Metal Carbonyl Clusters: Synthesis and Catalysis
Metal Carbonyl Clusters: Synthesis and Catalysis
Manuscript ID CS-REV-02-2021-000161.R2
Complete List of Authors: Cesari, Cristiana; University of Bologna, Chimica Industriale "Toso
Montanari"
Shon, JongHwa; UC Davis, Chemistry
Zacchini, Stefano; Universita di Bologna, Dipartimento di Chimica
Industriale "Toso Montanari"
Berben, Louise; UC Davis, Chemistry
Page 1 of 76 Chemical Society Reviews
1 Department of Chemistry, The University of California, Davis CA, 95616, United States
2 Dipartimento di Chimica Industriale "Toso Montanari", Università di Bologna, Viale
Risorgimento 4, 40136 Bologna, Italy
Corresponding Authors
[email protected], [email protected]
Abstract. In this review article, we discuss advances in the chemistry of metal carbonyl clusters
(MCCs) spanning the last three decades, with an emphasis on the more recent reports and those
involving Groups 8 – 10 elements. Synthetic methods have advanced and been refined, leading to
higher-nuclearity clusters and a wider array of structures and nuclearities. Our understanding of the
electronic structure in MCCs has advanced to a point where molecular chemistry tools and other
advanced tools can probe their properties at a level of detail that surpasses that possible with other
nanomaterials and solid-state materials. MCCs therefore advance our understanding of structure–
property–reactivity correlations in other higher-nuclearity materials. With respect to catalysis, this
article focuses only on homogeneous applications, but it includes both thermally and
electrochemically driven catalysis. Applications in thermally driven catalysis have found success
where the reaction conditions stabilise the compounds toward loss of CO. In more recent years,
MCCs, which exhibit delocalised bonding and possess many electron-withdrawing CO ligands,
have emerged as very stable and effective for reductive electrocatalysis reactions since reduction
often strengthens M–C(O) bonds and since room-temperature reaction conditions are sufficient for
driving the electrocatalysis.
1. Introduction
A review article describing clusters and their analogies with surfaces was published in 1979 by
Muetterties and Rhodin and coworkers; the second page states, “With such riches, a comprehensive
review of chemical bonding in metal clusters compared to metal surfaces is neither practical nor
appropriate here. ... The largest subgroup comprises the metal carbonyl clusters and the size range
1
Chemical Society Reviews Page 2 of 76
within this subgroup extends through the full, established range for clusters of 3 to 30.”1 Now 42
years later, the size range of homoleptic metal carbonyl clusters (MCCs) has expanded, ranging
from 3 to 56; since then, hundreds of papers reporting the activity targeting the many aspects of
MCC chemistry have been published. Numerous other reviews have appeared since that
time,234567891011121314 and a comprehensive review is certainly not practical or appropriate in the
foregoing pages of this article. Hence, we will focus our insights on synthetic methods for obtaining
Groups 8 – 10 MCCs and on their use as homogeneous catalysts and as homogeneous
electrocatalysts. We will feature work published in the most recent decades, and we will provide a
view of future interesting directions in the field. We have decided to exclude MCCs of Groups 5 – 7
because they span a smaller range of clusters, and examples of their synthesis and catalysis are
primarily older work.
The work of Hieber defined the beginning of modern metal-carbonyl chemistry and at the
same time, of modern organometallic hydride chemistry with the reports of reactions surrounding
Fe(CO)5 and the synthesis of H2Fe(CO)4 in 1931.15 The chemistry of MCCs developed rapidly in
the 1970s and 1980s, whereas there was a marked decline in publications on this topic in the 1990s.
A renewed interest started in the noughties,16 due to the intense research in the more general field of
molecular nanoclusters and ultra-small metal nanoparticles.171819202122 Within this framework, the
recent research in MCC chemistry has focused on the synthesis and structural characterization of
metal carbonyl nanoclusters (high-nuclearity MCCs), the investigation of their physical properties
in relationship with their molecular structure and composition, and their applications in catalysis
and electrocatalysis. Because high temperatures may initiate CO ligand loss, electrocatalysis
enables catalysis at room temperature, where MCC decomposition is less likely. Electrochemical
reduction reactions in MCCs are particularly promising because the many CO ligands are very
efficient at stabilising any electrons added to the cluster core to initiate reductive electrocatalytic
reaction cycles. Some of these aspects will be covered in this review. General references, covering
the full scope of MCC chemistry, may be found in the cited literature.
In the first part of this review, we discuss in detail the synthesis of MMCs from Groups 8 –
10; after general principles, we illustrate representative examples from recent work. In the
Electronic Supporting Information (ESI), we also tabulate many of the known MCCs with citations
to their syntheses. To further narrow the scope of the review, we mainly focus our attention on the
syntheses of homometallic and homoleptic MCCs. Only a few examples of significant
heterometallic and heteroleptic MCCs and of MCCs containing main-group elements will be
discussed. The readers can find pertinent references throughout the text. In the second part of this
review, we discuss both homogeneous catalysis and homogeneous electrocatalysis, along with the
2
Page 3 of 76 Chemical Society Reviews
relative advantages of using MCCs to promote these processes. To clarify, homogeneous refers to
the fact that the (electro)catalysts are in solution-phase along with reactants and other additives;
unless specified, "catalysis" refers to reactions requiring thermal activation while "electrocatalysis"
requires an applied potential. Heterogeneous catalysis, where the electrocatalyst is not in the same
phase as the reactants, falls outside the scope of this review; however, we recognise that MCCs
have been supported by various matrices to modify their catalytic activities, and some of that work
has been summarised in review articles or books.2324252627 For instance, Pt12 and Pt13 clusters have
been selectively synthesised within dendrimer templates and show dramatically different catalytic
activities in the oxygen reduction reaction, which occurs at the cathode of fuel cells.28 MCCs have
also been employed for the preparation of Pt clusters of controlled sizes inside metal–organic
frameworks and employed for catalysis and electrocatalysis.29
THERMAL METHODS
(a) Pyrolysis (solid)
DIRECT CARBONYLATION (b) Thermolysis (solvent)
CO
METAL
METAL METAL
CARBONYL CARBONYL
METAL SALT CO REDOX METHODS CLUSTER
OR COMPLEX Reducing agent (a) Reduction
(b) Oxidation
(c) Redox condensation
REDUCTIVE CARBONYLATION
(d) Disproportionation
CHEMICAL
MODIFICATION
OTHER CHEMICALLY
INDUCED METHODS
Scheme 1. General methods for the synthesis of MCCs.5 MCCs can be obtained from non-carbonyl
precursors by direct or reductive carbonylation. Low nuclearity metal carbonyls can be transformed
into larger MCCs through thermal and redox methods. Moreover, MCCs can be used as starting
3
Chemical Society Reviews Page 4 of 76
material in thermal, redox or other chemically induced methods for the preparation of further
MCCs.
It must be remarked that the distinction among these four categories is mainly based on the
experimental conditions adopted; from a mechanistic point of view (when a mechanism may be
devised), they are often overlapping. For instance, heating a metal carbonyl in the presence of a
base is usually classified as a thermal method, even if the mechanism may also involve redox
reactions. Syntheses of category (1) start from a non-carbonyl precursor, whereas those of
categories (2)–(4) employ carbonyl species.
Compared to C–C bonds, M–M interactions are weaker and non-directional. As a
consequence, MCCs adopt a rich variety of structures even in the case of species with similar
compositions. The products obtained are highly dependent on the experimental conditions, which,
therefore, must be very carefully controlled. A tailored synthesis of a particular structure is still a
dream. Nonetheless, it is nowadays possible to outline some general guidelines for the preparation
of MCCs with a given composition and for inducing their growth up to the nanometric level.2,3,5,6,8-
11,30 Moreover, isolated MCCs can be modified and functionalised with ancillary ligands to induce
new properties, to be anchored onto supports, or to be transformed into nanostructured
catalysts.2,3,5,6,8-11,30
Infrared (IR) monitoring of the reactions is a very useful tool. Reactions are often
accompanied by the formation of side products, such as small carbonyl complexes (e.g., Ni(CO)4,
[Co(CO)4]–, Ni(CO)3(PPh3), [Rh(CO)4]–), bulk metals, metal salts, and other MCCs. Separation and
purification of the major product is mainly achieved using solvent extraction, owing to the
solubility differences of the product and side products in water and organic solvents. In the case of
anionic MCCs, this may be enhanced by choosing suitable tetra-substituted-ammonium or
-phosphonium cations. For less charged and neutral MCCs, chromatography is sometimes
employed. The resulting yields vary case by case, but many MCCs can be obtained on the
hundreds-of-milligram or even gram scale.
Direct (reductive) carbonylation. Direct and reductive carbonylation are the major methods
for the preparation of metal carbonyls starting from non-carbonyl precursors.2,3,5 Direct
carbonylation of a finely divided metal under CO works only in a few cases, such as with Fe, Ni,
and Co. Reductive carbonylation is a more general procedure, which requires the reduction of a
metal salt under CO atmosphere. The reducing agent (H2, Na, CO itself, etc.), temperature, and CO
pressure employed depend on the metal. In some cases, harsh conditions are required, and in other
cases, the reactions may be performed at room temperature (RT) with 1 atm of CO. For instance,
4
Page 5 of 76 Chemical Society Reviews
ambient reductive carbonylation may be used for the synthesis of Rh4(CO)12,31 Ir4(CO)12,32
[Rh(CO)4]–, 33 [Ir(CO)4]–,33 and the [Pt3n(CO)6n]2– (n = 1–10).34 Even if Ru3(CO)12 may be prepared
with CO at atmospheric pressure,35 the high-pressure synthesis is often more convenient.36 The use
of CO as a reducing agent is summarised by the half-reaction (Equation 1); a suitable amount of
base is used to tune the reducing ability of CO.
Reductive carbonylation may also be used in the solid state.37 Moreover, reductive carbonylation of
Pd salts in the presence of PR3 has been extensively used by Dahl and Mednikov for the synthesis
of Pd-CO-PR3 clusters (see the ESI: Scheme S1).13
Thermal methods. CO elimination from a MCCs is an endothermic process with a
favourable entropic term; therefore, it is usually promoted by thermal methods, in the solid phase
(pyrolysis) and in solution (thermolysis). Other than decomposition, four possible events may
follow CO elimination from a MCC, with event (4) as the most common:2
1) No reaction. In very few cases, the same cluster cage can be stabilised by different numbers
of CO ligands, as in the case of [Rh12Sn(CO)27–x]4– (n = 0, 1, 2) 38 and [Ni9CoC2(CO)16–x]3–
(x = 0, 1).39
2) Permutation. Elimination of two CO ligands from the trigonal bipyramidal (TBP)
[PtRh4(CO)14]2– generates the TBP [PtRh4(CO)12]2– with the migration of Pt from an
equatorial to an apical position.40
3) Intramolecular condensation. In some cases, CO elimination induces a structural
rearrangement of the cluster with retention of the nuclearity. This is exemplified well by the
trigonal prismatic (TP) MCCs, [M6C(CO)15]2– and [Co6N(CO)15]– (M = Co, Rh), which are
transformed into octahedral (Oh) clusters, [M6C(CO)13]2– and [Co6N(CO)13]–, after
elimination of two carbonyl ligands.41 The number of M–M bonds increases from 9 (TP) to
12 (Oh), and because of this, the process may be referred to as an intramolecular
condensation where M–CO bonds are replaced by M–M bonds.
4) Intermolecular condensation. This is the most common process observed after heating a
MCC. In this case, CO elimination produces an unsaturated species, which condenses with a
second cluster fragment resulting in a higher-nuclearity MCC. Several examples will be
given in the following sections.
5
Chemical Society Reviews Page 6 of 76
Heating a MCC may also promote CO dismutation to C and CO2, resulting in carbide
MCCs.2,3,5,6,8-11,30 Thermal reactions of neutral clusters are often unselective, leading to mixtures of
products. As demonstrated by Chini, the use of an anionic cluster as starting material or the addition
of a base to a neutral carbonyl in thermal decomposition often provides better selectivity.42
Redox methods. Redox reactions used for the synthesis of MCCs include reduction,
oxidation, redox condensation, and disproportionation.2,3,5,6,8-11,30 Multivalent MCCs can undergo
reversible redox reaction without any major structural rearrangements;6 examples include [Ni22–
n– (n = 5–8),43 [Pt33(CO)38]n– (n = 0–9), and [Pt40(CO)40]n– (n = 4–11).44 In general,
xPd20+x(CO)48]
redox reactions on non-multivalent (electron-precise) MCCs or the use of excess redox reagents on
multivalent MCCs lead to irreversible processes yielding new MCCs or complete decomposition.
Reduction of a MCC can be performed by (1) direct addition of electrons, for instance, using
alkali metals; or (2) nucleophilic attack of the C-atom of a coordinated carbonyl by OH– ions. In the
latter case, the first product is a metallacarboxylic acid [MCOOH]–, which is readily transformed
into a hydridocarbonylate [MH]– upon CO2 elimination or a metallacarboxylate [MCOO]2– upon
deprotonation.45 Following further deprotonation or CO2 elimination, [MH]– and [MCOO]2– are
both transformed into the dianion [M]2–, which can be stable or undergo redox condensation to
larger MCCs (Scheme 2). This approach was developed by Hieber,15 who demonstrated that
different Fe carbonyl cluster anions could be obtained from the treatment of Fe(CO)5 with base.
This method is very versatile; it can be used for the preparation of low-nuclearity carbonylates and
hydridocarbonylates as well as larger MCCs.
+ OH– - CO2
M-CO [MCOOH]– [MH]–
- CO2
[MCOO]2– [M]2–
+ M-CO
[Mn(CO)x]z–
Scheme 2. Condensation induced by reduction. The process is initiated by OH– nucleophlic attack
on a CO ligand and eventually leads to the hydride [MH]– or the dianion [M]2–. This, in turn, can
undergo to a condensation reaction resulting in the higher nuclearity metal carbonyl [Mn(CO)x]z–.
6
Page 7 of 76 Chemical Society Reviews
When reduction of the MCC is achieved by the direct addition of electrons, the M–M and
M–CO bonds often cleave, with the resulting species undergoing fragmentation or condensation to
larger species. Alternatively, reduction can simply result in the replacement of a CO ligand with
two negative charges with or without structural rearrangement. Indeed, [M]2– may be stable or
undergo redox condensation (Scheme 2).
Oxidation of MCCs with innocent (non-coordinating) reagents, such as tropilium
tetrafluoroborate or ferrocenium ions, may result in the formation of new M–M bonds with or
without CO loss, as in the dimerisation of [Ir6(CO)15]2– to [Ir12(CO)26]2–,46 [Pt19(CO)22]4– to
[Pt38(CO)44]2–,47 or [Co6C(CO)15]2– to [Co11C2(CO)23]n– (n = 1–3).48 Using coordinating oxidants
such as H+ and [MLx]n+, the reaction may proceed through oxidation or formation of a Lewis-type
acid–base adduct. In order to increase the nuclearity of MCCs of a few metal atoms, the latter
process has been widely employed through the addition of [ML]+ fragments (M = Cu, Ag, Au; L =
neutral ligand).3,9,30 In some cases (vide infra), the reaction of MCCs with metal salts or complexes
may result in larger MCCs via redox condensation. The use of H+ ions (from acids or hydrolysis of
[M(H2O)x]n+ ions) may result in the oxidation of the MCC or the formation of isostructural MCCs
containing hydride ligands. For instance, [Rh7(CO)16]3– is oxidised by MCl2·xH2O salts (M = Ni,
Zn, Cd) to [HRh14(CO)25]3–, [Rh15(CO)27]3–, [Rh15(CO)25(MeCN)2]3–, and [Rh17(CO)37]3–.49
Although oxidation usually results in larger clusters, in a few cases oxidative degradation is
observed. For example, treatment of [H2Co20Pd16C4(CO)48]4– with strong acids affords
[HCo15Pd9C3(CO)38]2– and, eventually with excess acid, [Co13Pd3C3(CO)29]–.505152 In several cases,
polyhydride clusters may be interconverted by simple acid–base reactions. For instance, protonation
of [Co6(CO)15]2– affords [HCo6(CO)15]–,53 while stepwise protonation of [Co15Pd9C3(CO)38]3–
results in [H2Co15Pd9C3(CO)38]2–, [HCo15Pd9C3(CO)38]–, and H3Co15Pd9C3(CO)38; these reactions
may be reversed by the addition of base.
Redox condensation was introduced and rationalised by Paolo Chini, and it represents one of
the most powerful and versatile procedures for MCC preparation.54 Redox condensation involves a
comproportionation reaction between an MCC anion and a more oxidised species (not necessarily a
cluster nor even a carbonyl compound), which can be either cationic, neutral, or anionic (Scheme 3;
also see ESI: Figure S1). Redox condensation can be used for the preparation of homo- and hetero-
metallic MCCs, even when the second metal does not form a carbonyl compound, such Pd, Cu, Ag,
Au, Hg, or main-group metals.
7
Chemical Society Reviews Page 8 of 76
+ Pt(Et2S)2Cl2
[Co6C(CO)15]2- [Co8Pt4C2(CO)24]2-
+ Pd2+
[HRu3(CO)11]- [Ru6Pd6(CO)24]2-
+ [AuCl4]-
[Ni6(CO)12]2- [Ni32Au6(CO)44]6-
+ GaCl3
[Ni6(CO)12]2- [N12Ga(CO)22]3-
+ SnCl2
[Rh7(CO)16]3- [Rh12Sn(CO)27]4-
+ BiCl3
[Rh7(CO)16]3- [Rh12Bi(CO)27]3-
+ SbCl3
[Rh7(CO)16]3- [Rh21Sb2(CO)38]5-
Scheme 3. Some representative examples of redox condensation. Balanced equations have been
written only when the complete stoichiometry of all the reagents and products are known.
