Ref LDF OMC IC 2017

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

REVIEWS

London dispersion forces in sterically


crowded inorganic and organometallic
molecules
David J. Liptrot and Philip P. Power
Abstract | London dispersion forces are the weakest component of Van der Waals interactions.
They arise from attractions between instantaneously induced dipoles on neighbouring atoms.
Their relative weakness, in particular for light atoms, such as hydrogen, has led to their
importance being largely ignored in discussions of molecular stability and reactivity. This Review
highlights the influence of these attractive forces — usually between C–H moieties in ancillary
ligands — on the physical and chemical properties of organometallic and inorganic molecules.
We feature recent examples of organic species that have informed current thinking and follow
with a discussion of several prominent inorganic and organometallic complexes wherein
dispersion forces have been explicitly identified or calculated. These forces strongly influence the
behaviour of such complexes and often have a defining structural role. Attention is also drawn to
several compounds in which significant attractive dispersion forces are probably present but
have not been investigated.

Terphenyls In his later career, Johannes Diederik van der Waals’ that could lead to decomposition or transformation of
In this Review, a terphenyl favourite dictum was reportedly ‘all matter displays the compounds. This stabilization is caused by the sub-
ligand consists of a central aryl attraction’ (REF. 1). Fritz London’s exposition of part of this stituents occupying the space surrounding a reactive
ring substituted by two further attractive force — now known as the London dispersion centre7–9 so that access to it is prevented or, alternatively,
aryl rings at the ortho (that is,
flanking) positions relative to
force (LDF)2–4 — was fundamental in supporting this by shielding that space through the use of bowl-shaped
the carbon atom (ipso) through precept. It is therefore remarkable that over the past six or concave ligands. Examples of these ligands are ter-
which the terphenyl ligand is decades, work in synthetic organic, organometallic and phenyls10–12 or N‑heterocyclic carbenes13–17 (NHCs; FIG. 1)
attached to the reactive centre. inorganic chemistry has mostly ignored the importance bearing aryl substituents, as well as various bidentate
They are denoted by the
of these attractive forces. Although workers in protein ligands, such as β‑diketiminates, amidinates and
abbreviation ArRn, where the
superscript R refers to the type
science5, as well as those in supramolecular 6 and theo- guanidinates, which have similar steric properties18–21.
of substituents on the aryl rings retical chemistry, have readily embraced the significance In virtually all cases, the bulk is provided by one or
and the numeral indicates the of dispersion forces, most synthetic molecular chemists more hydrocarbon-substituted groups of various sizes.
number of R substituents have clung boldly to the simplification that non-bonded The overlap between the electron clouds of these groups
present: for example,
ArMe6 = C6H3-2,6-(C6H2-
atoms repel each other, as exemplified by steric repulsion. leads to the Pauli repulsive interaction, which may
2,4,6‑Me3)2 and Nevertheless, work over the past decade has increasingly distort bond lengths and angles. However, in a grow-
Ari-Pr4 = C6H3-2,6-(C6H3- revealed the limitations of such a model and, although ing body of work on hydrocarbon-substituted sterically
2,6-iPr6)2. organic chemists are becoming aware of the importance of crowded molecules, it has been shown that, in addi-
LDF effects, acceptance in other synthetic sub­disciplines tion to the aforementioned steric repulsion, attractive
remains lacking. This Review seeks to remedy this by dispersion forces (BOX 1) between C–H moieties often
highlighting a range of organometallic and inorganic produce results that are difficult to explain purely on
Department of Chemistry,
compounds wherein dispersion forces have been shown steric grounds22–24. Until recently, it has generally been
University of California,
One Shields Avenue, Davis, to have a key role in their chemical behaviour. assumed that the attractions arising from the LDF are
California 95616, USA. The use of large, sterically crowded substituents is individually weak and essentially negligible, a misconcep-
Corresponding to P.P.P fundamental to much synthetic work in both organic tion that is fuelled by the validity of the assumption when
pppower@ucdavis.edu and inorganic chemistry. It is usually assumed that the applied to a single pair of light atoms. However, in larger
doi:10.1038/s41570-016-0004 main purpose served by the substituents is to hinder the molecules, LDF effects can be of considerable importance
Published online 11 Jan 2017 atom that they are attached to from engaging in reactions in collectively inducing stabilization on the order of tens

