multinuclear metal complexes can be subdivided into four
categories (Fig. 1): (A) multinuclear metal complexes with a
covalently-linked macrocyclic ligand, (B) multinuclear metal
complexes forming a macrocyclic skeleton, (C) self-assembled
supramolecular coordination complexes, and (D) metal–
organic frameworks (MOFs).
Multinuclear metal complexes with a covalently-linked
macrocyclic ligand (category A) are created by combining an
elaborately designed macrocyclic ligand with metal ions. The
coordination of multiple metal ions to different chelating sites
Published on 21 November 2019. Downloaded on 1/2/2020 10:29:18 PM.

in the macrocyclic cavity gives homo- or heteronuclear metal


complexes. Multinuclear metal complexes forming a
macrocyclic skeleton (category B) are created by the self-
assembly of metal ions and polydentate ligands. The Fig. 1 Macrocyclic multinuclear metal complexes classified into four
macrocyclic skeletons are constructed through multiple metal– categories (A–D) and their functions and features.
ligand coordination bonds. Specific structures characterized by
multiple metal centers and macrocyclic frameworks lead to
excellent catalytic activities. Heterometallic complexes that are 1.2 Advantages of macrocyclic multinuclear metal complexes
similar to those in nature are also created.9–12 Self-assembled
supramolecular coordination complexes (category C) are Macrocyclic multinuclear metal complexes have both
molecular cages with a vacant inner space, and some of them promising advantages and challenging aspects as follows: (i)
are excellent supramolecular homogeneous catalysts.13–19 On synthetic challenges for the construction of macrocyclic
the other hand, metal–organic frameworks (MOFs) (category D) structures, (ii) specific sites for substrate binding and
provide another effective approach to create a sophisticated catalysis, (iii) synergetic effects based on the metal–metal or
coordination space. MOFs are three-dimensional lattices metal–ligand cooperation, (iv) variations of (mixed) metal
composed of metal ions and bridging organic ligands, having elements, (v) good reactivity, selectivity, and robustness
high porosity, a large surface area, dense active sites, high resulting from macrocyclic structures, and (vi) unlimited
adsorption capacity, and good thermal stability.20–39 MOFs can combinations of metals and ligands.
work as excellent heterogeneous catalysts. There may be room Macrocyclic multinuclear metal complexes are sometimes
for improvement of the above classification; for example, difficult to prepare because of the lack of versatile synthetic
molecular knots with multiple metal ions,40–44 which might methods. The macrocyclization step sometimes requires
belong to category A, are not considered here because the use rigorous tuning of reaction conditions and suffers from low
of molecular knots as catalysts has not been reported. yields. The careful design of organic ligands is key to the
successful construction of specific molecular structures with
good synthetic accessibility.
The binding of a substrate molecule to a metal center is a
Tadashi Ema was born in Gifu, crucial step in catalysis. Therefore, the catalytic activity can
Japan in 1966. He received his be significantly improved if the complex offers a specific
bachelor's degree from Kyoto coordination site to bind and activate the substrate molecule.
University in 1989, and he The cooperative effect of multiple metal centers is important
obtained his master's degree in for high catalytic performance.45–51 Cooperative catalysis may
1991 and doctor's degree in 1994 originate from two or more metal ions in close proximity.
from Kyoto University under the The presence of different metals is beneficial for excellent
supervision of Professor catalytic activities. Macrocyclic ligands with multiple
Hisanobu Ogoshi. He was chelating sites should be well designed to systematically
appointed as an Assistant introduce different metal elements.
Professor at Okayama University Another significant feature of macrocyclic multinuclear
in 1994 and was promoted to an metal complexes is robustness, which may come from the
Tadashi Ema Associate Professor in 2002 and sophisticated and tight assembly of metal ions and ligands
a Full Professor in 2012. His with multiple coordinations. Although macrocyclic metal
research interest is the design and synthesis of unique molecular complexes can be created by designing macrocyclic ligands,
catalysts. the self-assembly of metal ions and ligands may be assisted

This journal is © The Royal Society of Chemistry 2020 Catal. Sci. Technol., 2020, 10, 12–34 | 13
View Article Online

Perspective Catalysis Science & Technology

by more complicated and subtle metal–ligand and ligand– the copolymerization of cyclohexene oxide (CHO) and CO2
ligand interactions. Combinations of metal ions and ligands (Scheme 1). A robust catalyst 1 catalyzed the
sometimes give rise to unusual robustness. copolymerization of CHO and CO2 (1 atm) with a catalyst
Macrocyclic multinuclear metal complexes show a variety of loading of 0.1 mol% at 80–100 °C to form polyIJcyclohexene
physical and chemical properties based on structural diversity.52,53 carbonate), PCHC.96 The turnover number (TON) and
In the area of host–guest chemistry, they can act as host molecules turnover frequency (TOF) values reached 430–530 and 18–25
for the multipoint recognition of guest molecules.7,8 They play h−1, respectively. Increasing the CO2 pressure to 10 atm at
significant roles in catalysis,14–19,27–39,54–56 and they can mimic the 100 °C increased both the TON (838) and TOF (38 h−1). When
active site of metalloenzymes.57 They also find applications in the catalyst loading was reduced to 0.01 mol%, the TON and
molecular sensing,58–60 separation,61–68 and ionic/proton TOF values increased up to 3350 and 140 h−1, respectively. A
conductivity.69–71 They can be used in biochemical applications72 mechanism was proposed on the basis of kinetic studies
and drug delivery.73,74 Taking advantage of high porosity and vacant (Scheme 2).97 DFT calculations suggested that the rate-
inner space, they can store,75–78 separate, and purify a gas.63–67 determining step is the epoxide-ring opening of CHO by the
nucleophilic attack of the carbonate anion in the propagation
Published on 21 November 2019. Downloaded on 1/2/2020 10:29:18 PM.

1.3 Scope of this perspective step.98


Dinuclear Co complexes 2 and 3 exhibited comparable
Despite the wide range of catalytic applications of
catalytic activity at 1 atm CO2 pressure.99,100 Complex 2 with
macrocyclic multinuclear metal complexes in organic
a mixed valence CoIJII)/CoIJIII) core showed 20 times higher
synthesis, only a limited number of reviews have been
TOF (500 h−1) at 100 °C than complex 1. At a higher CO2
published so far, all of which are related to self-assembled
pressure (10 atm), both 2 and 3 showed much higher
supramolecular coordination complexes or MOFs (categories
catalytic activities, and complex 2 displayed a TOF of 3700
C and D in Fig. 1).14–19,27–39 In contrast, there are few reviews
h−1 at 100 °C. Mg complex 4 with a catalyst loading of 0.01
on categories A and/or B in Fig. 1. Here we provide, for the
mol% showed a TOF of 730 h−1 at 100 °C at 12 atm CO2
first time, an overview of homogeneous catalysis with
pressure.101 The Mg–carbonate bond may be more
macrocyclic multinuclear metal complexes in categories A–C
nucleophilic because of the decreased Lewis acidity of MgIJII)
in Fig. 1, illuminating the characteristics and advantages of
as compared with ZnIJII), which can facilitate the ring opening
each category. For example, it can be seen that the catalysts
of the epoxide. FeIJIII) complex 6 exhibited 8 times higher
in categories B and C show quite different activities,
activity at 10 atm CO2 pressure than at 1 atm CO2
selectivities, and mechanisms although the latter is
pressure.102
structurally defined as a three-dimensional version of the
Williams' group also reported the first heterodinuclear
former. In addition, it is also shown that all of them catalyze
Mg/Zn complex 7,103 and complex 8 was the first isolated
a wide range of useful organic reactions. In most cases, it is
impossible to clearly compare catalytic performances between
macrocyclic/multinuclear and non-macrocyclic/mononuclear
systems because of the lack of data although structural
features and proposed mechanisms strongly suggest the
advantages and benefits of macrocyclic multinuclear metal
complexes. Each achievement is concisely summarized one
by one using reaction schemes with representative catalyst
structures to help the readers easily understand the recent
progress and great potential of this research field.

2. Multinuclear metal complexes with


a covalently-linked macrocyclic
ligand (category A)
2.1 CO2 fixation
Carbon dioxide (CO2) is a renewable raw material for the
production of value-added chemicals, and a number of
important catalytic reactions have been developed.79–86 A
typical example can be seen in the copolymerization of
epoxides and CO2 to form polycarbonates,87–95 and several
multinuclear metal complexes were developed for this
purpose.
Williams and co-workers developed several macrocyclic
multimetallic complexes showing high catalytic activity for Scheme 1 Copolymerization of CO2 and CHO.

14 | Catal. Sci. Technol., 2020, 10, 12–34 This journal is © The Royal Society of Chemistry 2020
View Article Online

Catalysis Science & Technology Perspective

Fig. 3 Structure of CeZn3 complex 14.

modified structures were also synthesized, and their catalytic


Published on 21 November 2019. Downloaded on 1/2/2020 10:29:18 PM.

Scheme 2 Mechanism of copolymerization. activities for the copolymerization of CHO and CO2 were
studied by online ATR-IR measurements.111 Surprisingly, the
in situ IR spectroscopic measurements revealed that the TOF
heterometallic complex synthesized from a macrocyclic values reached up to 23 300 h−1 for complex 11 at 30 bar CO2
ligand.104 Heterodinuclear Ti/Zn complex 9 was also pressure and 100 °C. Complexes 12 and 13 containing Cl and
synthesized.105 Importantly, mixed metal catalyst 7 was CF3 groups, respectively, were expected to have metal centers
remarkably more effective than either homodinuclear Zn with enhanced Lewis acidity. Indeed, complex 13 exhibited
complex 1 or Mg complex 4. The catalytic activity of exceptionally high catalytic activity for the copolymerization
heterodinuclear complex 8 was about 5 times higher than of CHO and CO2 with a TOF of up to 155 000 h−1.
that of a 1 : 1 mixture of homodinuclear Zn and Mg Okuda, Mashima, and co-workers synthesized several
complexes. 8 was more than twice as active as Mg complex 5, heterometallic tetranuclear complexes, among which CeZn3
while a Zn analog of 5 showed no activity at all. Clearly, the complex 14 (Fig. 3) was the most effective for the
enhanced catalytic activity of heterodinuclear Mg/Zn copolymerization of CHO and CO2 at 100 °C, showing a TOF
complexes 7 and 8 is due to the synergistic effect of the two of 330 h−1 even at 0.6 MPa CO2 pressure with >99%
different metal ions. In addition, 8 formed the polycarbonate carbonate-linkage selectivity.112 The molecular weight of the
linkages highly selectively (>99%), and 8 with a catalyst copolymer was controlled by adding tetrabutylammonium
loading of 0.01 mol% showed a TOF of 624 h−1. More acetate as a chain-transfer agent. A rapid exchange of acetate
recently, when the axial ligand (Br) in 8 was replaced with anions is a key factor in the control of telomerization.
p-trifluoromethylbenzoate to prepare 10, a maximal TOF of Jing and co-workers examined chiral dinuclear Co
8830 h−1 was obtained under optimized reaction conditions complexes 15 and 16 in the kinetic resolution of propylene
(20 bar CO2, 120 °C).106 A series of dimetallic complexes were oxide (PO) with CO2 using phenyltrimethylammonium
characterized by X-ray crystal structural analysis.107 tribromide (PTAT) as a nucleophilic co-catalyst (Scheme 3).113
Based on DFT calculations on the copolymerization of (R,R,R,R,R,R)-15 showed higher enantioselectivity (s value 5)
CHO and CO2 with 1, showing the importance of the than (S,R,R,S,R,R)-16 (s value 3). The matched pair of two
shuttling mechanism (Scheme 2),98,108 Rieger and co-workers chiral moieties, BINOL and 1,2-cyclohexanediamine, is
synthesized 11 containing two β-diketiminate Zn complexes important for the enantioselectivity.
(Fig. 2).109,110 Complex 11 showed a TOF of 9130 h−1 at 100
°C and 40 bar CO2 pressure. The catalytic activity decreased
at higher temperature (120 °C) probably due to
decomposition of the catalyst. Several other catalysts with

Fig. 2 Structures of 11–13. Scheme 3 Kinetic resolution of epoxides.