8
Page 9 of 76 Chemical Society Reviews
Mx(CO)y + L
(L = tetrahydrofuran, ketones, amines)
(a) (c)
[ML62+][M(CO)n-]2 (b) L
mononuclear ions M M(CO)n
L
or L
[ML6 ][Ma(CO)b2-]
2+
xenophilic M-M bond
metal cluster ion L
L
(CO)nM C O M O C M(CO)n
L L
isocarbonyl linkage
Scheme 4. Disproportionation of neutral metal carbonyls can result in (a) ionic couples, (b)
isocarbonyl adducts, or (c) xenophilic clusters (M, black: metal in formal oxidation state = 0; M,
red: formal oxidation state > 0; M, blue: formal oxidation state < 0).55
9
Chemical Society Reviews Page 10 of 76
CO
[Ni10Rh2C(CO)20]2- [Ni6Rh8(C2)2(CO)24]4-
Figure 1. CO-induced condensation of [Ni10Rh2C(CO)20]2– to [Ni6Rh8(C2)2(CO)24]4– (green, Ni;
blue, Rh; grey, C). CO ligands have been omitted for clarity.58
As a final remark, main-group elements are most often introduced into MCCs by redox
condensation using suitable EXn compounds (E = main-group element; X = halide). Chalcogens can
also be introduced using elemental O2, S8, Se, or Te, or using chalcogen-containing EOx2–
molecules under reducing conditions.2,3,5,6,8-12,30,59,60 Suitable reagents for the synthesis of nitride-
containing MCCs from metal carbonyls are NO+, NO2–, N3, and NCO–. Carbide MCCs represent the
largest class of MCCs containing a main-group element. They can be prepared by three main
methods: 2,3,5,6,8-11,30,61
1) Thermal disproportionation of CO to C and CO2. This method is very effective in the case of
Re, Fe, Ru, and Os carbonyls.
2) CO splitting by the reduction of a bridging CO ligand. In this case, a CO ligand that C-
bonded to a metal is activated by O-coordination to a Lewis acid, such as RCO+ (generated
from RCOCl). The resulting M–C–O–C(O)R is then reduced to a metal carbide following
elimination of RCOO–. This method has been applied mainly to Fe and Co clusters.
3) Reaction of MCC anions with carbon halides (e.g., CCl4, C2Cl4, C2Cl6, C3Cl6). This method
is largely employed for the synthesis of Co-, Rh-, and Ni-carbide MCCs.
In the following sections, sample syntheses of MCCs from the Groups 8–10 metals will be
illustrated. For each metal, the synthesis of MCCs that can be further employed as precursors for the
preparation of new species will be summarised with key details. These include neutral clusters,
homometallic MCCs, species containing main-group elements, and heterometallic MCCs. Some
representative (not exhaustive) examples on how these species might be employed for the
preparation of new MCCs will then be briefly outlined.
10
Page 11 of 76 Chemical Society Reviews
Fe(CO)5 can be obtained in very large scale by the direct carbonylation of Fe metal. It is the best
commercially available starting material for the preparation of other iron carbonyls (Scheme 6). The
photolysis of Fe(CO)5 in acetic acid affords the almost insoluble Fe2(CO)9,62 which was probably
the first structurally characterised carbonyl compound containing a direct M–M bond, even if
supported by bridging carbonyls. Fe3(CO)12 has been obtained by three different methods:63 (a)
thermal treatment of Fe2(CO)9, (b) oxidation of [HFe(CO)4]–, and (3) reaction of [HFe3(CO)11]–
with strong acids.
[HFe(CO)4]-
KOH
[Fe(CO)4]2- NH3/H2O Fe2(CO)9 [Fe2(CO)8]2-
Na hv in CH3COOH
Fe(CO)5
100 °C in DMF or py
80 °C in NEt3/H2O
H2SO4 H2SO4
[HFe4(CO)13]- [Fe4(CO)13]2- [HFe3(CO)11]- Fe3(CO)12
KOH KOH
H2SO4 [Fe3(CO)11]2-
[H2Fe4(CO)12]2- [HFe4(CO)12]3-
CF3SO3CH3
[HFe5(CO)14]3-
11
Chemical Society Reviews Page 12 of 76
with some species having multiple synthetic routes. Scheme 6 shows only some of the most
significant examples. Reduction of Fe(CO)5 with Na/benzophenone in THF affords the Collman's
reagent Na2[Fe(CO)4]. Depending on the specific reaction conditions, treatment of Fe(CO)5 with
OH– conveniently yields [HFe(CO)4]– or [HFe3(CO)11]–; the former is produced with NH3 in water
at RT, while the latter with NEt3 in water at 80 °C. In contrast, [Fe4(CO)13]2– is prepared in high
yields from the disproprtionation of Fe(CO)5 in either DMF or pyridine (py) at 100 °C (see ESI:
Scheme S2).
Isonuclear carbonylferrates and hydridocarbonylferrates are easily interconverted by simple
acid–base reactions. For instance, [Fe3(CO)11]2– is quantitatively protonated to [HFe3(CO)11]– by
stoichiometric amounts of strong acids, and the reaction is reversed by using 1.2-M KOH in
methanol. Interestingly, the reaction of [HFe3(CO)11]– or [Fe4(CO)13]2– with 6-M KOH results in
[HFe4(CO)12]3–.67 The unique hydride of [HFe4(CO)12]3– is 3-coordinated to a triangular face of the
cluster. This coordination is retained after the addition of two [AuPPh3]+ fragments, yielding
[HFe4(CO)12(AuPPh3)2]–;68 however, further addition of [AuPPh3]+ produces the neutral
HFe4(CO)12(AuPPh3)3, and concomitantly, the hydride migrates from the triangular face to the
tetrahedral cavity (Scheme 7).
Au(PPh3)Cl Au(PPh3)Cl
Iron Carbide Carbonyl Clusters. By heating Fe(CO)5 in quinoline at 160–180 °C, both CO
and metal disproportionation are observed, resulting in the carbide cluster [Fe6C(CO)16]2– (Equation
12
Page 13 of 76 Chemical Society Reviews
2).69 The same cluster can be alternatively obtained by heating Fe(CO)5 and [Fe(CO)4]2– in diglyme
at 160 °C.
[Fe6C(CO)16]2– is a very versatile reagent for the preparation of other Fe carbide carbonyl
clusters (Scheme 8). Its reaction with NaOH in dimethylsulfoxide (DMSO) affords the highly
reduced [Fe6C(CO)15]4–.70 Moreover, [Fe6C(CO)16]2– is oxidised by Fe3+ ions in H2O to form
Fe5C(CO)15, which, in turn, is transformed into [Fe5C(CO)14]2– after reaction with [Fe(CO)4]2–.71 In
contrast, the reaction of [Fe6C(CO)16]2– with Fe3+ in MeOH results in the 4-
(methoxycarbonyl)methylidyne cluster [Fe4(CO)12C(CO2Me)]–;72 which can be further transformed
into Fe4C(CO)13 by reaction with CF3SO3H or into [HFe4C(CO)12]– by reaction with BH3·THF.7374
The related [Fe4C(CO)12]2– dianion can be obtained using a strong base to deprotonate
[HFe4C(CO)12]– or using Na to reduce the [Fe4(CO)12(COC(O)CH3)]– adduct, which is prepared
from the reaction of [Fe4(CO)13]2– and CH3COCl.61
Fe Fe 2-
Fe Fe [Fe(CO)4]2-
Fe Fe Fe Fe 2-
C C Fe
Fe Fe Fe Fe
Fe5C(CO)15 [Fe5C(CO)14]2- Fe
[Fe4(CO)13]2-
Fe3+ in H2O
CH3COCl
O -
Fe 2- Fe 4-
160 °C NaOH C
Fe(CO)5 Fe Fe Fe Fe O CH3
quinoline C DMSO C
Fe Fe Fe Fe C
Fe Fe
Fe Fe Fe
[Fe6C(CO)16]2- [Fe6C(CO)15]4-
Fe
Fe 3+
in CH3OH [Fe4(CO)12(COC(O)CH3)]-
Na
O OCH3 -
C
C C
-
Fe Fe Fe Fe
C C 2-
Fe Fe
CF3SO3H BH3·THF NaOH Fe Fe
Fe Fe Fe Fe
Fe Fe Fe Fe
H
Fe4C(CO)13 [Fe4(CO)12C(CO2CH3)]-
[HFe4C(CO)12]- [Fe4C(CO)12]2-
Scheme 8. Synthesis of Fe carbide carbonyl clusters. All the species have been isolated and fully
characterized.61,69-74 [Fe6C(CO)16]2– is prepared directly from Fe(CO)5 by thermal treatment. Other
Fe carbide carbonyl clusters can be obtained from [Fe6C(CO)16]2– using redox reactions.
13
Chemical Society Reviews Page 14 of 76
Alternatively, [Fe4C(CO)12]2– can be prepared starting from [Fe4(CO)13]2– in a two steps reaction by
CO scission. CO ligands have been omitted for clarity.
O O -
2- O
Fe 2- Fe S O Fe S OMe Fe S
e-
RT
2-
Fe S -
Fe Fe
Fe Fe e- Fe Fe e- Fe Fe
C C S C S
Fe Fe Fe Fe Fe Fe
Fe Fe Fe
2- -
[Fe6C(CO)14(S)] [Fe6C(CO)16(S)] Fe6C(CO)16(S)
Iron carbonyl clusters containing other main-group elements. Nitride atoms can be
introduced into Fe carbonyl clusters by the reduction of NO+ ions (see ESI: Scheme S3).3 The
thermal reaction of [NO][BF4] and [Fe2(CO)8]2– in diglyme is highly dependent on the temperature;
14
Page 15 of 76 Chemical Society Reviews
it affords [Fe4N(CO)12]– at 130 °C, whereas it produces [Fe5N(CO)14]– at 145 °C. [Fe4N(CO)12]– can
be alternatively obtained from the redox condensation of Fe3(CO)12 and [Fe(CO)3(NO)]–. Both
[Fe4N(CO)12]– and [Fe5N(CO)14]– are protonated by strong acids, forming the related neutral mono-
hydrides HFe4N(CO)12 and HFe5N(CO)14. The redox condensation of [Fe4N(CO)12]– with
[Fe2(CO)8]2– results in the fully interstitial nitride cluster [Fe6N(CO)15]3–, from which an Fe atom
can be removed by oxidation, affording [Fe5N(CO)14]–.
Oxygen is by far one of the most difficult main-group elements to incorporate into low-
valent metal carbonyl clusters because of its oxidising power, hard base nature, and higher affinity
for high-valent metals (see ESI: Table S2). One of the most significant exceptions is represented by
the 3-oxo carbonyl cluster [Fe3(3-O)(CO)9]2–, which is obtained from the reaction of
[Fe3(CO)11]2– with dry air.79 Indeed, this is a rare case in which letting (dry) air into a Schlenk tube
containing an anionic metal carbonyl cluster results in the selective synthesis of a new species. It is
noteworthy that protonation of [Fe3(3-O)(CO)9]2– occurs on the O atom affording the 3-hydroxo
cluster [Fe3(3-OH)(CO)9]–, whereas [AuPPh3]+ adds to the Fe3 metal cage affording [Fe3(3-O)(3-
AuPPh3)(CO)9]– (Scheme 10).8081 This is due to the fact that the hard H+ acid prefers to add to the
hard oxo base, whereas the soft [AuPPh3]+ acid prefers to interact with the softer base site of the
metal cage of the cluster. The coordination of the oxo ligand can be further expanded by reacting
[Fe3(3-O)(CO)9]2– with metal fragments, such as [Mn(CO)3(MeCN)3]+ and Ru3(CO)10(MeCN)2,
which results in [Fe3Mn(4-O)(CO)12]–82 and [Fe2Ru3(4-O)(CO)14]2–.83
15
Chemical Society Reviews Page 16 of 76
H -
O
-
Fe Fe O
[Fe3(3-OH)(CO)9] - Fe Mn
Fe Fe
Fe
[Mn(CO)3(MeCN)3]+
H+ [Fe3Mn(4-O)(CO)12]-
O 2-
O2
[Fe3(CO)11]2- Fe Fe
Fe
[Fe3(3-O)(CO)9]2- Ru3(CO)10(MeCN)2
Au(PPh3)Cl Ru 2-
Fe Ru
O - Fe Ru
O
Fe Fe
Fe [Fe2Ru3(4-O)(CO)14]2-
Au
PPh3
[Fe3(3-O)(3-AuPPh3)(CO)9]-
Scheme 10. Reactivity of [Fe3(3-O)(CO)9]2–. This compound is obtained from the reaction of
[Fe3(CO)11]2– with dry air,79 and is protonated on the oxo ligand resulting in [Fe3(3-OH)(CO)9]–.80
The metal cage can be expanded by condensation with metal fragments such as Au(PPh3)Cl,
[Mn(CO)3(MeCN)3]+ and Ru3(CO)10(MeCN)2.81-83 CO ligands have been omitted for clarity.
Other chalcogen atoms (S, Se, Te) can be introduced into Fe carbonyl clusters by various
routes:8485
(1) Reacting Fe carbonyl anions with reagents containing the chalcogen in a positive oxidation
state, such as EO32–, EO2, and SCl2;
(2) Reacting neutral Fe carbonyls—usually Fe(CO)5 or Fe3(CO)12—with En2– polychalcogenide
anions;
(3) Reaction of Fe(CO)5 with the elemental chalcogen in the presence of a base.
Iron carbonyl clusters containing other main-group elements (e.g., As, Bi, Sb, Pb, Tl, Sn)
can be generally obtained from the reaction of iron carbonyl anions with main-group halides or
oxides (see ESI: Scheme S4).86
Heterometallic Fe-based carbonyl clusters. Heterometallic Fe–M carbonyl clusters can be
prepared by the reactions of Fe carbonyl anions with metal salts and complexes.8788 For instance,
reactions of the Collman’s reagent Na2[Fe(CO)4]·2THF with M(NHC)Cl, where M are Cu, Ag, or
16
Page 17 of 76 Chemical Society Reviews
CO CO
+ M(NHC)Cl + M(NHC)Cl OC CO
Na2[Fe(CO)4] (NHC)M Fe CO Fe
(NHC)M M(NHC)
OC CO CO
M, M' = Cu, Ag, Au
+ M'(NHC)Cl
NHC = IMes, IPr
Me
Me i CO
Pr
N Me i
Pr OC CO
IMes = N Fe
IPr =
N Me N i
(NHC)M M'(NHC)
Pr CO
Me i
Pr
Me
Scheme 11. Reactions of Na2[Fe(CO)4] with M(NHC)Cl (M = Cu, Ag, Au; NHC = IMes, IPr; IMes
= C3N2H2(C6H2Me3)2; IPr = C3N2H2(C6H3iPr2)2). One or two [M(NHC)]+ fragments as well as two
different [M(NHC)]+ and [M'(NHC)]+ fragments can be added, resulting in [Fe(CO)4{M(NHC)}]–,
Fe(CO)4{M(NHC)}2 and Fe(CO)4{M(NHC)}{M'(NHC)}, respectively.89-91
Molecular gold nanoclusters protected by Fe–CO fragments have been prepared by redox
condensation of [Fe3(CO)11]2– with [AuCl4]–. The resulting [Au22{Fe(CO)4}12]6–,
[Au21{Fe(CO)4}10]5–, [Au28{Fe(CO)3}4{Fe(CO)4}10]8–, and [Au34{Fe(CO)3}6{Fe(CO)4}8]10– clusters
are reminiscent of thiolate-protected Au nanoclusters, comprising Au cores protected by
organometallic staple motifs (Figure 2).95
17
Chemical Society Reviews Page 18 of 76
Os3(CO)10(MeCN)2 affords different clusters than reactions using Os3(CO)12 (see ESI; Scheme S5).
Moreover, reaction of Os3(CO)10(MeCN)2 with the triosmiumfuryne complex Os3(CO)9(3-2-
C4H2O)(-H)2, which contains a bridging furanyl ligand, produces the bis-triosmium complex
Os3(CO)9(-H)2(3-2-2,3--2-4,5-C4HO)Os3(CO)10(-H)2; CH activation occurs at the
uncoordinated C=C double bond of the furanyl (Scheme 13).104 Further Ru and Os carbonyl clusters
can be obtained by reduction of M3(CO)12, as well as by redox reactions on various preformed
clusters.
Os3(CO)12
Scheme 12. Synthesis of Os carbonyl clusters by pyrolysis of Os3(CO)12. The obtained products
depend on temperature and/or the addition of some water to solid Os3(CO)12 during the pyrolysis.
When obtained in mixture, different products can be separated by chromatography.98
Os3(CO)10(MeCN)2
O H
Os H
H H Os Os Os
Os Os
H Os Os H
O H Os
Os3(CO)10(MeCN)2
H + O +
O
H Os 80 °C
Os Os H Os
H Os
H Os
Os Os Os Os Os
1 H H H H
Os
2 3 4
Scheme 13. The reaction of Os3(CO)10(MeCN)2 with the triosmiumfuryne complex Os3(CO)9(3-
2-C4H2O)(-H)2. Os3(CO)9(3-2-C4H2O)(-H)2 (1); Os3(CO)9(-H)2(3-2-2,3--2-4,5-
C4HO)Os3(CO)10(-H)2 (2); Os3(CO)9(-H)2(3-2-2,3--2-4,5-C4O)Os3(CO)9(-H)2 (3);
Os3(CO)9(-H)2(3-2-3-2-C-C-C)Os3(CO)9(-H)2 (4).104 The products can be separated by
chromatography. CO ligands have been omitted for clarity.
19
Chemical Society Reviews Page 20 of 76
Ru and Os carbide carbonyl clusters. Ru and Os carbide carbonyl clusters can be obtained
through CO disproportionation at high temperature.105106107 In particular, Ru6C(CO)17 is
conveniently prepared from the thermal treatment of Ru3(CO)12 at 165 °C under 30 atm of C2H4.
Ru6C(CO)17 is quantitatively transformed into Ru5C(CO)15 after exposure to CO (80 atm).
