You are on page 1of 77

A Journal of

Accepted Article
Title: Transition Metal Complexes With (C-C)-->M Agostic Interactions

Authors: Benjamin Grant Harvey and Richard D. Ernst

This manuscript has been accepted after peer review and appears as an
Accepted Article online prior to editing, proofing, and formal publication
of the final Version of Record (VoR). This work is currently citable by
using the Digital Object Identifier (DOI) given below. The VoR will be
published online in Early View as soon as possible and may be different
to this Accepted Article as a result of editing. Readers should obtain
the VoR from the journal website shown below when it is published
to ensure accuracy of information. The authors are responsible for the
content of this Accepted Article.

To be cited as: Eur. J. Inorg. Chem. 10.1002/ejic.201600989

Link to VoR: http://dx.doi.org/10.1002/ejic.201600989


European Journal of Inorganic Chemistry 10.1002/ejic.201600989

Transition Metal Complexes With (C-C)M Agostic Interactions

Benjamin G. Harvey*[a] and Richard D. Ernst*[b]

Keywords: Agostic bonds / Non-classical bonding / Olefin metathesis/


Metallacyclobutane / C-C bond activation

Abstract

The coordination of sigma bonding electron density to a Lewis acid had for some
time been largely the domain of boron hydride chemistry. Subsequently it was
found that C-H bonds could undergo intramolecular coordination to metal centers,
forming what are now known as “agostic bonds” and “agostic complexes”.
Thereafter even the H-H bond in molecular hydrogen was found capable of
coordination, leading to what are commonly referred to as “sigma complexes”.
Eventually, coordination by much more shielded C-C bonds has been established,
which is of particular significance relative to the interest in C-C bond activation.
This review focuses on the key developments in this field, including the nature of
the bonding interactions in these and some related species, as revealed by
spectroscopic, structural, and theoretical studies, their involvement in important
chemical reactions such as alkene polymerizations and metathesis, and
comparisons to the other types of electron deficient bridge bonding.

Biographies

Benjamin G. Harvey is a Senior Research Chemist at the Naval Air Warfare


Center, Weapons Division located in China Lake, CA. He received his Ph.D. in
inorganic chemistry from the University of Utah in 2005, studying with Professor
Richard D. Ernst. His graduate studies were focused on the synthesis and
characterization of early transition metal organometallic cage structures and
metallacyclobutanes. Many of the compounds studied in his research exhibited
unique C-C agostic bonding interactions. His current areas of interest include
catalysis, advanced biofuels, high temperature thermosetting resins, and green
chemistry.

Richard D. Ernst is a Professor of Chemistry at the University of Utah. He


received his B.S. in chemistry degree from the University of California in Berkeley
in 1973, and his Ph.D. degree in inorganic chemistry from Northwestern University

This article is protected by copyright. All rights reserved.


European Journal of Inorganic Chemistry 10.1002/ejic.201600989

in 1977. His major areas of interest include organometallic chemistry, particularly


of metal-pentadienyl compounds, coordination chemistry, and catalysis of the
Fischer-Tropsch and Water Gas Shift reactions.

Introduction

For quite some time, coordination of ligands to transition metals had only
been observed to take place via lone pair donation, or via the formation of single or
multiple bonds between the metal and one or more ligand atoms. Subsequently, 
coordination of alkenes and other multiply bonded species to metals was
established, beginning with Zeise’s salt, [PtCl3(C2H4)]-. The next step in this
progression was the observation that even  bonds could coordinate to transition
metals. The first, and most abundant, examples involved C-H bonds that were part
of a ligand that was bound to the metal center through one or more additional
interaction(s).[1] For such species, the term “agostic” interaction has been coined,
from the Greek, with the meaning of “to hold close to oneself”. Though
specifically meant at the time to apply only to C-H bonds,[1a] the definition also
matches observations for other sigma bonds, and thus its use is often extended to
them as well.

Later examples were found in which a ligand was attached to a metal only
through the  bond interaction, the most notable examples involving H2 as the
ligand.[2] In this review, an examination of the agostic interactions involving
specifically one or more C-C bonds will be presented. Quite clearly such
interactions are of utmost importance in effecting the activation of C-C bonds.[3,4]

The bonding between a C-C bond and a metal center would have two
components, a donor and an acceptor interaction, analogous to those in related C-
H[1] and H-H[2] species. For C-C agostic complexes, these would correspond to the
pictures below. Initially the prospects of isolating complexes with agostic (C-C)
M interactions appeared unlikely. This could be expected in part due to the

This article is protected by copyright. All rights reserved.


European Journal of Inorganic Chemistry 10.1002/ejic.201600989

significant p character present in typical C-C single bonds, which renders the C-C
electron density more directional and thus less accessible, as shown below.[3,5-7]

In accord with this, it has been observed that in the classic complexes with (C-
H)M agostic interactions, the metal center makes a much more substantial
interaction with the hydrogen side of the bond.[8] Nonetheless, in 1988 a
theoretical study related to a compound possessing an agostic (Si-C)Ti
interaction revealed that an analogous (C-C)Ti interaction would also be present
for the hypothetical compound in which the Si atom was replaced by C.[9] The
interaction was expected to be only about half as strong, and accompanied by weak
(C-H)Ti interactions. Additionally, early theoretical studies of olefin
insertion/polymerization processes suggested the intermediacy of species with C-
C and, especially, C-C agostic interactions.[10,11] Actually, since 1 would be an
intermediate or transition state for (C-C) bond activation [Equation (1)], it would
likewise be an intermediate for the reverse (insertion) step in olefin
polymerization.[12a]

(1)

It is of historical note that an “agostic” interaction involving a cyclopropane


ring and platinum was originally proposed to be present in the complexes [PtCl2(c-
C3H6)]2 and PtCl2(c-C3H6)(py)2 (py = pyridine),[13] but subsequent studies revealed
the complexes to be platinacyclobutanes, such as 2.[5] The notion that such
coordination could be feasible derives from the facts that in the strained
cyclopropane ring, the C-C bonding electron density is directed outwardly from the
ring, the potential donor orbital energy is increased, and the potential acceptor
orbital energy decreased, relative to an unstrained C-C bond. While a true

This article is protected by copyright. All rights reserved.


European Journal of Inorganic Chemistry 10.1002/ejic.201600989

cyclopropane complex was not isolated in this case, due to the occurrence of C-C
oxidative addition, theoretical and mechanistic studies suggested that such species

are intermediates in olefin cyclopropanation[14] as well as skeletal rearrangements


of substituted metallacyclobutanes.[15-17] Notably, somewhat later a lithium
cyclopropoxide complex (3, Figure 1) was isolated which displayed a weak C-C
agostic interaction.[18] This could be seen from the C2-C3 distance of 1.519(3) Å,
which was slightly longer than the average of the other two C-C bonds (1.499(2)
and 1.508(3) Å).

Alkyne-derived C-C Agostic Complexes

The first documented examples involving C-C agostic interactions with a


transition metal were isolated from the coupling reactions of
(cyclooctadienyl)titanium complexes with 2 or 3 equivalents of C6H5C2SiMe3 (4,
5, Figures 2, 3).[19] These reactions led to the construction of carbon atom cages

4 5

This article is protected by copyright. All rights reserved.


European Journal of Inorganic Chemistry 10.1002/ejic.201600989

around the titanium centers, thus helping to shield the metal from additional ligand
coordination as well as potential interactions with C-H bonds, resulting in formally
very electron deficient complexes (both 14 electrons). This then resulted in the
metal only being able to achieve further coordination from the C-C bonds, via one
or more agostic interactions. Notably, quite different products were isolated,
depending on whether the starting Ti(C5H5)(c-C8H11)(PR3) complex had R = CH3
or C2H5. In the former case, the expected stronger binding of the smaller PMe3
ligand led to its retention in the product, and only seemed to allow for the
incorporation of two equivalents of the alkyne.[12,19] In contrast, for the PEt3
complex, the phosphine was not retained, and three equivalents of alkyne were
incorporated.[19] In each case, the alkynes underwent coupling reactions with the
cyclooctadienyl ligand, but the isolated products were quite different.
For the PEt3 complex, the three alkynes led to a complex having the titanium
center coordinated by C5H5, diene, and alkyl ligands (5). However, the 14 electron
titanium center was observed to engage in a formally nonbonded 2.293(7) Å Ti-
C(7) interaction. This distance is shorter than the average Ti-C distances for the
Cp carbon atoms (range, 2.333(7)-2.384(6) Å), and for the diene (or enediyl)
ligand’s two internal carbon atoms, C4 and C5 (2.299(6) and 2.338(6) Å). This
requires either nonbonded contacts that are shorter than bonding ones, or else the
presence of the (C-C)Ti interactions. Indeed, the long C6-C7 bond, 1.596(8) Å,
seems an especially likely candidate for the latter possibility. This would be
consistent with the large C7-H coupling constant, 149 Hz.,[1a,12] and was supported
via a theoretical study which indicated substantial loss of C-C bonding electron
density as well as stabilization arising from the agostic interactions. Surprisingly,
three other C-C bonds were found to be involved simultaneously in these
interactions, C2-C3, C2-C7, and C7-C8. Nonetheless, the greatest interaction was
still found to involve the C6-C7 bond. Support for this was obtained from an
INADEQUATE NMR study, which provided the 13C-13C coupling constants.
Typically for a C-C single bond, these values should fall in the range of 30-40 Hz.
The values found for the four bonds in question were all below this range, being
smallest for the C6-C7 bond, and progressively greater for the C7-C8, C2-C3, and
C2-C7 bonds (17.9, 21.4, 24.8, and 29.6 Hz., respectively). The strengths of the
interactions estimated from a theoretical study correlated with these reasonably
well (20.2, 12.2, 14.3, and 10.3 kcal/mole, respectively).

This article is protected by copyright. All rights reserved.


European Journal of Inorganic Chemistry 10.1002/ejic.201600989

Some significantly weaker interactions were observed with three C-H bonds.
Two of the three involved backside interactions with (C-H) bonding density.
Interestingly, an alternative type of backside interaction, metal-to-(C-C * orbital),
has been proposed to promote C-C bond activations. These cases could involve
either pure metal orbital donation or a combination of metal/ligand orbital donation
to the * orbital.[4]

The incorporation of 2 equivalents of alkyne into the Ti(C5H5)(c-C8H11)-


(PMe3) complex led to a strikingly different coupling reaction. In this case, the
titanium center in the product 4 was coordinated by C5H5, olefin, -allyl, and PMe3
ligands, again providing for a 14 electron configuration, and for a carbon atom
cage surrounding the remaining part of the coordination sphere.[12,12a] The presence
of sigma allyl coordination at first appeared surprising, as  coordination would
have led to a 16 electron configuration, and metal complexes with -allyl and
olefin coordination involving such a bicyclic ring system were known.[20-22]
However, as in the case of complex 5, there were again some short, formally
nonbonded Ti—C contacts present, especially with C5 (2.539(3) Å). Initially an
agostic interaction with the C5-H bond was suspected, but the observation of a
large rather than small J(C5-H) value, 131 Hz., demonstrated otherwise. This left
the C4-C5 and C5-C10 bonds as possible candidates for agostic interactions. Their
respective 13C-13C coupling constants were found to be 31.1 and 22.7 Hz., the latter
value clearly implicating the presence of a (C5-C10)Ti agostic interaction.
However, this should have led to increased coupling constants for the C5-H and
C4-C5 (31.3 Hz.) bonds, but such was not the case. Their relatively normal values
could best be explained by the presence of accompanying but weaker agostic
interactions involving the C4-C5 and C5-H bonds, much as four interactions were
observed in the previous complex (5). The extra interactions help provide a
rationalization for the titanium center’s preference for the agostic coordination(s)
over -allyl coordination, though it may be that a small contribution of -allyl
character could also contribute. In any event, while it has been demonstrated that a
single (C-H)M interaction could be more favorable than olefin coordination, it
would seem much less likely that a single (C-C)M interaction would be more
favorable than olefin coordination. However, it may be that the sum of several
agostic interactions, as proposed in these complexes, may prevail over alternative
olefin coordination or C-H agostic interactions.

This article is protected by copyright. All rights reserved.