This journal is © The Royal Society of Chemistry 2020 Catal. Sci. Technol., 2020, 10, 12–34 | 15
View Article Online

Perspective Catalysis Science & Technology

it showed no activity for 1,4-cyclohexadiene oxide. On the other


hand, complex 4 was almost equally effective for both
epoxides.118,119 A synergistic effect between the two different
metal centers in heterodinuclear metal complex 8 was observed;
8 was 20 times more active than 5.104
The mechanism of epoxide/anhydride copolymerization is
proposed as follows: initially, a metal alkoxide is formed
through epoxide-ring opening, which then rapidly reacts with
the anhydride to form a metal carboxylate. The carboxylate
reacts with the epoxide, producing a metal alkoxide.93,118,120

2.3 Olefin polymerization


Since the discovery of Ziegler–Natta catalysts, numerous
Published on 21 November 2019. Downloaded on 1/2/2020 10:29:18 PM.

multinuclear metal complexes capable of catalyzing olefin


polymerization have been reported so far.121 Several macrocyclic
multinuclear metal complexes also worked as excellent
catalysts.122–126 Li and co-workers reported that in the presence
Scheme 4 Formation of oxalate from CO2. of a modified methylaluminoxane (MMAO) cocatalyst, trinuclear
FeIJII) complex 19 showed excellent catalytic activity for the
formation of polyethylene (PE) from ethylene as compared with
Murray and co-workers reported macrocyclic trinuclear Cu a non-macrocyclic mononuclear counterpart. This fact clearly
complexes, Cu3IJμ3-S)L 17 and Cu3IJμ3-Se)L 18, which showed indicates the importance of the multiple metal centers in the
catalytic activity for the selective reduction of CO2 to oxalate macrocyclic skeleton (Scheme 6).122 The significance of the
in the presence of a reductant (Scheme 4).114 To characterize macrocyclic ligand was also demonstrated by a kinetic study. It
the reactivity of 17 and 18, single-turnover reactions were was obvious that the macrocyclic ligand stabilized the
carefully conducted. Anionic complex [Cu3EL]− was generated catalytically active Fe center during polymerization and
by the one-electron reduction of Cu3EL with various suppressed the chain transfer reaction to produce a high
reductants such as [K(18-crown-6)]ijC10H8] in THF, and this molecular-weight polymer. A maximum catalytic activity of 4300
anionic species successively reduced CO2 to (CO2K)2. Single kg molFe−1 h−1 bar−1 was achieved by 19 at 0 °C and 1 atm
crystals of [K(THF)3]ijCu3EL] complexes were obtained by ethylene pressure. Both the catalytic activity of 19 and molecular
reduction of 17 or 18 with KC10H8 in THF, and X-ray analysis weight of PE were temperature-dependent, decreasing with an
revealed that [K(THF)3]+ interacted with one β-diketiminate increase in reaction temperature.
moiety of the [Cu3EL]− ion in an η5 fashion (Scheme 4). This Ma and co-workers also reported the steric protection of
crystal structure clearly illustrates how the solvent and cation the metal centers by the macrocyclic backbone in trinuclear
stabilize the key intermediate. The pseudo-first-order rate Ni catalyst 20 during ethylene polymerization (Fig. 4).123
constant for the reaction of [K(18-crown-6)]ij18] with CO2 was β-Hydrogen elimination and the chain-transfer process were
much greater than the corresponding value for [K(18-crown-
6)]ij17]. These complexes and previously reported ones are
promising catalysts for CO2 reduction to oxalate.115,116

2.2 Ring-opening copolymerization


Epoxide/anhydride copolymerization is an important area of
research.93,117 Williams and co-workers prepared several
copolymers by the ring-opening copolymerization (ROCOP) of
phthalic anhydride (PA) and epoxides (Scheme 5).104,118,119
Although complex 1 was effective for PA/CHO copolymerization,

Scheme 5 Copolymerization of PA and CHO. Scheme 6 Ethylene polymerization.

16 | Catal. Sci. Technol., 2020, 10, 12–34 This journal is © The Royal Society of Chemistry 2020
View Article Online

Catalysis Science & Technology Perspective


Published on 21 November 2019. Downloaded on 1/2/2020 10:29:18 PM.

Fig. 5 Structures of 22–29.

Dinuclear Fe and Co complexes 24–29 were synthesized to


examine the polymerization activity at various temperatures
and pressures in the presence of a MMAO co-catalyst.126
Fig. 4 Structures of 20 and 21.
24–29, which were thermodynamically more stable than the
corresponding monometallic analogs, produced polymers
with high molecular weights at room temperature and 1 atm
suppressed by the steric protection of the metal centers to ethylene pressure. Complex 24 showed the highest activity
produce PE with a high molecular weight and a low branch (975 g mmolFe−1 h−1 atm−1) at 100 °C and 5 atm ethylene
density. Complex 20 showed a 4-fold greater polymerization pressure, while complex 27 showed the lowest activity (54.8 g
activity than dinuclear analog 21. The same trend was mmolFe−1 h−1 atm−1) under the same reaction conditions.
observed for propylene polymerization. The most effective Complex 25 produced a higher molecular-weight polymer (Mn
catalyst 20 provided the highest TOF at 15 °C although a 62 700) at room temperature than the corresponding
polymer with the highest molecular weight was obtained at monometallic complex (Mn 8930). The cooperative interaction
−10 °C. Similar to ethylene polymerization, the macrocyclic between the growing polymer and the second metal center
steric environment around the nickel centers in 20 was might retard the undesirable chain transfer and/or catalyst
suggested to control the overall polymerization, which deactivation during the polymerization.
preferred sterically less hindered 1,2-insertion of propylene.
The cooperativity of two metal centers for olefin
polymerization was demonstrated by Takeuchi, Osakada, and 2.4 Aldol reaction
co-workers. Double-decker-type Ni catalyst 22 (Fig. 5) could Tanaka and co-workers reported asymmetric aldol reactions
copolymerize ethylene and bifunctional co-monomers such catalyzed by metal complexes with chiral macrocyclic ligands
as dienes and unsaturated esters in the presence of a co- L1, L2, and L3 under solvent-free conditions in a ball mill
catalyst, NiIJcod)2, which served as a phosphine scavenger.124 (Scheme 7).127 The most effective catalyst for the reaction of
The synergistic effect of the two Ni centers accelerated the cyclohexanone with 4-nitrobenzaldehyde was obtained from a
polymerization. The growing polymer chain at one Ni center mixture of L1 and CoBr2 in a 1 : 2 ratio. The anti-aldol
is stabilized by the coordination of the functional group to product was obtained as a major isomer in 82% yield and
the other Ni center. Therefore, macrocyclic dinuclear Ni 93% ee. Other catalytic systems with CuBr2, ZnIJOAc)2, and
complex 22 was more effective than a monometallic NiCl2 in combination with L1 gave 90, 79, and 75% ee,
counterpart. respectively.
The Takeuchi group studied the cooperative effect of the
two Ni centers of 22 and 23 on ethylene polymerization in
the presence of NiIJcod)2 (Fig. 5).125 The catalytic activity at 2.5 Henry reaction
room temperature decreased in the following order: 22 > 23 The Tanaka group employed similar catalytic systems in the
> a monometallic analog. The molecular weight of the enantioselective Cu-catalyzed Henry reaction of aldehydes with
polymers depended on the cooperativity of the two Ni nitromethane under solvent-free conditions (Scheme 8).128 The
centers, and higher-molecular-weight polymers were formed reactions with aromatic aldehydes provided β-hydroxy
from di-activated catalysts, whereas lower-molecular-weight nitroalkanes with up to 87% ee, and those with aliphatic
polymers were obtained by the mono-activated catalysts. aldehydes offered the products with up to 93% ee.

This journal is © The Royal Society of Chemistry 2020 Catal. Sci. Technol., 2020, 10, 12–34 | 17
View Article Online

Perspective Catalysis Science & Technology


Published on 21 November 2019. Downloaded on 1/2/2020 10:29:18 PM.

Scheme 10 Cleavage of phosphate esters.

Scheme 7 Asymmetric aldol reaction.

cleavage of the P–O bond is an important area of


research.131–134 Here, the hydrolysis of phosphate esters is
briefly described in view of the potential significance in
organic synthesis. Zhao and co-workers developed several
symmetrical and unsymmetrical macrocyclic bimetallic
Scheme 8 Asymmetric Henry reaction. systems to investigate the hydrolysis of phosphate diesters
and monoesters.135,136 Among six possible forms of dinuclear
Zn hydrates, trans-[Zn2LIJOH)IJH2O)]+ form 31 was considered
Savoia and co-workers developed an effective catalytic
to be the most active catalyst for the hydrolysis of
system for the asymmetric Cu-catalyzed Henry reaction of
phosphodiester BNPP (bisIJ4-nitrophenyl)phosphate)
aldehydes with nitromethane (Scheme 9).129,130 The structure
(Scheme 10).135 DFT calculations revealed that a ping-pong
of complex 30, prepared from CuIJOAc)2·H2O and a C2-
mechanism is most likely to occur, which involves a stepwise
symmetric macrocyclic ligand, was confirmed by X-ray
SN2-type nucleophilic addition–substitution reaction with
analysis. The reactions at room temperature gave a high
inversion of the configuration at the P atom (Fig. 6). The two
enantioselectivity of up to 95% ee. The excellent performance
ZnIJII) ions cooperatively facilitate the nucleophilic attack of
of 30 as compared with a non-macrocyclic counterpart clearly
the metal-bound OH−, stabilizing the anionic transition state.
demonstrates the essential role of the macrocyclic framework
Mechanistic investigations on the hydrolytic cleavage of
in the highly enantioselective Henry reaction.
phosphate monoester NPP (4-nitrophenyl phosphate) were
carried out by using unsymmetrical dinuclear catalysts 32 and
2.6 Hydrolysis of phosphate esters
33.136 Zn catalyst 32 followed an energetically favorable
The biological roles of phosphate esters in living things are pathway involving a concerted SN2-type addition–substitution
well known, and mechanistic studies on the hydrolytic reaction, while Mg catalyst 33 followed a stepwise SN2-type
addition–substitution reaction.
Kandaswamy and co-workers synthesized macrocyclic
dinuclear Ni complexes 34–39 and Zn complexes 40–44

Scheme 9 Asymmetric Henry reaction. Fig. 6 Cooperative P–O bond cleavage.

18 | Catal. Sci. Technol., 2020, 10, 12–34 This journal is © The Royal Society of Chemistry 2020
View Article Online

Catalysis Science & Technology Perspective

Scheme 11 Synthesis of PCHC and CHC.


Fig. 7 Macrocyclic dinuclear complexes.
Published on 21 November 2019. Downloaded on 1/2/2020 10:29:18 PM.