Ru6C(CO)17 and Ru5C(CO)15 are transformed into [Ru6C(CO)16]2– and [Ru5C(CO)14]2–,
respectively, by Na2CO3 in MeOH. The reaction of Ru3(CO)12 with CaC2 affords the dicarbide
[Ru10(C)2(CO)24]2–,108 whereas the monocarbide [Ru10C(CO)24]2– is obtained from the redox
condensation of [Ru6C(CO)16]2– and Ru3(CO)12.109 In general, Ru carbide clusters may be further
transformed by CO substitution and by the addition of neutral or cationic fragments.110111
Ru and Os carbonyl clusters containing other main-group elements. The preparation of Ru
nitride carbonyl clusters relies on the addition of N3– or NO+ to preformed Ru carbonyls. The
stepwise addition of Ru(CO)2 fragments to [Ru4N(CO)12]– affords the larger [Ru5N(CO)14]–,
[Ru6N(CO)16]–, and [Ru10N(CO)24]– clusters. [Ru6N(CO)16]– is readily prepared from Ru3(CO)12
and N3– at 80 °C. Reactions of [H3M4(CO)12]– with NOBF4 afford H3Ru4N(CO)11 and
[Os4N(CO)12]–.3
Ru and Os carbonyl clusters containing S, Se, and Te may be prepared by various methods
employing the elemental chalcogen, organic chalcogen compounds, or EO32– as sources (Scheme
14).112 The reaction of Ru3(CO)12 with Na2Te2 in water at 110 °C in the presence of [PPh4]Cl
affords the Te-rich [Ru6(Te2)7(CO)12]2– cluster (see ESI: Scheme S6).113 Moreover, reactions of
Ru3(CO)12 with K2TeO3 in the presence of Cu(I) salts under a variety of experimental conditions
result in Te–Ru–Cu carbonyl clusters, which have been recently reviewed.59,60,114
20
Page 21 of 76 Chemical Society Reviews
Os3(3-S)(CO)10
PhSH CO
h, benzene
, toluene
Os6(4-S)(CO)17 Os6(3-S)(CO)19
Scheme 14. The reaction of Os3(CO)12 with PhSH (blue, Os; yellow, S; red, O; grey, C; white,
H).112 Hydrogens on organic groups have been omitted for clarity. Light or thermal treatment are
indicated by hν and Δ, respectively.
Similar procedures may be applied to other main-group elements.115 For instance, Ru–Bi
and Os–Bi carbonyl clusters may be obtained by reacting M3(CO)12, as well as other Ru and Os
carbonyls, with NaBiO3. Using THF as solvent, the reaction of [HRu3(CO)11]– with SbPh2Cl affords
Ru3(CO)10(-H)(-SbPh2), whereas using CH2Cl2 results in Ru6(CO)20(-H)2(-SbPh2)2 (see ESI:
Scheme S7).116
Six different Ru–Ge carbonyl clusters have been obtained by heating Ru3(CO)12 and
tBuGeH in heptane (Scheme 15): Ru3(CO)9(3-GetBu)2, Ru2(CO)6(-GetBuH)3, Ru4(CO)10(4-
3
21
Chemical Society Reviews Page 22 of 76
Ru3(CO)12 + tBuGeH3
Heptane 98 °C
Ru Ru
H
t
H
Bu Ge Ge H
Ru Ge tBu Ge tBu Ge
t
Bu t
Bu
Ru Ru
Ru3(CO)9(3-GetBu)2 Ru2(CO)6(-GetBuH)3
H H H H
Ge Ge
t
Bu Ge Ru t
Bu
t
Bu Ge Ru t
Bu
t t t
Bu Ge Ru Bu Ge Ge Bu
H
Ru Ge t Ru Ru
Bu
Ge t
Ru Ru Bu
t t
Ru4(CO)10(4-Ge2 Bu2)(-Ge BuH)2 Ru4(CO)8(4-Ge2 Bu2)(-GetBuH)2(3-GetBu)(H)
t
t t
Bu Bu t
Bu
Ge Ru Ge Ru Ge
t
Bu Ru Ru
Ru Ge H
H Ge tBu
Ru Ru Ru Ru
H
Ge Ge t
Ru t Ru Bu
Bu
Ru5(CO)12(3-GetBu)2(4-GetBu)(H) Ru6(CO)12(3-GetBu)4(H)2
Scheme 15. The products of the thermal reaction of Ru3(CO)12 and tBuGeH3. The different products
have been separated by thin layer chromatography (TLC).117 CO ligands have been omitted for
clarity.
By reacting carbonyls with boranes, boron may be incorporated into Ru and Os carbonyl
clusters. Heating Ru3(CO)12 in toluene at 75 °C with THF·BH3 affords the octahedral
HRu6B(CO)17, whereas the redox condensation of Ru3(CO)10(MeCN)2 and [Ru3(CO)9(B2H5)]–
yields the trigonal prismatic cluster [H2Ru6B(CO)18]–.118 The reaction of H2Os3(CO)10 with B2H6
produces the carbonylborylidyne cluster H3Os3(CO)9(BCO), which is transformed into a mixture of
HOs4(CO)12(BH2) and HOs5B(CO)12 after heating in toluene at 110 °C.119 The cluster
Os3(CO)10(MeCN)2 may be used for the cage opening of carboranes; its reaction with closo-o-
C2B10H10 yields two interconvertible isomers: Os3(CO)9(3-4,5,9-C2B10H8)(-H)2 and Os3(CO)9(3-
3,4,8-C2B10H8)(-H)2 (Figure 3).120
22
Page 23 of 76 Chemical Society Reviews
Heterometallic Ru- and Os-based carbonyl clusters. Heterometallic Ru–M and Os–M
carbonyl clusters can be obtained by redox condensation of preformed Ru and Os carbonyls with
metal salts, complexes, and carbonyls. For instance, the reaction of [Ir(CO)4]– with 1 equiv.
(equivalent(s)) of Ru3(CO)12 yields the tetrahedral cluster [IrRu3(CO)13]–, whereas employing 2
equiv. of Ru3(CO)12 results in [IrRu6(CO)23]– (Scheme 16).121 Moreover, the redox condensation of
[Os10C(CO)24]2– with an excess of [Pd(MeCN)4][BF4]2 results in the large cluster,
[Os18Pd3C2(CO)42]2– (see ESI: Figure S3).122
Ru3(CO)12 Ru3(CO)12
[Ir(CO)4]-
1 equiv 2 equiv
[IrRu3(CO)13]-
[IrRu6(CO)23]-
Scheme 16. The reactions of Ru3(CO)12 and [Ir(CO)4]– (yellow, Ir; blue, Ru; red, O; grey, C).
Depending on the stoichiometry of the reaction, [IrRu3(CO)13]– or [IrRu6(CO)23]– are formed.121
23
Chemical Society Reviews Page 24 of 76
unbridged isomers with D3d and D2d symmetry. Co-crystallization with C60 traps the unbridged D3d
isomer in the Co2(CO)8·C60 adduct.123
Homometallic cobalt carbonyl clusters. The tetrahedral Co4(CO)12 cluster is obtained after
heating Co2(CO)8 in low-polarity solvents such THF, dioxane, or iPrOH. The reaction of Co4(CO)12
with halides or pseudo-halides results in the [Co4(CO)11X]– mono-anions (X = Br, I, SCN).
Octahedral Co6(CO)16 can be prepared by oxidation of [Co6(CO)15]2– under a CO atmosphere.
Co2(CO)8 is quantitatively reduced by Na/naphthalene or Na/Hg to Na[Co(CO)4], which
contains the very versatile cobaltate carbonyl anion. Alternatively, [Co(CO)4]– can be obtained as
the [Co(MeOH)6]2+ salt by disproportionation of Co2(CO)8 in methanol at 50 °C. In both cases, a
variety of quaternary ammonium or phosphonium salts of [Co(CO)4]– are readily prepared by
metathesis of the Na+ or [Co(MeOH)6]2+ cations.
[Co(CO)4]– may be considered a pseudo-halide and forms the HCo(CO)4 hydride, which
possesses acidic character. Well-defined isocyano analogues of HCo(CO)4 have been recently
investigated.124 The reaction of [Co(CO)4]– with M+ (M = Cu, Ag, Au) or Hg2+ in a 2:1
stoichiometric ratio affords the linear [M{Co(CO)4}2]– and Hg{Co(CO)4}2 complexes reminiscent
of [MX2]– and HgX2 (X = Cl, Br, I, CN).125 Transmetalation reactions involving 1:2 molar ratio of
Ln metal (Ln = Yb, Eu) and Hg{Co(CO)4}2 in Et2O afford the isocarbonyl polymeric arrays,
[{(Et2O)3Ln[Co4(CO)11]}, which contain the reduced [Co4(CO)11]2– anion (see ESI: Figure S4).126
Co2(CO)8 undergoes different disproportionation/condensation equilibria depending on the
polarity of the solvent, the presence of base, and the application of vacuum. In very polar solvents
under nitrogen, only disproportionation to Co2+ and [Co(CO)4]– is observed (see ESI: Scheme S8).
In contrast, heating Co2(CO)8 in dry ethanol at 60 °C under vacuum affords the octahedral
[Co6(CO)15]2– cluster as the result of the partial condensation between Co2+ and [Co(CO)4]–. The
stoichiometric reaction between Co2(CO)8 and pyridine in hexane results in the triangular
[Co3(CO)10]– anion.127
[Co6(CO)15]2– is protonated by strong acids to [HCo6(CO)15]–, which contains a fully
interstitial hydride at the centre of the octahedral cage of the cluster. Reduction of [Co6(CO)15]2–
with alkali metals affords the [Co6(CO)14]4– tetra-anion (Scheme 17).
24
Page 25 of 76 Chemical Society Reviews
H+ Na
Cobalt carbide carbonyl clusters. Several homometallic Co carbide carbonyl clusters are
known, and most of them can be prepared starting from the trigonal prismatic [Co6C(CO)15]2–
cluster128 via thermal or redox reactions (Scheme 18).48 In addition, Co3(3-CCl)(CO)9 is obtained
by refluxing Co2(CO)8 in CCl4.
[Co11(C2)(CO)22]3–
[Co6C(CO)13]2– Ox
[Co13C2(CO)24]4– [Co13C2(CO)24]3–
Red Ox
Ox [Co10(C2)(CO)21]2–
[Co8C(CO)17]4–
Red
[Co6C(CO)14]– [Co9(C2)(CO)19]2–
[Co6C(CO)12]3–
Ox
Red Ox
[Co11C2(CO)23]3– [Co11C2(CO)23]2– [Co11C2(CO)23]–
Scheme 18. Synthesis of Co carbide carbonyl clusters (Ox = oxidation; Red = reduction; =
thermal reaction). Several homometallic Co carbide carbonyl clusters are prepared by thermal or
redox reactions starting from [Co6C(CO)15]2–.48 See text for details.
25
Chemical Society Reviews Page 26 of 76
[Co6C(CO)15]2– can be prepared from the reaction of [Co(CO)4]– with CCl4 at a 6:1 molar
ratio or from [Co(CO)4]– and Co3(3-CCl)(CO)9 at a 3:1 molar ratio. By performing the latter
reaction at a 2:1 molar ratio, the paramagnetic [Co6C(CO)14]– cluster is obtained. [Co6C(CO)14]– can
be alternatively obtained by mild oxidation of [Co6C(CO)15]2– or from the reaction of [Co6(CO)15]2–
with MeCOCl via CO scission induced by the addition of MeCO+ to a coordinated CO ligand. In
the presence of even weak bases, [Co6C(CO)14]– is converted to [Co6C(CO)15]2–.
Gently heating [Co6C(CO)15]2– in THF results in an intramolecular condensation to yield
[Co6C(CO)13]2– via the loss of two CO ligands and the rearrangement of the Co6 cage from trigonal
prismatic to octahedral geometry. Under more severe thermal conditions, larger clusters such as
[Co13C2(CO)24]4– are obtained. The stepwise reduction of [Co6C(CO)15]2– with Na/naphthalene
affords, in sequence, the highly reduced clusters [Co7C(CO)15]3– and [Co8C(CO)17]4–, and
eventually, [Co6C(CO)12]3–.48
Several heterometallic Co–M carbide carbonyl clusters have been reported. Although they
will not be reviewed in detail, some general strategies for their synthesis, along with pertinent
examples, are listed below:
1) Reaction of Co3(3-CCl)(CO)9 with metal carbonyl complexes or clusters. Co3(3-
CCl)(CO)9 reacts with [Rh(CO)4]– yielding [Co2Rh4C(CO)13]2–,129 whereas it reacts with
Mo(CO)3(MeCN)3 producing [Mo3Co3C(CO)18]–.130 Interestingly, [Co3Ni9C(CO)20]3– results
from the addition of Co3(3-CCl)(CO)9 to [Ni6(CO)12]2–,131 whereas a different product,
[Co6Ni2C2(CO)16]2–, is obtained by doing the opposite—adding [Ni6(CO)12]2– to Co3(3-
CCl)(CO)9.132 C-C coupling and redox condensation are observed in the reaction of Co3(3-
CCl)(CO)9 with [Ni9C(CO)17]2–, which affords the [Co3Ni7(C2)(CO)15]3– mono-acetylide.133
The Ni–Co hexacarbide carbonyl cluster [H6–nNi22Co6C6(CO)36]n– (n = 3–6) has been
obtained by the redox condensation of [Ni10(C2)(CO)16]2– and Co3(3-CCl)(CO)9 (Scheme
19). Thermolysis of [H2Ni22Co6C6(CO)36]4– in THF affords the larger octa-carbide
[Ni36Co8C8(CO)48]6–.39,134
26
Page 27 of 76 Chemical Society Reviews
Co3(3-CCl)(CO)9 [Ni10(C2)(CO)16]2-
[Ni36Co8C8(CO)48]6-
Scheme 19. Synthesis of [H2Ni22Co6C6(CO)36]4– and [Ni36Co8C8(CO)48]6– (green, Ni; blue, Co;
yellow, Cl; red, O; grey, C). Redox condensation of [Ni10(C2)(CO)16]2– and Co3(3-CCl)(CO)9
affords [H2Ni22Co6C6(CO)36]4–, which is thermally converted into [Ni36Co8C8(CO)48]6–.39,134
2) Redox condensation involving a Co carbide carbonyl cluster and a metal complex, salt or
cluster. Some recent examples are represented by the reactions of [Co6C(CO)15]2– with
Pt(Et2S)2Cl2 and with Pd(Et2S)2Cl2, which result in the formation of [Co8Pt4C2(CO)24]2– 135
and [H6–nCo20Pd16C4(CO)48]n– (n = 3–6), respectively (Scheme 20).50-52
27
Chemical Society Reviews Page 28 of 76
Pd(Et2S)2Cl2 HBF4
[Co6C(CO)15]2-
CH2Cl2
[H6-nCo20Pd16C4(CO)48]n- [H3-nCo15Pd9C3(CO)38]n-
(n = 3-6) (n = 2, 3)
HBF4
CH2Cl2
HBF4 excess
CH2Cl2
[Co13Pd3C3(CO)29]- [H3-nCo15Pd9C3(CO)38]n-
(n = 0, 1)
Scheme 20. Synthesis and reactivity of [H6–nCo20Pd16C4(CO)48]n– (n = 3–6) (orange, Pd; blue, Co;
red, O, grey, C). This compound is obtained from the redox condensation of [Co6C(CO)15]2– and
Pd(Et2S)2Cl2. [H6–nCo20Pd16C4(CO)48]n– (n = 3–6) is transformed into [H3–nCo15Pd96C3(CO)38]n– (n
= 0–3) and, eventually, [Co13Pd3C3(CO)29]– upon reaction with increasing amounts of strong
acids.50-52
3) Addition of cationic metal fragments to an anionic Co carbide carbonyl cluster anion with
the formation of a Lewis-type acid–base adduct. This method is represented well by the
reaction of Au(PPh3)Cl with [Co6C(CO)15]2–, which affords several Co carbide carbonyl
clusters decorated by AuPPh3 fragments, depending on the experimental conditions (see
ESI: Scheme S9).136
29
Chemical Society Reviews Page 30 of 76
affording Rh6(CO)16. Meanwhile, the analogous Ir6(CO)16 is conveniently obtained from the
oxidation of [Ir6(CO)15]2–.142
Homometallic Rh and Ir carbonyl clusters. Neutral Rh and Ir carbonyls are suitable
precursors for the synthesis of larger clusters (see ESI: Tables S6 and S7). Moreover, CO ligands
can be replaced by phosphines or other soft nucleophiles, resulting in heteroleptic clusters.
Rh4(CO)12 and Rh6(CO)16 disproportionate in the presence of bases, such as amines. For instance,
the ionic couple [Rh(CO)2(py)]+[Rh5(CO)13(py)2]– is formed in pyridine.143 The most powerful
technique for obtaining anionic Rh and Ir carbonyl clusters involves the reduction of M4(CO)12 (M
= Rh, Ir).144 Several different species may be obtained, depending on the nature (alkali metals or
alkali hydroxides) and amount of reducing agent, the temperature, and the reaction atmosphere (N2,
CO, H2, vacuum).