European Journal of Inorganic Chemistry 10.1002/ejic.201600989

The relative energetics of these interactions would seem to be of some


importance in olefin polymerization processes. As equation (1) reveals, the
migratory insertion process entails the formal replacement of olefin coordination
by a C-C agostic interaction. This may help explain why many of the most active
catalysts involve early transition metals, for which olefin coordination is not
especially favorable, thus leading to smaller barriers than would likely be
encountered for most later metals.[12] The potential presence of additional agostic
interactions in 1 could lead to further reductions in the barrier.
Subsequently, an even more complicated tetra(alkyne) coupling product was
isolated, 6 (Figure 4).[23] Though it was isolated as a byproduct with 4, it actually

can best be imagined to derive from 5 by the addition of another alkyne equivalent,
suggesting that the reaction leading to 4 also produced some 5 as well. To a first
approximation, the titanium coordination sphere in 6 included C5H5 as well as
asymmetric -allyl coordinations, and a titanacyclobutene fragment (C6-8). As
was found for the di and tri coupling products, a 14 electron configuration then
resulted. Due to the complexity of the molecule, 13C-13C coupling constants were
not obtained for the complex; however, structural data were in accord with the
presence of agostic (C-C)Ti interactions, though weak. An examination of the
titanacyclobutene structural parameters in this and related species led to the
conclusion that olefin coordination in this fragment was minimal, and perhaps
secondary to agostic interactions involving C-C bonds. Structural data for this
fragment were equivocal, though a short Ti-C7 distance of 2.316(2) Å was
observed. There was also some support for weak agostic interactions with the C8-
C9 (1.570(3) Å) and C9-C10 (1.562(2) Å) bonds, which may be viewed as being
part of a titanacyclobutane fragment. The Ti-C9 distance of 2.457(2) Å is in fact

This article is protected by copyright. All rights reserved.


European Journal of Inorganic Chemistry 10.1002/ejic.201600989

shorter than some values in 16 electron titanacyclobutanes for which agostic C-


CTi interactions have been established (vide infra).

Another example of a complex appearing from structural data to have a (C-


C)Ti agostic interaction was isolated from the incorporation of a diyne,
(CH2)4(C2Et)2, into the titanocene coordination sphere (7a, Figure 5). This resulted

7a 7b

in the diyne ligand’s coupling with one of the C5H5 ligands, accompanied by
significant rearrangement of the carbon atom framework, due to C-C bond
breakage in addition to formation. This ultimately yielded the 16 electron tricyclic
species shown.[24,25a] A long distance was observed for the bridgehead bond (C4-
C5), which was observed to exceed all of those in the previously discussed
complexes, at 1.614(7) Å. Interestingly, a related species, 7b, could be isolated
from a coupling reaction involving Me3SiC2C5H4N.[25b] In this case, the additional
coordination by a pyridine fragment was compensated for by the conversion of the
allyl coordination from  to . Despite the fact that this complex also formally has
a 16 electron configuration, one does not observe convincing structural evidence
for a C-C agostic interaction, as the C1-C6 distance in this case is 1.529(3) Å, and
no other C-C bond is longer than 1.553(4) Å. It is not clear why there would seem
to be an agostic interaction in one 16 electron complex but not the other, though
the presence of a relatively strongly donating pyridine ligand could contribute.

More recent studies have expanded upon the breadth of these coupling
reactions, leading to additional examples of complexes that may well contain
agostic coordination of C-C bonds to titanium.[25c,d] Thus, incorporation of a 1:1
combination of C2(SiMe3)2 and C2H2 into the titanocene coordination sphere led to
7c, which was not structurally characterized, but whose identity could be

This article is protected by copyright. All rights reserved.


European Journal of Inorganic Chemistry 10.1002/ejic.201600989

spectroscopically distinguished from other possibilities. Likewise, the use of a


diborylated alkyne led to a complex which was characterized as 7d.

7c 7d

The recognition of the presence of (C-C)Ti agostic interactions in the


above species correlates with the occurrence of C-C activation reactions taking
place in titanium compounds formed via coupling reactions between their 2,4-
dimethylpentadienyl and diyne ligands [diyne = (HC2)2X; X = (CH2)2-4, (CH2)2O,
(CH2)2NR for R = Me or n-Bu, and (CH2)2C(CO2R’)2 for R’ = Me or Et], e.g., 8
(Figure 6 for X = (CH2)2). These reactions led to close Ti-C contacts (e.g.,

2.731(1) Å) with some saturated carbon atoms, even though clear lengthenings of
the C-C bonds were not evident.[26] The rearrangements of these compounds
typically occurred slowly at room temperature.

In the related complex 9, isolated from the 5 + 2 + 2 coupling reaction of a


zirconium-coordinated 6,6-dmch (dmch = dimethylcyclohexadienyl) ligand and
two equivalents of C6H5C2SiMe3 (Figure 7), one does observe a slight lengthening
of two C-C bonds (those connecting the bridgehead carbon atoms to the diene
termini), to 1.567(4) and 1.574(4) Å, accompanied by Zr-C distances of 2.697 and
2.744 Å for the two saturated carbon atoms.[27] The fact that zirconium leads to
stronger agostic interactions than titanium may be correlated with the fact that

This article is protected by copyright. All rights reserved.


European Journal of Inorganic Chemistry 10.1002/ejic.201600989

zirconium tends to form stronger bonds than titanium.[28] This trend will be seen to
play a role in some other agostic species (vide infra).

Some interesting potential agostic complexes have been reported to arise


from something of an inverse approach, not unlike some of the alkyne reactions
described above, through which the carbon atoms in the ultimate agostic
interaction were initially isolated, but were then brought together to form a C-C
bond in the vicinity of an electron deficient metal center. The species in question
arose from the incorporation of Me2Si(C2SiMe3)2 units into a Ti(C5MenH5-n)2
coordination sphere (n = 3-5), leading to complexes such as 10 (Figure 8), whose
formations required the migration of the trimethylsilyl groups from one carbon
atom to another.[29] The C-C bond lengths in these species ranged from 1.902(6)-
1.933(2) Å, while the value for a Ti(C5Me4H)2[C=C(SiMe3)]2SiPh2 analogue was
1.821(4) Å. The C-C separations are clearly closer to an expected bonded contact
of ca. 1.54 Å than a non-bonded contact of 3.4 Å, making the C-C agostic
description attractive. These would appear to be the longest potentially agostic C-C
bonds to have yet been observed, and additional investigations would be highly
desirable for assessing the actual situation here.

10

This article is protected by copyright. All rights reserved.


European Journal of Inorganic Chemistry 10.1002/ejic.201600989

Given that the degrees of lengthening in the C-C bond distances in these
early examples were small enough that they might be overlooked in less well-
determined structural studies, it was suggested that other examples of agostic C-
CM complexes may have been reported even earlier, but not recognized as
such.[12] Indeed, that has turned out to be the case. As one example, in 1984 the
complex Mo(COCH3)(S2CNMe2)(CO)(PMe3)2, 11 (Figure 9), was reported as
possessing an agostic structure involving an acyl C-H bond, consistent with a Mo-
C2-C3 angle of 90.9(8)º.[30a] Subsequently, analogous
Mo(COCH3)[S2C(PMe3)(OCH2CF3)](CO)(PMe3)2 and
Mo(COCH2SiMe3)(S2CNMe2)(CO)(PMe3)2 species were reported,[30b] and found to
be in equilibrium with dihaptoacyl structures. Notably, theoretical results for the
agostic isomers indicated that the agostic interaction also involved a contribution,

11

seemingly dominant, from the C2-C3 bond. Lengthening of the C-C bonds was not
especially evident from the structural studies due to the esds being on the order of
0.01-0.02 Å, but the theoretical calculations consistently indicated the bonds to be
lengthened, to 1.585-1.595 Å.

“Pincer Complexes”

The very interesting class of metal “pincer” complexes had its origins in the
mid-1970s, predominantly through the efforts of the Shaw, van Koten, and Kaska
groups, and it is likely that over a thousand such compounds are known.[31,32] In
these species, bidentate coordination typically involving lone pairs of meta-phenyl-
tethered pairs of phosphines, amines, ethers, or thioethers leads to the metal center
being in such close proximity to the intervening phenyl carbon atom and its
substituent that oxidative addition across that bond is quite common.[33,34]

This article is protected by copyright. All rights reserved.


European Journal of Inorganic Chemistry 10.1002/ejic.201600989

However, in some situations, such as when oxidative addition is not favored,


a metal center may simply undergo formal electrophilic addition to the intervening
carbon atom. This would result in the metal center being put in very close
proximity to a C-H or even C-C bond, as in the case of complex 12 (X = I; Figure
10), whose preparation [Equation (2)] and structure were reported in 1978.[34b] At
that early date, there would have been no reason to suspect any coordination of a
C-C bond, though today such a situation, involving an electron deficient metal
center, would require some consideration of this possibility. Given that the
potential C-C bond involved in the interaction was found to have a length of
1.55(1) Å, one could not definitively claim that such an interaction was present. In
fact, a subsequent structural study on a species having X = o-tolyl revealed a
corresponding distance of 1.51(1) Å, which would seem even further out of line
with any significant agostic interaction.[34d] Similar distances of ca. 1.54(1) Å were
additionally found for some benzyl analogues. [34c]

Notably, some interesting reactivity studies carried out by the authors nicely
supported their actual formulation for these complexes as -arenium species,
analogous to [C6(CH3)7]+. Thus, reactions of 12 with small nucleophiles (Nu-; Nu
= OH, C2C6H5) led to addition to one of the ring carbon atoms ortho to the site of

(2)

12

Pt substitution, while larger nucleophiles (Nu = CH(CO2Me)2) led to addition to


the para carbon atom, in each case yielding a diene complex, 12a or 12b. With
methoxide ion, both addition pathways could be realized.[34e] Also interesting is

12a 12b

This article is protected by copyright. All rights reserved.


European Journal of Inorganic Chemistry 10.1002/ejic.201600989

the fact that reactions of 12 (X = I) with halide ions led to C-CH3 bond cleavage,
and formation of [PtX(C6H3(CH2NMe2)2)-o,o’] species (X = Cl, Br, I), via loss of
CH3I, essentially the reverse of Eq. (2), thus providing a very early example of
formal reversible C-C bond cleavage/formation.

In 1999, Milstein, et al. began reporting on related -arenium complexes of


rhodium,[7,35-38] of the general type shown as 13, having R = H or Me, R’ = H, Me,
or CH2Ph, and R” = H or Me. For these complexes, agostic (C-C)M
interactions were proposed to exist as a result of the 16 electron rhodium centers
being forced into proximity with C-C bonds. Structural data for the complex
having R = R’ = R” = Me (Figure 11), including the respective Rh-C3 and Rh-C13

13

distances of 2.354(3) and 2.817(3) Å, and the Rh-C3-C13 angle of 90.25(16)º,


could be considered to be consistent with the presence of the proposed interaction.
Asymmetry in M-C bond lengths for such agostic interactions is quite common, as
the C-C bonds may not have sufficient freedom to make an optimal approach to the
metal center. While the difference here is notable, in an earlier study of a very
similar compound for which R = H, a prominent C-H agostic interaction had been
evident.[35a] These appear to be the first proposed examples of agostic C-C
interactions for pincer complexes, and it would be of great interest to have
available more detailed spectroscopic and theoretical data to provide a clearer
picture of the nature of the interaction between their metal centers and the
proximate C-C bonds.

Subsequently, related pincer compounds, such as 14, were prepared which


incorporated a hybrid amine/phosphine ligand.[39,40] At least the cationic complex
was formulated as having a (C-C)Rh agostic interaction, as were some naphthyl
analogues, e.g., 15.[41,42]

This article is protected by copyright. All rights reserved.


European Journal of Inorganic Chemistry 10.1002/ejic.201600989

14 15

Thiophosphoryl variations of the pincer ligand motif were also found to lead
to significant differences relative to the aforementioned pincer compounds.[43,44]
Notably, the ligand incorporation was accompanied by C-C oxidative addition,
generating the expected methyl complex 16a. Although an intermediate agostic
species was not observed, a theoretical study indicated for the model complex 16b,

16a 16b

having L = formaldehyde or cis-2-butene, the involvement of a 3-(C-C-H) agostic


interaction. This may be considered analogous to the (C-C-C) interactions in
metallacyclobutanes (vide infra), and perhaps as well to complexes involving even
more than two agostic interactions, as found in some of the alkyne coupling
products discussed earlier. This also provides some support for the proposals of
agostic C-C agostic interactions in the other pincer complexes.

A related complex was obtained from the reaction of methyl triflate with
complex 17 [Equation (3)].[45] The product, 18, was not characterized structurally,
but several spectral observations led to its formulation as a (C-C) Pt agostic
complex. First, a 2J(Pt-H) coupling constant of 92 Hz. was taken as an indication
that the platinum-bound methyl group was opposite to a very weak ligand, given
that a value of 47 Hz. was found for the starting material 17. In turn, a parallel
increase in the 195Pt-13CH3 coupling constant, from 636 to 982 Hz., was observed.

This article is protected by copyright. All rights reserved.


European Journal of Inorganic Chemistry 10.1002/ejic.201600989

The 195Pt-13C coupling constants for the aryl and methyl carbon atoms of the
presumed agostic C-C bond, 14 and 56 Hz., respectively, also indicated quite
modest interactions with the metal center, consistent with the presence of a weak
agostic interaction.