NATURE REVIEWS | CHEMISTRY VOLUME 1 | ARTICLE NUMBER 0004 | 1


REVIEWS

of kilocalories per mole25. This is sufficient to markedly a b c


influence their physical and chemical properties.
Currently, the importance of LDF effects between
C–H moieties and the resulting stabilization are not
widely recognized, despite attractive interactions between Si
N Si
closed-shell heavy element species (for example, d 10, s 2
E E
and others) having been intensively studied for many
E N Si
years, and ample structural and spectroscopic evidence
for LDF effects26. Nonetheless, the consequences of the
C–H attractions have recently become apparent through
the experimental and theoretical work of several groups
discussed later (FIG. 2). In addition, the development of Inaccessible area
relatively inexpensive quantum chemical methods27 Accessible area
that can estimate LDF effects has provided impor-
tant, often crucial, corroboration, as well as relatively Figure 1 | Examples of large, sterically crowded organic
Nature Reviews | Chemistry
accurate estimates of their magnitude. and inorganic substituents. Skeletal structures of a bulky
Although non-covalent intermolecular or intra- terphenyl ligand10–12 (part a), a N‑heterocyclic carbene13–17
molecular forces are not well addressed by density (part b) and a bulky alkyl ligand (part c). The grey shading
functional theory (DFT), several approaches28 have been highlights the qualitative effect of a concave structure
developed to remedy this problem while preserving (parts a and b) on shielding the area surrounding a reactive
the advantages of DFT. The most common approach centre, E, or of an alkyl ligand (part c), which occupies a
similar volume. The blue shading qualitatively represents
involves a ‘correction’ potential function29–33, such as
the coverage of the surface of atom E.
that developed by Grimme and co-workers29,30, which is
added to the exchange-correlation functional of choice
and is often denoted by the suffix ‘-D’ after the functional
name. It is also possible to fit the density functionals in H···H contacts. Such interactions are pictorially repre-
an empirical way so that they include the interactions sented when the atoms are within the sum of their van
attributable to dispersion forces34,35. Furthermore, several der Waals radii (2rH = 2.40 Å)50. These crystal structures
advanced functionals that inherently describe the dis- feature close arrangements of hydrogen atoms, despite
persion interactions have been developed36–39 and have the repulsive nature of these effects, locally speaking.
been used for the prediction of, for example, crystal Although the stability of molecules with effective steric
structures. The dispersion-corrected functionals, which repulsion requires additional rationalization, it is possi-
are subject to continuous review and assessment, are ble that dispersion forces have a contributing role. It is
generally accurate in predictions of structure and then reasonable to assume that the large number of such
reactivity in many different systems40–49. contacts indicates a strong attraction due to LDFs, hence,
This Review focuses primarily on weak, dispersive the H···H contacts are highlighted.
interactions between the C–H moieties in substituents,
groups or ligands in organometallic and inorganic com- Dispersion effects in hydrocarbon molecules
plexes. Addressing all cases in which dispersion force Among recent results on molecular hydrocarbon species,
effects may be significant in organometallic compounds three can be selected to provide cogent illustrations of
is not our major purpose. Instead, our approach is to the stabilizing effects of attractive dispersion forces. The
draw attention to several striking examples of organic, first example concerns the apparent instability of hexa-
inorganic and organometallic species wherein dispersion phenylethane, which exists as persistent ∙CPh3 radicals
force effects are likely to be of key importance to structure first reported by Gomberg in 1900 (REFS 51,52). On asso-
and reactivity. Obviously, we will only discuss systems in ciation, these radicals do not dimerize to the expected
which dispersion force effects have been explicitly iden- C–C bonded substituted ethane Ph3CCPh3, but instead
tified, because there are numerous compounds for which afford a head‑to‑tail methylenecyclohexadiene unit 53
these effects may be significant but remain unanalysed. (FIG. 3a; top structure). By contrast, the more crowded
We begin our discussion by presenting some organic tert-butyl substituted analogue, the tris(3,5‑di‑tert-
compounds in which the structure and reactivity has butylphenyl)methyl radical, does lead to the symmetric
been found to be significantly affected by the LDF. ethane derivative (3,5-tBu2H3C6)3C–C(C6H3-3,5-tBu2)3
The following sections are organized around the main (REFS 54,55) (FIG. 3a; bottom structure), which is a stable,
divisions of the periodic table: s block (groups 1 and 2), colourless, crystalline solid (melting point = 214 °C) and
p block (groups 13–18), d block (the transition elements) has a lengthened C–C bond of 1.67(3) Å (REF. 55).
and f block (lanthanide or actinide). The division of the A computational analysis of the structure by Grimme
compounds into main group, transition metal and lantha- and Schreiner 24 confirmed the greater steric crowding,
nide derivatives is mainly one of convenience, but much but also indicated that the meta tert-butyl groups on the
similarity can exist in the properties among some classes aryl rings did not add significantly to the Pauli repul-
of compounds from different blocks of the periodic table. sion already present in hexaphenylethane because of
This Review presents several crystallographically their location at the meta positions of the phenyl rings.
defined structures, most of which feature interligand However, the computations incorporating dispersion