The reduction of Rh4(CO)12 with increasing amounts of NaOH or KOH under CO at
atmospheric pressure affords [Rh12(CO)30]2–, [Rh6(CO)15]2–, [Rh7(CO)16]3–, [Rh4(CO)11]2–, and
[Rh(CO)4]–. Again with increasing base, the reduction of the analogous Ir4(CO)12 produces
[HIr4(CO)11]–, [Ir4(CO)11]2–, [HIr4(CO)10]3–, [Ir4(CO)10]4–, and [Ir(CO)4]– (see ESI: Scheme S10).33
Careful dosage of NaOH during thermolysis of Rh4(CO)12 under different atmospheres enables the
more or less selective preparation of several high-nuclearity Rh carbonyl clusters (see ESI: Scheme
S11), including [H5-nRh13(CO)25]n– (n = 2, 3), [Rh22(CO)37]4–, [H8-nRh22(CO)35]n– (n = 4, 5),
[Rh19(CO)31]5–, and [Rh33(CO)47]5– (Figure 5).145146 Similarly, the base-induced condensation of
Ir4(CO)12 may lead to products such as [Ir8(CO)22]2–, [Ir9(CO)20]3–, [HIr9(CO)19]4–, and
[Ir10(CO)21]2–, depending on the solvent and reaction conditions.147 Moreover, [Ir6(CO)15]2– may be
obtained from the reductive carbonylation of K2IrCl6 in 2-methoxy-ethanol/water or the redox
condensation of Ir4(CO)12 and [Ir(CO)4]–; this is a valuable precursor for the preparation of larger Ir
clusters. For instance, oxidation of [Ir6(CO)15]2– with [Cp2Fe]+ (Cp = cyclopentadienyl) affords
[Ir14(CO)27]– (Figure 5), which is the largest homometallic Ir carbonyl cluster reported to date.148
30
Page 31 of 76 Chemical Society Reviews
(a) (b)
Figure 5. Molecular structures of (a) [Rh33(CO)47]5– and (b) [Ir14(CO)27]–, the largest homometallic
homoleptic Rh and Ir carbonyl clusters (yellow, Ir; blue, Rh; red, O; grey, C).145-147
The robustness of the Rh–Rh and Ir–Ir bonds also allows the isolation of unusual species,
such as the -1-1-peroxo Ir4(CO)5(PR3)3(O2)2 149 and the Rh6(CO)12(3-GaCp*)4 and Rh6(CO)16-
x(3-InCp*)x (x = 1, 2) derivatives containing ECp* ligands (E = Ga, In; Cp* =
pentamethylcyclopentadienyl) (Figure 6).150
(a) (b)
Figure 6. Molecular structures of (a) Rh6(CO)12(3-GaCp*)4 and (b) Rh6(CO)15(3-InCp*) (blue,
Rh; yellow, Ga (a) or In (b); red, O; grey, C; white, H).150 These compounds have been obtained
upon reactions of Rh6(CO)16 with GaCp* and Rh6(CO)15(MeCN) with InCp*, respectively (Cp* =
pentamethylcyclopentadienyl).
contacts, rather than allowing interstitial heteroatoms to replace them with Ir–E bonds. Indeed, as
also observed for Pt, a few heteronuclear Ir carbonyl clusters have been reported, but they contain
an iridium core with the heteroatoms on the surface.151152 An exciting advancement in this field is
the recent discovery of the 15-vertex deltahedral cluster [(3-Bi3)2(IrCO)6(4-Bi)3]3–, which is
obtained from Ir(CO)2(acac) (acac = acetylacetonate) and K5Bi4 (Scheme 21).153
Ir(CO)2(acac) + K5Bi4
[(3-Bi3)2(IrCO)6(4-Bi)3]3-
Scheme 21. Synthesis of the 15-vertex deltahedral cluster [(3-Bi3)2(IrCO)6(4-Bi)3]3– (yellow, Ir;
orange, Bi; red, O; grey, C).153
A general strategy for the synthesis of Rh carbonyl clusters containing heavy main-group
elements such as Ge, Sn, Sb, and Bi is represented by the reaction of [Rh7(CO)16]3– with EXn
compounds.11,38,154155 Conversely, lighter elements, P and As, have been introduced into
[Rh10As(CO)22]3–,156 [Rh10P(CO)22]3–, and [Rh9P(CO)21]2–157 by reductive carbonylation of Rh(I)
complexes at high pressure and temperature in the presence of EPh3 (E = P, As). [Rh10S(CO)22]2–
was obtained from Rh4(CO)12 and SCN–.158
[Rh6C(CO)15]2– may be obtained from the reductive carbonylation of RhCl3 at atmospheric
pressure and RT in MeOH in the presence of CHCl3, or from the reaction of [Rh(CO)4]– and
CCl4.159 Refluxing the trigonal prismatic [Rh6C(CO)15]2– in iPrOH generates the octahedral
[Rh6C(CO)13]2– by the removal of two CO ligands.41 The self-assembly of [Rh6C(CO)15]2– with Ag+
ions affords oligomeric and polymeric species, such as [Ag{Rh6C(CO)15}2]3– and
[{AgOC4H8{Rh6C(CO)15}AgOC4H8}pyz] (pyz = pyrazine).160 Oxidation of [Rh6C(CO)15]2– with
Fe3+ ions results in Rh12C2(CO)25, whereas heating [Rh6C(CO)15]2– at 70 °C in the presence of
H2SO4 produces [Rh12C2(CO)24]2–. This latter cluster may be further transformed into
[Rh12C2(CO)23]4– by treatment with alkali hydroxides or into [Rh12C2(CO)20{Au(PPh3)}4] and
[Rh12C2(CO)18{Au(PPh3)}4] after reaction with Au(PPh3)Cl. In turn, several multivalent clusters
based on the Rh10(C)2Au4-6 framework may be obtained via the reduction of
[Rh12C2(CO)20{Au(PPh3)}4].161
32
Page 33 of 76 Chemical Society Reviews
The chemistry of Rh nitride carbonyl clusters is quite rich. [Rh6N(CO)15]3– can be prepared
from the reaction of [Rh6(CO)15]2– with NOBF4, or by treating [Rh7CO)16]3– with base under
NO/CO atmosphere.3 [Rh6N(CO)15]3– undergoes nucleophilic attack when reacted with OH–,
affording the hydride [HRh6N(CO)14]3–.162 Thermal treatment of [Rh6N(CO)15]3– under different
experimental conditions (e.g., temperature, solvent, time, pH) results in larger nitride and
polynitride clusters, such as [HRh12N2(CO)23]3–, [HRh12N2(CO)24]–, [Rh12N2(CO)24]2–,
[Rh14N2(CO)25]2–, [Rh23N4(CO)38]3–, and [H6-nRh28N4(CO)41]n– (n = 4, 5).163 Redox condensation of
[Rh6N(CO)15]3– with [PtRh4(CO)14]2– and with [M(CO)4]– (M = Co, Ir) results in the heterometallic
nitrides [PtRh10N(CO)21]3– and [MRh6N(CO)15]2–, respectively.164
Anionic Rh and Ir carbonyl clusters may be used in combination with metal salts and
complexes for the preparation of heterometallic clusters via redox condensation.
33
Chemical Society Reviews Page 34 of 76
+ NaOH +H2O
Ni(CO)4 [Ni5(CO)12]2- [Ni6(CO)12]2-
+ CO
+H+
+NH4Cl
+H3PO4 +H2O
[Ni9(CO)18]2-
+H2O
+H+
+Base +Base
[H2Ni12(CO)21]2- +
[HNi12(CO)21]3- [Ni12(CO)21]4-
+H +H+
Scheme 22. Synthesis of homoleptic homometallic Ni carbonyl clusters (green, Ni; red, O; grey, C).
The reaction of Ni(CO)4 with NaOH in DMF or DMSO affords [Ni5(CO)12]2–, whose hydrolysis
results in [Ni6(CO)12]2–, [HNi12(CO)21]3– or [H2Ni12(CO)21]2– depending on the pH.166,167
The highest nuclearities, which are obtained for Ni–Pd and Ni–Pt clusters, can exceed 40 metal
atoms.
(a) (b)
(c)
Figure 7. The molecular structures of (a) [Ni22-xPd20+x(CO)48]6– (x = 0.62) (green, Ni; purple, Pd;
yellow, Ni/Pd 16:84; blue, Ni/Pd 85:15) (b) [Ni32Au6(CO)44]6– (green, Ni; yellow, Au), (c)
[Ni32Pt24(CO)56]n– (green, Ni; purple, Pt; blue, Ni/Pt 66:34). In all structures, grey and red
represent C and O, respectively. These are representative high nuclearity heterometallic Ni-M
carbonyl clusters obtained from redox condensation of [Ni6(CO)12]2– with metal salts and
complexes.170-173
Nickel carbonyl clusters containing main-group elements. A similar trend has been
observed with post-transition metals; redox condensation yielding heteronuclear Ni–E clusters has
been observed with Ga, Ge, Sn, Sb, and Bi, whereas Lewis-type acid–base adducts are found for In,
which behaves like Cd. In addition, non-metallic main-group elements, such as C, P, As, Se, and
Te, can be included in the metal cage of Ni carbonyl clusters.
Several Ni carbide carbonyl clusters may be obtained from reactions of [Ni6(CO)12]2– with
halocarbons (CCl4, C2Cl4, C2Cl6, C3Cl6, C4Cl6); they contain isolated C atoms or tightly bonded C2
35
Chemical Society Reviews Page 36 of 76
units (see ESI: Table S10). Ni carbide carbonyl anions can be further reacted with metal salts or
complexes, affording heterometallic Ni–M carbide carbonyl clusters.174175176177178 Alternatively,
such compounds can be prepared starting from a preformed M carbide carbonyl and a Ni (carbonyl
or non-carbonyl) compound.
The reaction of [Ni6(CO)12]2– with GaCl3 in CH2Cl2 under nitrogen atmosphere results in a
mixture of [Ni12+xGa(CO)22+x]3– (x = 0–3) clusters, which is transformed into [Ni12Ga(CO)22]3– after
exposure to CO atmosphere (see ESI: Figure S5).179 Several Ni–Sb carbonyl clusters can be
obtained by the redox condensation of [Ni6(CO)12]2– with SbCl3 using various experimental
conditions (see ESI: Figure S5).180 Moisture must be avoided in order to prevent hydrolysis of the
ECl3 reactant. Indeed, anhydrous PCl3 and POCl3 are required for obtaining the Ni–P carbonyl
clusters [Ni11P(CO)18]3–, [Ni14P2(CO)22]2–, [Ni22-xP2(CO)29-x]4– (x = 0.84), [Ni22P6(CO)30]2–, [Ni23-
4– (x = 0.84), [H6–nNi31P4(CO)39]n– (n = 4, 5), and [Ni39P3(CO)44]6– (Figure 8).181
xP2(CO)30-x]
(a) (b)
(c) (d)
Figure 8. The molecular structures of (a) [Ni14P2(CO)22]2–, (b) [Ni22P6(CO)30]2–, (c) [H6–
n– (n = 4 and 5), and (d) [Ni39P3(CO)44]6– (green, Ni; purple, P; grey, C; red, O).181
nNi31P4(CO)39]
These have been obtained from the reactions of [Ni6(CO)12]2– with PCl3 or POCl3 under different
experimental conditions (e.g., stoichiometry, solvent).
36
Page 37 of 76 Chemical Society Reviews
Na2PtCl6
+ NaOH or CH3COONa CO (1 atm) / CH3OH / RT
OX
OX
RED
RED OX
OX
RED
OX
RED
RED OX
RED
Figure 9. Synthesis of Chini clusters, [Pt3n(CO)6n]2– (n = 2–8), using reductive carbonylation of
Na2PtCl6 and their redox interconversion (purple, Pt; red, O; grey, C).8,10,182,183,184 OX and RED
represent oxidation and reduction, respectively.
The reactions of Chini clusters may be grouped into two main categories, depending on
whether the resulting clusters do (A) or do not (B) retain the trigonal prismatic structure of the
parent species (Table 1 and Figure 10).
37
Chemical Society Reviews Page 38 of 76
Redox reactions of Chini clusters (A-1). Chini clusters may be easily and reversibly
interconverted by means of redox reactions (A-1), with full retention of their trigonal prismatic
structures. Reduction of [Pt3n(CO)6n]2– under CO atmosphere results in the gradual decrease of
cluster nuclearity, whereas oxidation leads to higher nuclearities (Figure 9). The oxidation process
eventually results in insoluble, infinite, and conductive molecular Pt carbonyl wires composed of
continuous stacks of Pt3(-CO)3(CO)3 units.182-184
Reactions of Chini clusters with soft nucleophiles: Non-redox substitution (A-2) and
redox fragmentation (B-1). The reactions of Chini clusters with soft nucleophiles may result in
heteroleptic Chini clusters via non-redox substitution (A-2) or in zero-valent neutral Pt–(CO)–L
species via redox fragmentation (B-1), where L is a phosphine-based ligand (Figure 11). The
outcome of the reaction depends on the nature of the nucleophile, the nuclearity of the Chini cluster,
and the stoichiometry of the reaction. In the case of phosphines, non-redox substitution is favoured
by the stoichiometric addition of the phosphine to lower-nuclearity Chini clusters, whereas redox
fragmentation is observed for larger clusters and in the presence of excess of ligand.185186187
38
Page 39 of 76 Chemical Society Reviews
Pt4(CO)4(P^P)2 [Pt12(CO)22(PPh3)2]2-
L = P^P
L = PPh3
L = dppm +L +L L = PTA
L = dppBz
CO L = dppe
Pt6(CO)6(dppm)3 [Pt12(CO)24]2-
[Pt12(CO)20(PTA)4]2-
[Pt9(CO)18]2-
[Pt12(CO)20(dppe)2]2-
Pt(dppBz)2
Figure 11. Reactions of Chini clusters with phosphines (L) include redox fragmentation (B-1) and
non-redox substitution (A-2). P^P = H2C=C(PPh2)2; dppm = CH2(PPh2)2; dppBz = 1,2-C6H4(PPh2)2;
PTA = C6H12N3P; dppe = Ph2PCH2CH2PPh2. In the examples reported in this Figure, 1-4 CO
ligands of [Pt12(CO)24]2– can be substituted with phosphines leading to heteroleptic Chini-type
clusters via non-redox substitution.185-187 Alternatively, redox fragmentation leads to the more
reduced homoleptic [Pt9(CO)18]2– cluster and neutral complexes containing Pt(0).
Formation of Lewis acid–base adducts based on Pt Chini clusters (A-3). The external
triangular faces of Chini clusters have been predicted to behave as Lewis bases. Nonetheless, this
feature has only been exploited in two cases, both of which produce Lewis acid–base adducts (A-3).
The reaction of [Pt9(CO)18]2– with CdCl2 affords 1-D {[Pt9(CO)18(3-CdCl2)2]2–}∞ superwires,188
and the reactions of [Pt3n(CO)6n]2– (n = 2, 3) with Ag(IPr)Cl (IPr = C3N2H2(C6H3iPr2)2) result in the
neutral adducts [Pt3n(CO)6n(AgIPr)2] (see ESI: Figure S6).189 Otherwise, the reactions of Chini
clusters with other Lewis acids proceed via oxidation to [Pt3(n+1)(CO)6(n+1)]2– species (A-1) or
formation of surface-decorated clusters (B-3).
Surface-decorated Pt carbonyl clusters (B-3). This reaction path may be considered as an
oxidation accompanied by the loss of CO ligands, resulting in the formation of a Ptn(CO)m core with
MXx fragments decorating its surface. For instance, [Pt3n(CO)6n]2– (n = 2–5) clusters react with
SnCl2 resulting in [Pt8(CO)10(SnCl2)4]2–, [Pt5(CO)5{Cl2Sn(OR)SnCl2}3]3– (R = H, Me, Et, iPr),
39
Chemical Society Reviews Page 40 of 76
(d) (e)
Figure 12. Some examples of surface-decorated Pt carbonyl clusters produced from Chini clusters
via reaction pathway B-3: (a) [Pt6(CO)6(SnCl2)2(SnCl3)4]4–, (b) [Pt8(CO)10(SnCl2)4]2–, (c)
[Pt10(CO)14{Cl2Sn(OH)SnCl2}2]2–, (d) [Pt13(CO)12Cd10Br14(DMF)6]2–, (e) [H2Pt26(CO)20(CdBr)12]8–
(purple, Pt; yellow, Cd; orange, (a-c) Sn and (d,e) Br; blue, N; red, O; grey, C; white, H).190-193
Thermal reactions of Chini clusters (B-2). Thermal treatment of Chini clusters under
controlled conditions causes CO loss and condensation via the formation of additional Pt-Pt bonds
(B-2), eventually leading to Pt–CO globular molecular nanoclusters, often referred to as Pt browns
due to their colour in solution.47194195 The full list of homoleptic homometallic Pt carbonyl clusters
(including Pt brown) that are structurally characterised to date is available in the ESI (Table S11);
some representative examples are given in Figure 13.
40
Page 41 of 76 Chemical Society Reviews
Figure 13. The molecular structures of some platinum browns: (a) [Pt14(CO)18]4– (bcc), (b)
[Pt40(CO)40]6– (bcc), (c) [Pt19(CO)22]4– (pp), (d) [Pt26(CO)32]– (hcp), (e) [Pt38(CO)44]2– (ccp) and (f)
[Pt44(CO)45]n– (ccp/hcp) (purple, Pt; red, O; grey, C). The structure of the metal core of the cluster is
given in parentheses: bcc, body-centred cubic; pp, pentagonal prismatic; hcp, hexagonal close
packed; ccp, cubic close packed. All these clusters have been obtained from thermal decomposition
(B-2) of Chini clusters (see ref. 194, 195 and Table S11 in ESI).
Heterometallic clusters (B-4). Heteronuclear Pt–M carbonyl clusters can be prepared by the
redox condensation of Chini clusters with metal salts, complexes, or carbonyls (B-4). Alternatively,
as described in the previous sections, such Pt–M clusters can be obtained from the redox
condensation of carbonyl anions of the second metal and Pt salts or complexes. For instance,
[Co8Pt4C2(CO)24]2– can be synthesised from [Co6C(CO)15]2– and [Pt6(CO)12]2– or more
conveniently, from [Co6C(CO)15]2– and Pt(Et2S)2Cl2.135
for both the homogeneous catalysis and electrocatalysis promoted by MCCs is the ability of
substrates, CO ligands, and hydride ligands or protons to migrate around the metal–metal bonds of
the cluster core. In most cases, careful probes for the formation of nanoparticles, using Hg
poisoning196 or measurements of solution dispersivity,197,198 are required to ensure that nanoparticle
side-products are not responsible for observed reactivity.