(3)

17 18

Very recently another apparent (C-C) agostic complex was isolated from the
pincer strategy.[46] The complex was prepared via oxidative addition of CH3I to an
Ir(I) complex, during which a N2 ligand was displaced. Heating the initial product
brought about a subsequent trade of a methyl group for a hydrogen atom, leading
to the final product, 19 [Equation (4)], which was characterized structurally

(4)

19

and spectroscopically. Variable temperature NMR spectroscopy revealed the


hydride chemical shift to be strongly temperature dependent, which was interpreted
as an indication of an equilibrium involving agostic and non-agostic contributions.
The spectroscopic as well as some theoretical data also suggested that the agostic
interaction involved both a C-H bond from the methyl group, as well as the C-C
bond attaching it to the cyclohexyl ring. The limiting chemical shifts for the two
species were estimated, and used to determine thermodynamic parameters for their
equilibrium, with Hº = -2.71 +/- 0.04 kcal/mole and Sº = 8.0 +/- 0.2 cal/(mol-K)
corresponding to the isomerization to the agostic form. The proposed weakness of
the agostic interaction was noted to be consistent with structural data, which
showed that any lengthening of the proposed 1.553(4) agostic C-C bond to be of a
magnitude comparable to statistical uncertainties.

This article is protected by copyright. All rights reserved.


European Journal of Inorganic Chemistry 10.1002/ejic.201600989

It is interesting that the use of pincer complexes in the above examples


accomplished much the same result as obtained from the alkyne coupling
reactions, that is, forcing a metal center, ideally electron deficient, into close
proximity of at least one C-C bond. As noted earlier, additional spectroscopic and
theoretical studies of promising pincer complexes would be of great interest.

Metallacyclobutanes

As described above, the earliest reported examples of (C-C)M agostic


complexes were fairly complicated species, generated from alkyne coupling
reactions. These reactions led to carbon atom cages surrounding part of the
metals’ coordination spheres, thus favoring the agostic interactions. Somewhat
later it was recognized that electron deficient metallacyclobutane complexes could
also possess these agostic interactions, in some cases involving competitive C()-H
agostic interactions.[12,47] The presence of these interactions in much simpler
species than the alkyne coupling products greatly facilitated detailed study of the
former complexes, which led to the realization that these interactions are important
in metal-catalyzed olefin metathesis.

Indeed, some of these alkyne coupling products having (C-C)M agostic


interactions contained fragments related to titanacyclobutanes. It is therefore
notable that a theoretical study on RuCl2(PH3)2(CH2)3 and RuCl2(PH3)(CH2)3 was
carried out, using these as models for active catalyst precursors for the olefin
metathesis reaction. The results indicated that only the latter 14 electron species
contained (C-C)M agostic interactions, which were reflected by the relative
calculated C-C bond distances of 1.511 and 1.587 Å for the 16 and 14 electron
ruthenacyclobutanes, respectively. It was proposed that these interactions could
play an important role in the olefin metathesis reaction.[47]

Structural and spectroscopic studies of M(C5R5)2[(CH2)2CRR’] species (20,


Figure 12: M = Ti, Zr, Mo, W; R, R’, R’’ = H, alkyl) provided the first
unambiguous experimental support for the proposal that agostic (C-C)M
interactions were present in electron deficient metallacyclobutane species (i.e., for
M = Ti, Zr).[12] Thus, for the 16 electron titanium and zirconium complexes,
greater C-C-C’ angles, 108.4(2)-113.6(2)º vs. 98.75(15)-99.9(3)º, and
correspondingly shorter M-C distances, were observed compared to the 18

This article is protected by copyright. All rights reserved.


European Journal of Inorganic Chemistry 10.1002/ejic.201600989

20

electron Mo and W complexes. Additionally, the C-C distances for the electron
deficient complexes were longer than for the 18 electron species, 1.540(3)-1.583(7)
Å vs. 1.528(2)-1.532(4) Å. Spectroscopic data paralleled the structural results.
First of all, significant differences in the 13C chemical shift patterns were evident.
For the 18 electron species, the C resonance appeared ca. 50 ppm downfield of the
C resonances, whereas for the electron deficient complexes, the C resonances
appeared ca. 60-80 ppm upfield of those for the C atoms. As had been observed
for the alkyne coupling products, 13C-13C coupling constants, determined by
INADEQUATE NMR spectroscopy, provided especially useful insight. Thus,
while relatively normal values were found for the 18 electron complexes (32.1 Hz
for M = Mo; 31.2 Hz for M = W), dramatically lower values were found for the
electron deficient titanium (CR2 = C(H)(i-C3H7), 24.2 Hz; CR2 = C(CH3)2, 22.0
Hz) and zirconium (CR2 = C(H)(CH2CH=CH2), 20.7 Hz.) compounds. That
zirconium led to a smaller J value than titanium correlated with expectations of a
stronger interaction for zirconium.[28] Some data also indicated that the presence of
a hydrogen atom on C led to a competing (C-H)M interaction for the electron
deficient species. This was evident from the J(13C-13C) values in the two titanium
complexes, and from the lengthening observed for the C-H bond (1.075(19) Å) in
the zirconium complex. The data suggested, however, that the (C-H)M
interaction was of secondary importance due to the greater M-Cdistance. The
presence of (C-C) agostic interactions in the electron deficient complexes led to the
proposal that these interactions play an important role in olefin metathesis.
Notably, earlier titanacyclobutane structures had been reported in which C-C bond
lengths were slightly lengthened, and the potential presence of C-C agostic
interactions not surprisingly went unrecognized.[48,49] In other cases, the C-C
distances did not show obvious lengthenings, and thus did not draw attention.[50-53]

This article is protected by copyright. All rights reserved.


European Journal of Inorganic Chemistry 10.1002/ejic.201600989

This is not surprising given the relative inaccessibility of C-C bonding electron
density,[3,5-7] leading to weaker interactions than for C-H or H-H bonds, and much
more modest degrees of bond lengthening. This is compounded by the fact that a
lengthening of only 0.15 Å leads to a 50% weakening in a C-C bond.[54] While the
previously reported electron deficient species are nearly certain to contain C-C
agostic interactions, additional data would be of value.

An experimental X-ray diffraction electron density study for


Ti(C5H4CH3)2(CH2)2C(CH3)2 (20, C5R5 = C5H4CH3; R’ = R” = CH3) provided
additional data reflecting the presence of (C-C)M agostic interactions.[55] The
complex exhibited lengthened C-C bonds, with distances that averaged 1.575(1)
Å. Topological analysis of the charge density also revealed these bonds to have
the lowest electron density at their bond-critical-points (1.50(2) e/Å3). These
bonds also exhibited the least negative Laplacian value (-8.3(1) e/Å5). Ellipticity
profiles[56] for the titanacyclobutane C-C bonds demonstrated a dramatic elongation
of their electron densities toward the titanium atom, whereas any elongation was
minimal for an 18 electron molybdenum analogue.

Further examples of (C-C)M agostic metallacyclobutanes that were not


initially recognized as such involved some of the complexes that were active
catalysts for olefin metathesis. Thus, structures of two
W(NR)(OC(CF3)2R’)2(CH2CHR”CHR”) (R = 2,6-C6H3(i-Pr)2; R’ = CH3 or n-
C3F7); R” = SiMe3 or H, respectively) complexes, e.g., 21 (Figure 13), reported

21

by the Schrock group, displayed lengthened C-C bonds.[57] These species


adopted structures which are nearly trigonal bipyramidal, with the imido and one

This article is protected by copyright. All rights reserved.


European Journal of Inorganic Chemistry 10.1002/ejic.201600989

of the alkoxide ligands being axial. For the former, more accurately determined
structure, having R’ = CH3 and R” = SiMe3, these bonds averaged 1.616(11) Å.
Additionally, the W-C distances were found to be only 0.27-0.29 Å longer than
the W-C bonds, compared to ca. 0.60 Å for nonagostic situations. The proposed
presence of the agostic interactions in these species,[12] but not for complexes with
more typical alkoxide ligands, can be attributed to the electron withdrawing
abilities of the polyfluoroalkoxide ligands. As it has been found that alkoxide
ligands can serve as 5 (as well as 3 or 1) electron donors,[58] these tungsten
complexes need not be electron deficient, and thus agostic interactions would seem
unnecessary. The polyfluoroalkyl groups, however, can be expected to severely
reduce the  donor ability of the alkoxide ligands, enhancing the electron
deficiencies of their complexes, providing the driving force for the generation of
agostic interactions, and thereby leading to increased activity in olefin metathesis.

Somewhat more insightful information was subsequently reported.


Especially interesting was the observation of equilibria between the square
pyramidal and trigonal bipyramidal coordination geometries for some systems.
Thus, in the case of W(NR)(OR’)2(CH2)3 [R = 2,6-C6H3(i-Pr)2; R’ = CMe2CF3], at
-40º C the 13C NMR spectrum contained peaks for the former isomer at 43.6 ppm
for C, and 24.2 ppm for C atoms, with J(13C-13C) = 28 Hz.[59] In contrast, for the
latter isomer the respective values were found to be 98.8 ppm, -3.6 ppm, and 13
Hz. The (C)-(C) values were observed to correlate well with the W-C
distances in various related structures, such that a larger chemical shift difference
was accompanied by a shorter M-C (and also M-C) distance. Although it seems
not to have been appreciated at the time, perhaps the most significant difference
was actually that involving the J(13C-13C) values. The value of 28 Hz. is not
unusual, especially for a small ring, but that of 13 Hz. clearly stands out as being
unexpectedly small. Based on the results for agostic species described above, this
small value is clearly indicative of an agostic (C-C)W interaction in the trigonal
bipyramidal arrangement. The trigonal bipyramidal form was observed to be
disfavored when strongly electron donating alkoxide ligands were present, which
was attributed to enhanced -donor competition between the trans-positioned
ligands in a tbp arrangement. This correlates also with the observation that the tbp
isomer is the one having the agostic interaction, which can easily be attributed to
the lower basicities of its fluoro-substituted alkoxides, thus increasing the electron

This article is protected by copyright. All rights reserved.


European Journal of Inorganic Chemistry 10.1002/ejic.201600989

deficiency of the metal center. Additionally, the trigonal bipyramidal isomers are
more active in the metathesis reactions.

A structural study of the square pyramidal


W(NR)[OC(CH3)2CF3]2[CH2CH(t-Bu)CH2], 22 (R = 2,6-C6H3(i-Pr)2), revealed that
the imido ligand occupied the axial position (Figure 14), thus presumably greatly
reducing the competition for -bonding.[59] For this complex the C-C bond
distances were not significantly lengthened, being 1.55(1) and 1.57(1) Å. The W-
C distance was found to be 2.79(1) Å, as compared to the W-C distances of

22 23

2.14(1) and 2.17(1) Å. These parameters revealed a dramatic difference relative to


the trigonal bipyramidal complex 21.

A subsequent study reported a W(NR)(OR’)(2,5-Me2Pyr)(CH2)3 complex


(23, R’ = (R)-3,3’-dibromo-2’-(t-butyldimethylsiloxy)-5,5’,6,6’,7,7’,8,8’-
octahydro-1,1’- binaphthoxide) in which one of the alkoxy groups was replaced by
a 1-2,5-dimethylpyrrolide ligand (2,5-Me2Pyr).[60] Only one isomer was evident
from the NMR data, and as can be seen in Figure 15, this is the trigonal
bipyramidal isomer, having the imido and the alkoxy ligands in the axial sites. The
13
C NMR data again reflect agostic (C-C)W interactions, with J(13C values
in the range of 154-157 Hz. In accord with the spectral data, a structural study
revealed the C-C distances to be somewhat lengthened, at 1.585(7) and 1.591(7)
Å, respectively, together with W-C distances of 2.040(4) and 2.057(4) Å, and a

This article is protected by copyright. All rights reserved.


European Journal of Inorganic Chemistry 10.1002/ejic.201600989

W-C distance of 2.349(5) Å. The related 2,3,5,6-tetraphenylphenoxide complex


was also reported. A structural study revealed similar features for its
tungstacyclobutane portion, with C-C distances of 1.603(3) and 1.597(3) Å, W-
C distances of 2.057(2) and 2.058(2) Å, and a W-C distance of 2.368(2) Å.

A surprising variation in structural pattern was found for


W(NR)(OR’)(pyrrolyl)(CH2)3 (R’ = 2,6-bis(2,5-diphenylpyrrolyl)phenoxide).[61]
Although NMR data suggested the presence of a trigonal bipyramidal geometry, an
X-ray structural study revealed the complex to be square pyramidal, though
involving an entirely new arrangement, 24. In this case, the idealized axial
position was occupied by one of the tungstacyclobutane carbon atoms (Figure 16).
Particularly curious is the fact that the metallacyclobutane geometry was very
much like those in the trigonal bipyramidal structures, rather than the non-agostic
square pyramidal species (SP), leading to the new coordination environment being
designated as SP’. Especially pertinent in this regard are the clearly lengthened
C-C bond distances, 1.603(3) and 1.590(3) Å. These are accompanied by W-C
distances of 2.035(2) and 2.083(2) Å, and a W-C distance of 2.370(2) Å. Notably,
calculations for species containing aryloxide and pyrrolide ligand-containing

24

tungstacyclobutane compounds suggested that the SP’ isomer was closest to the
transition state for loss of olefin from the metallacyclobutane, while the SP isomer
was farthest. A subsequent study of W(NR)(OR’)(2,5-dimethylpyrrolyl)(CH2)3 (R
= 2,6-di(2,4,6-mesityl)phenyl; R’ = OSiPh3) revealed it to have an intermediate
geometry, being approximately 60% trigonal bipyramidal in character. [62] This
species also displayed slightly lengthened C-C bonds, of 1.593(3) and 1.600(3) Å,

This article is protected by copyright. All rights reserved.