(Fig. 7) to study the hydrolytic cleavage of NPP and Kleij and co-workers reported trinuclear complexes 48–50
DNA.137,138 The pH-dependent formation of the nucleophilic for the reaction of epoxides with CO2 (Scheme 13).141
OH− group by deprotonation of one of the metal-coordinated Complex 48 gave 97% conversion even after 5 repeated uses
water molecules in the complexes was suggested to accelerate of the recycled catalyst, and both the yield (>93%) and
the hydrolysis of phosphate esters via nucleophilic attack on selectivity (>99%) remained high. The metal-bound I− is
the P atom. Interestingly, symmetrical Ni complex 34 showed dissociated upon heating to become a nucleophile, which
higher catalytic performance for the cleavage of NPP than increases the Lewis acidity of the metal centers to facilitate
unsymmetrical complexes 35–41. In the hydrolytic cleavage of the coordination and activation of the epoxide.
DNA, macrocyclic Zn complexes exhibited higher catalytic Takaishi, Ema, and co-workers synthesized macrocyclic
performance than Ni analogs, which was attributed to the multinuclear complexes Ni4L3 51 and Zn5L3 52 by the self-
more Lewis acidic ZnIJII) ion capable of tightly binding the assembly of a chiral bipyridyl–binaphthyl ligand and metal
phosphate group of DNA. The phosphate group of DNA is salts (Scheme 14).142 X-ray analysis revealed unique
activated in a cooperative manner by the two ZnIJII) ions, complex@complex structures as represented by Ni@Ni3L3
where the ZnIJII)-bound OH− group attacks the P atom to and Zn2@Zn3L3. They exhibited dual catalytic activities for
break one of the P–O bonds of the DNA. Among the CO2 fixations under solvent-free conditions. The reaction of
unsymmetrical Zn complexes, the best DNA cleavage activity epoxides with CO2 at 120 °C gave cyclic carbonates, and 51
was observed for 44.

3. Multinuclear metal complexes


forming a macrocyclic skeleton
(category B)
3.1 CO2 fixation
Ko and co-workers reported macrocyclic dinuclear Co
complex 45 and Ni complex 46 for the reaction of epoxides
with CO2 (Scheme 11).139 Ni complex 46 exhibited high
polymerization activity (TOF 265 h−1) at 150 °C for the
formation of PCHC (Mn 20 800, carbonate linkage >99%)
from CHO and CO2 in the absence of a co-catalyst. In the
presence of a co-catalyst, Bu4NBr (TBAB), Co complex 45
produced cyclohexene carbonate (CHC) (cis-selectivity >99%)
from CHO and CO2 at 120 °C with a TOF of 174 h−1.
Liu and co-workers synthesized several anion-induced 3d–
4f coordination clusters Zn2Ln2L4 (Ln = Nd, Eu, Tb, Er, Yb)
and Zn4Ln2L4 (Ln = Nd, Tb).140 Zn2Tb2 complex 47 showed a
TON of up to 9900 and a TOF of up to 700 h−1 (>99%
selectivity) for the conversions of various epoxides and CO2
into cyclic carbonates under solvent-free conditions
(Scheme 12). Scheme 12 Synthesis of cyclic carbonates.

This journal is © The Royal Society of Chemistry 2020 Catal. Sci. Technol., 2020, 10, 12–34 | 19
View Article Online

Perspective Catalysis Science & Technology

Scheme 13 Synthesis of cyclic carbonates.


Published on 21 November 2019. Downloaded on 1/2/2020 10:29:18 PM.

and 52 could be recycled at least 5 times without loss of


catalytic activity. Zn5L3 complex 52 also acted as an excellent
catalyst for the temperature-controlled N-functionalization
reactions of amines with CO2 and PhSiH3; N-formylation
products and N-methylation products were selectively
obtained at 30 °C and 100 °C, respectively.
Maverick and co-workers synthesized well-designed
macrocyclic dinuclear Cu complexes 53 and 54 capable of
Scheme 15 Reduction of CO2 to oxalate.

reducing CO2 to oxalate (Scheme 15).143 Reduction of the


CuIJII) ions with a reductant (ascorbate) gave a CuIJI) species,
which reduced CO2 to form the C–C bond. Acid treatment
regenerated the starting CuIJII) species with the release of
oxalic acid.
Homometallic or heterometallic metal–metal bonding
produces unique catalytic activity and selectivity.45,144–146 Pairing
of Ni0 with group 13 metal ions forms heterometallic complexes
characterized by a strong Ni-to-MIJIII) dative bond. Because of the
withdrawal of the electron density by the Lewis acidic MIJIII) ion,
Ni becomes more electron-deficient, enabling the binding of
small molecules such as H2. Lu and co-workers reported the
hydrogenation of CO2 with NiGaL complex 55 to give formate at
ambient temperature in the presence of a strong base
(Scheme 16).147 The TON and TOF values reached 3150 and
6900 h−1, respectively. The essential role of the GaIJIII) ion was

Scheme 14 Multitasking catalysts. Scheme 16 Hydrogenation of CO2 and olefin.

20 | Catal. Sci. Technol., 2020, 10, 12–34 This journal is © The Royal Society of Chemistry 2020
View Article Online

Catalysis Science & Technology Perspective

clearly demonstrated by the fact that a Ni complex lacking the


GaIJIII) ion exhibited no catalytic activity. They performed an
intensive mechanistic study to identify the most favorable
reaction pathway: (i) binding of H2 to the Ni center to form
adduct (η2-H2)NiGaL, (ii) deprotonation of the adduct with a
base to generate a [HNiGaL]− species, (iii) production of formate
adduct [(η1-HCO2)NiGaL]− through CO2 insertion, and (iv) the
release of formate and regeneration of catalyst 55 (Scheme 17).148
On the other hand, Takaya and Iwasawa synthesized a
macrocyclic Pd–Al complex from a different acyclic ligand,
which acted as a catalyst for the hydrosilylation of CO2 (1 atm)
to give silyl formate with a TOF of 19 300 h−1 at 25 °C.149
Scheme 18 Hydrogenation of levulinic acid.
3.2 Olefin hydrogenation
Published on 21 November 2019. Downloaded on 1/2/2020 10:29:18 PM.

Lu and co-workers applied bimetallic complexes 55–57 to


3.4 Ethylene polymerization
olefin hydrogenation (Scheme 16).150 Although Ni/In complex
57 showed modest activity for the hydrogenation of olefins, Wang, Chen, and co-workers prepared several dinuclear Ni
Ni/Al complex 56 was not effective at all. Ni/Ga complex 55 complexes 60–63 for ethylene polymerization (Scheme 19).152
displayed the highest catalytic performance, delivering >99% Iminopyridyl complexes 60 and 61 exhibited higher catalytic
yield under mild reaction conditions. The formation of a (H2) activity than aminopyridyl complexes 62 and 63, and the
Ni–M adduct and then a HNi(μ-H)M species was likely to be highest activity of up to 2200 kg molNi−1 h−1 was observed for
key to the smooth olefin hydrogenation. 60 at 20 °C. Dinuclear complex 60 showed higher thermal
stability and catalytic activity than the corresponding
3.3 Hydrogenation of ketones mononuclear analog, the former of which produced PE with
a lower branching density than the latter. The cooperative
Darkwa, Makhubela, and co-worker developed hexanuclear effect of the two metal ions is considered to slow down
Ru4Zn2 complexes 58 and 59 with phosphino–carboxylate β-hydride elimination and the chain-walking process. 60 and
bridges as precatalysts for the hydrogenation of levulinic acid to 62 with the isopropyl groups were catalytically more active
γ-valerolactone using formic acid as a hydrogen source than 61 and 63 with the methyl groups, respectively, and 60
(Scheme 18).151 These multimetallic complexes showed superior and 62 produced PE with a higher molecular weight. It is
catalytic activities over the corresponding monometallic suggested that the isopropyl groups on the aniline rings
counterparts, one of which gave a maximum TOF of 540 h−1. efficiently suppressed chain-transfer reactions. The steric
Catalyst 58 was slightly more effective than 59. The high bulkiness of the substituents on the aniline rings also
catalytic activity was ascribed to the presence of the ZnIJII) ions, affected the branching density of the polymers.
which are considered to decrease the electron density of the Ru
centers to promote the H–H bond activation.
3.5 Friedel–Crafts alkylation
The Friedel–Crafts alkylation of aromatic compounds is
extensively used in organic synthesis. Kostakis and co-

Scheme 17 Catalytic cycle. Scheme 19 Ethylene polymerization.

This journal is © The Royal Society of Chemistry 2020 Catal. Sci. Technol., 2020, 10, 12–34 | 21
View Article Online

Perspective Catalysis Science & Technology

workers synthesized a series of isoskeletal tetranuclear


heterometallic clusters 64–70 (Scheme 20).153 X-ray analysis
revealed the close proximity of the four metal ions in all
cases. Each of them served as an excellent catalyst for the
Friedel–Crafts alkylation of indoles with aldehydes, and
Zn2Y2 complex 64 showed the highest catalytic activity. The
ZnIJII) and YIJIII) ions are considered to activate indole and
aldehyde molecules, respectively, in a cooperative manner.
Tetranuclear clusters also catalyzed the Friedel–Crafts
reaction of indoles with trans-β-nitrostyrenes at room
temperature.154 The maximum yield of 99% was obtained by
using Zn2Dy2 complex 69.

3.6 Domino reaction Scheme 21 Domino reaction.


Published on 21 November 2019. Downloaded on 1/2/2020 10:29:18 PM.

Kostakis and co-workers prepared three isoskeletal tetranuclear


complexes 71–73, among which Ni2Dy2 complex 73 (1 mol%)
was the most effective for the domino condensation/ring- exhibited excellent catalytic activity for the Suzuki–Miyaura
opening/electrocyclization of amines and 2-furaldehyde at room reaction of both electron-rich and electron-poor aryl bromides
temperature to produce trans-4,5-diaminocyclopent-2-enones with phenylboronic acids (Scheme 22).157 Although all the
(Scheme 21).155 A series of Ni2Ln2 complexes 74–78 analogous complexes exhibited high catalytic efficiency, rigid macrocyclic
to 73 were further synthesized, and it was found that a lower dithiolates 79 and 80 were more effective than the octanuclear
amount of Ni2Y2 complex 74 (0.5 mol%) catalyzed the reaction complex 81. With a trace amount of catalyst 79 (0.000004
to afford the products in a quantitative yield.156 The catalytic mol%), the product was obtained in 85% yield (TON 21 million).
activity was mainly induced by the YIJIII) or 4f ions rather than
the NiIJII) ions. 3.8 Heck reaction

3.7 Suzuki–Miyaura reaction Macrocyclic dithiolate complexes 79–81 also showed excellent
catalytic activity in the Heck reaction of aryl bromides with
Dey and co-workers prepared self-assembled tetranuclear and
alkenes; the TON and TOF reached 410 000 and 25 625 h−1,
octanuclear Pd-based macrocyclic dithiolates 79–81, which
respectively (Scheme 23).158 As observed for the Suzuki–
Miyaura reaction, the catalytic activity of the three palladium
complexes decreased in the following order: 79 > 80 > 81.

3.9 Michael addition


Endo, Shibata, and co-workers developed self-assembled
oligomeric multinuclear Cu/Zn complexes 82 and 83 as
asymmetric catalytic systems (Scheme 24).159 The specific
ligand scaffold was essential for the generation of the metal
cluster that could achieve highly enantioselective asymmetric

Scheme 20 Friedel–Crafts alkylation. Scheme 22 Suzuki–Miyaura reaction.