The geometric and electronic structures of MCCs (Figure 14) fall between those of single-
site metal coordination complexes (molecular catalysts) and those of extended solids
(heterogeneous catalysts). When discussing heterogeneous materials, we use a band diagram as a
model for the energy levels that make up the valence band and conduction band. When we talk
about molecular catalysts, we refer to a molecular orbital diagram. However, as molecules become
larger as in the case of clusters, their molecular orbitals become closer in energy and approach the
band-diagram model. The size and electronic structure of both metal clusters and nanomaterials are
found between those of single-site metal coordination complexes and of heterogeneous materials.
When MCCs are large enough, it has been demonstrated that their properties—such as diffusion
coefficients in solutions and even their size—more closely resemble those of a nanoparticle,
quantum dot, or fullerene than that of a molecule. Many of the more convincing links between the
structure and reactivity of MCCs with those of nanomaterials have been elucidated from
comparisons of their electrochemical properties, e.g., electron-transfer rate constants. Thus, one
motivation for studying MCCs in thermally driven catalysis and electrocatalysis is that we can
characterise their properties using the tools of molecular chemistry, which include single-crystal X-
ray diffraction (SC-XRD), NMR, and IR spectroscopy; these provide atom-level insights into
reactivity, which can also inform our understanding of the reactivity and properties of
nanomaterials.
42
Page 43 of 76 Chemical Society Reviews
Figure 14. Relative energy spacings in the electronic structures of a single-site metal coordination
complex (molecular catalyst), MCC, nanomaterial, and extended solid (heterogeneous material or
electrocatalysts). MCC = metal carbonyl cluster. Reproduced with permission from Ref. 238.
Unique to the homogeneous catalysis of MCCs, the MCC in many cases—but not all—
serves as a catalyst precursor, and therefore, debates regarding the active catalyst remain in many
examples. Fragmentation of the cluster to form a mononuclear complex is one possibility when
ligands are added as co-catalyst and when they include protons, phosphine, silane, and MeCN.
While this article focuses on carbonyl clusters, clusters where one or two of the carbonyl ligands are
replaced by another ligand, such as phosphine or amine, are also discussed. Fragmentation of
clusters was well-characterised in some mechanistic proposals for homogeneous catalysis with
MCCs,199 and it is consistent with the fact that most reactions catalysed by MCCs are thermally
initiated. In the following discussion of homogeneous catalysis by MCCs, we include clusters of
Groups 8, 9, and 10 that are accompanied by various additives used to initiate the catalytic reaction.
Many examples of stoichiometric reaction chemistry have been explored on the MCC framework,
and a review of that work falls outside the scope of this manuscript.77,78,200 MCCs have also served
as precursors to heterogeneous catalysts, which also falls outside the scope of this review article.
Electrocatalytic reactions using MCCs has seen increased interest since about 2011, and
efforts have focused on reduction chemistry—primarily hydride-transfer chemistry—using Fe and
Co clusters containing interstitial atoms. These clusters represent some of the most stable known
species in the transition series. Consequently, their product profiles and the mechanistic details of
their catalysis have been thoroughly characterised, primarily employing the electrochemical
technique of cyclic voltammetry (CV). The two-electron reduction of two protons to yield H2 is the
simplest reaction performed by Fe and Co MCCs. This simple reaction has enabled a mechanistic
study for understanding hydride formation and hydride-transfer reactivity as mediated by MCCs.
The resulting insights have made it possible to tailor hydride transfer to other substrates, such as
CO2 for formate production.
43
Chemical Society Reviews Page 44 of 76
containing headspaces have also been shown to enhance MCC stability in electrocatalytic reactions
promoted by MCCs (vide infra). Pettit et al. made a comparison of various metal (Fe, Rh, Ru, Os,
Ir, and Pt) carbonyl complexes and clusters as catalysts for the hydroformylation reaction and the
water-gas shift reaction (Scheme 23). In one of the simpler proposed mechanisms for syngas
reactivity, CO is replaced by a hydride to initiate hydride-transfer chemistry. In the foregoing
paragraphs, we give more specific details of homogeneous catalysis by MCCs, organizing the
reactions by group, i.e., Group 8, 9, and 10.
CO2 CO2
-
[HMx(CO)y-1] OH- OH-
[HMx(CO)y-1]-
RCH=CH
2
CO
[HMx(CO)y]- CO
O
H2 O -
OH- CO RCH2CH2C[Mx(CO)y-1]
RCH2CH2CHO H2 O
H2+OH-
Catalysis by MCCs of Group 8 (Fe, Ru, Os). A variety of MCC structure types were
explored for catalytic applications in the early development of MCC chemistry, and much of the
work focused on Group 8 compounds due to their increased stability relative to MCCs with other
transition elements (Chart 1, Table 2).201,202,203,204,205 Casey and coworkers have shown catalytic
performances of the iron cluster Fe3(CO)12 on the isomerization reaction of the alkene in the 1970s;
more specifically, Fe3(CO)12 catalysed the isomerization of the ethyl pentene substrates. Deuterium-
labeling studies revealed that the olefin isomerization reaction occurs via intramolecular hydrogen
shift.206 Geoffroy and coworkers utilised the radical state of the iron cluster, [Fe3(CO)11]−•, to
perform the reduction of nitrobenzene to aniline. The Fe radical anion was produced from the
disproportionation reaction of the halide anion with Fe3(CO)12. While the reaction is stoichiometric,
PhNO2 convert to the azo-azoxybenzene product.65
In more recent examples using MCCs as catalysts, additives are often used, and their role in
the chemistry varies. Nagashima and coworkers207 reported hydrosilylation of tertiary amides to
amine functional groups with Fe3(CO)12 as the catalyst and with 1,1,3,3-tetramethyl disiloxane
(TMDS) or polymethyl-hydro siloxane (PMHS) as the reducing agent. Nearly concurrently, Beller
44
Page 45 of 76 Chemical Society Reviews
and coworkers demonstrated the catalytic activity of iron- and ruthenium-based MCCs toward
hydrosilylation chemistry.208,209,210 In a report by Enthaler and coworkers, a combination of
Fe3(CO)12 and stoichiometric amounts of silyl reagent results in the activation of a sulfoxide,
yielding sulfide product.211
M M -
H CO
M M
M M
M3(CO)12 [HM3(CO)11]-
M = Fe, Ru, Os M = Fe, Ru, Os
Ru Ru
H H H H
Ru Ru Ru Ru
Ru H Ru H
H2Ru4CO13 H4Ru4CO12
Chart 1. Homogeneous MCC Catalysts from Group 8. Terminal CO ligands have been omitted for
clarity.
45
Chemical Society Reviews Page 46 of 76
The transformation of amide functional groups into nitriles has been explored using the
hydrogenated iron carbonyl cluster, [Et3NH][HFe3(CO)11]. Using various silyl additives such as
Ph2SiH2, (EtO)3SiH, and (EtO)2MeSiH, the amide group is transformed into a nitrile group within
3 h with above 90% yield. The same transformation was achieved in 97% yield using Fe2(CO)9
although the reaction time was longer (20 h);213 83% yield was obtained using Fe3(CO)12 with a 2-h
reaction time. In general, the amide substrates employed in this work were primary carboxylic
amides, which include alkyl, aromatic, heteroaromatic, and aliphatic amides. Reactions with
activated phenyl carboxylic amides achieved completion in 4 h, whereas unactivated substrates took
30 h.
Asymmetric transfer hydrogenation (ATH) converts a ketone or amine into an alcohol or
imine, respectively. Various iron-containing catalysts, including mononuclear catalysts and the iron
carbonyl clusters [Et3NH][HFe3(CO)11] and Fe3(CO)12, show high yields and selectivity in ATH;
however, there are significant differences in reaction outcomes depending on the cluster employed,
as described below. The hydrogenation of imines was successfully achieved and described by Beller
and coworkers using systems with chiral ligands also acting as co-catalysts.210 The scope of the
imine substrates included a ketamine, aromatic imine, and hetero-aromatic imine, while the co-
substrates for ATH included iPrOH and chiral P2N2 ligands, which are tetradentate ligating at two P
and two N centres. Since these observations were made under conditions employing chiral ligand
additives, it was initially inferred that the chiral ligand likely promoted the formation of a chiral
mono-iron coordination complex, which serves as the active catalyst when used with Fe3(CO)12.
However, a recent report by Gao and coworkers has shown that varying the iron carbonyl clusters
for ATH does change the reaction outcomes. As an example, Fe3(CO)12 fully converts
acetophenone to (S)-1-phenylethanol with 97% ee (enantiomeric excess), while
[Et3NH][HFe3(CO)11] does not activate the substrate. One possible explanation for these different
catalytic performances might arise from the varying coordination of the chiral ligands directly onto
different Fe3 cluster cores or perhaps to a fragment of the initial cluster.213 In another report by Gao
and coworkers, the catalytic reaction was monitored by IR spectroscopy, verifying that
[Et3NH][HFe3(CO)11] stayed intact during the reaction. This may serve as evidence for
[Et3NH][HFe3(CO)11] playing the role of active catalyst, or it may be that a very small fraction of
[Et3NH][HFe3(CO)11] is converted to the active catalyst although the event is not detected by IR.212
Taken together, these studies do not yet define a clear role for Fe3(CO)12 or [Et3NH][HFe3(CO)11] in
the catalytic ATH conversion of amines to imines. Shieh and coworkers have recently reported
46
Page 47 of 76 Chemical Society Reviews
47
Chemical Society Reviews Page 48 of 76
The second area of interest for Ru carbonyl clusters involves reactions with silane
additives. Some of the most successful implementations of this approach include the transformation
of esters to alkyl silyl acetals, which are hydrolysed to aldehyde. The proposed mechanism of this
reaction involves oxidative addition of silane to a metal in Ru3(CO)12, and mechanistic work
suggests that an intermediate silane-ligated Ru cluster is important for Si–H bond activation, which
leads to subsequent reaction with olefin.222 Also using Ru3(CO)12, hydrosilylation of acetophenone
has faster turnover than olefin formation for 1-octene using (EtO)3SiH (TON: 220 and 60,
respectively).217 By replacing the silane with (EtO)3Si(CH2)3NH2, the Ru catalyst becomes more
active for the hydrogenation of olefin substrates.223 Murai and coworkers demonstrated the
carbonylative [4+1] cycloaddition of alkyl and aryl α,β-unsaturated imines with CO using
Ru3(CO)12. Although they were not able to definitively identify the active species, a mechanism has
been proposed (Scheme 24).216
48
Page 49 of 76 Chemical Society Reviews
Ru(CO)4
NR NR
O +CO
NR NR
(OC)3Ru Ru(CO)4
O
Scheme 24. Proposed mechanism for carbonylative [4+1] cycloaddition using Ru3(CO)12.
Fuchikami and coworkers have reported reduction of an ester to aldehyde via a proposed
silyl acetal intermediate that is hydrolyzed to afford the aldehyde.218 The role of the silane addition
is different in this example because it reacts with the substrate directly, rather than activating
Ru3(CO)12 by direct reaction with a Ru centre.220 Nagashima has found that coordination between
Ru3(CO)12 and acenaphthylene enables catalytic hydrosilylation of ketone and reduction of an
amide to an amine group in a presence of silane additive.219
Compared with Fe and Ru carbonyl clusters, there are far fewer reports of catalysis using
Os carbonyl clusters (Table 4).205,224 Muetterties and coworkers investigated the reactivity of
Os3(CO)12 and Ir4(CO)12 towards hydrocarbons, which led to the discovery that Os3(CO)12 can
effect hydrogen–deuterium exchange. Years later, Adams proposed multiple molecular structures
of intermediates to better understand the catalytic reaction of Os3(CO)12 with isocyanides and
related functional groups, using a series of stoichiometric reactions where activation of the nitrile
CN bond was observed.225
49
Chemical Society Reviews Page 50 of 76
Catalysis by MCCs of Group 9 (Co, Rh, Ir). Pittman and coworkers studied phosphine-
substituted Co carbonyl clusters, Co3(CO)9CPh and Co4(CO)10(PPh)2.226 These clusters promote the
hydroformylation reaction of pentene, which was the first example of homogeneous catalysis with
Co clusters (Chart 2, Table 5). Both catalysts completely converted 1- or 2-pentene to three
different types of formylated products within 24 h. A discussion of possible fragmentation
pathways to afford the active catalyst was presented, but no definitive active catalyst could be
identified in this work.
Ph Ph
CO
CO
C P
M M
M M M M M M
M P
Co2(CO)8 Ph
Co3(CO)9(CPh)
Co4(CO)10(PPh)2
M = Rh M = Ir
OC M
M
CO
O M
C M M
M M
M M M M
C
M
O
M
CO
OC
M
OC
Rh4(CO)12
Rh6(CO)16 Ir4(CO)12
Chart 2. Homogeneous MCC Catalysts from Group 9. Terminal CO ligands have been omitted for
clarity.
50
Page 51 of 76 Chemical Society Reviews
NaY zeolite, catalytic CO hydrogenation yields a C2–C4 carbon product. Gates and coworkers have
also investigated catalytic performances of various other MCCs supported by MgO or zeolite.232,233
The encapsulated Ir cluster was prepared by the adsorption of [Ir(CO)2(acac)] into the cage of NaY
zeolite and thermal treatment. IR and extended X-ray absorption fine structure (EXAFS) techniques
were adopted to verify the structure of the final catalyst because Ir6(CO)16 has a distinct signature at
1730 cm−1. Using a feedstock of CO and H2, zeolite-encaged Ir cluster yields various hydrogenated
products, which were detected by gas chromatography (GC). The continuous reaction time of up to
8 days reveals high stability for Ir6(CO)16 within the zeolite cage. In the reactions starting with
MgO-supported Ir6(CO)16, IR spectroscopy detected the structure of Ir4(CO)12, and it is believed
that this cluster forms on the MgO surface. In each of these reports, it is believed that the supporting
substrate enhances the stability of the MCCs.
Catalysis by MCCs of Group 10 (Pd and Pt). Likely due to their limited stability, the
carbonyl clusters of Ni are not represented in the literature for catalysis. As for Pt carbonyl clusters
(Chart 3, Table 8), the first catalytic reaction was reported by Pettit and coworkers in 1977, and that
report described hydroformylation and water-gas shift reactions.202 The next known catalytic
reaction was reported in the late 1990s. It is challenging to stabilise the MCCs of Group 10 under
ambient conditions, especially under catalytic conditions; for this reason, all the examples involve
catalysts supported on mesoporous silica or zeolite. Various evidence for the identity of the active
catalyst in each of these cases is obtained from in-situ IR spectroscopy and EXAFS. In an
alternative mechanism for catalysis in Group 10, Bhaduri and coworkers have employed redox
52
Page 53 of 76 Chemical Society Reviews
pathways where a reduced form of Pt9(CO)18 is used as a reductant to reduce nicotinamide adenine
dinucleotide (NAD+). In this reaction, the platinum cluster takes the role of electron transporter
from the sacrificial electron-donor molecules, which are hydroxide ions (OH−) recovered from
ferric cyanide. In an alternative redox reaction, methylene blue, Safranine O, and methyl viologen
were used as sacrificial electron acceptors, which were then reduced by the Pt cluster.235
2-
O Pt
O C CO
C Pt
CO
OC
Pt Pt OPt Pt Pt
C
CO
C C
O n O
Pt Pt
{[Pt3(CO)6]2-}n C
n = 1-10 O
Chart 3. Homogeneous MCC Catalysts from Group 10. Terminal CO ligands have been omitted
for clarity.
materials that are commonly deposited on the electrode. We will focus only on homogeneous
electrocatalysis, where the most basic requirements for the electrocatalyst include stability in a polar
solvent system and a redox potential that is accessible without oxidising or reducing the solvent.
These criteria alone limit the number of MCCs that might be suitable electrocatalysts, but those
MCCs that do meet these criteria have many advantages over single-metal-ion homogeneous
electrocatalysts, as described below. Considered from another angle, there are advantages for using
electrocatalysis instead of thermal activation as a means to control the reactivity of MCCs. One big
advantage of electrochemical catalysis is that energy can be added to the system at RT and with
very mild conditions; as a result, the stability of MCCs can be maintained throughout the catalytic
reaction and for many turnovers. In most instances, it is also possible to include a partial
atmosphere of CO (if it is needed), imparting even greater stability to the MCC; the mole-fraction
of CO in the headspace can be tuned at will.
The advantages of using suitable MCCs as homogeneous electrocatalysts, rather than using
single-site metal–ligand catalysts or heterogeneous electrocatalysts, have come to light in recent
years. Suitable MCCs and their reduced forms are readily protonated, allowing the formation of
hydride intermediates at low applied potential; in one instance, catalysis was promoted at a rate of
109 s-1 with an applied potential as low as –0.86 V vs. SCE (versus the saturated calomel
electrode).238 Multiple sites for protonation on the clusters lead to very fast rates of proton transfer
(PT), facilitating diffusion-limited PT chemistry, which can be applied to hydride-transfer reactions.
It is thought that the fast rates arise from a statistical effect of having so many possible protonation
sites and hydride migration sites. Another advantage of MCCs as electrocatalysts involve the easy
substitution reactions at the cluster surface, where CO ligands can be replaced by phosphines
containing an array of functional groups. Those substitutions can be used to tune the secondary
coordination sphere of a catalyst or to tune the reactivity properties by changing the electronic
properties of the cluster core. Substitutions of metal ions within the MCC core or of an interstitial
atom within the cluster core can also change the electronic and reactivity properties. A more
detailed discussion of synthetic modifications to MCCs and their effect on catalysis is included
below.