European Journal of Inorganic Chemistry 10.1002/ejic.201600989

along with W-C distances of 2.063(2) and 2.050(2) Å, and a W-C distance of
2.366(2) Å.

Other examples, involving niobium[63] (25, Figure 17) and tantalum[64] (26,
Figure 18, Ar = 2,6-C6H3(i-Pr)2) were also formulated as (C-C) agostic species
after the fact, in addition to titanocene and (imido)tungsten complexes such as
those described above.[65,66] Some of these differed notably from the other
examples in that the agostic interactions occurred asymmetrically, with one C-C
bond being longer than the other. Thus, in 25, the C-C distances were 1.65(3)
and 1.56(3) Å, vs. 1.577(8) and 1.614(9) Å for its tantalum analogue, while

25 26

for 26, the distances were 1.546(22) and 1.553(21) Å. Theoretical calculations
estimated these bonds for 25 and 26 to range from 1.570-1.642 Å. [65] Mayer M-C
bond orders for these species, and for model structures of electron deficient
ruthenacyclobutanes, 27 and 28, related to metathesis catalysts, reflected the
presence of the (C-C)M agostic interactions. Species in which the phosphine
ligand was replaced by a more strongly donating heterocyclic carbene ligand, 2,5-
X2-2,5-C3N2H4 or 2,5-X2-2,5-C3N2H2 (X = H, F, CH3), have also been examined
theoretically, and all displayed lengthened C-C bonds (1.58-1.59 Å).[67] In contrast
to a more typical catalyst for metathesis (e.g., X = 1,3,5-mesityl), these species
could have their carbene ligands adopt either a parallel or a perpendicular
orientation to the ruthenacyclobutane ring. Either orientation led to the expected
agostic interactions, but with the perpendicular orientation of the 2,5-C3N2H6

27 28

This article is protected by copyright. All rights reserved.


European Journal of Inorganic Chemistry 10.1002/ejic.201600989

ligand, significant delocalization of the electron densities in the agostic bonding


and carbene ligand  system orbitals was found, which was neither observed for
the parallel orientation, nor for either orientation of the 2,5-C3N2H4 ligand.

The calculations for the ruthenium models suggested the presence of long
C-C bonds, and also revealed the ruthenacyclobutanes to be intermediates rather
than transition states[68] in the olefin metathesis reaction. This was proposed to be
a result of the fact that one has more than one agostic interaction, as had been
proposed earlier as the driving force for the general preference observed in some
titanium complexes for agostic vs. olefin coordination.[12] QTAIM (Quantum
Theory of Atoms in Molecules) analyses were carried out, but failed to provide
evidence for any Ru-C interactions (vide infra). At least one earlier study had also
found lengthened C-C bonds for ruthenacyclobutane intermediates in the olefin
metathesis reaction, though the presence of agostic C-C interactions did not seem
to be appreciated.[69] These calculations, however, also found transition states for
cyclopropane reductive elimination in which C-C distances of ca. 1.8-1.9 Å were
present, reminiscent of the bonding in 10.

A follow up theoretical study examined a number of agostic and nonagostic


ruthenacyclobutanes and tungstacyclobutanes.[65a] In the case of the ruthenium
complexes, both 14 and 16 electron species were examined. The 14 electron
species contained two halide ligands in addition to an amine, carbene, or phosphine
ligand, leading to five-coordinate Ru(IV) species. A trigonal bipyramidal isomer,
with apical halides and agostic interactions with the C-C bonds of the
metallacyclobutane ring, was observed as the ground state in all but one of 16
cases. In the lone exception, a square pyramidal arrangement was observed. The
16 electron complexes were six coordinate species, in which either two phosphines
or a phosphine and a carbene ligand were present, giving distorted octahedral
structures. In these cases the halide ligands adopted a cis orientation. Agostic
interactions were found in the ground state for only one of five cases. For the
agostic species studied, a nearly linear dependence was observed between the
calculated Ru-C and C-C bond orders. A QTAIM analysis of the 14 electron
species failed in all cases to show a bond path between the metal and C, though
one was observed for the 16 electron complex with an agostic ground state.

This article is protected by copyright. All rights reserved.


European Journal of Inorganic Chemistry 10.1002/ejic.201600989

Comparative calculations were carried out for tungsten analogues having


one imide ligand along with either two alkoxides or a combination of one alkoxide
with one pyrrolyl ligand. Ten species were found to adopt trigonal bipyramidal
geometries, and contained agostic interactions, while one species was found to be
square pyramidal, and lacked such interactions.[65a]

Although structural data for electron deficient ruthenacyclobutanes were not


at hand, spectroscopic data for a 14 electron species, 29, were obtained[70,71] and
revealed the unusual chemical shift and coupling constant trends that had been
observed for the 16 electron (C-C) agostic titana- and zirconacyclobutanes.[12,19]
Most telling was the observation of a J(13C-13C) value of 15.0 Hz. – much lower
than those for the 16 electron titanium and zirconium analogues, but comparable to
the 13 Hz. measured for W(NR)[OC(CH3)2CF3]2(CH2)3 (R = 2,6-C6H3(i-Pr)2).[59]
This was ascribed to a relatively flat four-membered ring with a significant degree
of M-C interaction, in accord with Schrock’s correlation of (C)-(C)

29

values with the M-C distance. The nature of the interaction was also proposed to
involve agostic (C-C)Ru bonding,[12a,71] though a subsequent MO study
concluded that it was more of a direct Ru-C interaction.[72] In addition, the C-C
bond distances were calculated to be 1.56 Å, which does not seem to indicate a
great deal of C-C activation, and might be consistent with the proposal of a direct
Ru-C interaction.

Notably, as in the cases of 25 and 26,[63-65] an unsymmetrically activated


species, 30, could also be obtained.[73] In this complex, the less substituted C-C
bond was found to have a J(13C-13C) value of 23 Hz., with the other value being 8.5
Hz. The average is close to the 15.0 Hz. value observed for 29, and the 13 Hz.
value observed for W(NR)[OC(CH3)2CF3]2(CH2)3 (R = 2,6-C6H3(i-Pr)2).[59] The
calculated C-C distances for the two bonds in 30 were 1.55 and 1.65 Å,

This article is protected by copyright. All rights reserved.


European Journal of Inorganic Chemistry 10.1002/ejic.201600989

respectively.[72] The 0.1 Å bond elongation for the latter was attributed to the ca.
6.2 kcal/mole ring strain brought about by a five-membered ring. Perhaps this
resulted in some localization of the agostic interaction in the more substituted C-C

30

bond. Spectroscopic data have also indicated the presence of C-C agostic
interactions in related 14 electron substituted ruthenacyclobutane complexes, again
having trigonal bipyramidal coordination spheres with two axial chloride
ligands.[73, 74]

Subsequently, a series of formally 16 electron molybdena- and


tungstacyclobutane complexes, based on the M(C5Me5)(NO) fragment, has been
reported by Legzdins, et al. The molybdenum complexes were prepared from the
reactions of the transient neopentylidene complex Mo(C5Me5)(NO)(CHCMe3) with
five-, six-, seven-, and eight-membered cycloalkenes,[76,77] and included species of
the general type shown below. Initially the products involved a cis orientation of
the cycloalkene-derived atoms in the metallacyclobutane ring, but for the six- and
larger-membered ring systems (31a-d, respectively, n = 1-4), conversion to the
trans isomers occurred, either in solution or in the solid state. Further
rearrangements to other species, such as allyl or diene complexes, also occurred in
some cases. Structural studies of the five- and eight-membered ring complexes

31

This article is protected by copyright. All rights reserved.


European Journal of Inorganic Chemistry 10.1002/ejic.201600989

revealed striking elongation of some of the metallacyclobutane C-C bonds. For the
former (Figure 19), as the cis isomer, the lengths were 1.561(2) and 1.693(3) Å,
while for the larger ring, as the trans isomer, they were 1.608(3) and 1.635(3) Å.
Notably, for the smaller ring, the longer, more activated bond was part of the ring
system, while the opposite was true for the larger ring. The 1.693(3) Å bond is
especially notable in that it is close to the point at which it is claimed that a C-C
bond becomes only half as strong.[54] The unusually long C-C distances correlated
well with the (C)-(C) values, which when large have been found to indicate
short M-C distances. The values for the molybdenum and tungsten compounds
were unusual, being as large as 134 ppm (cf., the already large value of ca. 92 ppm
for a ruthenacyclobutane).[70] Further, J(13C-H) values of at least 156 Hz. were
observed for the metallacyclobutane rings, consistent with a C-C agostic
interaction.[55] A related species was prepared from norbornene, and found to be
far less reactive.

At least some of the analogous tungsten complexes could also be isolated,


via the same general route, though higher temperatures were required.
Spectroscopic analysis suggested that the cyclopentene product adopted the same
cis orientation as found for the molybdenum analogue, while the six- and eight-
membered ring products appeared to adopt the trans orientation.[76,78] A subsequent
study found that the anticipated metallacyclobutane, 32, could be prepared using
2,5-dihydrofuran.[79] A structural study revealed that the expected cis orientation
was adopted, as had been the case for cyclopentene. Once again the two
metallacyclobutane C-C bonds were significantly lengthened, at 1.659(3) and
1.586(3) Å, the longer one again residing in the five-membered ring. More
complicated chemistry was observed for 3,4-dihydro-2H-pyran and 1,2,3,6-
tetrahydropyridine.

32

This article is protected by copyright. All rights reserved.


European Journal of Inorganic Chemistry 10.1002/ejic.201600989

Agostic Interactions Involving Cyclopropyl Rings

Another class of (C-C) agostic complexes has been discovered, in which the
C-C bonds of strained cyclopropyl groups are involved. The strain enhances the
chances of such an interaction both due to spatial and electronic reasons. The 60
degree angles within the ring result in the C-C bond density being projected
outward from the ring, and thereby their electron densities can much more
efficiently interact with a metal center. Further, the strain in the ring increases the
energies of the C-C bonding orbitals, leading them to be better donors
electronically, while the antibonding orbitals become better acceptors. In fact, as
noted above, in 1996 a lithium cyclopropoxide complex (3, Figure 1) was found to
have an interaction between a C-C bond and the formal lithium ion.[18]

As noted earlier, it seems ironic that the first proposal of a metal complex
involving coordination by a C-C bond involved platinum compounds obtained
from the reaction of H2PtCl6 with cyclopropane, which instead had yielded
platinacyclobutanes, e.g., 2.[13] Given the substantial strain in a cyclopropane
molecule, which could be greatly reduced via oxidative addition reactions, this
would seem to greatly reduce the chances of isolating complexes containing
agostic interactions between cyclopropyl groups’ C-C bonds and metal centers.
However, for transition metals with d0 or possibly even d1, configurations, or at
least for metals in high enough oxidation states for which oxidative addition would
be unfavorable, these agostic complexes might indeed be stabilized. One such
example suggested by theoretical studies involves the potential intermediacy of an
agostic C-C interaction with cyclopropane in a -hydride abstraction reaction of
Zr(C5H5)2(c-C3H5)2.[80]

The first such isolable complex to be reported was the formal Nb(III)
complex, Nb(Tp’)Cl(c-C3H5)(MeC2Me) (33, Tp’ = hexamethylpyrazolylborate)
(Figure 20).[81-83] The purported 16 electron complex showed no indication of an
agostic interaction between the metal center and any hydrogen atoms, but instead
structural data revealed an interaction with the coordinated C1-C3 bond. This can
be seen from the Nb-C1-C(2,3) angles, 131.4(2) and 109.7(2), and the C1-C(2,3)
distances, 1.490(4) and 1.539(4) Å (vs. C2-C3, 1.478(5) Å). While in general it
would be unusual that a C-C bond could be a more effective donor than a C-H

This article is protected by copyright. All rights reserved.