2 | ARTICLE NUMBER 0004 | VOLUME 1 www.nature.com/natrevchem


REVIEWS

energy disclosed a very large, about 70 kcal mol−1, stabi- recent calculations have indicated that the parent tetra-
lization energy arising from attractive LDF between the hedrane should also be kinetically stable61. Nonetheless,
C(C6H3-3,5-tBu2)3 fragments. This energy is more than the synthesis of (CtBu)4 was deemed astonishing at
sufficient to make the structure stable against dissoci- the time, because the tert-butyl substituents lower the
ation (bond dissociation energy (De) = 35.2 kcal mol−1). thermo­stabilization of the C–C bond, and it was noted
The calculations reproduced the structural parameters that the valence-bond isomerization to cyclobutadiene
accurately, including the key central C–C bond length or fragmentation to two molecules of the corresponding
of 1.661 Å, in agreement with the experimental value of acetylene is orbitally forbidden. The proposal that LDF
1.67(3) Å. By contrast, calculations without a dispersion effects stabilize (CtBu)4 is therefore an appealing one;
correction predict instability with a negative value for De calculations on this species support this view 22.
of −26.4 kcal mol−1 — a difference of over 60 kcal mol−1
from the De of +35.2 kcal mol−1 given above. Main group species
A further powerful example illustrating the effect of Importance of LDF effects in molecules with single
the LDF can be found in the work of Fokin, Schreiner bonds between main group elements. Realization of
and co-workers on ethane derivatives bearing highly the importance of LDF effects on single bond lengths
rigid diamondoid substituents such as those based on originated from studies of main group compounds
adamantyl (–C10H15)22,23,25,56,57. Among these is a molecule with very bulky ligands, particularly those featuring
— stable up to 300 °C — featuring a central C–C bond of supersilyl (–SitBu3) and related groups62. Experimental
1.647 Å between two diamantane units. Another example studies on the synthesis, structure and properties of
is 2-(1‑diamantyl)[121]tetramantane, which has a central cyclo-tert-butyltrisilane63 c-(SitBu2)3 and hexa-tert-
C–C bond, here between diamantane and triamantane butyldisilane64 tBu3SiSitBu3 (FIG. 4a) in the 1980s have
units, that is even longer (1.71 Å; FIG. 3b)23. The structures shown that these compounds have extremely long
of these and related species are well-reproduced by DFT Si–Si and Si–C bonds: for example, 2.511(3) Å and
calculations using dispersion-corrected hybrid func- 1.99 Å in c-(SitBu2)3 and 2.697 Å and 1.970(5) Å in the
tionals31–35. The high thermal stability of the structures disilane (predicted Si–Si and Si–C single bond lengths
results from numerous attractive interactions between for the disilane are 2.34 and 1.88 Å, respectively)65. In
neighbouring hydrogen‑terminated diamond-like the cyclotrisilane, the highlighted distances correspond
surfaces25,56,57, and such forces also greatly increase the to Pauling bond orders66,67 of 0.60 and 0.75, respectively.
barrier to rotation around the central C–C bond from It was estimated that the Si–Si distance was equivalent
about 7 to 33 kcal mol−1. to a C–C distance of 1.70 Å (REF. 61) (similar to the C–C
The third example is tetra(tert-butyl)tetrahedrane distance of 1.71 Å in 2-(1‑diamantyl)[121]tetramantane
(CtBu)4 (REFS 58–60), a molecule with a high strain noted earlier; FIG. 3b)23. By contrast, the Si–Si distance in
energy of 129–137 kcal mol−1. The stability of this species the disilane corresponds to a lower Pauling bond order
was originally interpreted on the basis of the interac- of 0.45 (REF. 63). Despite the long element–element bonds
tion between tert-butyl substituents. It was argued that that suggest instability, both compounds displayed sur-
the decomposition of the molecule by stretching a tet- prisingly high thermal robustness, were stable towards
rahedrane C–C bond, promoted by repulsion between air and moisture, and had relatively high melting points
tert-butyl groups, is counterbalanced by the attractive of 182 °C for c-(SitBu2)3 and 162 °C for tBu3SiSitBu3. In
interactions between the stretched unit and the remain- the early 1990s, Bock and co-workers68 reported the
ing tert-butyl groups — a so‑called corset effect. More related disilane (Me3Si)3SiSi(SiMe3)3, which had a cen-
tral Si–Si bond length of 2.405(5) Å. However, they noted
close contacts between the peripheral SiMe3 groups and
Box 1 | Origin of London dispersion force suggested these to be responsible for the surprising
The London dispersion force (LDF)2–4 arises from the attraction of instantaneous stability of this molecule.
molecular multipoles originating from electron correlation. It is one of the three major Further work on the related germanium and tin
attractive components of the van der Waals forces between molecules, the remaining analogues supported this view 69,70. The structures of
two being: the force between two permanent dipoles (Keesom force) and the force c-(GetBu2)3 and tBu3GeGetBu3 (REF. 69) feature the longest
between a permanent dipole and a corresponding induced dipole. The stabilization known Ge–Ge and Ge–C single bonds: 2.563(1) and
energy produced by the attractive LDF between two atoms, A and B, can be 2.056(3) Å for c-(GetBu2)3 and corresponding values
approximated on the basis of their polarizability volumes (α), first ionization potentials of 2.710 and 2.076 Å in tBu3GeGetBu3. In addition, the
(I) and their separation (R). Ge–Ge and Si–Si distances in the tBu3EEtBu3 (E = Si or
3 IAIB αAαB
Ge) species — 2.697 and 2.705(1) Å, respectively — were
disp
EAB ≈− almost equal, despite the different radii of these ele-
2 IA + IB R6
ments (Si = 1.17 Å, Ge = 1.22 Å)65. These findings further
The LDF is highly dependent on distance, evident from the R−6 factor in this emphasize the importance of LDF stabilization as being
approximation. It is considered the weakest of intermolecular forces, typically being distinct from stabilization attributable to covalent bond-
≤1 kcal mol−1 for interactions between atoms (and lower for hydrogen atoms). However, ing between the E sites of the –EtBu3 moieties. As with the
this increases with atomic number, molecular size, contact area and polarizability. The silicon analogues, the germanium compounds are hydro-
LDF is present intramolecularly and its collective effect can be substantial (tens of lytically and thermally stable. On the other hand, the
kcal mol−1) in large molecules. Thus, the LDFs can be significant enough to considerably distannane tBu3SnSntBu3 (REF. 70) features a Sn–Sn bond
influence physical and chemical properties of large molecules.
(2.894(1) Å) that is only moderately elongated relative to

NATURE REVIEWS | CHEMISTRY VOLUME 1 | ARTICLE NUMBER 0004 | 3


REVIEWS

a normal Sn–Sn single bond (2.80 Å)65. This is probably ‘halves’ are bonded not only by a weak Al–Al covalent
due to tin being much larger than germanium and silicon, interaction, but also by LDFs between the peripheral CH3
which relieves steric repulsion in the present case. substituents73. Tetra-tBu3Si-dimetallane derivatives of
The study of other compounds incorporating indium and thallium also feature moderately length-
–SitBu3 substituents provides further evidence of the ened M–M (M = In or Tl) distances 76. The gallium
importance of LDF effects62,71. The reaction of NaSitBu3 derivative has a more complex chemistry owing to the
with Sn[N(SiMe3)2]2 or Sn(OtBu)2 at −25 °C affords increased steric crowding caused by the smaller size
the tristannaallene Sn[Sn(SitBu3)2]2. On warming, this of this element 65. This does not permit formation of
species isomerizes to cyclotristannene c-Sn3(SitBu3)4, (tBu3Si)2GaGa(SitBu3)2, but instead affords the radical
which can also be prepared by carrying out the reaction species (tBu3Si)GaGa(SitBu3)2 (REF. 77).
at room temperature71. The Sn3 ring structure is nota-
ble since its Sn–Sn double bond is 2.59(1) Å in length, LDF effects on compounds with multiple bonds. The
which is the shortest known Sn–Sn distance in a sta- study of compounds with multiple bonds between main
ble compound, including the nominally triple-bonded group elements, particularly between the heavier group
distannynes (about 2.64–2.67 Å)72. members, continues to be a highly dynamic research area
It was noted that in the ring structure of c-Sn3(SitBu3)4, that constitutes a major portion of current investigative
the four –SitBu3 groups roughly occupy the edges of a work on main group molecular species78,79. It is now
tetrahedron, thus adopting a close packing arrangement, widely accepted that the majority of such bonds differ
presumably LDF-assisted, which obscures the Sn3 ring fundamentally from those of their lighter congeners,
almost completely 71. as indicated, for example, by non-planar or non-linear
Another example of the stabilizing effects arising coordination in the heavier element analogues of ethyl-
from the –SitBu3 ligand can be found in derivatives of ene and acetylene. In the past, the multiple bonding in
the neighbouring group 13 metals. The reaction of three such compounds was often characterized by argument
equivalents of NaSitBu3 with AlX3 (X = Cl or Br) affords and controversy, mainly centred on the concept of bond
the dialane(4) tetrasilyl derivative (tBu3Si)2AlAl(SitBu3)2 order 78. The eventual outcome of this debate was that
(REF. 73) (FIG. 4b). The compound features two planar coor- Pythagorean bond orders (that is, those that can be
dinated aluminium atoms with coordination planes that written as whole numbers or as fractions thereof) prev-
are staggered with respect to each other. The Al–Al bond alent in organic chemistry have limited applicability in
length is 2.751(2) Å, which is unusually long and about heavier-element species. Ironically, more recent findings
0.1 Å longer than those in other tetraorgano dialanes have shown that the orbital overlap between the heavier
(R2AlAlR2, where R is CH(SiMe3)2 (REF. 74) or C6H2- elements in many of the multiply bonded species can be
2,4,6-iPr3 (REF. 75)). In solution it exists in equilibrium of lower strength than the attractions arising from the
with a very low concentration of the radical Al(SitBu3)2, LDF between the ligands, thereby reducing the impor-
and its treatment with H2 or I2 yields (tBu3Si)2AlX (X = H tance of the arguments regarding bond orders. As it is
or I). The monomer–dimer equilibrium is strongly now described, the recent literature has provided several
tilted towards the dimeric form, because the Al(SitBu3)2 examples of compounds that illustrate this point.