22 | Catal. Sci. Technol., 2020, 10, 12–34 This journal is © The Royal Society of Chemistry 2020
View Article Online

Catalysis Science & Technology Perspective

Scheme 23 Heck reaction.


Published on 21 November 2019. Downloaded on 1/2/2020 10:29:18 PM.

Scheme 24 Asymmetric conjugate addition. Scheme 26 Michael and Henry reactions.

conjugate addition of organozinc reagents to enones. A small exciting research area. Schelter, Walsh, and co-workers
amount (0.5 mol%) of 83 was enough to promote the studied both Michael and Henry reactions catalyzed by
reactions at −20 °C to furnish products in high yields with Shibasaki's catalysts 84–86 containing rare earth (RE), alkali
almost complete enantiomeric purities (up to >98% ee), and metal (M), and BINOLate (Scheme 26).163 Dynamic ligand
even 0.05 mol% of 83 could catalyze the same reactions with exchange behaviors were clarified by 2D EXSY NMR
slightly lower enantioselectivity. A wide range of enones measurements; both alkali metal cations and BINOLate
bearing various electron-donating or electron-withdrawing ligands in the RE/M frameworks were rapidly exchanged
groups were tolerated. under catalytically relevant conditions. In addition, the first
They also reported that multinuclear complexes prepared X-ray crystal structure of a substrate-bound complex of 86
from a regioisomeric ligand (SP1 or SP2), Et2Zn, and a Cu salt was determined, and using this crystal, stoichiometric and
showed opposite enantioselectivity in the asymmetric conjugate catalytic reactivity studies were done to observe that substrate
addition (Scheme 25).160 The oligomeric multinuclear Cu/Zn deprotonation by the catalyst framework was necessary for
complex with SP1 provided the (S)-isomer with 90% ee, while the high enantioselectivity. Rather than a trisIJBINOLate)
that with SP2 afforded the (R)-isomer with 95% ee. A similar
catalytic system with a self-assembled multinuclear Pd/Zn
complex was successfully applied to the asymmetric alkylative
ring-opening reaction of oxabicyclic alkenes.161
Shibasaki and co-workers made a great contribution to the
development of asymmetric catalysis with multinuclear metal
complexes.162 Further studies are continued to explore this

Scheme 25 Inversion of enantioselectivity. Scheme 27 Transesterification.

This journal is © The Royal Society of Chemistry 2020 Catal. Sci. Technol., 2020, 10, 12–34 | 23
View Article Online

Perspective Catalysis Science & Technology


Published on 21 November 2019. Downloaded on 1/2/2020 10:29:18 PM.

Scheme 28 Ring-opening polymerization.

species, a bisIJBINOLate) or oligomeric heterobimetallic


species may be the catalytically active species.

3.10 Transesterification
Ohshima and co-workers reported the transesterification
between methyl benzoate and cyclohexanol using tetranuclear Scheme 29 Oxidation of alkane.
Zn cluster 87 in the presence of an N-heteroaromatic ligand
as an additive (Scheme 27).164,165 The yield was dramatically
improved by the addition of the N-heteroaromatic ligand.
X-ray analysis of single crystals obtained from 87 and meta- of macrocyclic Mn clusters, such as Mn12O12IJOAc)16IJH2O)4
and congeners, using O2 as an oxidant and TEMPO
bisIJimidazolylmethyl)benzene revealed that an active species
was a coordination polymer containing macrocyclic dinuclear (2,2,6,6-tetramethylpiperidine 1-oxyl) as a co-catalyst.169,170
complexes as a repeating unit. The best catalytic activity (94% yield, TON 470) was
observed for mixed-metal cluster CeMn6O9IJOAc)9IJNO3)IJH2-
O)2 95 (Scheme 30). This performance must be due to
3.11 Ring-opening polymerization
the presence of the heterometallic MnIJIV)/CeIJIV) structure
Redshaw and co-workers investigated the ring-opening intimately connected via oxide bridges (CeMn6O9 core) in
polymerization of ε-caprolactone (ε-CL), δ-valerolactone (δ-VL), the cluster.
or rac-lactide (rac-LA) and the ring-opening copolymerization of Photosystem II (PSII) in higher plants contains a Mn4Ca
ε-CL and rac-LA in the presence of macrocyclic Zn clusters cluster for water oxidation.9–12 Because water oxidation is an
88–92 (Scheme 28).166,167 Fluorine-containing catalysts 90 and inorganic reaction, being outside the range of this
91 were more active for the homopolymerizations, whereas all perspective, we just briefly note that there are several
of them showed comparable activities for the copolymerization excellent classes of macrocyclic multinuclear metal
of ε-CL and rac-LA. complexes for photocatalytic water oxidation.171–177 They are
potential photocatalysts for organic reactions.
3.12 Oxidation of alkanes
Selective oxidation of alkanes is a challenging subject. Liu,
Cui and co-workers synthesized triple-stranded cluster
helicates with a heptametallic dicubane core 93 and 94 from
a hexadentate Schiff-base ligand (Scheme 29).168 Cu species
93 catalyzed the peroxidative oxidation of alkanes by a
radical-type mechanism. Selective C–H oxidation of
cyclopentane, cyclohexane, cyclooctane, toluene, and
ethylbenzene proceeded to give the corresponding ketones
along with a trace amount of alcohols.

3.13 Oxidation of alcohols


Christou and co-workers investigated the selective
oxidation of benzyl alcohol to benzaldehyde with a series Scheme 30 Oxidation of alcohol.

24 | Catal. Sci. Technol., 2020, 10, 12–34 This journal is © The Royal Society of Chemistry 2020
View Article Online

Catalysis Science & Technology Perspective

4. Self-assembled supramolecular
coordination complexes (category C)
4.1 Diels–Alder reaction
A variety of self-assembled supramolecular coordination
complexes have been synthesized and used as catalysts for
organic reactions, taking advantage of the vacant inner space,
positive or negative charges, and peripheral catalytic
functional groups.14–19,178,179 Fujita has done pioneering
work, creating self-assembled supramolecular coordination
complexes as exemplified in Fig. 8. Supramolecular complex
96 promoted the Diels–Alder reaction of anthracenes with Scheme 31 Diels–Alder reaction.
N-cyclohexylmaleimide to furnish syn-adducts with unusual
regioselectivity (Scheme 31).180 Although the formation of a
Published on 21 November 2019. Downloaded on 1/2/2020 10:29:18 PM.

9,10-Diels–Alder adduct was favored in the absence of 96, the volume of 462 Å3), with N-tert-butylmaleimide in D2O at 100
1,4-regioselective Diels–Alder reaction was mediated by 96. °C (46% yield).183
This 1,4-regioselectivity results from steric restrictions inside
96. Unusual regioselectivity was also observed for the Diels–
Alder reaction of alkyl-substituted naphthalenes with 4.2 [2 + 2] olefin cross photoaddition
N-cyclohexylmaleimide in 96.181,182 No Diels–Alder reaction The first efficient [2 + 2] cross photoaddition reaction of
occurred in the absence of 96, clearly indicating the essential fluoranthenes with N-cyclohexylmaleimide proceeded in the
role of 96 as a molecular flask to bring the reactants into cavity of 96 to give the [2 + 2] adducts with high
close proximity inside the cavity. The precise tuning of the regioselectivity (Scheme 32).184,185 Interestingly, chiral cage
cavity volume of the coordination cage as well as the 98 achieved an asymmetric induction of 40–50% ee, whereas
substrate sizes can lead to the enhancement in reactivity. chiral cage 99 resulted in an asymmetric induction of 20%
Indeed, 97 (cavity volume of 380 Å3) offered suitable space ee. These results originate from the chiral deformation of the
inside the cavity to promote the Diels–Alder reaction of cavity induced by the different chiral diamino ligands located
unsubstituted naphthalene, which was inert in 96 (cavity at the remote peripheral positions of the cage.

4.3 Knoevenagel condensation


Knoevenagel condensation of aromatic aldehydes was
accelerated in the hydrophobic cavity of cationic cage 100 in
water under neutral conditions (Scheme 33).186 Anionic
intermediates are effectively stabilized by the Pd2+ centers
located at every corner of the cage. Owing to the host–guest
size discrepancy, the product was spontaneously released
from the cage, and the empty cavity allowed for further
incorporation of a substrate molecule to continue the
catalytic cycle. Only a small amount of 100 (1 mol%) was
enough to produce the condensation product under mild

Fig. 8 Self-assembled supramolecular coordination complexes. Scheme 32 [2 + 2] cross photoaddition.

This journal is © The Royal Society of Chemistry 2020 Catal. Sci. Technol., 2020, 10, 12–34 | 25
View Article Online

Perspective Catalysis Science & Technology

Scheme 33 Knoevenagel condensation.

Scheme 35 Nucleophilic substitution.


conditions. Little or no condensation occurred in the absence
of 100.
Using different supramolecular coordination cages, Zhang, preferentially attacked the less protected terminal site of the
Published on 21 November 2019. Downloaded on 1/2/2020 10:29:18 PM.

Lin, Wang, and co-workers investigated the substrate- allylic chlorides.


dependent catalytic activity for the Knoevenagel condensation
of malononitrile with various aldehydes.187 The reaction was 4.6 Organometallic transformation
switched on by the size/shape-selective substrate recognition
Metal–metal bonds are generally photolabile, which imposes
and the electrostatic modulation of the binding pocket via
some restriction in photochemical transformations. Fujita
cationic regulators.
and co-workers demonstrated that the photocleavage of the
Ru–Ru bond of a dinuclear Ru complex could be successfully
4.4 Alkyne hydration suppressed by the tight encapsulation of the complex in 96
(Scheme 36).190 The remarkable photostability of
Coordination cage 96, which contains an electron-deficient
[(Me4Cp)RuIJCO)2]2 by the cage effect enabled CO/alkyne
triazine core and metal-coordinated pyridyl moieties, can act
photosubstitution to generate a Ru–alkyne π complex. This π
as a good electron acceptor as well as a molecular flask.
complex was a metastable species, which was rearranged into
Indeed, this cage served as a photosensitizing molecular flask,
diruthenacyclopentenone by intramolecular CO insertion
facilitating a cage-mediated guest-to-host electron transfer
upon extraction from the cage with CH2Cl2. Interestingly,
leading to new reactivity and selectivity (Scheme 34).188 Under
sterically less demanding alkyne, 2-butyne, was directly
UV-light irradiation, the photo-hydration of internal aryl
converted into the corresponding diruthenacyclopentenone
alkynes occurred in the hydrophobic cavity of 96 to give
product in 96 probably because of the loose packing of the
benzyl ketones with anti-Markovnikov selectivity. After the
Ru–alkyne π complex intermediate in the cavity.
formation of benzyl ketones, 96 acted as a protection chamber
to suppress further photochemical reaction or degradation of
the ketone. 4.7 Oxidation/Diels–Alder cascade reaction
Fujita and co-workers developed supramolecular capsules
[Pd12L24]24+ 102 and 103 bearing TEMPO and MacMillan's
4.5 Nucleophilic substitution catalyst, respectively (Scheme 37).191 A solution of the
Coordination cage 101 non-covalently protected the internal substrate, 102, 103, PhIIJOAc)2, and CF3CO2H in CD3NO2/D2O
reaction site of aryl-substituted allylic chlorides through (98 : 2) was kept at −10 °C for 72 h. The two-step cascade
encapsulation to control the regioselectivity of the nucleophilic reaction, allylic oxidation followed by the intramolecular
substitution reaction (Scheme 35).189 The incoming nucleophile asymmetric Diels–Alder reaction, proceeded successfully,

Scheme 34 Anti-Markovnikov hydration of alkyne. Scheme 36 Photosubstitution.