European Journal of Inorganic Chemistry 10.1002/ejic.201600989

33

bond, the above-mentioned outward projection and strain of cyclopropyl C-C


bonds, which would make them higher in energy, would together make these bonds
better donors, as confirmed theoretically. For purposes of comparison, related
cyclobutyl, cyclopentyl and cyclohexyl analogues were synthesized, and found
spectroscopically to be primarily  C-H agostic species, though minor rotamers
were observed, believed to be  C-H agostic species, which became more abundant
with increasing ring size. Theoretical studies were consistent with the observed
relative favorabilities for the agostic interactions found in the cyclopropyl,
cyclobutyl, and cyclopentyl complexes, and were also in good accord with their
structural data. The cyclohexyl complex appeared thermally unstable, paralleling
observations on electron deficient titanium and zirconium (8,9) complexes having
organic 9-membered rings, serving as both diene and allyl ligands, fused to other
rings varying from four to six carbon atoms.[26]

A subsequent structural study[84,85] of Nb(Tp’)(C6H5)(c-C3H5)(MeC2Me) was


unable to establish whether there was again an agostic C-C interaction, but
INADEQUATE NMR studies did reveal that for this species, as well as its Cl
(above) and CH3 analogues, one C-C bond in the cyclopropyl ligand possessed a
13
C-13C coupling constant that could be concluded to have a value less than 3 Hz.,
as compared to values of about 9-15 Hz. for the other bonds. The authors were
able to show that DFT studies could reasonably reproduce these coupling
constants, and values for the C-C bonds involved in the agostic interactions were
all calculated to be less than 3 Hz. Due to potentially serious difficulties in
obtaining very small C-C coupling constants experimentally, theoretical studies
may provide a convenient alternative for assessing the presence of C-C agostic
interactions in these types of structures. Structural data for a mesityl analogue

This article is protected by copyright. All rights reserved.


European Journal of Inorganic Chemistry 10.1002/ejic.201600989

have been reported,[86] which again appeared to reflect the presence of a C-C
agostic interaction involving the cyclopropyl group.

Additional cases of cyclopropyl-derived C-C agostic interactions were found


in rhodium and iridium complexes related to BINOR-S, derived from the
dimerization of norbornadiene. In these complexes, such as 34 (Figure 21) and 35,
typically one of the cyclopropane rings in BINOR-S had undergone oxidative
addition to the metal, making the ligand a formal dianion. One series of complexes
with the general formula [Rh(PR3)(C14H16)]+ (R = i-C3H7, c-C5H9, c-C6H11),
incorporated the relatively non-coordinating B[3,5-C6H3(CF3)2]4- as the

34 35

counterion, and thus may be regarded as being Rh(III)[87,88] complexes. The


formation of the rhodacyclobutane fragment brought the metal center into close
proximity to the remaining intact cyclopropane ring, bringing about C-C agostic
coordination by one of the C-C bonds. The formal donation of the two extra
electrons would lead to a 14 electron complex, though what appeared to be much
weaker C-H agostic interactions might also be present. Each structure displayed a
lengthened C-C bond (1.604(4), 1.607(4), and 1.608(3) Å) as compared to the non-
interacting bonds, whose lengths ranged from 1.502(4)-1.521(4) Å. The Rh-C
single bond distances ranged from 2.025(2)-2.042(3) Å, whereas those involving
the agostic carbon atoms ranged from 2.349(3)-2.387(2) Å. While two cyclopropyl
hydrogen atoms were judged to be close enough to the metal center for C-H
agostic interactions to take place, there was no evidence for these interactions, in
line with the results for the niobium complexes described above.

In contrast to results for some other agostic species, whether involving C-H
or C-C bonds, a theoretical study found bond paths between the rhodium center

This article is protected by copyright. All rights reserved.


European Journal of Inorganic Chemistry 10.1002/ejic.201600989

and both agostic carbon atoms in the i-Pr complex. However, a subsequent
experimental electron density study[89] of the R = c-C6H11 complex, as the
[HCB11Me11]- salt, only revealed one such path, though indications from ellipticity
and other parameters were consistent with agostic interactions involving both
carbon atoms. The outward orientation of cyclopropyl groups’ C-C bonding
electron density would seem likely to greatly facilitate the formation of bond paths,
much as bond paths have even shown up in cases involving repulsive interactions
(vide infra). The fact that one bond path did not show up even for an agostic
interaction with a cyclopropyl ring, predisposed both spatially and energetically to
such interactions, does not generate great confidence in the QTAIM approach for
assessing these interactions in general. However, a theoretical electron density
study of this same complex did find bond paths for both carbon atoms engaged in
the expected C-C agostic interaction.

Variable temperature (VT) NMR studies of the R = i-Pr rhodium complex


demonstrated that a fluxional process occurred through which the two sides of the
C14H16 ligand became equivalent, leading to oxidative addition of the agostic
cyclopropane ring, accompanied by reductive elimination of the formal carbanions
of the rhodacyclobutane fragment. Lineshape analysis revealed a Hvalue of
9.6(2) kcal/mole, S value of -4.5(10) cal/mole-K, and G value of 10.7(10)
kcal/mole (at 298 K).[88] Based on theoretical calculations, a pathway involving a
Rh(V) intermediate was favored for this process.

Studies aimed at isolating additional examples of these rhodium complexes


led to new species with P(t-C4H9)2(CH3) and P(c-C6H11)2(C6H5) ligands, but
attempts with either larger or smaller phosphines resulted in decomposition.[90] It
was then proposed that these complexes were only stable when the phosphine’s
cone angle fell in the range of 160-170º. The two new complexes could again be
formulated as possessing agostic C-C interactions based on lengthened C-C bonds
(1.608(10) and 1.616(5) Å, respectively). It is notable that these BINOR-S
complexes did not display any agostic interactions in their rhodacyclobutane
fragments, despite their electron deficiencies. This is consistent with the authors’
earlier study of the formally 16 electron (chloro)bis(phosphine)rhodacyclobutane
complex, 36, which had C-C bond lengths of 1.521(3) and 1.524(3) Å, thus also
providing no indication of a C-C agostic interaction.[91]

This article is protected by copyright. All rights reserved.


European Journal of Inorganic Chemistry 10.1002/ejic.201600989

36

Subsequently the related iridium complex [Ir(PR3)(C14H16)]+ (R = i-C3H7), as


the B(C6H3(CF3)2)4- salt, was reported[92]. As with the rhodium analogues such as
34, fluxionality could be observed in the NMR spectra, leading the C14H16 ligand to
display C2v symmetry. As this higher symmetry was observed even at lower
temperatures, the barrier to the process would have to be smaller, in accord with
expectations that an Ir(V) intermediate would be more accessible than Rh(V),
which was implicated as being involved in the previous fluxional process.
Theoretical calculations were in accord with this, yielding an estimate of the
activation energy of only about 1 kcal/mole. In addition, significant changes in
chemical shifts were observed for the iridium, but not rhodium, complex as a
function of temperature, and this was attributed to an observable equilibrium
between the Ir(III) and Ir(V) species. Structural studies for 35 revealed the
compound to be isomorphous with its rhodium analogue, and also demonstrated
significant changes with temperature (100, 150, 200, and 250 K). Thus, the C8-
C10 distance increased from 1.704(5) Å at 100 K to 1.833(13) Å at 250 K, while
the Ir-P distance increased from 2.2884(8) Å to 2.3545(15) Å. The Ir-C1-C2-C3
fragment did not show significant structural changes with temperature, remaining
an iridacyclobutane throughout, with reasonably constant Ir-C(1,3) and C1-C3
distances (ca. 2.05 and 2.20 Å, respectively).

However, accompanying the increase in the C8-C10 separation with


temperature was a decrease in the Ir-C8 and Ir-C10 distances, consistent with the
proposal that the second species in the equilibrium was a formal Ir(V) tetraalkyl,
with two iridacyclobutane fragments. The changes taking place for the C8-C10
fragment could be modeled as a two component disorder, and the Ir(V) component
could then be estimated to increase in abundance from 17(2)% at 100 K to 45(2)%
at 250 K. From these data, the thermodynamic parameters were found to be H =
0.461(7) kcal/mole and S = 1.48(5) cal/mole-K. One unexplained point relates to
the relative Ir-C8 and Ir-C10 distances from the data in which the two components
were modeled separately. For the primary Ir(III) component, the Ir-C8 distances

This article is protected by copyright. All rights reserved.


European Journal of Inorganic Chemistry 10.1002/ejic.201600989

tended to be shorter than those for the less abundant Ir(V) species, whereas the
reverse was true for the Ir-C10 distances. At 250 K, the components were in
nearly equal abundance, and the Ir-C8 distances were 2.18(2) and 2.27(3) Å for the
Ir(III) and Ir(V) species, respectively, whereas the corresponding Ir-C10 distances
were 2.26(2) and 2.14(2) Å. In addition, while the Ir-C(1,3) distances averaged
about 2.05 Å, the Ir-C(8,10) distances for the purported Ir(V) species were
substantially longer, ranging from ca. 2.14-2.31 Å. Furthermore, the Ir-C9
distance of ca. 2.99 Å in the Ir(V) component appeared quite different from the Ir-
C2 distance of ca. 2.66 Å. It is not clear if these inconsistencies were a reflection
of unappreciated complications in the disorder or in the actual nature of the
individual species.

37

An interesting complex with two cyclopropyl groups bridging yttrium and


lithium, [Y(C5Me5)2(-c-C3H5)2Li(THF)], 37 (Figure 22), has been reported,[93] and
found to have agostic interactions involving both metals, each with a different
cyclopropyl ligand. For the lithium ion, the tilt of the organic ring toward the
metal could be judged from the Li-C3 and Li-C5 distances (2.473(11) and 3.01 Å),
as well as from the Li-C1-C3 and Li-C4-C5 angles (82.9(4)º and 110.3(5)º). The
C1-C3 bond appeared to be longer than the C1-C2 and C2-C3 bonds (1.546(8) vs.
1.515(7) and 1.474(8) Å), as compared to the agostic length of 1.519(3) Å in 3, but
that did not translate to a significant difference in the Li-C1 and Li-C4 distances
(2.131(11) vs. 2.127(11) Å). For yttrium, the trends were generally similar. Thus,
the cyclopropyl ligand tilt was evidenced by the Y-C4 and C6 distances (2.432(5)
and 2.910(5) Å) and the Y-C4-C6 and Y-C1-C2 angles (91.4(3) and 127.4(4)º).
The C4-C6 bond appeared to be longer than the C4-C5 and C5-C6 bonds (1.539(7)
vs. 1.514(7) and 1.486(7) Å), though the differences are subject to significant
statistical uncertainties. In contrast to lithium, however, there does seem to be a

This article is protected by copyright. All rights reserved.


European Journal of Inorganic Chemistry 10.1002/ejic.201600989

perceptible difference between the Y-C1 and Y-C4 distances, 2.477(5) and
2.432(5) Å. Perhaps the best indication for a stronger interaction with yttrium can
be seen in the differing cyclopropyl ligand tilts, as estimated from the differing Li-
C-C’ (tilt, 27.4º) and Y-C-C’ angles (tilt, 36.0º) given above.

VT NMR studies revealed a facile fluctional process leading the two


cyclopropyl ligands to be equivalent down to 183 K, consistent with theoretical
calculations that indicated similar energies for the symmetric, nonagostic and the
unsymmetric, agostic structures. Although the 13C-13C coupling constants were not
determined experimentally, the theoretical calculations did reveal significant
reductions compared to cyclopropanes, especially for the C4-C6 bond, whose
interaction involved the yttrium center. The calculations also indicated the
interaction with lithium to be primarily electrostatic, with that for yttrium being
markedly covalent. In addition, a significant -CH agostic interaction, involving
yttrium and a hydrogen atom on C6, was indicated, which had not been apparent
from the other data.

While not a cyclopropyl complex, there is an example of a deltate ion


(C3O32-) complex in which an agostic interaction involves a strained carbon-carbon
bond in its three-membered ring. Thus, the reaction of the U(III) complex U[1,4-
-C8H6(Si(i-Pr)3)2](C5Me5)(THF) with CO has been found to give rise to the
dimeric U(IV) complex 38 (Figure 23), which has incorporated three CO

38 (R = Si(i-Pr)3)

equivalents as the formal deltate ion.[94] The coordination of this ion involved one
oxygen atom being bound to one of the uranium centers, while the other uranium
center experienced coordination by the other two oxygen atoms, as well as by the

This article is protected by copyright. All rights reserved.


European Journal of Inorganic Chemistry 10.1002/ejic.201600989

C-C bond linking them. This bond was found to have a length of 1.436(7) Å,
longer than the other two adjacent C-C bonds, 1.377(6) and 1.381(6) Å. Spin-
restricted DFT calculations led to an optimized structure similar to that observed.
The C-C agostic interaction was found to be involved as the principal component
arising from the interaction of one of the deltate dianion’s filled molecular orbitals
with a uranium center. At room temperature a fluctional process, by which the two
uranium centers would switch environments, was found to be rapid on the NMR
timescale.

Assessment of C-C Agostic Interactions

In general, agostic interactions between C-C bonds and metal centers tend to
be more difficult to recognize than their C-H or H-H counterparts. Part of this is
due to a weaker interaction expected for the C-C bonds, as compared to C-H or H-
H bonds, leading to smaller changes in bond distances; furthermore, the
measurements of 13C-13C coupling constants, which should decrease dramatically
from agostic C-C coordination, are made difficult by the low abundance of these
nuclei. Nonetheless, either through the use of labeled compounds[12,70] or
INADEQUATE NMR spectroscopy, these values may be obtained for C-C agostic
complexes.[12,19,84] In typical aliphatic compounds, a 13C-13C coupling constant will
normally be in the range of about 30-40 Hz., though ca. 28-29 Hz. for
cyclobutanes, and 12.4 Hz. for cyclopropane.[95] However, an indirect indication
of the interaction might be obtained from the J(13C-H) values for the particular
carbon atoms. As in the case of C-H agostic interactions, the weakening of the
coordinated sigma bond leads to strengthening of any other bonds for the involved
atoms. Thus, if there are hydrogen atoms attached to at least one of the two carbon
atoms, one typically observes anomalously high J(13C-H) values (cf., 4, 5, 31).