Electrocatalytic reactions where MCCs have been employed are all reduction reactions;
low-valent clusters can stabilise additional electron density with their delocalised electronic
structures and multiple -acceptor CO ligands. Electrons added to the molecular-orbital manifold of
MCCs often strengthen the bonding between the metal and carbon atoms of the M–CO moiety. In a
typical electrocatalytic reduction reaction—including those mediated by MCCs, catalysis is initiated
by electron transfer (ET) to the MCC as potential is applied; this electrochemical elementary step is
54
Page 55 of 76 Chemical Society Reviews
abbreviated as “E” in the standard nomenclature of the field. After the initiation step, if the
intermediate has enough energy to react with the substrate, it will react chemically, and that
chemical reaction is abbreviated as “C” (Scheme 25). Alternatively, the initial E event can be
followed by a second E event before any chemical steps proceed. Thus, a series of elementary steps
in an electrocatalytic reaction will be described as a series of E or C steps. The most common
reaction mechanisms are ECEC and EECC. The latter is considered equivalent to the ECCE
mechanism; while they differ in their starting points in the catalytic cycle, they have identical
electrochemical responses when CV is used to monitor their reactions.239
H2 [Mx(CO)y]n H2 [Mx(CO)y]n
+ e- + e-
+ + e-
+H
+ e- + H+ + H+ + H+
[HMx(CO)y]n [HMx(CO)y]n
In any electrochemical reaction, the energy difference between the electron-donor and -
acceptor is an important parameter, and this difference is expressed by their redox potentials with
respect to a reference electrode. The IUPAC (International Union of Pure and Applied Chemistry)
recommends use of the ferrocene/ferrocenium couple as a reference in organic solvent;240 other
commonly employed references include the SCE and the natural hydrogen electrode (NHE).
Provided that the experimental conditions are clearly described, conversion between the different
referencing schemes is generally trivial, allowing catalysts to be compared.241 In this review, we use
SCE referencing throughout. Metrics used to assess the utility of electrocatalytic reactions include
the Faradaic efficiency (FE) and the turnover frequency (TOF). The FE is the percentage of the
charge that is converted into product in a preparative-scale experiment. The TOF is equivalent to
kobs, the observed rate constant, and it is reported in units of s-1; as long as factors such as the
applied potential or the overpotential () are accounted for, it can be used to compare the rates of
reactions.
Two electrochemical reactions have been studied using MCCs as the electrocatalyst:
(a) the two-electron reduction of two protons to H2, and
55
Chemical Society Reviews Page 56 of 76
Scheme 26. Schematic showing the two possible reaction pathways for a metal hydride – which can
(a) react with protons to form H2 or (b) react with CO2 to form formate.
Electrocatalysis by MCCs of Group 8 (Fe). It was first reported in 2011 that [Fe4N(CO)12]−
can cause the evolution of H2 from protons or the evolution of formate from CO2, depending on the
reaction conditions 244 (Chart 4, Table 9). The first attempt for the electrocatalytic hydrogen-
evolution reaction (HER) and CO2-reduction reaction (CO2RR) was carried out under 1 atm of N2
or CO2, respectively, using organic acids as the proton source. In the proposed mechanism,
[Fe4N(CO)12]− is first reduced at −1.23 V (vs. SCE), and then PT affords [HFe4N(CO)12]−, which
can go on to react with either H+ or CO2 following an ECCE mechanism. There are very few
molecular electrocatalysts that produce formate selectively,242 and [Fe4N(CO)12]− performs this
reaction at modest rates and a low applied potential. In subsequent work reported in 2013,
replacement of organic acids with water buffered at pH 7 afforded formate with 97% FE at –1.23 V
of applied potential.246 The solvent-dependence reflects a lowering of the activation barrier for
hydride transfer due to solvation effects,243 which will not be discussed in detail here. In addition to
the ability of [Fe4N(CO)12]− to perform catalytic small-molecule reduction chemistry effectively,
these reports have demonstrated that it is stable in water and that it is stable under an applied
potential over reaction times of up to 24 h, although longer reactions were not investigated.
56
Page 57 of 76 Chemical Society Reviews
Fe - 2-
Fe Fe
X C
Fe Fe Fe
Fe Fe
Fe
-
X = N, [Fe4N(CO)12] [Fe5C(CO)15]2-
X = C, [Fe4C(CO)12]2-
2-
Fe
Fe Fe
C
Fe Fe
Fe
[Fe6C(CO)18]2-
Table 9. Homometallic clusters of Fe and Co and their reactivity for HER and CO2RR.
not a long-term thermodynamic product. This study also enabled a comparison of the nitride-
centred [Fe4X(CO)12]− and carbide-centred [Fe4C(CO)12]2− clusters. Protonation of [Fe4C(CO)12]2−
is concerted with ET and is therefore pH-dependent; meanwhile, [Fe4N(CO)12]− undergoes
sequential ET and PT reactions. The difference in reactivity is attributed to the varied charges on
the clusters. For both [Fe4N(CO)12]− and [Fe4C(CO)12]2−, the protonation site is located along the
bridge of two iron atoms, and the interstitial atom does not get protonated. Another effect of the
increased charge on [Fe4C(CO)12]2- is that PT rates are significantly higher, so H2 evolution is
faster.
Electrocatalysis by MCCs of Group 9 (Co). A cobalt carbonyl cluster has recently been
reported as an electrocatalyst for fast H2 evolution; this cluster has 13 Co atoms,
[Co13C2(CO)24]4−.238 Many of its properties, including its ET and PT rate constants and diffusion
coefficient, are very similar to those of nanoparticles or quantum dots, illustrating the ability of
MCCs to behave like nanoparticles and like heterogeneous electrocatalysts. The larger size of
[Co13C2(CO)24]4−, its multiple Co–Co bonding motifs, and its many electron-withdrawing CO
ligands are just some of the structural features that result in its observed physical properties. When
studying the reduction of protons to hydrogen or the transfer of hydride to CO2 to form formate, the
ability to detect and monitor reaction intermediates is very important: Co carbonyl clusters offer a
unique system for probing protonation chemistry using IR spectra, as was first described by
Zacchini and coworkers.48 When plotting the CO absorption band energy, CO, obtained from IR
data against the x/z values (where x is the number of Co atoms and z is the cluster charge), a linear
relationship is observed for a wide range of clusters—specifically, from 6 to 20 Co atoms. The
charge on the clusters, z, can be varied either by redox events or by protonation of a Co cluster. For
a given cluster charge, protonation events can be monitored accurately using IR spectroscopy,
providing a powerful tool for determining the reaction mechanisms of hydride formation and
hydride transfer to substrates.
Co
Co OC Co
Co CO
OC COCO
OC Co Co Co
Co C C
Co CO
CO Co CO
OC
OC Co
Co Co
58
Page 59 of 76 Chemical Society Reviews
Using [Co13C2(CO)24]3−, the reduction of protons to hydrogen was studied using an applied
potential as the source of electrons (Scheme 26). PT rates were measured at 109 M−1s−1 for the
chemical reaction where electrochemically generated [Co13C2(CO)24]5− reacts with one proton. The
fast PT chemistry supported the HER with an observed rate of 2.3 109 M−1s−1. These PT rates
appear to be limited only by diffusion or the mass transport of protons in solution. The ability of
[Co13C2(CO)24]5− to mediate fast rates of PT was attributed to a statistical effect, which arises from
the multiple Co–Co bonds available for protonation; in heterogeneous catalysis, a similar effect
arises from the multiple PT sites on the electrode surface. Statistical effects that enhance the rates of
PT have also been observed in older work for single-site metal catalysts, i.e., mononuclear
coordination complexes that can act as homogeneous catalysts. In these cases, PT-rate enhancement
is derived from the addition of Lewis bases to the supporting ligands; the multiple Lewis-base sites
deliver protons to the catalytic active site on the metal centre.247 Studies were performed to
elucidate the mechanism for HER using [Co13C2(CO)24]4−. When either H2O or D2O was added to
the CV experiment of [Co13C2(CO)24]3−, an electrochemical kinetic-isotope effect was observed.
The shift in peak potential for the reduction of [Co13C2(CO)24]4− to [Co13C2(CO)24]5− when D2O was
used is indicative of a PT event that occurs in a concerted process with an ET event. Based on the
currently available information, the observed Co–H bond making/breaking process could be either a
protonation of the cluster that is concerted with ET or an intracluster proton migration that is
concerted with ET; the electrochemical data cannot distinguish between these two possibilities.
Both scenarios reflect the ability of large MCCs to interact effectively with protons and to facilitate
proton mobility on the surface of the cluster.
59
Chemical Society Reviews Page 60 of 76
Scheme 26. Top: Proposed mechanism for proton reduction to H2 in [Co13C2(CO)24]3−; the C and
CO ligands are not shown to simplify the mechanistic scheme. Reproduced with permission from
ref. 238. Bottom: Protonation of [Co13C2(CO)24]3− and migration of CO around the cluster core
(blue, Co; grey, C; red, O).
Atom substitution in the MCC Core: Effects on electrocatalysis. Starting in 2017, Hogarth
and coworkers published several studies on tri-iron carbonyl clusters (Chart 6, Table 10), which
were functionalised by replacing CO ligands or by adding capping ligands to the three-iron
cluster.248,249 Ruthenium clusters were also studied in some cases. For each cluster, two one-electron
reduction events were observed, and H2 evolution was observed at the most negative potential.
Hydride equivalents generated at the first reduction event were not hydridic enough to transfer
hydride to protons, which would have liberated H2. Thiol-capped [Fe3(CO)9(μ-SR)(μ-H)] (where R
= iPr, tBu) was investigated using organic acids, CF3CO2H or HBF4∙Et2O; H2 evolution was
observed at –0.84 (R = iPr) or –1.0 (R = tBu) V vs. SCE, respectively, with the mechanism
assigned as ECEC.249 Although they are less electron-rich, telluride-capped tri-iron clusters also
reduce protons to H2, but with at slower HER rates. Following phosphine substitution, Te-
substituted clusters showed enhanced reactivity. Their proposed mechanism shows that it is possible
to evolve hydrogen with a pathway that starts from the cluster’s first or second reduced state.237
60
Page 61 of 76 Chemical Society Reviews
R
H N S
N
M M Fe Fe
M H Fe H
[Fe3(CO)9(3-SR)(-H)]
[Fe3(CO)9(3-pyNPh)(-H)]
(M = Fe, Ru) (R = iPr, tBu)
Te
Mn
M M M O
Fe Fe
Te Fe
M3(CO)9(3-Te)2 [Fe3MnO(CO)12]-
(M = Fe, Ru)
Chart 6. Heterometallic and capped Fe carbonyl clusters used as electrocatalysts. CO ligands have
been omitted for clarity.
Table 10. Heterometallic and ligand-capped Fe and Ru clusters and their reactivity for HER.
61
Chemical Society Reviews Page 62 of 76
observed due to its reduced ability for hydride donation, which was determined independently using
IR spectroelectrochemical studies.
Ligand substitution on the MCC Core: Effects on electrocatalysis. Ligand substitution on
[Fe4N(CO)12]−, where CO ligands are replaced by phosphine ligands, has been used to study the
mechanism of H2 and formate evolution. Ligand substitution has also been used to modify the
reactivity of [Fe4N(CO)12]−. For these studies, the control compound is [Fe4N(CO)11(PPh3)]−, which
contains a phosphine ligand but does not contain a reactive functional group.253 The phosphine-
substituted compounds (including the control [Fe4N(CO)11(PPh3)]−) have a reduction potential of
around –1.45 V, while that of [Fe4N(CO)12]− is –1.2 V. Phosphine ligands with hydroxyl,
methoxyphenyl, pyridyl, and N,N-dimethylaniline groups have been used to afford
[Fe4N(CO)11(PPh2(CH2)2OH)]−, [Fe4N(CO)11(PPh2(MeOPh))]−, [Fe4N(CO)11(PPh2py)]−, and
[Fe4N(CO)11(PPh2(N,N-Me2NPh))]−, respectively (Chart 6). For each of these examples (Table 11),
MeCN/H2O (95:5) was used for the mechanistic or other electrochemical studies because it
provided adequate solubility; MeCN has a large electrochemical window, and the small amount of
H2O serves as a proton source and as a source of stabilization for the transition state during hydride
transfer. Using multiple approaches, electrochemical mechanistic studies have shown that water
stabilises a charge-separated transition state and enhances the rates of hydride transfer from [H-
Fe4N(CO)12]− to CO2 by five orders of magnitude relative to the rates of hydride transfer observed
in the absence of water. Other research groups working with single-site hydride-transfer catalysts
have also shown using density functional theory (DFT) studies that water may play a role in the
stabilization of the hydride-transfer transition state via a hydrogen-bonding network.242
L = CO
Fe L PPh3
PPh2(CH2)2OH
N
PPh2(2-Me2NPh)
Fe Fe PPh2(2-MeOPh)
Fe
PPh2py
62
Page 63 of 76 Chemical Society Reviews
63
Chemical Society Reviews Page 64 of 76
also be used to modify the overall charge of a cluster or its solubility, making a cluster more
suitable for catalytic applications. Another challenge facing the chemistry of ligand-substituted
MCCs is how to better predict which cluster–ligand combinations will afford stable ensembles.
Outlook for Tuning the Material Properties of MCCs: Another area needing further
development is the synthesis and characterization of heterometallic and alloy molecular
nanoclusters. Mixing different metals with atomic control may lead to nanomaterials with new
physical or chemical properties. Alloying and metal-doping of molecular nanoclusters is gaining
more and more interest,258,259 and heterometallic MCCs may contribute to our understanding of the
synergetic effects induced by alloying with atomic precision.87 This might also be of some
significance in interpreting the metal migration, chemiadsorption, and catalytic behaviour of alloy
nanoparticles.
The electronic properties of molecular nanoclusters can be experimentally investigated using
spectroscopic and electrochemical techniques as well as magnetic measurements. For additional
insights, they can also be compared to theoretical models. As the size of the cluster increases, an
incipient metallisation of its metal core should be observed.6 Indeed, CV measurements indicate a
decrease of the energy gap (E) between consecutive redox couples in multivalent MCCs as their
size increases. At a certain size, E should be comparable to the thermal energy at RT, leading to
auto-disproportionation and equilibria among isostructural MCCs with different charges. This new
electronic status should also influence the magnetic properties of larger MCCs and nanoclusters in
general. Paramagnetism of even-electron MCCs has been debated for a long time, but recently it has
been clearly demonstrated that they can have unpaired electrons in both the ground and excited
electronic configurations.135 Magnetic measurements of molecular nanoclusters require ultra-pure
samples in order to rule out the presence of paramagnetic impurities, but they can add significant
information to our knowledge of nanomaterials. In addition, by understanding and learning to tune
the electronic and magnetic properties of MCCs, it will be possible to produce suitable candidates
for molecular nanocapacitors, superparamagnetic quantum dots, and nanomagnets.
Outlook for Catalysis with MCCs: From a structural point of view, a regular and simple
structure–size relationship has not been established in the case of MCCs and molecular
nanoclusters. This is likely due to the fact that at these length scales, M–ligand and M–M
interactions contribute similarly to the overall energetic properties. Thus, we can expect that
introducing defects in the metal cages should lead to rearrangements, as shown in the case of some
Pt MCCs. Indeed, there is evidence that the metal cages of some MCCs are fluxional, and generally
speaking, they are soft and deformable. A deeper knowledge of the dynamic behaviour of the metal
cage and ligand shell of molecular clusters is required, and this might have some relevance also to
65
Chemical Society Reviews Page 66 of 76
heterogeneous catalysis with ultra-dispersed metals. The size, structure, and composition of
molecular nanoclusters strongly affect their catalytic behaviour.
There are still significant opportunities for studying MCCs of different sizes and
compositions in the activation of small molecules, in homogeneous catalysis, as precursors of
heterogeneous nanostructured catalysts, and in homogeneous electrocatalysis. An overview of
homogeneous catalysis and homogeneous electrocatalysis by MCCs has been given in this review,
and other recent reviews have touched on related aspects of catalysis with MCCs.23, 260, 261 General
ideas that can be drawn from these examples include the knowledge that thermal activation of
catalytic cycles often results in MCC decomposition and that many of the most effective thermal
catalytic reactions are those that occur under a MCC-stabilising CO atmosphere. Moving forward, it
is essential that we develop better predictive methods for knowing which MCCs will be stable to
various reaction conditions.
Electrocatalysis has recently provided a method where energy, in the form of the applied
potential, can be added to the system without raising the temperature. This has enabled RT catalytic
reactions, including reactions proceeding via a hydride intermediate. As an example, the reduction
of CO2 into formate can proceed in water for over 24 h when [Fe4N(CO)12]– is electrolyzed at –1.2
V vs. SCE.245 Future work in electrocatalysis with MCCs should take advantage of the tunable
surface of MCCs, where CO ligands can be replaced with phosphine ligands. Inclusion of
functionalised phosphine ligands has the potential to modulate reaction rates by orders of magnitude
and to modulate product selectivity when the correct functional groups are included in the
secondary coordination sphere and installed in appropriate locations. Another frontier in
electrocatalysis using MCCs is the study of larger and larger clusters. Recent work has shown that
larger MCCs, containing 13 cobalt atoms, enhance the rates of proton transfer by many orders of
magnitude compared to the four-iron cluster, [Fe4N(CO)12]–, until they resemble those observed for
heterogeneous electrocatalysts. Proton- and hydride-migration chemistry on these tridecanuclear
cobalt clusters also resembles the behaviour of hydrogen atoms on a surface where hydrogen atoms
are not localised at one place.238 The combination of these nearly diffusion-limited reaction rates
with the ability to use molecular chemistry techniques to characterise structure, reactivity, and
reaction kinetics and mechanism in precise detail is a powerful driving force for the future
development of electrocatalytic reactions promoted by MCCs.
5. Conclusions
The ongoing, new insights into the electronic properties, structural dynamics, and catalytic
mechanisms of MCCs show no sign of abating. Nowadays, we possess a solid and broad knowledge
66
Page 67 of 76 Chemical Society Reviews
of the synthesis of MCCs, which can be exploited for future developments and applications as
outlined in the Summary and Outlook. The relevance of electrocatalysis to fundamental chemistry
and industrial applications is growing; and electrochemical techniques have provided a much-
needed RT approach, enabling MCCs to be stable throughout chemical and catalytic
transformations. Molecular nanoclusters, in general, and MCCs, in particular, can add new
perspectives to electrocatalysis. Being at the cusp of the nanodomain, they can contribute to our
knowledge of nanochemistry and solid-state materials.