26 | Catal. Sci. Technol., 2020, 10, 12–34 This journal is © The Royal Society of Chemistry 2020
View Article Online

Catalysis Science & Technology Perspective


Published on 21 November 2019. Downloaded on 1/2/2020 10:29:18 PM.

Scheme 38 Aza-Cope rearrangement and aza-Prins and Prins


cyclizations.

They also reported the bimolecular aza-Prins cyclization of


amines with formaldehyde via the formation of an iminium
intermediate in the cavity of 104, which was followed by an
unexpected 1,5-hydride transfer (Scheme 38).193 The
Scheme 37 Cascade reaction. piperidine products with no methyl group on the nitrogen
atom were obtained; they are unusual as compared to the
products in bulk solution without 104. Deuterium labeling
providing a bicyclic product with high enantio- and experiments strongly suggested that the aza-Prins
diastereoselectivity. Control experiments indicated that only cyclization product underwent 1,5-hydride transfer from the
the combined use of 102 and 103 accelerated the cascade N-methyl group to the tertiary carbocation. This reactivity is
reaction, clearly demonstrating cooperative catalysis. It the result of the constrictive binding of the intermediate in
should be noted that the M12L24 spherical framework 104 that favors a more compact transition state. The
protected both of the active sites, TEMPO and MacMillan's carbocation of the intermediate is preorganized in close
catalyst, from undesirable deactivation, realizing the smooth proximity to the N-methyl C–H bonds, which facilitates a
one-pot cascade reaction (site-isolation strategy). 1,5-hydride transfer.
Synthesis of enantiopure supramolecular tetrahedral cage
[Ga4L6]12− 104 is challenging because a racemic mixture of
4.8 Aza-Cope rearrangement and aza-Prins and Prins 104 is usually formed. Toste, Bergman, Raymond, and co-
cyclizations workers solved this problem by using a ligand with
Bergman, Raymond, and co-workers synthesized and evaluated additional chiral amide groups; enantiopure cluster K12Ga4L6
supramolecular tetrahedral cage [Ga4L6]12− 104 as a chiral 105 was obtained as a single diastereomer, which was
catalyst for the enantioselective aza-Cope rearrangement strongly controlled by the chiral terephthalamide-based
(Scheme 38).192 A catalytic amount of enantiopure ΔΔΔΔ-104 ligand (Scheme 38).194 105 was stable in both the solid and
was used to induce the reactivity of a series of enammonium solution states at elevated temperature and in aerobic D2O,
tosylates. The aza-Cope rearrangement product, the iminium in contrast to 104 that was sensitive to temperature and
ion, was hydrolyzed to neutral aldehydes with up to 78% ee. oxidation. The additional amide group of 105 stabilized the
The enammonium substrate is considered to be tightly assembly via intramolecular hydrogen bonding with the
encapsulated in the cage to adopt a chiral reactive catecholate oxygen atom, preventing the catecholate ligand
conformation. The enantioselectivities were found to depend on from being oxidized and decomposed. ΔΔΔΔ-105 presented a
the size and shape of the substrate molecules as well as on the rare example of the chiral host-catalyzed enantioselective
reaction temperature. Prins cyclization of carbonyl–ene compounds, and 105

This journal is © The Royal Society of Chemistry 2020 Catal. Sci. Technol., 2020, 10, 12–34 | 27
View Article Online

Perspective Catalysis Science & Technology

displayed better catalytic performance (catalyst loading of 2.5


mol%) than 104 (10 mol%).195

4.9 Cross coupling


Alkyl–alkyl cross coupling processes often suffer from poor
turnover and undesired side reactions because of sluggish
reductive elimination of sp3 fragments from transition
metals. Bergman, Raymond, Toste, and co-workers reported
that the alkyl–alkyl reductive elimination from high-valent
transition metal complexes was greatly accelerated by the
supramolecular microenvironment strategy (Scheme 39).196
The reductive elimination of ethane from AuIJIII) and PtIJIV)
complexes in the presence of 104 (10 mol%) was apparently
80 000-fold and 2300-fold faster, respectively, than the
Published on 21 November 2019. Downloaded on 1/2/2020 10:29:18 PM.

uncatalyzed reactions. The kcat/kuncat values determined by


kinetic measurements were 19 000 000 and 26 000 for AuIJIII)
and PtIJIV) complexes, respectively. Furthermore, when
supramolecular capsule 105 and (Me3P)2PtMe2 were used as
dual catalysts for the reaction of MeI with Me4Sn, the alkyl–
alkyl cross coupling reaction proceeded to give ethane
(Scheme 39). Both catalysts were essential for this reaction,
but the role of 105 in the acceleration of the reductive
elimination is quite important for the catalytic turnover. The
same group also studied the oxidative addition of aryl halides
that was facilitated by 104.197 The microenvironment inside
the cage of 104 was useful for the encapsulation of guest
molecules, increasing the local substrate concentration, and
for the stabilization of cationic organometallic intermediate.

4.10 Epoxidation
Liu, Cui, and co-workers constructed five chiral single- or
mixed-linker tetrahedral coordination cages 106–110 to achieve
the sequential asymmetric epoxidation/epoxide ring-opening Scheme 40 Cascade reaction.
reaction (Scheme 40).198 Four bowl-like Cp3Zr3IJμ3-O)IJμ2-OH)3
clusters located in the vertices are linked by six dicarboxylate
ligands of Mn(salen), Cr(salen), and/or Fe(salen) to form a cationic cage {[Cp3Zr3IJμ3-O)IJμ2-OH)3]4IJML)6}Cl6. Mn/Cr mixed-
linker cage 108 (0.2 mol%) catalyzed the sequential reaction in
the presence of 2-(tert-butylsulfonyl)iodosylbenzene (sPhIO) and
TMSN3 in CH2Cl2 at 0 °C; Mn(salen) and Cr(salen) moieties
acted as catalysts for alkene epoxidation and epoxide-ring
opening, respectively. In contrast, each of the single-linker cages
106 and 107 exhibited no catalytic activity. Interestingly, Mn/Cr
cage 108 showed higher activity and enantioselectivity than a
mixture of two catalytic components. The increased stability of
the metallosalen units, the increased reactant concentration,
and the efficient sequential reaction in the single cavity led to
the high catalytic efficacy of 108.
Bruin and co-workers synthesized a self-assembled cubic
molecular cage with an encapsulated manganese porphyrin
catalyst and conducted the epoxidation of styrene derivatives in
50% acetonitrile aqueous media. The epoxide products were
obtained with a TON of up to 319, and substrate size-selectivity
was observed.199 The cage not only acted as a phase-transfer
catalyst but also improved the stability of the encapsulated
Scheme 39 Acceleration of alkyl–alkyl reductive elimination. metalloporphyrin catalyst.

28 | Catal. Sci. Technol., 2020, 10, 12–34 This journal is © The Royal Society of Chemistry 2020
View Article Online

Catalysis Science & Technology Perspective

4.11 Kemp elimination constructed elegant supramolecular multi-component catalyst


The catalytic efficiency of a supramolecular cage catalyst 112 in situ (Scheme 42).202 The chiral Rh complex was
depends on the ability to recognize and bind substrate encapsulated in a supramolecular metallocage, the latter of
molecules and to accelerate the reaction by increasing the local which was constructed by the self-assembly of the
reactant concentrations and/or stabilizing the transition state. macrocyclic Pd complex and Zn porphyrin complex. The Rh
In addition, the ability to expel the product is also important for catalyst entrapped in the cage showed much higher regio-
a high catalytic turnover. Hunter, Williams, Ward, and co- and enantioselectivity in the hydroformylation of styrenes
workers studied the Kemp elimination catalyzed by cubic than the non-encapsulated one. The supramolecular cage
coordination cage [Co8L12]16+ 111 (Scheme 41).200 acted as a second coordination sphere, enabling a through-
Deprotonation of benzisoxazole at the 3-position with hydroxide space chirality control around the catalyst center. The Reek
results in ring-opening to generate a 2-cyanophenolate anion. group also achieved the size-selective hydroformylation of
Hydrophobic binding of benzisoxazole in the cavity and polar terminal alkenes with a Rh catalyst encapsulated in the cavity
binding of hydroxide ions at the cage surface accelerated the of a supramolecular tetrahedral cage.203
reaction with a rate enhancement (kcat/kuncat) of 200 000 at pD
Published on 21 November 2019. Downloaded on 1/2/2020 10:29:18 PM.

8.5. The poor binding of the anionic product in the 4.13 Michael addition
hydrophobic cage allowed easy ejection from the cage,
providing a smooth catalytic turnover. The effect of different A substrate molecule can be bound with a supramolecular
surface-bound anions on the reactivity of the cavity-bound catalyst via hydrogen bonding. Mukherjee and co-workers
benzisoxazole was also reported.201 reported that supramolecular trigonal prism 113 with
multiple urea groups acted as a catalyst for the Michael
addition reaction of nitro-olefins with dimethylbarbituric acid
4.12 Hydroformylation reaction in aqueous medium (Scheme 43).204 Water-insoluble nitro-
Encapsulation of a transition metal complex in a self- olefins that are hydrogen-bonded with the urea moiety of 113
assembled coordination cage may form an efficient undergoes the conjugate addition of the enolate form of
supramolecular catalyst. Costas, Ribas, Reek, and co-workers dimethylbarbituric acid to generate a water-soluble nitronate
intermediate, which is spontaneously released from 113 upon
protonation because of the large size of the product.
The Mukherjee group also reported self-assembled
tetrahedral cage Pt12L4 114 capable of accelerating the
Michael addition of indole with nitrostyrenes (Scheme 44).205

Scheme 41 Kemp elimination reaction. Scheme 42 Hydroformylation.

This journal is © The Royal Society of Chemistry 2020 Catal. Sci. Technol., 2020, 10, 12–34 | 29
View Article Online

Perspective Catalysis Science & Technology

organic reactions are catalyzed by these beautiful


coordination systems. A variety of metal clusters can be
formed in well-designed macrocyclic polydentate ligands,
while metal coordinations can be used to construct
macrocyclic frameworks and metal clusters at the same time.
It has been demonstrated that the macrocyclic skeletons are
useful for the construction, protection, and activation of
metal clusters. Both homometallic and heterometallic clusters
inside the macrocycles show excellent catalytic activity
originating from the synergetic effect. On the other hand, self-
assembled supramolecular coordination complexes with a
hollow cavity can be constructed by the self-assembly of metal
ions and polyvalent ligands. Some of the cage complexes are
positively or negatively charged by an unbalanced
Published on 21 November 2019. Downloaded on 1/2/2020 10:29:18 PM.