It has also been suggested that there should be some evidence of an empty
metal orbital properly disposed to overlap with the C-C bond in question, and that
there should be some evidence of weakening of the C-C bond.[96] The weakening
might be recognized from a theoretical study, as suggested by the authors, though
careful X-ray diffraction studies as well as 13C NMR studies, can also be
employed. Ideally, every one of these techniques should be applied, if at all
possible. Some use of topological analysis of electron density distributions has
also been made, but the results often were inconsistent with the other methods

This article is protected by copyright. All rights reserved.


European Journal of Inorganic Chemistry 10.1002/ejic.201600989

outlined above. In particular, bond paths have not always been evident in species
which by all other standards meet the criteria above for having one or more C-C
agostic interactions, whether the analysis was applied to experimental or
theoretical electron density data. Although the QTAIM (Quantum Theory of
Atoms in Molecules[97]) method has proven to be of great value in helping to
understand many types of chemical bonds, it is important to recognize any
limitations there may be, and then attempt to apply the method accordingly.

The QTAIM approach has generated cases in which bonds were found when
they should not exist, as well as cases in which bonds were not found but should
exist. As examples of the first case, the QTAIM approach points to anion/anion
bonding in a variety of halide salts,[98] and bonding between Cs+ ions in CsSrF3 or
CsBaF3. Likewise, QTAIM bonds are found between oxygen atoms of adjacent
nitro groups in the C(NO2)3- ion,[99] and as well between hydrogen atoms forced to
be in close proximity to one another, as in polycyclic aromatics,[100,101] as well as
between halogen atoms in polyhalogenated organic molecules,[102] between oxygen
and fluorine atoms in protonated fluoroacetones,[103] and between nitrogen atoms in
diaminoguanidinium cations.[104] It may be that in these situations, the systems are
working to reduce repulsions by having the interacting electron densities
concentrate themselves more between the formally nonbonded pairs of atoms,
thereby mimicking chemical bonds and generating the bond paths, despite the
interaction being unfavorable.

Structures in which expected chemical bonding is not perceived via the


QTAIM approach are even more difficult to reconcile. Numerous examples such
as some Ga2H4 isomers[105] as well as various classical organometallic complexes
are known,[106,107] but more pertinent here are other examples which involve agostic
bonding. Especially relevant is the observation made by Popelier, et al., regarding
the formal CH3TiCl2 cation[108]: “In going from the staggered to the eclipsed
conformation, the C-H3 bond becomes longer, the Ti-C-H3 angle reduces to
almost 90º, the b value is reduced, and the electron population of H3 increases.
These effects are collectively reminiscent of the formation of an agostic bond
(potentially between Ti and H3), although there is clearly no BCP present. So
although both geometrical effects and AIM properties point in the direction of an
-agostic interaction, we do not find a bond expressing this interaction”. In fact,

This article is protected by copyright. All rights reserved.


European Journal of Inorganic Chemistry 10.1002/ejic.201600989

Crabtree, et al. have observed that close approaches of a metal to a C-H bond were
only found when the metal was electron deficient, thus providing a clear indication
of the presence of agostic interactions in such situations as this.[8] Similar
observations were made for the known compound Ti(C2H5)Cl3(dmpe) (dmpe =
1,2-bis(dimethylphosphino)ethane). Despite the fact that one of the C-H bonds
was bent dramatically toward the Ti atom, only weak QTAIM evidence for a bond
was seen; however, substantial curvature of the Ti-C bond path was observed, and
it was proposed that this could serve as a general and more robust indicator of -
agostic interactions.[56,109] In addition, the agostic C-H bond was bent away from
the metal. Another useful proposal came out of a study of a 16 electron
[Ir(C5H4MeEt)(PPh3)2]+ complex,[1b, 110] which despite its electron deficiency and a
C-C bond proximate to the metal center, was concluded not to have an agostic
interaction.[96] This led to the suggestion that support for the existence of a
potential C-C agostic interaction should include verification of the presence of an
empty metal orbital capable of overlap with the C-C bond(s) in question, as well as
some evidence of weakening of the C-C bond(s).

A similar conflict has also been observed for a complex containing an


apparent C-N agostic interaction. Thus, the formally 16 electron complex, Zr(2,4-
C7H11)[(i-Pr)NC(H)(Ph)CH2C(Me)=C(H)C(Me)=CH2] (C7H11 =
dimethylpentadienyl), 39 (Figure 24)[111] exhibits a Zr-N(sp2)-C(i-Pr) angle of
102.87(4)º, much less than the minimal value of 120º expected, and much lower
than values such as 117.64(9)º in Ti(C5H5)[(i-
Pr)NC(H)(Ph)CH2C(Me)=C(H)C(Me)=CH2][112] or 130.4(4)º for Ti[C5H3(t-
C4H9)2][(CH3)NC(H)(Ph)CH2C(Me)=C(H)C(Me)=CH2] (40, Figure 25). By
QTAIM the zirconium center was found to form just three bond paths, one to a
terminal diene carbon atom, one to the central carbon atom of the dienyl ligand,

39 40

This article is protected by copyright. All rights reserved.


European Journal of Inorganic Chemistry 10.1002/ejic.201600989

and one to the amide nitrogen atom. Even beyond missing a number of Zr-C bond
paths, the QTAIM model found no bond to account for the substantial distortion of
the amide i-propyl group toward zirconium, reminiscent of the observation for the
CH3TiCl2 cation mentioned above. Furthermore, were the zirconium atom bound
to the dienyl ligand only by a bond to the central carbon atom, the geometrical
arrangement should be something like 41, whereas a 70º rotation of the Zr-C

41

vector has taken place, so that the metal center has been brought into bonding
proximity with the other four dienyl carbon atoms. Indeed, electron density plots
and ellipticities[111] both pointed to the presence of Zr-C bonding interactions with
all nine unsaturated dienyl and diene carbon atoms.

The nature of the agostic interaction in this complex remains to be clarified,


though the J[13C(H)(CH3)2] coupling constant, 128 Hz., is quite normal. There is a
noticeable decrease in the [15N-13CH(CH3)2] coupling constant as compared to that
in a related titanium complex, Ti(C5H5)[(i-
Pr)NC(H)(Ph)CH2C(Me)=C(H)C(Me)=CH2] (1.5 vs. 2.1 Hz.).[12b] For comparison,
the value for the precursor imine, (Ph)(H)C=15N-13CH(CH3)2, is 1.3 Hz, consistent
with an earlier estimate of <3 Hz[113]. Preliminary theoretical data, however,
suggest a more complicated agostic situation than might be expected.
Interestingly, the presence of a (C-N)Y agostic interaction has been proposed in
Y(C5Me5)(o-C6H4CH2NMe2)2,[114] but seems to have been unappreciated in the
literature. In this species, a small C-N-Y angle of 101.6(4)º, short Y-C (3.202(8)
Å) and Y-N (2.506(6) vs. 2.568(5) Å) bonds, and a lengthened C-N bond (1.55(1)
Å) all appear to reflect such an interaction. The authors noted that a previously
reported structure, that of Y(C5H5)2(o-C6H4CH2NMe2),[115] also exhibited related
structural characteristics, presumably due again to a (C-N) Y agostic interaction.
Other examples involving boron and silicon have been found as well.[116,117]

This article is protected by copyright. All rights reserved.


European Journal of Inorganic Chemistry 10.1002/ejic.201600989

In contrast to the above agostic complexes, (C-C)M bond paths have


sometimes appeared for interactions involving cyclopropyl groups. Given that
cyclopropyl bonding electron density is directed outward from the ring, it appears
reasonable that this, in addition to the energy changes for the C-C bonding and
antibonding orbitals due to ring strain, would lead to greater buildup of electron
density between a metal and the C-C bond. In addition, the cyclopropyl cases have
thus far involved a metal interacting with just one C-C bond, rather than with
several as has otherwise been more common, so that the bonding is not diluted in
two or more locations.

Despite some shortcomings in applying the QTAIM method to agostic


bonding interactions, it has been well demonstrated that electron density
distributions, including bond ellipticities, may provide valuable assessments
regarding the presence or absence of C-C agostic bonding. While these data
should be readily available from theoretical studies, it should also be of interest to
obtain experimental data for comparison and confirmation. With continued
advances in both the hardware and the software needed for these studies, it can be
hoped if not expected that detailed experimental studies will also become fairly
routine in the near future.

This article is protected by copyright. All rights reserved.


European Journal of Inorganic Chemistry 10.1002/ejic.201600989

Summary

Although the prospects for metal coordination by C-C single bonds might
have not appeared good initially, their existence as well as their importance has
now been well-established. These unusual bonds play important roles in both
olefin metathesis and polymerization reactions, not to mention C-C bond
activations in general. It is interesting that a number of compounds with these
interactions have been known for some time, while the presence of the interactions
went unrecognized, in large part due to their relatively small C-C bond
lengthenings, as compared to the bond distance esds. In some cases their presence
might be considered adventitious, arising from a forced proximity of C-C bonds to
the metal center, whereas in other cases, one sees a ligand distortion take place to
bring the C-C bond in closer to the metal center. It seems worth noting that C-C
agostic interactions have been found not only for transition metals, but also for
main group, lanthanide, and actinide metals as well. That all these types would be
recognized in such a short time speaks well for the future of this field, and many
more exciting discoveries can be expected.

Acknowledgment

Our research on C-C agostic bonding was most recently supported by grants
from the NSF (GOALI, CHE-9617268) and the Petroleum Research Fund,
administered by the American Chemical Society (39071-AC3). BGH would like
to thank the Office of Naval Research Independent Laboratory Internal Research
Program for support during the preparation of the article.

This article is protected by copyright. All rights reserved.


European Journal of Inorganic Chemistry 10.1002/ejic.201600989

TOC Entry
In contrast to the discoveries of agostic C-C Agostic Complexes
C-H and H-H bond coordination to Benjamin G. Harvey and
transition metal centers, analogous Richard D. Ernst
coordination involving C-C bonds has
only been more recently discovered. Transition Metal
Complexes
Already they have been found to play With (C-C)M Agostic
key roles in both olefin metathesis and Interactions
polymerization reactions. This review
summarizes the early work in this area,
focusing on the nature of these species,
and the recognition that such complexes
had been around but unrecognized for
some time.

This article is protected by copyright. All rights reserved.


European Journal of Inorganic Chemistry 10.1002/ejic.201600989

References

1. a) M. Brookhart, M. L. H. Green, G. Parkin, Proc. Natl. Acad. Sci. 2007,


104, 6908-6914; b) R. H. Crabtree, Angew. Chem. Int. Ed. 1993, 32, 789-
805; C. Hall, R. N. Perutz, Chem. Rev. 1996, 96, 3125-3146.

2. a) G. J. Kubas, “Metal Dihydrogen and -Bond Complexes”, Modern


Inorganic Chemistry Series, J. P. Fackler, Ed., Kluwer Academic, 2001; b)
G. J. Kubas, “Dihydrogen and other σ Bond Complexes”, Comprehensive
Organometallic Chemistry III, D. M. P. Mingos, R. H. Crabtree, Eds., 2007,
1, 671-698.

3. For an interesting review correlating activation and cleavage of C-C single


bonds, see K. Ruhland, Eur. J. Org. Chem. 2012, 2683-2706.

4. a) R. Hoffmann, P. Hofmann, 1976, 98, 2784-2797; b) C. M. Older, J. M.


Stryker, J. Am. Chem. Soc. 2000, 122, 2784-2797.

5. a) D. M. Adams, J. Chatt, R. G. Guy, N. Sheppard, J. Chem. Soc. 1961,


738-742; b) R. D. Gillard, M. Keeton, R. Mason, M. F. Pilbrow, D. R Russell,
J. Organometal. Chem. 1971, 33, 247-258; c) R. Z. Hinrichs, J. J. Schroden,
H. F. Davis, J. Am. Chem. Soc. 2003, 125, 860-861.

6. M. R. A. Blomberg, P. E. M. Siegbahn, U. Nagashima, J. Wennerberg, J.


Am. Chem. Soc. 1991, 113, 424-433.

7. B. Rybtchinski, D. Milstein, Angew. Chem. Int. Ed. 1999, 38, 870-883.

8. R. H. Crabtree, E. M. Holt, M. Lavin, S. M. Morehouse, Inorg. Chem. 1985,


24, 1986-1992.

9. N. Koga, K. Morokuma, J. Am. Chem. Soc. 1988, 110, 108-112.

10. H. Kawamura-Kuibayashi, N. Koga, K. Morokuma, J. Am. Chem. Soc. 1992,


114, 2359-2366.