Theory Computational Suggestion that the stability


Inorganic and organometallic Organic of (Me3Si)3SiSi(SiMe3)3 is Donor–acceptor singly bonded diplumbylene
related to the close contact synthesized and shown to be principally
M[N(SiMe3)2]3 (M = Sc, Eu) between SiMe3 substituents68 stabilized by dispersion attractions80
Van der Waals’ thesis structurally characterized;
defines intermolecular pyramidalization Development of DFT-D methods Increased dispersive interactions
interactions (German attributed to attractive allowing the accessible assessment and between the ligands gives
translation, 1877; English intramolecular reassessment of dispersion force effects access to M[N(SiMe3)Dipp]2
translation, 1890)1 dispersion interactions116 in molecular structures27,29,30,32–34,37,42 (M = Fe–Cu)131,133

1873 1930 1973 1986 1993 1995 1998 2004 2011 2012 2013 2016

London dispersion M(Cp*)2 (M = Sc, Eu) (tBu3Si)2AlAl(SitBu3)2 Hexaphenylethane Molecules with Heavier main group
forces are found to structurally characterized; synthesized, its stability, riddle solved using long C–C multiple bonding
contribute to van der bent structure observed despite weak Al–Al newly developed DFT-D bonds reassessed in light of
Waals interactions2 and calculated to be due bonding, is attributed methods indicating the stabilized by attractive London
to attractive to dispersion62 importance of dispersion are dispersion using
intramolecular dispersion intramolecular designed and DFT-D methods80,82,84,86
interactions97,98 dispersion interactions24 synthesized22,23

Figure 2 | Timeline of the relevant advances in the study of London dispersion forces in organic and inorganic
compounds. A timeline highlighting the milestones in theory and experiment that led to the current understanding of
Nature
the role of London dispersion forces in the structure and reactivity of sterically crowded inorganic andReviews | Chemistry
organometallic
molecules. Cp*, C5(CH3)5; Dipp, 2,6‑diisopropylphenyl; DFT-D, dispersion corrected (-D) density functional theory.

4 | ARTICLE NUMBER 0004 | VOLUME 1 www.nature.com/natrevchem


REVIEWS

a R
b (E = Ge, Sn or Pb; FIG. 5b) and their dissociation into
R
R E[CH(SiMe3)2]2 and E[N(SiMe3)2]2 monomers were
R
R=H H
H performed. When the alkyl derivatives are considered
H
R H without including a dispersion force correction, asso-
C C
R R H
ciation is only favoured by about 2.1 kcal mol−1 for the
R R
H H tin species, whereas dissociation is favoured for the
R
germanium and lead derivatives. However, inclusion
R R
R R R
of LDF correction by the Grimme method shows that
R R
R C–C = 1.71 Å association becomes much more strongly preferred, with
R LDF stabilization energies falling in the range of about
R R
c 15.2–28.7 kcal mol−1 for the three classes of compounds.
R t
C C Bu
R In addition, thermodynamic calculations indicated that
R = tBu dimerization of germanium and tin fragments (but not
R
R
R R
t
Bu t
Bu
t
Bu lead) is energetically favoured overall.
R Clearly, the LDF effects are of greater importance
C–C = 1.67 Å C–C = 1.495 Å (calculated) than covalent E–E bonding in the association ener-
gies of these monomers. Calculations for the amido-
Figure 3 | London dispersion force effects in organic molecules. a | Contrasting
Nature Reviews | Chemistry
association of the ∙CPh3 and ∙C(C6H3-3,5-tBu2)3 radicals. Unsubstituted ∙CPh3 leads to a
substituted analogues [(Me 3Si) 2N] 2EE[N(SiMe 3) 2] 2
quinoid-type methylenecyclohexadienyl structure (top)53, whereas the tert-butyl reveal LDF-based attraction between the C–H
substituted derivative leads to a C–C bonded ethane structure (bottom) with a long C–C groups of the –N(SiMe 3) 2 ligands, although mon-
bond of 1.67 Å (REF. 55). b | A diamondoid ethane derivative with bonding between omeric structures are still observed partly because
triamantane and diamantane moieties that has a C–C bond length of 1.71 Å (REF. 23). of the E–E bonding being weakened by the pres-
c | A substituted tetrahedrane with calculated central C–C bond lengths of 1.495 Å ence of more electronegative –N(SiMe3)2 substitu-
(REF. 59) or 1.484 Å (REF. 60). All of these structures have been shown to be stabilized by ents. In addition, the distannene (tBu 2MeSi) 2SnSn
London dispersive attractions22–25. (SiMetBu2)2 (REF. 83) has been shown to have a unique
structure with a very short Sn–Sn bond of 2.647 Å.
Although it has planar coordination at each tin centre,
The dimerization of a cyclic disilylated plumbylene which would normally suggest strong multiple-bonding
(a lead analogue of a carbene) is a reaction that occurs, character for the Sn–Sn bond, the coordination planes
according to calculations by Marschner, Müller and at each tin are twisted by 44° with respect to each other.
co-workers80, as a result of attractions arising from This is contrary to the traditional σ–π double bond model
the LDF (FIG. 5a). A plumbylene phosphine complex is having a 0° torsion angle and indicative of a weakened
formed by the reaction of 1,4‑dipotassio‑1,1,4,4‑tetra­ interaction. Calculations 84 instead show that LDF effects
kis(trimethylsilyl)-2,2,3,3‑tetra­methyltetrasilane with are key to the stability of the compound, stabilizing the
PbBr2 in the presence of PEt3. Treatment of the complex molecule by 21 kcal mol−1 (REFS 82).
with B(C6F5)3 removes PEt3 as Et3PB(C6F5)3 and causes Parallel calculations on the sterically related group
the plumbylene monomer to dimerize, with the lead 15 [(Me3Si)2HC]2EʹEʹ[CH(SiMe3)2]2 and [(Me3Si)2N]2­
atoms being linked through an unusual single donor– EʹEʹ[N(SiMe3)2]2 (Eʹ = P or As)84 dimers showed that
acceptor bond (3.0640(8) Å), resembling that observed without the inclusion of LDF effects, cleavage of the
in the distannene [Sn(C6H-2‑tBu‑4,5,6‑Me3)2]2 (REF. 81). Eʹ–Eʹ bond to form the monomeric radicals is favoured.
This type of association is much rarer than the double Inclusion of LDF effects shows that the dimeric
donor–acceptor element–element bond, which is more diphosphane or diarsane structures become favoured.
commonly observed in heavier group 14 dimetallene Moreover, it was shown that the attractions arising from
dimers78,79. The binding energy of the plumbylenes in LDFs are more than sufficient to overcome the so‑called
the dimer was computed to be 26.4 kcal mol−1, which Jack‑in‑the-box conformational change85 from syn, anti
is 11.5 kcal mol −1 stronger than that in the parent to syn, syn (FIG. 5b), a process that otherwise leads to
diplumbene (that is, H2PbPbH2). The predicted structure dissociation when the dimers are dissolved or vaporized.
was in good qualitative agreement with the experimental Despite the importance of LDF effects, a common
data, but the charge transfer between the plumbylene feature of all the dimeric species discussed before is that
units was calculated to contribute only about 25% to the E–E interactions, although weak, are always pres-
the bond energy. In effect, most of the binding energy ent, and their role in directing dimer formation in the
between isolated plumbylenes is LDF-based. The unu- first place cannot be discounted. Other computational
sual results are consistent with the absence of a correla- studies of triply bonded main group compounds86 have
tion between the out‑of‑plane torsion angle and bond also uncovered LDF effects, one example being from
Extended transition state– length in the heavier ditetrelenes, in which much of the the work of Ziegler and co-workers, which involved the
natural orbitals for chemical attraction between the two tetrylene ‘halves’ is due to bonding in heavier main group element alkyne ana-
valence LDF effects, rather than tetrel (that is, group 14) element logues ArEEAr (where Ar is a terphenyl ligand (FIG. 1a)
(ETS–NOCV). A scheme for the orbital overlap78,79. and E = Si, Ge (REF. 87), Sn (REF. 88) or Pb (REF. 89)). For the
analysis of chemical bonds
based on the decomposition of
More recently, dispersion-corrected DFT calcula- silicon, germanium and tin species, an extended transi-
the bonding on the basis of tions82 on the group 14 ethylene analogues [(Me3Si)2HC]2­ tion state–natural orbitals for chemical valence (ETS-NOCV)90
charge and energy90. EE[CH(SiMe 3) 2] 2 and [(Me 3Si) 2N] 2EE­[ N(SiMe 3) 2] 2 analysis based on doublet EAr fragments revealed that