6. Acknowledgement
L.A.B.’s work on MCCs has been supported by the Department of Energy, Office of Science, Basic
Energy Sciences with award number DE-SC0016395. S.Z. thanks the University of Bologna for
financial support.
References
1 E. L. Muetterties, T. N. Rhodin, E. BAND, C.F. Brucker, and W. R. Pretzer, Chem. Rev. 1979, 79, 91-137.
2 G. Schmid (Ed.), Clusters and Colloids, Wiley-VCH, New York, 1994.
3 P. Braunstein, L. A. Oro and P. R. Raithby (Eds.), Metal Clusters in Chemistry, Wiley-VCH, New York,
1999.
4 G. Schmid and D. Fenske, Phil. Trans. R. Soc. A, 2010, 368, 1207-1210.
5 S. Zacchini, Eur. J. Inorg. Chem., 2011, 4125-4145.
6 C. Femoni, F. Kaswalder, M. C. Iapalucci, G. Longoni and S. Zacchini, Coord. Chem. Rev., 2006, 250,
1580-1604
7 B. F. G. Johnson and J. S. McIndoe, Coord. Chem. Rev., 2000, 200-202, 901-932.
8 I. Ciabatti, C. Femoni, M. C. Iapalucci, G. Longoni and S. Zacchini, J. Clust. Sci., 2014, 25, 115-146.
9 I. Ciabatti, C. Femoni, M. C. Iapalucci, S. Ruggieri and S. Zacchini, Coord. Chem. Rev., 2018, 355, 27-38.
10 B. Berti, C. Femoni, M. C. Iapalucci, S. Ruggieri and S. Zacchini, Eur. J. Inorg. Chem., 2018, 3285-3296.
11 C. Femoni, M. C. Iapalucci, S. Ruggieri and S. Zacchini, Acc. Chem. Res., 2018, 51, 2748-2755.
12 K. H. Whitmire, Coord. Chem. Rev., 2018, 376, 114-195.
13 E. G. Mednikov and L. F. Dahl, Phil. Trans. R. Soc. A, 2010, 368, 1301-1322.
14 J. A. Cabeza and P. García-Álvarez, Chem. Soc. Rev., 2011, 40, 5389-5405.
15 W. Hieber, F. Leutert, Die Naturwissenschaften. 19, 1931, 360–361.
16 G. Hogarth, S. E. Kabir and E. Nordlander, Dalton Trans., 2010, 39, 6153-5174.
17 R. Jin, C. Zeng, M. Zhou and Y. Chen, Chem. Rev., 2016, 116, 10346-10413.
18 I. Chakraborty and T. Pradeep, Chem. Rev., 2017, 117, 8208-8271.
19 X. Kang, Y. Li, M.Zhu and R. Jin, Chem. Soc. Rev., 2020, 49, 6443-6514.
20 X. Du and R. Jin, Dalton Trans., 2020, 49, 10701-10707.
21 (a) Q. Yao, T. Chen, X. Yuan and J. Xie, Acc. Chem. Res., 2018, 51, 1338-1348; (b) S. Takano, S.
Hasegawa, M. Suyama and T. Tsukuda, Acc. Chem. Res., 2018, 51, 3074-3083.
22 (a) Z. Lei, X. -K. Wan, S. -F. Yuan, Z. -J. Guan and Q.- M. Wang, Acc. Chem. Res., 2018, 51, 2465-2474;
(b) K. Konishi, M. Iwasaki and Y. Shichibu, Acc. Chem. Res., 2018, 51, 3125-3133.
23 P. Buchwalter, J. Rosé and P. Braunstein, Chem. Rev., 2015, 115, 28-126.
24 R. D. Adams and F. A. Cotton (Eds.), Catalysis by Di- and Polynuclear Metal Cluster Complexes, Wiley-
67
Chemical Society Reviews Page 68 of 76
27 J. M. Thomas, B. F. G. Johnson, R. Raja, G. Sankar and P. A. Midgley, Acc. Chem. Res., 2003, 36, 20-30
28 (a) T. Imoaka, H. Kitazawa, W.-J. Chun, S. Omura, K. Albrecht and K. Yamamoto, J. Am. Chem. Soc.,
2013, 135, 13089-13095; (b) T. Imaoka, H. Kitazawa, W.-J. Chun and K. Yamamoto, Angew. Chem. Int.
Ed., 2015, 54, 9810-9815.
29 K. Kratzl, T. Kratky, S. Günther, O. Tomanec, R. Zbořil, J. Michalička, J. M. Macak, M. Cokoja and R. A.
Chini, G. Longoni and S. Martinengo, J. Am. Chem. Soc., 1974, 96, 2614-2616.
35 M. Fauré, C. Saccavini and G. Lavigne, Chem. Commun., 2003, 1578-1579.
36 M. I. Bruce, C. M. Jensen and N. L. Jones, Inorg. Synth., 1990, 28, 216-218.
37 (a) M. Ichikawa, Platinum Met. Rev., 2000, 44, 3-14; (b) A. Fukuoka, N. Higashimoto, Y. Sakamoto, M.
Sasaki, N. Sugimoto, S. Inagaki, Y. Fukushima and M. Ichikawa, Catal. Today, 2001, 66, 23-31.
38 (a) C. Femoni, M. C. Iapalucci, G. Longoni, C. Tiozzo, S. Zacchini, B. T. Heaton and J. A. Iggo, Dalton
4600.
40 A. Fumagalli, S. Martinengo, P. Chini, D. Galli, B. T. Heaton and R. Della Pergola, Inorg. Chem., 1984,
23, 2947-2954.
41 (a) V. G. Albano, D. Braga and S. Martinengo, J. Chem. Soc., Dalton Trans., 1986, 981-984; (b) V. G.
Albano, D. Braga and S. Martinengo, J. Chem. Soc., Dalton Trans., 1981, 717-720; (c) G. Ciani and S.
Martinengo, J. Organomet. Chem., 1986, 306, C49-C52.
42 P. Chini, J. Organomet. Chem., 1980, 200, 37-61.
43 B. Berti, C. Cesari, C. Femoni, T. Funaioli, M. C. Iapalucci and S. Zacchini, Dalton Trans., 2020, 49,
5513-5522.
44 E. Cattabriga, I. Ciabatti, C. Femoni, T. Funaioli, M. C. Iapalucci and S. Zacchini, Inorg. Chem., 2016, 55,
6068-6079.
45 J. J. Brunet, Chem. Rev., 1990, 90, 1041-1059.
46 R. Della Pergola, F. Demartin, L. Garlaschelli, M. Manassero, S. Martinengo and M. Sansoni, Inorg.
Angew. Chem. Int. Ed., 1979, 18, 80-81; (b) D. W. Hart, R. G. Teller, C.-Y. Wei, R. Bau, G. Longoni, S.
Campanella, P. Chini and T. F. Koetzle, J. Am. Chem. Soc., 1981, 103, 1458-1466.
54 P. Chini, Inorg. Chim. Acta, 1968, 2, 31-51.
55 B. R. Whittlesey, Coord. Chem. Rev., 2000, 206-207, 395-418.
56 (a) G. Fachinetti, G. Fochi, T. Funaioli and P. F. Zanazzi, J. Chem. Soc., Chem. Commun., 1987, 89-90;
(b) C. Mealli, D. M. Proserpio, G. Fachinetti, T. Funaioli, G. Fochi and P. F. Zanazzi, Inorg. Chem., 1989,
28, 1122-1127.
57 G. Kong, G. N. Harakas and B. R. Whittlesey, J. Am. Chem. Soc., 1995, 117, 3502-3509.
58 C. Femoni, M. C. Iapalucci, G. Longoni and S. Zacchini, Eur. J. Inorg. Chem., 2009, 2487-2495.
68
Page 69 of 76 Chemical Society Reviews
Am. Chem. Soc., 1974, 96, 4155-4159; (c) C. F. Campana, I. A. Guzei, E. G. Mednikov and L. F. Dahl, J.
Clust. Sci., 2014, 25, 205-224.
64 (a) P. J. Krusic, J. R. Morton, K. F. Preston, A. J. Williams and F. L. Lee, Organometallics, 1990, 9, 697-
700; (b) P. J. Krusic, J. San Filippo, B. Hutchinson, R. L. Hance and L. M. Daniels, J. Am. Chem. Soc.,
1981, 103, 2129-2131; (c) P. J. Krusic, J. Am. Chem. Soc., 1981, 103, 2131-2133.
65 F. Ragaini, J.-S. Song, D. L. Ramage, G. L. Geoffroy, G. A. P. Yap and A. L. Rheingold,
(b) C. Femoni, M. C. Iapalucci, G. Longoni and S. Zacchini, Dalton Trans., 2011, 40, 8685-8694.
68 M. Bortoluzzi, I. Ciabatti, C. Femoni, M. Hayatifar, M. C. Iapalucci, G. Longoni and S. Zacchini, Angew.
2017, 3135-3143.
71 A. Gourdon and Y. Jeannin, J. Organomet. Chem., 1985, 290, 199-211.
72 (a) J. S. Bradley, G. B. Ansell and E. W. Hill, J. Am. Chem. Soc., 1979, 101, 7417-7419; (b) J. S. Bradley,
Reina, L. Rodríguez, O. Rossell, M. Seco, M. Font-Bardia and X. Solans, Organometallics, 2001, 20,
1575-1579.
77 (a) C. Joseph, S. Kuppuswamy, V. M. Lynch and M. J. Rose, Inorg. Chem., 2018, 57, 20-23; (b) J.
McGale, G. E. Cutsail III, C. Joseph, M. J. Rose and S. DeBeer, Inorg. Chem., 2019, 58, 12918-12932.
78 L. Liu, T. B. Rauchfuss and T. J. Woods, Inorg. Chem., 2019, 58, 8271-8274.
79 A. Ceriotti, L. Resconi, F. Demartin, G. Longoni, M. Manassero and M. Sansoni, J. Organomet. Chem.,
Whitmire, M. R. Churchill and J. C. Fettinger, J. Am. Chem. Soc., 1985, 107, 1056-1057; (c) K. H.
Whitmire, C. B. Lagrone, M. R. Churchill, J. C. Fettinger and L. V. Biondi, Inorg. Chem., 1984, 23, 4227-
4232.
87 M. Ferrer, R, Reina, O. Rossell and M. Seco, Coord. Chem. Rev., 1999, 193-195, 619-642.
69
Chemical Society Reviews Page 70 of 76
88 B. Berti, M. Bortoluzzi, C. Cesari, C. Femoni, M. C. Iapalucci, L. Soleri and S. Zacchini, Inorg. Chem.,
2020, 59, 15936-15952.
89 (a) M. Bortoluzzi, C. Cesari, I. Ciabatti, C. Femoni, M. Hayatifar, M. C. Iapalucci, R. Mazzoni and S.
Zacchini, J. Clust. Sci., 2017, 28, 703-723; (b) B. Berti, M. Bortoluzzi, C. Cesari, C. Femoni, M. C.
Iapalucci, R. Mazzoni, F. Vacca and S. Zacchini, Eur. J. Inorg. Chem., 2019, 3084-3093.
90 B. Berti, M. Bortoluzzi, C. Cesari, C. Femoni, M. C. Iapalucci, R. Mazzoni, F. Vacca and S. Zacchini,
Inorg. Chem., 2020, 2191-2202; (b) B. Berti, M. Bortoluzzi, C. Cesari, C. Femoni, M. C. Iapalucci, R.
Mazzoni, F. Vacca and S. Zacchini, Inorg. Chem., 2020, 59, 2228-2240.
92 N. P. Mankad, Chem. Commun., 2018, 54, 1291-1302.
93 N. P. Mankad, Chem. Eur. J., 2016, 22, 5822-5829.
94 Y. Lakliang and N. P. Mankad, Organometallics, 2020, 39, 2043-2046.
95 C. Femoni, M. C. Iapalucci, G. Longoni, C. Tiozzo and S. Zacchini, Agew. Chem. Int. Ed. 2008, 47, 6666-
6669.
96 B. F. G. Johnson and J. Lewis, Inorg. Synth., 1972, 13, 92-95.
97 (a) F. Calderazzo and F. L’Eplattenier, Inorg. Chem., 1967, 6, 1220- 1224; (b) F. L’Eplattener, Inorg.
Chem., 1969, 8, 965-970; (c) B. F. G. Jonhson, J. Lewis and M. V. Twigg, J. Organomet. Chem., 1974,
67, C75-C76.
98 J. Lewis and P. R. Raithby, J. Organoment, Chem., 1995, 500, 227-237.
99 (a) C. R. Eady, B. F. G. Johnson and J. Lewis, J. Chem. Soc., Dalton Trans., 1977, 838-844; (b) S. A. R.
Knox, J. W. Koepke, M. A. Andrews and H. D. Kaesz, J. Am. Chem. Soc., 1975, 97, 3942-3947.
100 M. Yu. Afonin, B. Yu. Savkov, A. V. Virovets, V. S. Korenev, A. V. Golovin and V. A. Maksakov, Eur.
1275; (b) K, V. Kong, W. K. Leong and L. H. K. Lim, Chem. Res. Toxicol., 2009, 22, 1116-1122.
104 R. D. Adams, E. J. Kiprotich and M. D. Smith, Chem. Commun., 2018, 54, 3464-3467.
105 P. J. Bailey, B. F. G. Johnson and J. Lewis, Inorg. Chim. Acta, 1994, 227, 197-200.
106 J. P. Canal, A. A. Bengali, M. C. Jennings and R. K. Pomeroy, Inorg. Chem. Commun., 2014, 43, 31-34.
107 S. Takemoto and H. Matsuzaka, Coord. Chem. Rev., 2012, 256, 574-588.
108 M. I. Bruce, N. N. Zaitseva, B. W. Skelton and A. H. White, J. Chem. Soc., Dalton Trans., 2002, 3879-
3885.
109 M. P. Cifuentes, M. G. Humphrey, J. E. McGrady, P. J. Smith, R. Stranger, K. S. Murray and B.
Adams, M. D. Smith, J. D. Tedder and N. D. Wakdikar, Inorg. Chem., 2019, 58, 8357-8368.
111 (a) R. D. Adams, B. Captain, W. Fu, P. J. Pellechia and M. D. Smith, Angew. Chem. Int. Ed., 2002, 41,
1951-1953; (b) S. Saha, L. Zhu and B. Captain, Inorg. Chem., 2013, 52, 2526-2532.
112 (a) R. D. Adams, I. T. Horváth and H.-S. Kim, Organometallics, 1984, 3, 548-552; (b) R. D. Adams, I. T.
Horváth and P. Mathur, Organometallics, 1984, 3,623-630; (c) R. D. Adams and I. T. Horváth, Inorg.
Chem., 1984, 23, 4718-4722.
113 S.-P. Huang and M. G. Kanatzidis, J. Am. Chem. Soc., 1992, 114, 5477-5478.
114 M. Shieh, M.-H. Hsu, W.-S. Sheu, L.-F. Jang, S.-F. Lin, Y.-Y. Chu, C.-Y. Miu, Y.-W. Lai, H.-L. Liu and
Chem., 1987, 330, C5-C11; (b) M. D. Randles, A. C. Willis, M. P. Cifuentes and M. G. Humphrey, J.
Organomet. Chem., 2007, 692, 4467-4472.
116 Y.-Z. Li, R. Ganguly, W. K. Leong and Y. Liu, Eur. J. Inorg. Chem., 2015, 3861-3872.
117 S. Saha, D. Isrow and B. Captain, J. Organomet. Chem., 2014, 751, 815-820.
70
Page 71 of 76 Chemical Society Reviews
118 (a) F.-E. Hong, T. J. Coffy, D. A. McCarthy and S. G. Shore, Inorg. Chem., 1989, 28, 3284-3285; (b) C.
E. Housecroft, D. M. Matthews, A. L. Rheingold and X. Song, J. Chem. Soc., Chem. Commun., 1992,
842-843.
119 J.-H. Chung, D. Knoeppel, D. McCarthy, A. Columbie and S. G. Shore, Inorg. Chem., 1993, 32, 3391-
3392.
120 R. D. Adams, J. Kiprotich, D. V. Peryshkov and Y. O. Wong, Chem. Eur. J., 2016, 22, 6501-6504.
121 R. D. Adams, Q. Zhang and X. Yang, J. Am. Chem. Soc., 2011, 133, 15950-15953.
122 K.-F. Yung and W.-T. Wong, Angew. Chem. Int. Ed., 2003, 42, 553-555.
123 T. Y. Garcia, J. C. Fettinger, M. M. Olmstead and A. L. Balch, Chem. Commun., 2009, 7143-7145.
124 (a) A. E. Carpenter, A. L. Rheingold and J. S. Figueroa, Organometallics, 2016, 35, 2309-2318; (b) A. E.
Fachinetti, G. Fochi, T. Funaioli and P. F. Zanazzi, Angew. Chem. Int. Ed., 1987, 26, 680-681.
128 (a) V. G. Albano, P. Chini, S. Martinengo, M. Sansoni and D. Strumolo, J. Chem. Soc., Chem. Commun.,
1974, 299-300; (b) S. Martinengo, D. Strumolo, P. Chini, V. G. Albano and D. Braga, J. Chem. Soc.,
Dalton Trans., 1985, 35-41.
129 V. G. Albano, D. Braga, F. Grepioni, R. Della Pergola, L. Garlaschelli and A. Fumagalli, J. Chem. Soc.,
203-211; (b) I. Ciabatti, C. Femoni, M. Hayatifar, M. C. Iapalucci, A. Ienco, G. Longoni, G. Manca and S.
Zacchini, Inorg. Chem.2014, 53, 9761-9770; (c) M. Bortoluzzi, I. Ciabatti, C. Femoni, T. Funaioli, M.
Hayatifar, M. C. Iaplaucci, G. Longoni and S. Zacchini, Dalton Trans., 2014, 43, 9633-9646.
137 (a) A. Fumagalli, P. Ulivieri, M. Costa, O. Crispu, R. Della Pergola, F. Fabrizi de Biani, F. Laschi, P.