11. See also G. Lanza, I. L. Fragala, T. J. Marks, J. Am. Chem. Soc. 2000, 122,
12764-12777.

This article is protected by copyright. All rights reserved.


European Journal of Inorganic Chemistry 10.1002/ejic.201600989

12. a) B. G. Harvey, C. L. Mayne, A. M. Arif, R. Tomaszewski, R. D. Ernst, J.


Am. Chem. Soc. 2005, 127, 16426-15435 (corrn.: 2006, 128, 1770); b)
Benjamin G. Harvey, Ph.D. thesis, University of Utah, 2005; c) R. D. Ernst,
Comments Inorg. Chem. 1999, 21, 285-325; d) R. D. Ernst, “Metal
Pentadienyl Chemistry”, Tobinstock, Evanston, Ill., Nov. 7, 2004.

13. C. F. H. Tipper, J. Chem. Soc. 1955, 2045-2046.

14. C. Rodriguez-Garcia, A. Oliva, R. M. Ortuño, V. Branchadell, J. Am. Chem.


Soc. 2001, 123, 6157-6163.

15. C. P. Casey, D. M. Scheck, A. Shusterman, J. Am. Chem. Soc. 1979, 101,


4233-4236.

16. T. H. Johnson, J. Org. Chem. 1979, 44, 1356-1358.

17. R. J. Al-Essa, R. J. Puddephatt, P. J. Thompson, C. F. H. Tipper, J. Am.


Chem. Soc. 1980, 102, 7546-7553.

18. B. Goldfuss, P. von R. Schleyer, F. Hampel, J. Am Chem. Soc. 1996, 118,


12183-12189.

19. R. Tomaszewski, I. Hyla-Kryspin, C. L. Mayne, A. M. Arif, R. Gleiter, R.


D. Ernst, J. Am. Chem. Soc. 1998, 120, 2959-2960.

20. A. V. Rivera, G. M. Sheldrick, Acta Crystallogr. 1978, 34B, 1716-1718.

21. J. Blümel, N. Hertkorn, B. Kanellakopulos, F. Köhler, J. Lachmann, G.


Müller, F. E. Wagner, Organometallics 1993, 12, 3896-3905.

22. H.-J. Chung, J. B. Sheridan, M. L. Coté, R. A. Lalancette, Organometallics


1996, 15, 4575-4585.

23. B. G. Harvey, A. M. Arif, R. D. Ernst, J. Organometal. Chem. 2006, 691,


5211-5217.

24. U. Rosenthal, P.-M. Pellny, F. G. Kirchbauer, V. V. Burlakov, Acc. Chem.


Res. 2000, 33, 119-129.

This article is protected by copyright. All rights reserved.


European Journal of Inorganic Chemistry 10.1002/ejic.201600989

25 a) A. Tillack, W. Baumann, A. Ohff, C. Lefeber, A. Spannenberg, R.


Kempe, U. Rosenthal, J. Organometal. Chem. 1996, 520, 187-193; b)
U. Rosenthal, C. Lefeber, P. Arndt, A. Tillack, W. Baumann, R. Kempe, V.
V. Burlakov, J. Organometal. Chem. 1995, 503, 221-223; c) D. Thomas, N.
Peulecke, V. V. Burlakov, B. Heller, W. Baumann, A. Spannenberg, R.
Kempe, U. Rosenthal, R. Beckhaus, Z. Anorg. Allg. Chem. 1998, 624, 919-
924; d) K. Altenberger, P. Arndt, A. Spannenberg, W. Baumann, U.
Rosenthal, Eur. J. Inorg. Chem. 2013, 3200-3205.

26. a) A. M. Wilson, A. L. Rheingold, T. E. Waldman, M. Klein, F. G. West, R.


D. Ernst, J. Organometal. Chem. 2009, 694, 1112-1121; b) A. M. Wilson,
T. E. Waldman, A. L. Rheingold, R. D. Ernst, J. Am. Chem. Soc. 1992, 114,
6252-6254.

27. B. G. Harvey, A. M. Arif, R. D. Ernst, J. Mol. Struct. 2008, 890, 107-111.

28. D. J. Cardin, M. F. Lappert, C. L. Raston, Chemistry of Organo-zirconium


and –hafnium Componds; Ellis Horwood: Chichester, UK, 1986.

29. M. Horáček, P. Štěpnička, R. Gyepes, I. Císařová, J. Kubišta, L. Lukešová,


P. Meunier, K. Mach, Organometallics 2005, 24, 6094-6103.

30 a) E. Carmona, L. Sánchez, J. M. Marΐn, M. L. Poveda, J. L. Atwood, R. D.


Priester, R. D. Rogers, J. Am. Chem. Soc. 1984, 106, 3214-3222; b) G.
Ujaque, F. Maseras, A. Lledós, L. Contreras, A. Pizzano, D. Rodewald, L.
Sánchez, E. Carmona, A. Monge, C. Ruiz, Organometallics 1999, 18, 3294-
3305.

31. G. v. Koten, Top. Organomet. Chem., 2013, 40, 1-20.

32. P. A. Chase, R. A. Gossage, G. v. Koten, Top. Organomet. Chem. 2016, 54,


1-15.

33. M. Gozin, A. Weisman, Y. Ben-David, D. Milstein, Nature, 1993, 364, 699-


701.

This article is protected by copyright. All rights reserved.


European Journal of Inorganic Chemistry 10.1002/ejic.201600989

34. a) D. M. Grove, G. v. Koten, J. N. Louwen, J. G. Noltes, A. L. Spek, H. J. C.


Ubbels, J. Am. Chem. Soc. 1982, 104, 6609-6616; b) G. v. Koten, K.
Timmer, J. G. Noltes, J. Chem. Soc., Chem. Comm. 1978, 250-252; c) M.
Albrecht, A. L. Spek, G. van Koten, J. Am. Chem. Soc. 2001, 123, 7233-
7246; d) J. Terheijden, G. v. Koten, I. C. Vinke, A. L. Spek, J. Am. Chem.
Soc. 1985, 107, 2891-2898; e) D. M. Grove, G. v. Koten, H. J. C. Ubbels,
Organometallics 1982, 1, 1366-1370.

35. a) A. Vigalok, B. Rybtchinski, L. J. W. Shimon, Y. Ben-David, D. Milstein,


Organometallics 1999, 18, 895-905; b) A. Vigalok, O. Uzan, L. J. W.
Shimon, Y. Ben-David, J. M. L. Martin, D. Milstein, J. Am. Chem. Soc.
1998, 120, 12539-12544.

36. A. Vigalok, D. Milstein, Organometallics 2000, 19, 2341-2345.

37. A. Vigalok, D. Milstein, Acc. Chem. Res. 2001, 34, 798-807.

38. M. E. van der Boom, D. Milstein, Chem. Rev. 2003, 103, 1759-1792.

39. M. Gandelman, A. Vigalok, L. Konstantinovski, D. Milstein, J. Am. Chem.


Soc. 2000, 122, 9848-9849.

40. M. Gandelman, L. J. W. Shimon, D. Milstein, Chem. Eur. J. 2003, 9, 4295-


4300.

41. C. M. Frech, L. J. W. Shimon, D. Milstein, J. Am. Chem. Soc. 2006, 128,


12434-12435.

42. C. M. Frech, L. J. W. Shimon, D. Milstein, Chem. Eur. J. 2007, 13, 7501-


7509.

43. N. Montag, I. Efremenko, Y. Diskin-Posner, Y. Ben-David, J. M. L. Martin,


D. Milstein, Organometallics 2012, 31, 505-512.

44. N. Montag, I. Efremenko, G. Leitus, Y. Ben-David, J. M. L. Martin, D.


Milstein, Organometallics 2013, 32, 7163-7180.

45. B. L. Madison, S. B. Thyme, S. Keene, B. S. Williams, J. Am. Chem. Soc.


2007, 129, 9538-9539.

This article is protected by copyright. All rights reserved.


European Journal of Inorganic Chemistry 10.1002/ejic.201600989

46. K. L. Jonasson, A. V. Polukeev, R. Marcos, M. S. G. Ahlquist, O. F. Wendt,


Angew. Chem. Int. Ed. 2015, 54, 9372-9375.

47. a) C. H. Suresh, N. Koga, Organometallics 2004, 23, 76-80; b) See also S.


F. Vyboishchikov, M. Bühl, W. Thiel, Chem. Eur. Journal 2002, 8, 3962-
3975.

48. J. B. Lee, G. J. Gajda, W. P. Schaefer, T. R. Howard, T. Ikariya, D. A.


Straus, R. H. Grubbs, J. Am. Chem. Soc. 1981, 103, 7358-7361.

49. J. R. Stille, B. D. Santarsiero, R. H. Grubbs, J. Org. Chem. 1990, 55, 843-


862.

50. R. Beckhaus, S. Flatau, S. Trojanov, P. Hofmann, Chem. Ber. 1992, 125,


291-299.

51. P. H. P. Brinkmann, M.-H. Prosenc, G. A. Luinstra, Organometallics 1995,


14, 5481-5482.

52. J. L. Polse, A. W. Kaplan, R. A. Andersen, R. G. Bergman, J. Am. Chem.


Soc. 1998, 120, 6316-6328.

53. G. Greidanus, R. McDonald, J. M. Stryker, Organometallics 2001, 20, 2492-


2504.

54. K. K. Baldridge, Y. Kasahara, K. Ogawa, J. S. Siegel, K. Tanaka, F. Toda, J.


Am. Chem. Soc. 1998, 120, 6167-6168.

55. S. Scheins, M. Messerschmidt, M. Gembicky, M. Pitak, A. Volkov, P.


Coppens, B. G. Harvey, G. C. Turpin, A. M. Arif, R. D. Ernst, J. Am. Chem.
Soc. 2009, 131, 6154-6160.

56. For additional insightful comments on similar species, see W. Scherer, G. S.


McGrady, Angew. Chem. Intl. Ed. 2004, 43, 1782-1806.

57. R. R. Schrock, R. T. DePue, J. Feldman, C. J. Schaverien, J. C. Dewan, A.


H. Liu, J. Am. Chem. Soc. 1988, 110, 1423-1435.

58. a) T. V. Lubben, P. T. Wolczanski, G. D. Van Duyne, Organometallics


1984, 3, 977-983; b) R. Tomaszewski, A. M. Arif, R. D. Ernst, J. Chem.

This article is protected by copyright. All rights reserved.


European Journal of Inorganic Chemistry 10.1002/ejic.201600989

Soc., Dalton Trans. 1999, 1883-1890; c) J. A. Dobado, J. M. Molina, R.


Uggla, M. R. Sundberg, Inorg. Chem. 2000, 39, 2831-2836; d) R. Basta, A.
M. Arif, R. D. Ernst, Organometallics 2005, 24, 3982-3986.

59. J. Feldman, W. M. Davis, J. K. Thomas, R. R. Schrock, Organometallics


1990, 9, 2535-2548.

60. A. J. Jiang, J. H. Simpson, P. Müller, R. R. Schrock, J. Am. Chem. Soc.


2009, 131, 7770-7780.

61. M. R. Reithofer, G. E. Dobereiner, R. R. Schrock, P. Müller,


Organometallics 2013, 32, 2489-2492.

62. L. C. H. Gerber, R. R. Schrock, Organometallics 2013, 32, 5573-5580.

63. a) K. Mashima, M. Kaidzu, Y. Nakayama, A. Nakamura, Organometallics


1997, 16, 1345-1348; b) K. Mashima, M. Kaidzu, Y. Tanaka, Y. Nakayama,
A. Nakamura, J. G. Hamilton, J. J. Rooney, Organometallics 1998, 17,
4183-4195.

64. K. C. Wallace, A. H. Liu, J. C. Dewan, R. R. Schrock, J. Am. Chem. Soc.


1988, 110, 4964-4977.

65. a) C. H. Suresh, M.-H. Baik, J. Chem. Soc., Dalton Trans. 2005, 2982-2984;
b) P. R. Remya, C. H. Suresh, J. Chem. Soc., Dalton Trans. 2015, 17660-
17672.

66. C. H. Suresh, J. Organometal. Chem. 2006, 691, 5366-5374.

67. R. L. Lord, H. Wang, M. Vieweger, M.-H. Baik, J. Organometal. Chem.


2006, 691, 5505-5512.

68. For a case in which a 16 electron metallacyclobutane is a transition state


rather than an intermediate, see G. S. Hill, R. J. Puddephatt,
Organometallics 1998, 17, 1478-1486.

69. F. Bernardi, A. Bottoni, G. P. Miscione, Organometallics 2003, 22, 940-947.

70. P. E. Romero, W. E. Piers, J. Am. Chem. Soc. 2005, 127, 5032-5033.

71. P. E. Romero, W. E. Piers, J. Am. Chem. Soc. 2007, 129, 1698-1704.

This article is protected by copyright. All rights reserved.


European Journal of Inorganic Chemistry 10.1002/ejic.201600989

72. C. N. Rowley, E. F. van der Eide, W. E. Piers, T. K. Woo, Organometallics


2008, 27, 6043-6045.