NATURE REVIEWS | CHEMISTRY VOLUME 1 | ARTICLE NUMBER 0004 | 5


REVIEWS

a σ bond, a π bond perpendicular to the –CEEC– plane LMg(i)–Mg(i)L (L = HC[CMe{N(C6H2-2,6-iPr2)2}2])


and a slipped π bond in the molecular plane constitute highlights that it is held together mainly through
the triple bond. In contrast, ArPbPbAr was found to LDF effects91.
have a single Pb–Pb σ bond only with an ipso‑C–Pb–Pb More recently, the stabilization resulting from attrac-
angle close to 90°. The stabilizing influence of the iPr tive LDF effects has been evaluated for a series of mul-
groups arises from attractive LDFs between the C–H tiply bonded B–B, Si–Si and P–P species, and related
moieties, along with hyperconjugative donation into the compounds featuring bulky N‑heterocyclic carbene lig-
σ* orbitals of the iPr groups. The dispersive interactions ands92. Dispersion corrections upwards of 30 kcal mol−1
were calculated to be 27.5 kcal mol−1 (Si), 29.1 kcal mol−1 were calculated and it was concluded that LDF effects
(Ge), 26.2 kcal mol−1 (Sn) and 44 kcal mol−1 (Pb). The contribute critically to the thermodynamic stabilities of
larger value for lead is connected to the terphenyl sub- these compounds. The reactivity of these species towards
stituent used for that metal, which has a greater number small molecules and their redox chemistry has been
of iPr substituents (six versus four) on each terphe- experimentally assessed92. Computational interrogation
nyl ligand. The relatively high values of the calculated of these results indicated that chemical transformations
dispersion energies (≥26.2 kcal mol−1) reflect the stabiliza- at the central E–E moiety occurred with conservation
tion influence of these effects in the present compounds. of the dispersive shell: that is, with limited rearrange-
Similarly, the optimized crystal structure of Jones’ dimer ment of the ligands owing to their highly dispersive
nature. Dispersion force effects have also been implicated
in complexes of the related N‑heterocyclic silylene, which
a
yields a three‑coordinate silylene adduct of Fe[N(SiMe3)2]2
wherein the dispersion attraction was calculated to be
22.7 kcal mol−1 (REF. 93). Dispersion force effects have
also been implicated in driving the Wagner–Meerwein
rearrangement of oligosilane cations94.
Among the earliest papers proposing C–H···H–C
LDF-based interactions as a possible explanation for
the unusual structures of organometallic molecules in
both main group species and lanthanides were those
that reported the structures of η5-C5Me5 (Cp*) deriva-
tives, M(Cp*)2 (M = Mg, Ca, Yb, Sr or Ba)95–98. For such
compounds it was proposed96–99 and calculated100,101 that
LDF effects involving the Cp* rings cause the molecu-
lar geometries to be bent. The calculated and observed
structures corresponded closely. However, the differ-
b ences in energy between the bent and linear forms are
relatively small.
The probable influence of attractive dispersion forces
on molecular structure is also seen in the divalent chal-
cogenolato substituted tetrylenes (carbene analogues). A
series of heavier element carbene analogues of formula
E(SArMe6)2 (E = Si, Ge or Sn; ArMe6 = C6H3-2,6(C6H2-
2,4,6‑Me3)2), E(SAri-Pr 4)2 (Ari-Pr 4 = C6H3-2,6(C6H3-2,6-
i
Pr2)2), E(SAri-Pr 6)2 (Ari-Pr 6 = C6H3-2,6(C6H2-2,4,6-iPr3)2)
and E(SAr i-Pr 8) 2 (Ar i-Pr 8 = C 6H-2,6(C6H 2-2,4,6-iPr 3) 2-
3,5‑Pr2), in which the bulk of the terphenyl thiolato
substituent is enhanced by increasing the number and
size of the alkyl groups attached to the terphenyl substi-
tuted ring 102. However, contrary to steric expectations,
the increase in substituent size leads to a decrease, rather
than an increase, in the S–E–S angle at the central tet-
Carbon Hydrogen Short contacts, rel to below 90° (TABLE 1). In addition, it was also found
probable London that as the bending angle decreased, the torsion angle
Silicon Aluminium
dispersion interaction
between the central aryl rings of the terphenyl substitu-
ents also decreased. Experimental and structural data in
Figure 4 | London dispersion Nature
force effects in superilyl
Reviews | Chemistry conjunction with high-level DFT calculations, including
and related groups. a | Crystal structure of tBu3SiSitBu3
corrections for dispersion effects, led to the conclusion
with a long Si–Si bond length of 2.697 Å (REF. 64). b | X-ray
crystal structure of (tBu3Si)2AlAl(SitBu3)2 with a long Al–Al that LDFs between the alkyl substituents stabilized the
bond length of 2.751(8) Å (REF. 73). Both structures show observed acute interligand angles102.
selected hydrogen atoms where interligand distances are Another bulky ligand remarkable for the stabiliza-
shorter than the sum of the van der Waals hydrogen radii tion of many unusual compounds is the well-known
(≤2.4 Å)50. tris(trimethylsilyl)methyl group –C(SiMe 3) 3. Some