Zanello, P. Macchi and A. Sironi, Inorg. Chem., 2004, 43, 2125-2131; (b) A. Fumagalli, M. Costa, R.
Della Pergola, P. Zanello, F. Fabrizi de Biani, P. Macchi and A. Sironi, Inorg. Chim. Acta, 2003, 350, 187-
192.
138 C. S. Hong, L. A. Berben and J. R. Long, Dalton Trans., 2003, 2119-2120.
139 G. Ciani, A. Sironi, S. Martinengo, L. Garlaschelli, R. Della Pergola, P. Zanello, F. Laschi and N.
P. L. Bellon, F. Demartin, M. Manassero, M. Y. Chiang, C.-Y. Wei and R. Bau, J. Am. Chem. Soc., 1984,
106, 6664-6667.
143 (a) G. Fachinetti, T. Funaioli and P. F. Zanazzi, J. Organomet. Chem., 1993, 460, C34-C36; (b) K. J.
Bradd, B. T. Heaton, J. A. Iggo, C. Jacob, J. T. Sampanthar and S. Zacchini, Dalton Trans., 2008, 685-
690.
144 (a) P. Chini and S. Martinengo, Chem. Commun., 1969, 1092-1093; (b) S. Martinengo, A. Fumagalli and
P. Chini, J. Organomet. Chem., 1985, 284, 275-279; (c) L. Malatesta, G. Caglio and M. Angoletta, J.
Chem. Soc., Chem. Commun., 1970, 532-533.
71
Chemical Society Reviews Page 72 of 76
1994, 1501-1503.
148 R. Della Pergola, L. Garlaschelli, M. Manassero, N. Masciocchi and P. Zanello, Angew. Chem. Int. Ed.,
26, 461-466; (b) C. Femoni, G. Bussoli, I. Ciabatti, M. Ermini, M. Hayatifar, M. C. Iapalucci, S. Ruggieri
and S. Zacchini, Inorg. Chem., 2017, 56, 6343-6351.
155 (a) A. Boccalini, P. J. Dyson, C. Femoni, M. C. Iapalucci, S. Ruggieri and S. Zacchini, Dalton Trans.,
2018, 47, 15737-15744; (b) C. Femoni, T. Funaioli, M. C. Iapalucci, S. Ruggieri and S. Zacchini, Inorg.
Chem., 2020, 59, 4300-4310.
156 J. L. Vidal, Inorg. Chem., 1981, 20, 243-249.
157 (a) J. L. Vidal, W. E. Walker, R. L. Pruett and R. C. Shoening, Inorg. Chem., 1979, 18, 129-136; (b) J. L.
Dalton Trans., 1983, 2175-2182; (b) R. Della Pergola, A. Sironi, C. Manassero and M. Manassero, Eur.
J. Inorg. Chem., 2009, 4618-4621.
161 L. Cerchi, A. Fumagalli, S. Fedi, P. Zanello, F. Fabrizi de Biani, F. Laschi, L. Garlaschelli, P. Macchi and
1994, 471-475.
163 A. Fumagalli, S. Martinengo, G. Bernasconi, G. Ciani, D. M. Proserpio and A. Sironi, J. Am. Chem. Soc.,
Chini, Inorg. Chem., 1976, 15, 3029-3031; (c) A. Ceriotti, P. Chini, R. Della Pergola and G. Longoni,
Inorg. Chem., 1983, 22, 1595-1598.
167 (a) J. K. Beattie, A. F. Masters and J. T. Meyer, Polyhedron, 1995, 14, 829-868; (b) A. F. Masters and J.
Ed., 2002, 41, 3685-3688; (b) N. T. Tran, M. Kawano, D. R. Powell and L. F. Dahl, J. Chem. Soc., Dalton
Trans., 2000, 4138-4144.
72
Page 73 of 76 Chemical Society Reviews
171 (a) C. Femoni, M. C. Iapalucci, G. Longoni, P. H. Svensson and J. Wolowska, Angew. Chem. Int. Ed.,
2000, 39, 1635-1637; (b) C. Femoni, M. C. Iapalucci, G. Longoni, P. H. Svensson, P. Zanello and F.
Fabrizi de Biani, Chem. Eur. J., 2004, 10, 2318-2326; (c) C. Femoni, M. C. Iapalucci, G. Longoni and P.
H. Svensson, Chem. Commun., 2001, 1776-1777.
172 (a) C. Femoni, M. C. Iapalucci, G. Longoni and P. H. Svensson, Chem. Commun., 2004, 2274-2275; (b) I.
Ciabatti, C. Femoni, M. C. Iapalucci, G. Longoni, S. Zacchini, S. Fedi and F. Fabrizi de Biani, Inorg.
Chem., 2012, 51, 11753-11761.
173 N. T. Tran, M. Kawano, D. R. Powell, R. K. Hayashi, C. F. Campana and L. F. Dahl, J. Am. Chem. Soc.,
2017, 849-850, 299-305; (b) C. Femoni, M. C. Iapalucci, G. Longoni, S. Zacchini, S. Fedi and F. Fabrizi
de Biani, Dalton Trans., 2012, 41, 4649-4663; (c) C. Femoni, M. C. Iapalucci, G. Longoni and S.
Zacchini, Chem. Commun., 2008, 3157-3159.
175 (a) A. Bernardi, I. Ciabatti, C. Femoni, M. C. Iapalucci, G. Longoni and S. Zacchini, J. Organomet.
Chem., 2016, 812, 229-239; (b) A. Bernardi, I. Ciabatti, C. Femoni, M. C. Iapalucci, G. Longoni and S.
Zacchini, Dalton Trans., 2013, 42, 407-421; (c) I. Ciabatti, C. Femoni, M. C. Iapalucci, A. Ienco, G.
Longoni, G. Manca and S. Zacchini, Inorg. Chem., 2013, 52, 10559-10565.
176 C. Cesari, I. Ciabatti, C. Femoni, M. C. Iapalucci and S. Zacchini, J. Clust. Sci., 2017, 28, 1963-1979.
177 (a) M. Bortoluzzi, I. Ciabatti, C. Femoni, M. Hayatifar, M. C. Iapalucci, G. Longoni and S. Zacchini,
Dalton Trans., 2014, 43, 13471-13475; (b) A. Bernardi, C. Femoni, M. C. Iapalucci, G. Longoni and S.
Zacchini, Dalton Trans., 2009, 4245-4251; (c) A. Bernardi, C. Femoni, M. C. Iapalucci, G. Longoni, F.
Ranuzzi, S. Zacchini, P. Zanello and S. Fedi, Chem. Eur. J., 2008, 14, 1924-1934.
178 (a) A. Bernardi, C. Femoni, M. C. Iapalucci, G. Longoni and S. Zacchini, Inorg. Chim. Acta, 2009, 362,
1239-1246; (b) A. Bernardi, C. Femoni, M. C. Iapalucci, G. Longoni, S. Zacchini, S. Fedi and P. Zanello,
Eur. J. Inorg. Chem., 2010, 4831-4842.
179 C. Femoni, M. C. Iapalucci, G. Longoni and S. Zacchini, Eur. J. Inorg. Chem., 2010, 1056-1062.
180 C. Femoni, M. C. Iapalucci, G. Longoni, S. Zacchini, I. Ciabatti, R. G. Della Valle, M. Mazzani and M.
Chem., 2018, 57, 1136-1147; (b) C. Capacci, C. Cesari, C. Femoni, M. C. Iapalucci, F. Mancini, S.
Ruggieri and S. Zacchini, Inorg. Chem., 2020, 59, 16016-16026.
182 (a) C. Femoni, F. Kaswalder, M. C. Iapalucci, G. Longoni, M. Mehlstäubl and S. Zacchini, Chem.
5992-6004.
184 B. Berti, M. Bortoluzzi, A. Ceriotti, C. Cesari, C. Femoni, M. C. Iapalucci and S. Zacchini, Inorg. Chim.
52, 4384-4395; (b) I. Ciabatti, C. Femoni, M. C. Iapalucci, G. Longoni and S. Zacchini, Organometallics,
2013, 32, 5180-5189.
186 (a) C. Cesari, I. Ciabatti, C. Femoni, M. C. Iapalucci, F. Mancini and S. Zacchini, Inorg. Chem., 2017, 56,
2137.
189 M. Bortoluzzi, C. Cesari, I. Ciabatti, C. Femoni, M. C. Iapalucci and S. Zacchini, Inorg. Chem., 2017, 56,
6532-6544.
190 (a) A. Ceriotti, M. Daghetta, S. El Afefey, A. Ienco, G. Longoni, G. Manca, C. Mealli, S. Zacchini and S.
Zarra, Inorg. Chem., 2011, 50, 12553-12561; (b) M. Bortoluzzi, A. Ceriotti, I. Ciabatti, R. Della Pergola,
C. Femoni, M. C. Iapalucci, A. Storione and S. Zacchini, Dalton Trans., 2016, 45, 5001-5013.
73
Chemical Society Reviews Page 74 of 76
191 (a) M. Bortoluzzi, A. Ceriotti, C. Cesari, I. Ciabatti, R. Della Pergola, C. Femoni, M. C. Iapalucci, A.
Storione and S. Zacchini, Eur. J. Inorg. Chem., 2016, 3939-3949; (b) B. Berti, M. Bortoluzzi, C. Cesari,
C. Femoni, M. C. Iapalucci and S. Zacchini, Inorg. Chim. Acta, 2020, 503, 119432.
192 D. Bonincontro, A. Lolli, A. Storione, A. Gasparotto, B. Berti, S. Zacchini, N. Dimitratos and S.
2406-2409; (b) I. Ciabatti, C. Femoni, M. C. Iapalucci, G. Longoni, S. Zacchini and S. Zarra, Nanoscale,
2012, 4, 4166-4177.
194 (a) E. Cattabriga, I. Ciabatti, C. Femoni, M. C. Iapalucci, G. Longoni and S. Zacchini, Inorg. Chim. Acta,
2018, 470, 238-249; (b) F. Gao, C. Li, B. T. Heaton, S. Zacchini, S. Zarra, G. Longoni and M. Garland,
Dalton Trans., 2011, 40, 5002-5008.
195 D. M. Washecheck, E. J. Wucherer, L. F. Dahl, A. Ceriotti, G. Longoni, M. Manassero, M. Sansoni and
197 L. Liu, T. J. Woods and T. B. Rauchfuss, Eur. J. Inorg. Chem., 2020, 2020, 3460–3465.
198 N. Queyriaux, D. Sun, J. Fize, J. Pécaut, M. J. Field, M. C-. Kerlidou and V. Artero, J. Am. Chem. Soc.,
200 J. McGale, G. E. Cutsail, C. Joseph, M. J. Rose and S. DeBeer, Inorg. Chem., 2019, 58, 12918–12932.
202 H. C. Kang, C. H. Mauldin, T. Cole, W. Slegeir, K. Cann and R. Pettit, J. Am. Chem. Soc., 1977, 99,
8323–8325.
203 K. Cann, T. Cole, W. Slegeir and R. Pettit, J. Am. Chem. Soc., 1978, 100, 3969–3971.
204 D. J. Darensbourg, C. Ovalles and M. Pala, J. Am. Chem. Soc., 1983, 105, 5937–5939.
205 M. G. Thomas, B. F. Beier and E. L. Muetterties, J. Am. Chem. Soc., 1976, 98, 1296–1297.
206 C. P. Casey and C. R. Cyr, J. Am. Chem. Soc., 1973, 95, 2248–2253.
207 Y. Sunada, H. Kawakami, T. Imaoka, Y. Motoyama and H. Nagashima, Angew. Chem. Intl. Ed., 2009, 48,
9511–9514.
208 S. Zhou, K. Junge, D. Addis, S. Das and M. Beller, Angew. Chem., 2009, 121, 9671–9674.
209 S. Zhou, D. Addis, S. Das, K. Junge and M. Beller, Chem. Commun., 2009, 4883.
210 S. Zhou, S. Fleischer, K. Junge, S. Das, D. Addis and M. Beller, Angew. Chem. Intl. Ed., 2010, 49, 8121–
8125.
211 S. Enthaler, ChemCatChem, 2011, 3, 666–670.
212 Y.-Y. Li, S.-L. Yu, W.-Y. Shen and J.-X. Gao, Acc. Chem. Res., 2015, 48, 2587–2598.
213 Y. Li, S. Yu, X. Wu, J. Xiao, W. Shen, Z. Dong and J. Gao, J. Am. Chem. Soc., 2014, 136, 4031–4039.
214 (a) C.-N. Lin, C.-Y. Huang, C.-C. Yu, Y.-M. Chen, W.-M. Ke, G.-J. Wang, G.-A. Lee and M. Shieh,
Dalton Trans., 2015, 44, 16675–16679.; (b) M. Shieh, Y.-H. Liu, C.-C. Wang, S.-H. Jian, C.-N. Lin, Y.-
M. Chen and C.-Y. Huang, New J. Chem., 2019, 43, 11832–11843.
215 P. C. Ford, R. G. Rinker, C. Ungermann, R. M. Laine, V. Landis and S. A. Moya, J. Am. Chem. Soc.,
217 H. S. Hilal, S. Khalaf and W. Jondi, J. Organometallic Chem., 1993, 452, 167–173.
218 T. Morimoto, N. Chatani and S. Murai, J. Am. Chem. Soc., 1999, 121, 1758–1759.
219 H. Nagashima, A. Suzuki, T. Iura, K. Ryu and K. Matsubara, Organometallics, 2000, 19, 3579–3590.
220 M. Igarashi, R. Mizuno and T. Fuchikami, Tet. Lett., 2001, 42, 2149–2151.
221 W. Imhof, A. Göbel, L. Schweda, D. Dönnecke and K. Halbauer, J. Organometallic Chem., 2005, 690,
3886–3897.
222 G. Süss-Fink and J. Reiner, J. Mol. Catal., 1982, 16, 231–242.
223 H. S. Hilal, W. Jondi, S. Khalaf and R. Abu-Halawa, J. Organomet. Chem.,1993, 452, 161–165.
74
Page 75 of 76 Chemical Society Reviews
224 K. G. Caulton, M. G. Thomas, B. A. Sosinsky and E. L. Muetterties, Proc. Natl. Acad. Sci., 1976, 73,
4274–4276.
225 R. D. Adams, Acc. Chem. Res., 1983, 16, 67–72.
226 R. C. Ryan and C. U. Pittman, J. Am. Chem. Soc., 1977, 99, 1986–1988.
227 T. Murai, T. Sakane and S. Kato, J. Org. Chem., 1990, 55, 449–453.
229 I. Ojima, P. Ingallina, R. J. Donovan and N. Clos, Organometallics, 1991, 10, 38–41.
231 K. Kaneda, H. Kuwahara and T. Imanaka, J. Mol. Catal., 1994, 88, L267–L270.
232 S. Kawi, J. R. Chang and B. C. Gates, J. Am. Chem. Soc., 1993, 115, 4830–4843.
233 S. Kawi, J. R. Chang and B. C. Gates, J. Phys. Chem., 1994, 98, 12978–12988.
234 T. Yamamoto, T. Shido, S. Inagaki, Y. Fukushima and M. Ichikawa, J. Phys. Chem. B, 1998, 102, 3866–
3875.
235 S. Bhaduri, Current Science, 2000, 78, 1318.
236 S. Basu, H. Paul, C. Gopinath, S. Bhaduri and G. Lahiri, J. Catal., 2005, 229, 298–302.
238 (a) C. R. Carr, A. Taheri and L. A. Berben, J. Am. Chem. Soc., 2020, 142, 12299–12305. (b) S.
240 G. Gritzner and J. Kuta, Pure & appl. Chem., 1983, 56, 461–466
241 I. Noviandri, K. N. Brown, D. S.Fleming, P. T. Gulyas, P. A. Lay, A. F. Masters and L. Phillips, J. Phys.
Chem. Soc., 2017, 139, 3685–3696; (b) D. C. Cunningham and J. Y. Yang, Chem. Commun., 2020,
56,12965–12968.
243 A. Taheri, C. R. Carr and L. A. Berben, ACS Catal., 2018, 8, 5787–5793
244 (a) M. D. Rail and L. A. Berben, J. Am. Chem. Soc., 2011, 133, 18577–18579; (b) N. D. Loewen, T.
Neelakantan, L. A. Berben, Acc. Chem. Res. 2017,50, 2362-2370; (c) A. Taheri, N. D. Loewen, D. B.
Cluff, J. C. Fettinger, L. A. Berben, Organometallics, 2018, 37, 1087-1091.
245 A. Taheri, E. J. Thompson, J. C. Fettinger and L. A. Berben, ACS Catal., 2015, 5, 7140–7151.
246 (a) A. D. Nguyen, M. D. Rail, M. Shanmugam, J. C. Fettinger and L. A. Berben, Inorg. Chem., 2013, 52,
866.
248 S. Ghosh and G. Hogarth, J. Organomet. Chem., 2017, 851, 57–67.
249 S. Ghosh, S. Basak-Modi, M. G. Richmond, E. Nordlander and G. Hogarth, Inorg. Chim. Acta, 2018, 480,
47–53.
250 A. Rahaman, G. C. Lisensky, J. Browder-Long, D. A. Hrovat, M. G. Richmond, E. Nordlander, and G.
433.
75
Chemical Society Reviews Page 76 of 76
256 N. D. Loewen, S. Pattanayak, R. Herber, J. C. Fettinger, L. A. Berben, L. A. J. Phys. Chem. Lett. 2021,
12, 3066-3073.
257 M. J. Drance, C. C. Mokhtarzadeh, M. Melaimi, D. W. Agnew, C. E. Moore, A. L. Rheingold, J. S.
3114-3124.
259 S. Wang, Q. Li, X. Kang and M. Zhu, Acc. Chem. Res., 2018, 51, 2784-2792.
260 I. G. Powers and C. Uyeda, ACS Catal., 2017, 7, 936–958.
261 M. T. Nielsen, R. Padilla and M. Nielsen, J. Clust. Sci., 2020, 31, 11–61.
76