73. E. F. van der Eide, P. E. Romero, W. E. Piers, J. Am. Chem. Soc. 2008, 130,
4485-4491.

74. A. G. Wenzel, G. Blake, D. G. VanderVelde, R. H. Grubbs, J. Am. Chem.


Soc. 2011, 133, 6429-6439.

75. H. Wang, J. O. Metzger, Organometallics 2008, 27, 2761-2766.

76. P. M. Graham, M. S. A. Buschhaus, P. Legzdins, J. Am. Chem. Soc. 2006,


128, 9038-9039.

77. P. M. Graham, M. S. A. Buschhaus, C. B. Pamplin, P. Legzdins,


Organometallics 2008, 27, 2840-2851.

78. M. S. A. Buschhaus, C. B. Pamplin, I. J. Blackmore, P. Legzdins,


Organometallics 2008, 27, 4724-4738.

79. M. S. A. Buschhaus, C. J. Semiao, P. Legzdins, Organometallics 2009, 28,


1122-1126.

80. Y. Hu, N. Romero, C. Dinoi, L. Vendier, S. Mallet-Ladeira, J. E. McGrady,


A. Locati, F. Maseras, M. Etienne, Organometallics 2014, 33, 7270-7278.

81. J. Jaffart, M. Etienne, M. Reinhold, J. E. McGrady, F. Maseras, J. Chem.


Soc., Chem. Commun. 2003, 876-877.

82. J. Jaffart, M. L. Cole, M. Etienne, M. Reinhold, J. E. McGrady, F. Maseras,


J. Chem. Soc., Dalton Trans. 2003, 4057-4064.

83. M. Besora, F. Maseras, J. E. McGrady, P. Oulié, D. H. Dinh, C. Duhayon,


M. Etienne, J. Chem. Soc., Dalton Trans. 2006, 2362-2367.

84. C. Boulho, T. Keys, Y. Coppel, L. Vendier, M. Etienne, A. Locati, F.


Bessac, F. Maseras, D. A. Pantazis, J. E. McGrady, Organometallics 2009,
28, 940-943.

85. M. Etienne, J. E. McGrady, F. Maseras, Coord. Chem. Rev. 2009, 253, 635-
646.

This article is protected by copyright. All rights reserved.


European Journal of Inorganic Chemistry 10.1002/ejic.201600989

86. C. Boulho, L. Vendier, M. Etienne, A. Locati, F. Maseras, J. E. McGrady,


Organometallics 2011, 30, 3999-4007.

87. S. K. Brayshaw, J. C. Green, G. K.-K. Köhn, E. L. Sceats, A. S. Weller,


Angew. Chem. Intl. Ed., 2006, 45, 452-456.

88. S. K. Brayshaw, E. L. Sceats , J. C. Green, A. S. Weller, Proc. Natl. Acad.


Sci., 2007, 104, 6921-6926.

89. H. A. Sparkes, T. Krämer, S. K. Brayshaw, J. C. Green, A. S. Weller, J. A.


K. Howard, J. Chem. Soc., Dalton Trans. 2011, 10708-10718.

90. A. B. Chaplin, A. S. Weller, J. Organometal. Chem. 2013, 730, 90-94.

91. A. B. Chaplin, A. S. Weller, Organometallics 2010, 29, 2332-2342.

92. A. B. Chaplin, J. C. Green, A. S. Weller, J. Am. Chem. Soc. 2011, 133,


13162-13168.

93. Y. Escudié, C. Dinoi, O. Allen, L. Vendier, M. Etienne, Angew. Chem. Int.


Ed. 2012, 51, 2461-2464.

94. O. T. Summerscales, F. G. N. Cloke, P. B. Hitchcock, J. C. Green, N.


Hazari, Science 2006, 311, 829-831.

95. V. Wray, P. E. Hansen, Ann. Rep. NMR Spectrosc., C. A. Webb, Ed., 1981,
11A, 99.

96. F. Maseras, R. Crabtree, Inorg. Chim. Acta 2004, 357, 345-346.

97. R. F. W. Bader, J. Phys. Chem. A 1998, 102, 7314-7323.

98. Y. A. Abramov, J. Phys. Chem. A 1997, 32, 5725-5728.

99. J. Cioslowski, S. T. Mixon, W. D. Edwards, J. Am. Chem. Soc. 1991, 113,


1083-1085.

100. J. Cioslowski, S. T. Mixon, Can. J. Chem. 1992, 70, 443-449.

101. J. Cioslowski, S. T. Mixon, J. Am. Chem. Soc. 1992, 114, 4382-4387.

This article is protected by copyright. All rights reserved.


European Journal of Inorganic Chemistry 10.1002/ejic.201600989

102. J. Cioslowski, L. Edgington, B. B. Stefanov, J. Am. Chem. Soc. 1995, 117,


10381-10384.

103. S. C. Choi, R. J. Boyd, Can. J. Chem. 1986, 64, 2042-2047.

104. J. P. Ritchie, K.-Y. Lee, D. T. Cromer, E. M. Kober, D. D. Lee, J. Org.


Chem. 1990, 55, 1994-2000.

105. K. Lammertsma, J. Leszcyzyńki, J. Phys. Chem. 1990, 94, 5543-5548.

106. L. J. Farrugia, C. Evans, M. Tegel, J. Phys. Chem. A 2006, 110, 7952-7961.

107. E.-L. Zins, B. Silvi, M. E. Alikhani, Phys. Chem. Chem. Phys. 2015, 17,
9258-9281.

108. P. Popelier, J. Logothetis, J. Organometal. Chem. 1998, 555, 101-111.

109. W. Scherer, W. Hieringer, M. Spiegler, P. Sirsch, G. S. McGrady, A. J.


Downs, A. Haaland, B. Pedersen, J. Chem. Soc., Chem. Commun. 1998,
2471-2472.

110. R. H. Crabtree, R. P. Dion, D. J. Gibboni, D. V. McGrath, E. M. Holt, J. Am.


Chem. Soc. 1986, 108, 7222-7227.

111. S. Pillet, G. Wu, V. Kulsomphob, B. G. Harvey, R. D. Ernst, P. Coppens, J.


Am. Chem. Soc. 2003, 125, 1937-1949.

112. B. G. Harvey, A. M. Arif, R. D. Ernst, J. Chem. Cryst. 2010, 40, 783-787.

113. G. Binsch, J. B. Lambert, B. W. Roberts, J. D. Roberts, J. Am. Chem. Soc.


1964, 86, 5564-5570.

114. M. Booij, N. H. Klers, A. Meetsma, J. H. Teuben, W. J. J. Smeets, A. L.


Spek, Organometallics 1989, 8, 2454-2461.

115. M. D. Rausch, D. F. Foust, R. D. Rogers, J. L. Atwood, J. Organometal.


Chem. 1985, 265, 241-248.

116. M. Etienne, A. S. Weller, Chem. Soc. Rev. 2014, 43, 242-259.

117. W. Chen, S. Shimada, M. Tanaka, Science 2002, 295, 308-310.

This article is protected by copyright. All rights reserved.


European Journal of Inorganic Chemistry 10.1002/ejic.201600989

Figure Captions

Figure 1. Molecular structure of the lithium cyclopropoxide complex 3.

Figure 2. Structure of the di(1-phenyl-2-trimethylsilylacetylene) coupling


product, 4. The silicon-attached methyl groups have been omitted for
clarity.

Figure 3. Structure of the tri(1-phenyl-2-trimethylsilylacetylene) coupling


product, 5. The silicon-attached methyl groups have been omitted for
clarity.

Figure 4. Structure of the tetra(1-phenyl-2-trimethylsilylacetylene) coupling


product, 6. The silicon-attached methyl groups have been omitted for
clarity.

Figure 5. Structure of the diethyloctadiyne coupling product, 7a.

Figure 6. Structure of the hexadiyne coupling product, 8 (X = C2H4).

Figure 7. Structure of the di(1-phenyl-2-trimethylsilylacetylene) coupling


product, 9. The silicon-attached methyl groups have been omitted for
clarity.

Figure 8. Structure of the silylated complex 10.

Figure 9. Structure of the dihaptoacyl complex 11.

Figure 10. Structure of the arenium platinum complex 12 (X =I).

Figure 11. Structure of the arenium rhodium complex 13 (R = R’ = R” = CH3).

Figure 12. Structure of the 16 electron zirconacyclobutane complex 20 (R = CH3;


R’ = H; R” = CH2CH=CH2).

Figure 13. Structure of the tungstacyclobutane complex 21 (R’ = CH3). The


fluorine atoms have been reduced in size for clarity.

Figure 14. Structure of the tungstacyclobutane complex 22. The fluorine atoms
have been reduced in size for clarity.

Figure 15. Structure of the tungstacyclobutane complex 23.

This article is protected by copyright. All rights reserved.


European Journal of Inorganic Chemistry 10.1002/ejic.201600989

Figure 16. Structure of the tungstacyclobutane complex 24.

Figure 17. Structure of the niobacyclobutane complex 25.

Figure 18. Structure of the tantalacyclobutane complex 26 (Ar = 2,6-C6H3(i-


Pr)2).

Figure 19. Structure of the molybdenacyclobutane complex 31.

Figure 20. Structure of the cyclopropyl niobium complex 33.

Figure 21. Structure of the BINOR-S rhodium complex 34.

Figure 22. Structure of the cyclopropyl yttrium complex 37. The carbon atoms
of the THF ligand have been omitted for clarity.

Figure 23. Structure of the uranium deltate complex 38. The silicon-attached i-
Pr groups have been omitted for clarity.

Figure 24. Structure of the potentially C-N agostic zirconium complex 39.

Figure 25. Structure of the titanium amide complex 40. The methyl groups of
the t-Bu substituents have been omitted for clarity.

This article is protected by copyright. All rights reserved.


European Journal of Inorganic Chemistry 10.1002/ejic.201600989

Figure 1

This article is protected by copyright. All rights reserved.


European Journal of Inorganic Chemistry 10.1002/ejic.201600989

Figure 2

This article is protected by copyright. All rights reserved.


European Journal of Inorganic Chemistry 10.1002/ejic.201600989

Figure 3

This article is protected by copyright. All rights reserved.


European Journal of Inorganic Chemistry 10.1002/ejic.201600989

Figure 4

This article is protected by copyright. All rights reserved.


European Journal of Inorganic Chemistry 10.1002/ejic.201600989

Figure 5

This article is protected by copyright. All rights reserved.


European Journal of Inorganic Chemistry 10.1002/ejic.201600989

Figure 6

This article is protected by copyright. All rights reserved.


European Journal of Inorganic Chemistry 10.1002/ejic.201600989

Figure 7

This article is protected by copyright. All rights reserved.


European Journal of Inorganic Chemistry 10.1002/ejic.201600989

Figure 8

This article is protected by copyright. All rights reserved.


European Journal of Inorganic Chemistry 10.1002/ejic.201600989

Figure 9

This article is protected by copyright. All rights reserved.


European Journal of Inorganic Chemistry 10.1002/ejic.201600989

Figure 10

This article is protected by copyright. All rights reserved.


European Journal of Inorganic Chemistry 10.1002/ejic.201600989

Figure 11

This article is protected by copyright. All rights reserved.


European Journal of Inorganic Chemistry 10.1002/ejic.201600989

Figure 12

This article is protected by copyright. All rights reserved.


European Journal of Inorganic Chemistry 10.1002/ejic.201600989

Figure 13

This article is protected by copyright. All rights reserved.


European Journal of Inorganic Chemistry 10.1002/ejic.201600989

Figure 14

This article is protected by copyright. All rights reserved.


European Journal of Inorganic Chemistry 10.1002/ejic.201600989

Figure 15

This article is protected by copyright. All rights reserved.


European Journal of Inorganic Chemistry 10.1002/ejic.201600989

Figure 16

This article is protected by copyright. All rights reserved.


European Journal of Inorganic Chemistry 10.1002/ejic.201600989

Figure 17

This article is protected by copyright. All rights reserved.


European Journal of Inorganic Chemistry 10.1002/ejic.201600989

Figure 18

This article is protected by copyright. All rights reserved.


European Journal of Inorganic Chemistry 10.1002/ejic.201600989

Figure 19

This article is protected by copyright. All rights reserved.


European Journal of Inorganic Chemistry 10.1002/ejic.201600989

Figure 20

This article is protected by copyright. All rights reserved.


European Journal of Inorganic Chemistry 10.1002/ejic.201600989

Figure 21

This article is protected by copyright. All rights reserved.


European Journal of Inorganic Chemistry 10.1002/ejic.201600989

Figure 22

This article is protected by copyright. All rights reserved.


European Journal of Inorganic Chemistry 10.1002/ejic.201600989

Figure 23

This article is protected by copyright. All rights reserved.


European Journal of Inorganic Chemistry 10.1002/ejic.201600989

Figure 24

This article is protected by copyright. All rights reserved.


European Journal of Inorganic Chemistry 10.1002/ejic.201600989

Figure 25

This article is protected by copyright. All rights reserved.

You might also like