6 | ARTICLE NUMBER 0004 | VOLUME 1 www.nature.com/natrevchem


REVIEWS

examples of complexes bearing this ligand are now persistence of the anionic motif can also be seen in the
described, including the seminal work of Eaborn, copper(i) complex of tris(trimethylsilyl)methanide
Smith and co-workers103. The metals of the s block (FIG. 5c), a rare structurally characterized two‑coordi-
provide an ideal proving ground for dispersion force nate dialkylcuprate. This compound is isostructural to
effects, because the non-directional bonding of these the lithium salt, also suggesting the effects of LDF on
elements allows such small forces to have significant this structure109.
structural effects. In contrast to the usual associated Group 2 bis[tris(trimethylsilyl)methanide] com-
structures, the lithium and sodium derivatives of this pounds also exhibit noteworthy structural parameters,
ligand both exhibit an unusual charge-separated struc- the two extreme cases being the magnesium and cal-
ture of the form [M{C(SiMe3)3}2]− (M = Li or Na)104–106. cium derivatives. Mg[C(SiMe3)3]2 is linear and surpris-
Inspection of these congeners indicates that the elon- ingly inert for a dialkylmagnesium compound110,111, as
gated M–C bonds in the sodium derivative result in opposed to the calcium derivative that features a more
the rotation of trimethylsilyl substituents to maxim- notable bent C–Ca–C angle of 149.7(6)° (REF. 112). The
ise H···H contacts, suggesting a structural effect of presence of short interligand H···H contacts in the crystal
interligand dispersion interactions. Furthermore, structure of the latter suggests that the forces resulting
although this isomeric form persists for the slightly in this bent structure may be similar to that for Ca(Cp*)2
modified –C(SiMe3)2(SiMe2Ph) ligand for lithium107, (REF. 100) — that is, attractive LDFs.
the structural constraints imposed on the methyl sub- Similarly, attractive dispersion forces between ligands
stituents in –C(SiMe2Ph)3 perturbs the potential of are likely to contribute to the surprising stability of
the interligand dispersion forces to yield the solvated group 12 complexes incorporating two very bulky alkyl
lithium alkyl [Li{C(SiMe2Ph)3}(THF)] (REF. 108) rather ligands. Thus, Zn[C(SiMe3)3]2 (REF. 113) shows anoma-
than the dissociated ionic structure observed for the lously high stability for a dialkylzinc and can be steam
–C(SiMe3)3 or –C(SiMe3)2(SiMe2Ph) derivatives. The distilled under air. This is in contrast to the usual flam-
mability of this class of compounds in the presence of
oxygen. Dibenzylmercury was found to be around 4,500
a RR
b times more susceptible to unimolecular dissociation
Si
Si syn, syn syn, anti than Hg[C(SiMe2Ph)3](CH2Ph), behaviour that was
R
Si
Pb Pb R R R attributed to the transfer of energy from vibrations of
Si R Si Si R H R
H R the large ligand to bond dissociation114. In both of these
R Si Si R H
R R E E E cases, interligand attraction was not considered but is
H R
H R
R = SiMe3 R R likely to be a contributing factor to these anomalously
R R H
high stabilities115.
R = SiMe3 E = Ge, Sn, Pb
c R R Transition and lanthanide metal complexes
R
M It is probable that the first publication to attribute
R R
R structural distortion to London dispersion effects in
R = SiMe3 M = Li, Na, Cu
an inorganic molecular species concerned the three-
coordinate transition metal and lanthanide amido
derivatives M[N(SiMe 3) 2] 3 (M = Sc and Eu), which
displayed a pyramidalized rather than planar metal
coordination116. It was suggested that distortion from
planarity occurs on formation of the crystalline phase,
in which the metal ion is squeezed out of the coordina-
tion plane by a symmetrical contraction resulting from
intra­molecular–interligand van der Waals attractions.
However, dipole moment measurements suggest that
the pyramidal structure is not preserved in solution.
Another early possibility is the ytterbium species
Yb(Cp*)2 (REF. 94), which has a close structural resem-
Carbon Hydrogen Silicon Short contacts, probable blance to the group 2 bis(pentamethylcyclopentadienyl)
London dispersion interaction
Tin Lithium metal M(Cp*)2 (M = Mg, Ca, Sr or Ba)95–99 species. These
have bent structures (also in the vapor phase)95–97, as do
Figure 5 | London dispersion force effects on molecules with multiple bonds the divalent lanthanide derivatives Mʹ(Cp*)2 (Mʹ = Sm or
Nature Reviews bond
between main group elements. a | Unusual side‑on single donor–acceptor | Chemistry
in Eu)117–119. In the samarium species, the shortest distance
diplumbylene stabilized by London dispersion forces80. b | Association of group 14
between the methyl group carbons is 3.34(1) Å, which is
element monomers gives dimers with rotated substituents82 and the X-ray crystal
less than the sum of their van der Waals radii (3.4 Å)50.
structure82 of [(Me3Si)2HC]2SnSn[CH(SiMe3)2]2. c | The unusual structure of anionic
group 1 and group 11 metal complexes of the tris(trimethylsilyl)methane ligand104–109 The very crowded complex Sm(Cp*)3 (obtained from
and the X-ray crystal structure104 of the dialkyllithium [Li{C(SiMe3)3}2]− anion. The Sm(Cp*)2 by a disproportionation reaction with cyclo­
[Li(THF)4]+ counterion is not shown for clarity. Crystal structures in b and c show selected octatetrene) shows significant structural strain and a
hydrogen atoms where interligand distances are shorter than the sum of the van der fascinating chemistry in light of its probable stabilization
Waals hydrogen radii (≤2.4 Å)50. by LDFs between ligands120.