Scheme 43 Michael addition. combination of metal cations and anionic or neutral ligands.
The net charge and the hydrophobic cavity can attract
substrates effectively, and the specific cavity is useful for
reaction acceleration and selectivity. The design flexibility of
all these types of complexes will enable further success in
catalytic applications. In many cases, the macrocyclic
multinuclear metal complexes featured herein exerted
remarkable and/or unusual catalytic performances that the
corresponding non-macrocyclic/monometallic counterparts
could not achieve, which is well rationalized by the proposed
mechanisms; for example, see Schemes 2, 15, 17, 31, 36 and
39 and Fig. 6. Creation of artificial catalysts that are superior
to enzymes containing metal clusters in terms of catalytic
efficiency, reaction diversity, selectivity, substrate scope, and
robustness will be one of the most challenging and
fascinating goals. There are many enzymes and proteins
containing an unstable multiple-metal cluster, for which even
model systems are difficult to synthesize. A collection of the
macrocyclic multinuclear metal complex systems shown here
will be useful for a broad range of scientific studies, giving
hints or clues to the creation of a high level of coordination
systems for innovative breakthroughs.

Scheme 44 Michael addition.


Conflicts of interest
There are no conflicts to declare.
The substrates can be stabilized by π–π interactions with the
cage walls, and β-nitrostyrene can be activated by Notes and references
electrostatic interactions with the cationic charge at the cage
vertices, both of which accelerated the electrophilic 1 N. Elgrishi, M. B. Chambers, X. Wang and M. Fontecave,
substitution on indole to give the Michael adduct. Chem. Soc. Rev., 2017, 46, 761–796.
2 C.-Y. Lin and P. P. Power, Chem. Soc. Rev., 2017, 46,
5347–5399.
5. Conclusions and outlook 3 G. Tseberlidis, D. Intrieri and A. Caselli, Eur. J. Inorg.
Chem., 2017, 3589–3603.
In this perspective, we provide an overview of the recent 4 D. S. Nesterov, O. V. Nesterova and A. J. L. Pombeiro, Coord.
progress in homogeneous catalysis with macrocyclic Chem. Rev., 2018, 355, 199–222.
multinuclear metal complexes: (i) multinuclear metal 5 E. Y. Tsui, J. S. Kanady and T. Agapie, Inorg. Chem.,
complexes with a covalently-linked macrocyclic ligand 2013, 52, 13833–13848.
(category A), (ii) multinuclear metal complexes forming a 6 T. Cadenbach, J. R. Pankhurst, T. A. Hofmann, M. Curcio,
macrocyclic skeleton (category B), and (iii) self-assembled P. L. Arnold and J. B. Love, Organometallics, 2015, 34,
supramolecular coordination complexes (category C). Useful 2608–2613.

30 | Catal. Sci. Technol., 2020, 10, 12–34 This journal is © The Royal Society of Chemistry 2020
View Article Online

Catalysis Science & Technology Perspective

7 K. Omoto, S. Tashiro, M. Kuritani and M. Shionoya, J. Am. Bazin, F. Kapteijn, M. Daturi, E. V. Ramos-Fernandez, F. X.
Chem. Soc., 2014, 136, 17946–17949. Llabrés i Xamena, V. Van Speybroeck and J. Gascon, Chem.
8 T. Nakamura, Y. Kaneko, E. Nishibori and T. Nabeshima, Soc. Rev., 2017, 46, 3134–3184.
Nat. Commun., 2017, 8, 129. 35 F. N. Al-Rowaili, A. Jamal, M. S. Ba Shammakh and A. Rana,
9 J. S. Kanady, E. Y. Tsui, M. W. Day and T. Agapie, Science, ACS Sustainable Chem. Eng., 2018, 6, 15895–15914.
2011, 333, 733–736. 36 A. Dhakshinamoorthy, A. M. Asiri, M. Alvaro and H. Garcia,
10 C. Chen, C. Zhang, H. Dong and J. Zhao, Chem. Commun., Green Chem., 2018, 20, 86–107.
2014, 50, 9263–9265. 37 A. Dhakshinamoorthy, Z. Li and H. Garcia, Chem. Soc. Rev.,
11 C. Zhang, C. Chen, H. Dong, J.-R. Shen, H. Dau and J. 2018, 47, 8134–8172.
Zhao, Science, 2015, 348, 690–693. 38 A. Dhakshinamoorthy, A. M. Asiri and H. Garcia, ACS
12 C. Chen, Y. Chen, R. Yao, Y. Li and C. Zhang, Angew. Chem., Catal., 2019, 9, 1081–1102.
Int. Ed., 2019, 58, 3939–3942. 39 M. Liu, J. Wu and H. Hou, Chem. – Eur. J., 2019, 25,
13 M. M. J. Smulders, I. A. Riddell, C. Browne and J. R. 2935–2948.
Nitschke, Chem. Soc. Rev., 2013, 42, 1728–1754. 40 S. D. P. Fielden, D. A. Leigh and S. L. Woltering, Angew.
Published on 21 November 2019. Downloaded on 1/2/2020 10:29:18 PM.

14 M. Yoshizawa, J. K. Klosterman and M. Fujita, Angew. Chem., Int. Ed., 2017, 56, 11166–11194.
Chem., Int. Ed., 2009, 48, 3418–3438. 41 J.-F. Ayme, J. E. Beves, C. J. Campbell and D. A. Leigh,
15 C. J. Brown, F. D. Toste, R. G. Bergman and K. N. Raymond, Chem. Soc. Rev., 2013, 42, 1700–1712.
Chem. Rev., 2015, 115, 3012–3035. 42 M. Marenda, E. Orlandini and C. Micheletti, Nat. Commun.,
16 M. Otte, ACS Catal., 2016, 6, 6491–6510. 2018, 9, 3051.
17 C. M. Hong, R. G. Bergman, K. N. Raymond and F. D. 43 F. B. L. Cougnon, K. Caprice, M. Pupier, A. Bauzá and A.
Toste, Acc. Chem. Res., 2018, 51, 2447–2455. Frontera, J. Am. Chem. Soc., 2018, 140, 12442–12450.
18 C. Tan, D. Chu, X. Tang, Y. Liu, W. Xuan and Y. Cui, Chem. 44 N. C. H. Lim and S. E. Jackson, J. Phys.: Condens. Matter,
– Eur. J., 2019, 25, 662–672. 2015, 27, 354101.
19 Y. Fang, J. A. Powell, E. Li, Q. Wang, Z. Perry, A. Kirchon, X. 45 P. Buchwalter, J. Rosé and P. Braunstein, Chem. Rev.,
Yang, Z. Xiao, C. Zhu, L. Zhang, F. Huang and H.-C. Zhou, 2015, 115, 28–126.
Chem. Soc. Rev., 2019, 48, 4707–4730. 46 T. Shima, Y. Luo, T. Stewart, R. Bau, G. J. McIntyre, S. A.
20 T. R. Cook, Y.-R. Zheng and P. J. Stang, Chem. Rev., Mason and Z. Hou, Nat. Chem., 2011, 3, 814–820.
2013, 113, 734–777. 47 Y. Li, Y. Li, B. Wang, Y. Luo, D. Yang, P. Tong, J. Zhao, L.
21 A. G. Slater and A. I. Cooper, Science, 2015, 348, aaa8075. Luo, Y. Zhou, S. Chen, F. Cheng and J. Qu, Nat. Chem.,
22 S. Kitagawa, Acc. Chem. Res., 2017, 50, 514–516. 2013, 5, 320–326.
23 T. Kitao, Y. Zhang, S. Kitagawa, B. Wang and T. Uemura, 48 T. Shima, S. Hu, G. Luo, X. Kang, Y. Luo and Z. Hou,
Chem. Soc. Rev., 2017, 46, 3108–3133. Science, 2013, 340, 1549–1552.
24 R. Gaillac, P. Pullumbi, K. A. Beyer, K. W. Chapman, D. A. 49 S. Hu, T. Shima and Z. Hou, Nature, 2014, 512, 413–415.
Keen, T. D. Bennett and F.-X. Coudert, Nat. Mater., 2017, 16, 50 K. Wang, G. Luo, J. Hong, X. Zhou, L. Weng, Y. Luo and L.
1149–1154. Zhang, Angew. Chem., Int. Ed., 2014, 53, 1053–1056.
25 H.-C. Hu and B. Zhao, Chem. – Eur. J., 2018, 24, 51 G. Luo, Y. Luo, Z. Hou and J. Qu, Organometallics, 2016, 35,
16702–16707. 778–784.
26 M. J. Kalmutzki, C. S. Diercks and O. M. Yaghi, Adv. Mater., 52 E. J. L. McInnes, G. A. Timco, G. F. S. Whitehead and
2018, 30, 1704304. R. E. P. Winpenny, Angew. Chem., Int. Ed., 2015, 54,
27 H. Furukawa, K. E. Cordova, M. O'Keeffe and O. M. Yaghi, 14244–14269.
Science, 2013, 341, 1230444. 53 S. Castellanos, F. Kapteijn and J. Gascon, CrystEngComm,
28 A. H. Chughtai, N. Ahmad, H. A. Younus, A. Laypkov and F. 2016, 18, 4006–4012.
Verpoort, Chem. Soc. Rev., 2015, 44, 6804–6849. 54 A. Dhakshinamoorthy, M. Alvaro and H. Garcia, Chem.
29 A. Schoedel, Z. Ji and O. M. Yaghi, Nat. Energy, 2016, 1, Commun., 2012, 48, 11275–11288.
16034. 55 A. Dhakshinamoorthy, A. M. Asiri and H. Garcia, Chem.
30 C. A. Trickett, A. Helal, B. A. Al-Maythalony, Z. H. Yamani, Soc. Rev., 2015, 44, 1922–1947.
K. E. Cordova and O. M. Yaghi, Nat. Rev. Mater., 2017, 2, 56 S. Subudhi, D. Rath and K. M. Parida, Catal. Sci. Technol.,
17045. 2018, 8, 679–696.
31 L. Zhu, X.-Q. Liu, H.-L. Jiang and L.-B. Sun, Chem. Rev., 57 T. Joshi, B. Graham and L. Spiccia, Acc. Chem. Res.,
2017, 117, 8129–8176. 2015, 48, 2366–2379.
32 A. Dhakshinamoorthy, A. M. Asiri and H. Garcia, ACS 58 B. Roy, A. K. Ghosh, S. Srivastava, P. D'Silva and P. S.
Catal., 2017, 7, 2896–2919. Mukherjee, J. Am. Chem. Soc., 2015, 137, 11916–11919.
33 Y.-B. Huang, J. Liang, X.-S. Wang and R. Cao, Chem. Soc. 59 Z. Sun, M. Yang, Y. Ma and L. Li, Cryst. Growth Des.,
Rev., 2017, 46, 126–157. 2017, 17, 4326–4335.
34 S. M. J. Rogge, A. Bavykina, J. Hajek, H. Garcia, A. I. Olivos- 60 E. A. Dolgopolova, A. M. Rice, C. R. Martin and N. B.
Suarez, A. Sepúlveda-Escribano, A. Vimont, G. Clet, P. Shustova, Chem. Soc. Rev., 2018, 47, 4710–4728.