NATURE REVIEWS | CHEMISTRY VOLUME 1 | ARTICLE NUMBER 0004 | 7


REVIEWS

More recently, there has been a rapid increase in the presence of the isopropyl substituents stabilizes the mol-
number of publications on LDF effects on the stability, ecule by 20 kcal mol−1 relative to the CrAri-Pr4 monomers,
structure and reactivity of various transition metal as a result of LDF effects between these groups.
complexes. Compounds studied include Pd(PPh3)4 A characteristic of transition metals is their tendency
(REF. 121), Pd(PtBu3)2 (REF. 122), and group 8 species such to display a range of oxidation states that differ by unity
as Os(Cl)2(H2)(PiPr3)2 (REF. 123). Computational studies steps. This stems from the presence of several energeti-
were performed on ruthenium olefin metathesis cata- cally accessible d orbitals. The modest energies that often
lysts of formula Ru(Cl)2(PR3)2(NHC) (REF. 124), in which separate the stabilities of complexes in various oxidation
various basis sets and their effects on phosphine dissoci- states make them more susceptible to disproportionation.
ation barriers were also evaluated125. Furthermore, steric The use of –N(SiMe3)Dipp (Dipp = C6H3-2,6-iPr2)
attraction has also been shown to have decisive influence ligand has recently permitted preparation of the first
on bite angle in two‑coordinate transition metals and two-coordinate130 Cu(ii) complex, Cu[N(SiMe3)Dipp]2
a consequent effect on reactivity 126. In addition, H···H (REF. 131) (FIG. 6b). Its formation apparently takes place by
interactions in some metal hydride complexes were also a dispersion energy-influenced disproportionation of a
shown to arise from dispersion forces127. Cu(i) complex formed by the reaction of LiN(SiMe3)Dipp
In comparison to the sterically crowded main group with CuCl in non-donor solvents. Also, it is possible that
species discussed earlier, LDF effects are less studied in in the synthesis of the homoleptic vanadium(ii) amide
sterically crowded transition metal complexes with alkyl, V[N(H)Ari-Pr6]2 from a vanadium(iii) starting material132,
aryl or amido ligands. However, a few calculations have reduction to the vanadium(ii) product occurs through
indicated that dispersive interactions are of great impor- a disproportionation (analogous to the formation of
tance for the stabilization of transition metal complexes Cu[N(SiMe3)Dipp]2 (REF. 131); FIG. 6b) with concomi-
bearing such ligands86,128. In parallel with the heavier tant formation of an unidentified vanadium(iv) spe-
group 14 element alkyne analogues Ar i-Pr 4EEAr i-Pr 4 cies. Dispersion force effects in heteroatom-substituted
(E = Ge or Sn) and Ari-Pr 6PbPbAri-Pr 6 mentioned above, terphenyls have precedent 102, suggesting that this puta-
calculations 128 on the quintuply bonded species tive disproportionation is partly assisted by attractive
Ari-Pr 4CrCrAri-Pr 4 bearing the same ligand as the ger- dispersion forces between the bulky ligands.
manium and tin species129 (FIG. 6a) have confirmed five Stabilizing London dispersion interactions were
distinct bonding components to the Cr–Cr bond, as well also calculated for two-coordinate amido complexes of
as the presence of secondary Cr–C interactions involv- formula M(ii)[N(SiMe3)Dipp]2 (M = Fe, Co or Ni)133,
ing the ipso carbon of the terphenyl group. However, the as well as the Ni(i) tetramer [Ni{N(SiMe3)2}]4 (REF. 134).

Table 1 | Correlation between ligand size and structural properties in dithiolatotetrylenes


Ligand E(SAr)2 E–S bond length S–E–S bond Torsion angle (°)
(Å) angle (°) between ES2 and
central aryl ring plane
Si(SArMe6)2 2.158(3) 90.52(2) 44.9(1)
S
Ge(SArMe6)2 2.265(11) 88.68(2) 40(6)

Sn(SArMe6)2 2.479(5) 85.555(3) 38(10)

Si(SAri-Pr4)2 2.137(1) 85.08(5) 38.2(2)


iPr iPr
S
Ge(SAri-Pr4)2 2.284(4) 81.26(2) 32.3(1)
iPr iPr
Sn(SAri-Pr4)2 2.470(1) 78.63(3) 31.2(3)

iPr iPr iPr iPr


Si(SAri-Pr6)2 2.089(9) 84.8(1) 29(7)
S
Ge(SAri-Pr6)2 2.24(4) 79.6(2) 27(7)
iPr iPr
Sn(SAri-Pr6)2 2.46(6) 78.2(2) 34(6)

iPr iPr iPr iPr


Ge(SAri-Pr8)2 2.2940(6) 77.01(2) 3.6(2)
S
Sn(SAri-Pr8)2 2.5009(6) 73.09(2) 3.8(2)
iPr iPr
iPr iPr

This table shows the effects of increasing steric bulk of the terphenyl substituent on S–E bond length, S–E–S bond angle, and torsion
angle between ES2 and the plane of the central aryl ring of dithiolatotetrylenes owing to London dispersion forces102. ArMe6 = C6H3-
2,6(C6H2-2,4,6‑Me3)2; Ari-Pr4 = C6H3-2,6(C6H3-2,6-iPr2)2; Ari-Pr6 = C6H3-2,6(C6H2-2,4,6-iPr3)2; Ari-Pr8 = C6H-2,6(C6H2-2,4,6-iPr3)2-3,5‑Pr2.