This journal is © The Royal Society of Chemistry 2020 Catal. Sci. Technol., 2020, 10, 12–34 | 31
View Article Online

Perspective Catalysis Science & Technology

61 C. García-Simón, M. Garcia-Borràs, L. Gómez, T. Parella, S. 88 M. R. Kember, A. Buchard and C. K. Williams, Chem.


Osuna, J. Juanhuix, I. Imaz, D. Maspoch, M. Costas and X. Commun., 2011, 47, 141–163.
Ribas, Nat. Commun., 2014, 5, 5557. 89 X.-B. Lu and D. J. Darensbourg, Chem. Soc. Rev., 2012, 41,
62 X. Li, B. Wang, Y. Cao, S. Zhao, H. Wang, X. Feng, J. Zhou 1462–1484.
and X. Ma, ACS Sustainable Chem. Eng., 2019, 7, 4548–4563. 90 D. J. Darensbourg and S. J. Wilson, Green Chem., 2012, 14,
63 H. Sato, W. Kosaka, R. Matsuda, A. Hori, Y. Hijikata, R. V. 2665–2671.
Belosludov, S. Sakaki, M. Takata and S. Kitagawa, Science, 91 X.-B. Lu, W.-M. Ren and G.-P. Wu, Acc. Chem. Res., 2012, 45,
2014, 343, 167–170. 1721–1735.
64 K. S. Walton, Nat. Chem., 2014, 6, 277–278. 92 N. Ikpo, J. C. Flogeras and F. M. Kerton, Dalton Trans.,
65 A. Cadiau, K. Adil, P. M. Bhatt, Y. Belmabkhout and M. 2013, 42, 8998–9006.
Eddaoudi, Science, 2016, 353, 137–140. 93 S. Paul, Y. Zhu, C. Romain, R. Brooks, P. K. Saini and C. K.
66 H. Li, L. Li, R.-B. Lin, G. Ramirez, W. Zhou, R. Krishna, Z. Williams, Chem. Commun., 2015, 51, 6459–6479.
Zhang, S. Xiang and B. Chen, ACS Sustainable Chem. Eng., 94 Y. Zhu, C. Romain and C. K. Williams, Nature, 2016, 540,
2019, 7, 4897–4902. 354–362.
Published on 21 November 2019. Downloaded on 1/2/2020 10:29:18 PM.

67 Y. Ye, Z. Ma, R.-B. Lin, R. Krishna, W. Zhou, Q. Lin, Z. 95 S. J. Poland and D. J. Darensbourg, Green Chem., 2017, 19,
Zhang, S. Xiang and B. Chen, J. Am. Chem. Soc., 2019, 141, 4990–5011.
4130–4136. 96 M. R. Kember, P. D. Knight, P. T. R. Reung and C. K.
68 W.-Y. Zhang, Y.-J. Lin, Y.-F. Han and G.-X. Jin, J. Am. Chem. Williams, Angew. Chem., Int. Ed., 2009, 48, 931–933.
Soc., 2016, 138, 10700–10707. 97 F. Jutz, A. Buchard, M. R. Kember, S. B. Fredriksen and
69 K. Fujie, K. Otsubo, R. Ikeda, T. Yamada and H. Kitagawa, C. K. Williams, J. Am. Chem. Soc., 2011, 133, 17395–17405.
Chem. Sci., 2015, 6, 4306–4310. 98 A. Buchard, F. Jutz, M. R. Kember, A. J. P. White, H. S. Rzepa
70 M. Sadakiyo, T. Yamada and H. Kitagawa, J. Am. Chem. Soc., and C. K. Williams, Macromolecules, 2012, 45, 6781–6795.
2014, 136, 13166–13169. 99 M. R. Kember, A. J. P. White and C. K. Williams,
71 M. Sadakiyo, T. Yamada and H. Kitagawa, Inorg. Chem. Macromolecules, 2010, 43, 2291–2298.
Commun., 2016, 72, 138–140. 100 M. R. Kember, F. Jutz, A. Buchard, A. J. P. White and C. K.
72 A. Casini, B. Woods and M. Wenzel, Inorg. Chem., 2017, 56, Williams, Chem. Sci., 2012, 3, 1245–1255.
14715–14729. 101 M. R. Kember and C. K. Williams, J. Am. Chem. Soc.,
73 N. Ahmad, H. A. Younus, A. H. Chughtai and F. Verpoort, 2012, 134, 15676–15679.
Chem. Soc. Rev., 2015, 44, 9–25. 102 A. Buchard, M. R. Kember, K. G. Sandeman and C. K.
74 S. K. Samanta, D. Moncelet, V. Briken and L. Isaacs, J. Am. Williams, Chem. Commun., 2011, 47, 212–214.
Chem. Soc., 2016, 138, 14488–14496. 103 P. K. Saini, C. Romain and C. K. Williams, Chem. Commun.,
75 J. Ren, H. W. Langmi, B. C. North and M. Mathe, Int. J. 2014, 50, 4164–4167.
Energy Res., 2015, 39, 607–620. 104 J. A. Garden, P. K. Saini and C. K. Williams, J. Am. Chem.
76 B. Li, H.-M. Wen, W. Zhou, J. Q. Xu and B. Chen, Chem, Soc., 2015, 137, 15078–15081.
2016, 1, 557–580. 105 J. A. Garden, A. J. P. White and C. K. Williams, Dalton
77 G. Li, H. Kobayashi, J. M. Taylor, R. Ikeda, Y. Kubota, K. Trans., 2017, 46, 2532–2541.
Kato, M. Takata, T. Yamamoto, S. Toh, S. Matsumura and 106 G. Trott, J. A. Garden and C. K. Williams, Chem. Sci.,
H. Kitagawa, Nat. Mater., 2014, 13, 802–806. 2019, 10, 4618–4627.
78 M. Inukai, M. Tamura, S. Horike, M. Higuchi, S. Kitagawa 107 A. C. Deacy, C. B. Durr, J. A. Garden, A. J. P. White and
and K. Nakamura, Angew. Chem., Int. Ed., 2018, 57, C. K. Williams, Inorg. Chem., 2018, 57, 15575–15583.
8687–8690. 108 M. W. Lehenmeier, C. Bruckmeier, S. Klaus, J. E. Dengler,
79 I. Omae, Coord. Chem. Rev., 2012, 256, 1384–1405. P. Deglmann, A.-K. Ott and B. Rieger, Chem. – Eur. J.,
80 L. Zhang and Z. Hou, Chem. Sci., 2013, 4, 3395–3403. 2011, 17, 8858–8869.
81 M. Aresta, A. Dibenedetto and A. Angelini, Chem. Rev., 109 M. W. Lehenmeier, S. Kissling, P. T. Altenbuchner, C.
2014, 114, 1709–1742. Bruckmeier, P. Deglmann, A.-K. Brym and B. Rieger, Angew.
82 C. Maeda, Y. Miyazaki and T. Ema, Catal. Sci. Technol., Chem., Int. Ed., 2013, 52, 9821–9826.
2014, 4, 1482–1497. 110 S. Kissling, P. T. Altenbuchner, M. W. Lehenmeier, E.
83 Q. Liu, L. Wu, R. Jackstell and M. Beller, Nat. Commun., Herdtweck, P. Deglmann, U. B. Seemann and B. Rieger,
2015, 6, 5933. Chem. – Eur. J., 2015, 21, 8148–8157.
84 G. Fiorani, W. Guo and A. W. Kleij, Green Chem., 2015, 17, 111 S. Kissling, M. W. Lehenmeier, P. T. Altenbuchner, A.
1375–1389. Kronast, M. Reiter, P. Deglmann, U. B. Seemann and B.
85 B. Yu and L.-N. He, ChemSusChem, 2015, 8, 52–62. Rieger, Chem. Commun., 2015, 51, 4579–4582.
86 Q.-W. Song, Z.-H. Zhou and L.-N. He, Green Chem., 112 H. Nagae, R. Aoki, S. Akutagawa, J. Kleemann, R.
2017, 19, 3707–3728. Tagawa, T. Schindler, G. Choi, T. P. Spaniol, H. Tsurugi,
87 S. Klaus, M. W. Lehenmeier, C. E. Anderson and B. Rieger, J. Okuda and K. Mashima, Angew. Chem., Int. Ed.,
Coord. Chem. Rev., 2011, 255, 1460–1479. 2018, 57, 2492–2496.

32 | Catal. Sci. Technol., 2020, 10, 12–34 This journal is © The Royal Society of Chemistry 2020
View Article Online

Catalysis Science & Technology Perspective

113 L. Jin, Y. Huang, H. Jing, T. Chang and P. Yan, Tetrahedron: 141 M. V. Escárcega-Bobadilla, M. M. Belmonte, E. Martin, E. C.
Asymmetry, 2008, 19, 1947–1953. Escudero-Adán and A. W. Kleij, Chem. – Eur. J., 2013, 19,
114 B. J. Cook, G. N. Di Francesco, K. A. Abboud and L. J. 2641–2648.
Murray, J. Am. Chem. Soc., 2018, 140, 5696–5700. 142 K. Takaishi, B. D. Nath, Y. Yamada, H. Kosugi and T. Ema,
115 M. Rudolph, S. Dautz and E.-G. Jäger, J. Am. Chem. Soc., Angew. Chem., Int. Ed., 2019, 58, 9984–9988.
2000, 122, 10821–10830. 143 U. R. Pokharel, F. R. Fronczek and A. W. Maverick, Nat.
116 R. Angamuthu, P. Byers, M. Lutz, A. L. Spek and E. Commun., 2014, 5, 5883.
Bouwman, Science, 2010, 327, 313–315. 144 R. J. Eisenhart, L. J. Clouston and C. C. Lu, Acc. Chem. Res.,
117 J. M. Longo, M. J. Sanford and G. W. Coates, Chem. Rev., 2015, 48, 2885–2894.
2016, 116, 15167–15197. 145 N. P. Mankad, Chem. – Eur. J., 2016, 22, 5822–5829.
118 P. K. Saini, C. Romain, Y. Zhu and C. K. Williams, Polym. 146 I. G. Powers and C. Uyeda, ACS Catal., 2017, 7, 936–958.
Chem., 2014, 5, 6068–6075. 147 R. C. Cammarota, M. V. Vollmer, J. Xie, J. Ye, J. C. Linehan,
119 M. Winkler, C. Romain, M. A. R. Meier and C. K. Williams, S. A. Burgess, A. M. Appel, L. Gagliardi and C. C. Lu, J. Am.
Green Chem., 2015, 17, 300–306. Chem. Soc., 2017, 139, 14244–14250.
Published on 21 November 2019. Downloaded on 1/2/2020 10:29:18 PM.

120 R. C. Jeske, J. M. Rowley and G. W. Coates, Angew. Chem., 148 J. Ye, R. C. Cammarota, J. Xie, M. V. Vollmer, D. G. Truhlar,
Int. Ed., 2008, 47, 6041–6044. C. J. Cramer, C. C. Lu and L. Gagliardi, ACS Catal., 2018, 8,
121 M. Delferro and T. J. Marks, Chem. Rev., 2011, 111, 4955–4968.
2450–2485. 149 J. Takaya and N. Iwasawa, J. Am. Chem. Soc., 2017, 139,
122 J. Liu, Y. Li, J. Liu and Z. Li, Macromolecules, 2005, 38, 6074–6077.
2559–2563. 150 R. C. Cammarota and C. C. Lu, J. Am. Chem. Soc., 2015, 137,
123 Z. Chen, X. Zhao, X. Gong, D. Xu and Y. Ma, 12486–12489.
Macromolecules, 2017, 50, 6561–6568. 151 G. Amenuvor, J. Darkwa and B. C. E. Makhubela, Catal. Sci.
124 D. Takeuchi, Y. Chiba, S. Takano and K. Osakada, Angew. Technol., 2018, 8, 2370–2380.
Chem., Int. Ed., 2013, 52, 12536–12540. 152 C. Rong, F. Wang, W. Li and M. Chen, Organometallics,
125 D. Takeuchi, Y. Chiba, S. Takano, H. Kurihara, M. 2017, 36, 4458–4464.
Kobayashi and K. Osakada, Polym. Chem., 2017, 8, 153 K. Griffiths, P. Kumar, G. R. Akien, N. F. Chilton, A. Abdul-
5112–5119. Sada, G. J. Tizzard, S. J. Coles and G. E. Kostakis, Chem.
126 D. Takeuchi, S. Takano, Y. Takeuchi and K. Osakada, Commun., 2016, 52, 7866–7869.
Organometallics, 2014, 33, 5316–5323. 154 P. Kumar, S. Lymperopoulou, K. Griffiths, S. I. Sampani
127 K. Tanaka, A. Asakura, T. Muraoka, P. Kalicki and Z. and G. E. Kostakis, Catalysts, 2016, 6, 140.
Urbanczyk-Lipkowska, New J. Chem., 2013, 37, 2851–2855. 155 K. Griffiths, C. W. D. Gallop, A. Abdul-Sada, A. Vargas, O. Navarro
128 K. Tanaka and S. Hachiken, Tetrahedron Lett., 2008, 49, and G. E. Kostakis, Chem. – Eur. J., 2015, 21, 6358–6361.
2533–2536. 156 K. Griffiths, P. Kumar, J. D. Mattock, A. Abdul-Sada, M. B.
129 A. Gualandi, L. Cerisoli, H. Stoeckli-Evans and D. Savoia, Pitak, S. J. Coles, O. Navarro, A. Vargas and G. E. Kostakis,
J. Org. Chem., 2011, 76, 3399–3408. Inorg. Chem., 2016, 55, 6988–6994.
130 D. Savoia, A. Gualandi and H. Stoeckli-Evans, Org. Biomol. 157 K. V. Vivekananda, S. Dey, D. K. Maity, N. Bhuvanesh and
Chem., 2010, 8, 3992–3996. V. K. Jain, Inorg. Chem., 2015, 54, 10153–10162.
131 H. Lönnberg, Org. Biomol. Chem., 2011, 9, 1687–1703. 158 P. A. Mane, S. Dey and K. V. Vivekananda, Tetrahedron Lett.,
132 H. Korhonen, T. Koivusalo, S. Toivola and S. Mikkola, Org. 2017, 58, 25–29.
Biomol. Chem., 2013, 11, 8324–8339. 159 K. Endo, M. Ogawa and T. Shibata, Angew. Chem., Int. Ed.,
133 M. Zhao, H.-B. Wang, L.-N. Ji and Z.-W. Mao, Chem. Soc. 2010, 49, 2410–2413.
Rev., 2013, 42, 8360–8375. 160 K. Endo, D. Hamada, S. Yakeishi, M. Ogawa and T. Shibata,
134 C. I. Maxwell, N. J. Mosey and R. S. Brown, J. Am. Chem. Org. Lett., 2012, 14, 2342–2345.
Soc., 2013, 135, 17209–17222. 161 K. Endo, K. Tanaka, M. Ogawa and T. Shibata, Org. Lett.,
135 X. Zhang, X. Zheng, D. L. Phillips and C. Zhao, Dalton 2011, 13, 868–871.
Trans., 2014, 43, 16289–16299. 162 N. Kumagai, M. Kanai and H. Sasai, ACS Catal., 2016, 6,
136 X. Zhang, Y. Zhu, X. Zheng, D. L. Phillips and C. Zhao, 4699–4709.
Inorg. Chem., 2014, 53, 3354–3361. 163 J. R. Robinson, J. Gu, P. J. Carroll, E. J. Schelter and P. J.
137 S. Anbu, M. Kandaswamy and B. Varghese, Dalton Trans., Walsh, J. Am. Chem. Soc., 2015, 137, 7135–7144.
2010, 39, 3823–3832. 164 T. Ohshima, Chem. Pharm. Bull., 2016, 64, 523–539.
138 S. Anbu, S. Kamalraj, B. Varghese, J. Muthumary and M. 165 D. Nakatake, Y. Yokote, Y. Matsushima, R. Yazaki and T.
Kandaswamy, Inorg. Chem., 2012, 51, 5580–5592. Ohshima, Green Chem., 2016, 18, 1524–1530.
139 C.-Y. Yu, H.-J. Chuang and B.-T. Ko, Catal. Sci. Technol., 166 C. Redshaw and M. R. J. Elsegood, Angew. Chem., Int. Ed.,
2016, 6, 1779–1791. 2007, 46, 7453–7457.
140 R. Zhang, L. Wang, C. Xu, H. Yang, W. Chen, G. Gao and 167 Y. F. Al-Khafaji, M. R. J. Elsegood, J. W. A. Frese and C.
W. Liu, Dalton Trans., 2018, 47, 7159–7165. Redshaw, RSC Adv., 2017, 7, 4510–4517.

This journal is © The Royal Society of Chemistry 2020 Catal. Sci. Technol., 2020, 10, 12–34 | 33
View Article Online

Perspective Catalysis Science & Technology

168 Y. Fang, W. Gong, L. Liu, Y. Liu and Y. Cui, Inorg. Chem., 188 T. Murase, H. Takezawa and M. Fujita, Chem. Commun.,
2016, 55, 10102–10105. 2011, 47, 10960–10962.
169 A. J. Tasiopoulos, P. L. Milligan, Jr., K. A. Abboud, T. A. 189 Y. Kohyama, T. Murase and M. Fujita, Chem. Commun.,
O'Brien and G. Christou, Inorg. Chem., 2007, 46, 9678–9691. 2012, 48, 7811–7813.
170 G. Maayan and G. Christou, Inorg. Chem., 2011, 50, 190 S. Horiuchi, T. Murase and M. Fujita, Angew. Chem., Int.
7015–7021. Ed., 2012, 51, 12029–12031.
171 A. Sartorel, M. Bonchio, S. Campagna and F. Scandola, 191 Y. Ueda, H. Ito, D. Fujita and M. Fujita, J. Am. Chem. Soc.,
Chem. Soc. Rev., 2013, 42, 2262–2280. 2017, 139, 6090–6093.
172 J.-H. Xu, L.-Y. Guo, H.-F. Su, X. Gao, X.-F. Wu, W.-G. Wang, 192 C. J. Brown, R. G. Bergman and K. N. Raymond, J. Am.
C.-H. Tung and D. Sun, Inorg. Chem., 2017, 56, 1591–1598. Chem. Soc., 2009, 131, 17530–17531.
173 F. Evangelisti, R. Moré, F. Hodel, S. Luber and G. R. Patzke, 193 D. M. Kaphan, F. D. Toste, R. G. Bergman and K. N.
J. Am. Chem. Soc., 2015, 137, 11076–11084. Raymond, J. Am. Chem. Soc., 2015, 137, 9202–9205.
174 F. Evangelisti, R. Güttinger, R. Moré, S. Luber and G. R. 194 C. Zhao, Q.-F. Sun, W. M. Hart-Cooper, A. G. DiPasquale,
Patzke, J. Am. Chem. Soc., 2013, 135, 18734–18737. F. D. Toste, R. G. Bergman and K. N. Raymond, J. Am.
Published on 21 November 2019. Downloaded on 1/2/2020 10:29:18 PM.

175 N. S. McCool, D. M. Robinson, J. E. Sheats and G. C. Chem. Soc., 2013, 135, 18802–18805.
Dismukes, J. Am. Chem. Soc., 2011, 133, 11446–11449. 195 W. M. Hart-Cooper, K. N. Clary, F. D. Toste, R. G. Bergman
176 P. F. Smith, C. Kaplan, J. E. Sheats, D. M. Robinson, N. S. and K. N. Raymond, J. Am. Chem. Soc., 2012, 134,
McCool, N. Mezle and G. C. Dismukes, Inorg. Chem., 17873–17876.
2014, 53, 2113–2121. 196 D. M. Kaphan, M. D. Levin, R. G. Bergman, K. N. Raymond
177 S. Berardi, G. La Ganga, M. Natali, I. Bazzan, F. Puntoriero, and F. D. Toste, Science, 2015, 350, 1235–1238.
A. Sartorel, F. Scandola, S. Campagna and M. Bonchio, 197 T. A. Bender, M. Morimoto, R. G. Bergman, K. N. Raymond
J. Am. Chem. Soc., 2012, 134, 11104–11107. and F. D. Toste, J. Am. Chem. Soc., 2019, 141, 1701–1706.
178 Y. Inokuma, M. Kawano and M. Fujita, Nat. Chem., 2011, 3, 198 J. Jiao, C. Tan, Z. Li, Y. Liu, X. Han and Y. Cui, J. Am. Chem.
349–358. Soc., 2018, 140, 2251–2259.
179 K. Harris, D. Fujita and M. Fujita, Chem. Commun., 199 P. F. Kuijpers, M. Otte, M. Dürr, I. Ivanović-Burmazović,
2013, 49, 6703–6712. J. N. H. Reek and B. de Bruin, ACS Catal., 2016, 6,
180 M. Yoshizawa, M. Tamura and M. Fujita, Science, 2006, 312, 3106–3112.
251–254. 200 W. Cullen, M. C. Misuraca, C. A. Hunter, N. H. Williams
181 T. Murase, S. Horiuchi and M. Fujita, J. Am. Chem. Soc., and M. D. Ward, Nat. Chem., 2016, 8, 231–236.
2010, 132, 2866–2867. 201 W. Cullen, A. J. Metherell, A. B. Wragg, C. G. P. Taylor,
182 S. Horiuchi, T. Murase and M. Fujita, Chem. – Asian J., N. H. Williams and M. D. Ward, J. Am. Chem. Soc.,
2011, 6, 1839–1847. 2018, 140, 2821–2828.
183 Y. Fang, T. Murase and M. Fujita, Chem. Lett., 2015, 44, 202 C. García-Simón, R. Gramage-Doria, S. Raoufmoghaddam,
1095–1097. T. Parella, M. Costas, X. Ribas and J. N. H. Reek, J. Am.
184 Y. Nishioka, T. Yamaguchi, M. Kawano and M. Fujita, J. Am. Chem. Soc., 2015, 137, 2680–2687.
Chem. Soc., 2008, 130, 8160–8161. 203 S. S. Nurttila, W. Brenner, J. Mosquera, K. M. van Vliet, J. R.
185 S. Horiuchi, Y. Nishioka, T. Murase and M. Fujita, Chem. Nitschke and J. N. H. Reek, Chem. – Eur. J., 2019, 25,
Commun., 2010, 46, 3460–3462. 609–620.
186 T. Murase, Y. Nishijima and M. Fujita, J. Am. Chem. Soc., 204 P. Howlader, P. Das, E. Zangrando and P. S. Mukherjee,
2012, 134, 162–164. J. Am. Chem. Soc., 2016, 138, 1668–1676.
187 Y. Qiao, L. Zhang, J. Li, W. Lin and Z. Wang, Angew. Chem., 205 I. A. Bhat, A. Devaraj, P. Howlader, K.-W. Chi and P. S.
Int. Ed., 2016, 55, 12778–12782. Mukherjee, Chem. Commun., 2018, 54, 4814–4817.

34 | Catal. Sci. Technol., 2020, 10, 12–34 This journal is © The Royal Society of Chemistry 2020

You might also like

pFad - Phonifier reborn

Pfad - The Proxy pFad of © 2024 Garber Painting. All rights reserved.

Note: This service is not intended for secure transactions such as banking, social media, email, or purchasing. Use at your own risk. We assume no liability whatsoever for broken pages.


Alternative Proxies:

Alternative Proxy

pFad Proxy

pFad v3 Proxy

pFad v4 Proxy