8 | ARTICLE NUMBER 0004 | VOLUME 1 www.nature.com/natrevchem


REVIEWS

E E

E Bond order
Si, Ge, Sn 3
Pb 1
Cr 5

c
b

Carbon Hydrogen Nitrogen Short contacts, probable


London dispersion interaction
Silicon Chromium Copper Iron

Figure 6 | London dispersion force effects in transition and lanthanide metal complexes. aNature | Results of extended
Reviews | Chemistry
transition state–natural orbitals for chemical valence (ETS–NOCV) analysis of the bonding of terphenyl supported dimers
that indicated the crucial effects of London dispersion forces86,128, and the X‑ray crystallographically characterized
structure of Ari-Pr4CrCrAri-Pr4 (REF. 129) showing selected hydrogen atoms where interligand distances are shorter than the
sum of the van der Waals hydrogen radii (≤2.4 Å)50. b | Crystal structure of the two-coordinate Cu(ii) species
Cu[N(SiMe3)Dipp]2 showing selected hydrogen atoms where interligand distances are shorter than the sum of van der
Waals hydrogen radii131. c | Crystal structure of Fe(1‑norbornyl)4 (REFS 135,138) showing London dispersion interactions
between CH2···CH2 moieties136. Hydrogen atoms are omitted for clarity.

The M(ii) compounds are characterized by planar core DFT calculations136 indicated that the M–C bonds are
geometries with eclipsed structures. The M–N distances shortened and the complexes are stabilized by dispersion
are shortened, and the compounds, especially the nickel attractions between the 1‑norbornyl ligands (FIG. 6c). For
derivative Ni[N(SiMe3)Dipp]2 (REF. 133), have enhanced the iron complex, these effects resulted in a calculated
stability relative to the corresponding Ni[N(SiMe3)2]2 stabilization of 45.9 kcal mol−1, indicating that the degree
analogue, which decomposes at room temperature. of stabilization from dispersion energies is key in deter-
Calculations show that the effects of LDFs between mining the structure and stability of these complexes,
the ligands across the metal stabilize the complexes by and driving the disproportionation from the salts of
21.1–29.4 kcal mol−1. metals in the lower oxidation state.
The treatment of transition metal halides with organ- The behaviour discussed above is not unique to the
olithium reagents featuring the powerfully reducing metals of the d block; reaction of the bulky cyclopenta-
alkyl anion frequently affords products with different diene [(4‑nBu-C6H4)5C5H] (abbreviated as CpBIGH) with
oxidation states. A case in point is the 1972 synthesis Ln(iii) tribenzyl species (Ln = Yb or Sm) gives sponta-
of the norbornyl complexes M(1‑nor)4 (M = Ti, V, Cr, neous reduction to Lnii(CpBIG)2 — reactivity that was
Mn, Fe or Co; 1‑nor = bicyclo[2.2.1]hept‑1‑yl), which attributed to the presence of a weakly-bound C–H···C
feature metals in the unusually high 4+ oxidation state (π) hydrogen-bonding network137. Reaction of this
in the case of iron and cobalt 135. Dispersion-corrected ligand with a related system also suggests the importance

NATURE REVIEWS | CHEMISTRY VOLUME 1 | ARTICLE NUMBER 0004 | 9


REVIEWS

Box 2 | Characteristic signs of the existence of significant LDF effects in molecular species

• High stability of compounds despite lengthened bonds62


• Sterically counterintuitive structures102
• Lack of expected correlations of various structural parameters79
• Isomerization to different forms to maximize dispersive interactions94,104–107
• Tendency to display spontaneous disproportionation131,135,136,139
• Counterintuitive effects on equilibrium constants82,84
• Isolability of electronically disfavoured molecules58,61
• Well-defined products from dispersion stabilized structures with conservation of dispersion92
• Increased reactivity for catalytic centres with more sterically demanding substituents124,126
• Apparent mutual attraction of rigid multicyclic groups within molecules22,23,57
LDF, London dispersion force.

of these forces in this Schlenk-type equilibrium. from post-hoc rationalizations of LDF effects (BOX 2).
Reaction of the corresponding dibenzylcalcium species By contrast, the designed incorporation of disper-
with equimolar CpBIGH does not yield the sterically sion force effects into inorganic and organo­metallic
expected heteroleptic species but instead gives half molecules to alter their stability, structure or reactivity
an equivalent of Ca(CpBIG)2 and half an equivalent of has not yet occurred, except in a handful of examples131.
dibenzyl starting material. Despite the growing number of computational studies
focused on the understanding of LDF effects on sterically
Conclusions crowded compounds, there are still many experimen-
LDFs are always present in molecular species. However, tally described systems where these effects need compu-
chemists have only recently realized that such forces tational authentication. The use of both computational
can considerably influence the stability, structure and and experimental methods to incorporate LDF effects
reactivity of inorganic and organometallic derivatives into ligand design should permit a considerable expan-
bearing sterically large substituents. Until now, recog- sion in the range of known classes of compounds and
nizing the importance of these forces has come mainly chemical behaviour.

1. Tang, K.‑T. & Toennies, J. P. Johannes Diderik van der 14. Bourissou, D., Guerret, O., Gabbai, F. P. & 27. Brandenburg, J. G., Hocheim, M., Bredow, T. &
Waals: a pioneer in the molecular sciences and nobel Bertrand, G. Stable carbenes. Chem. Rev. 100, 39–92 Grimme, S. Low-cost quantum chemical methods for
prize winner in 1910. Angew. Chem. Int. Ed. 49, (2000). noncovalent interactions. J. Phys. Chem. Lett. 5,
9574–9579 (2010). 15. Valente, C. et al. Complexes for the most-challenging 4275–4284 (2014).
2. London, F. Zur Theorie und Systematik der cross-coupling reactions. Angew. Chem. Int. Ed. 51, 28. Fey, N., Ridgway, B. M., Jover, J., McMullin, C. L. &
Molekularkräfte. Z. Physik. 63, 245 (1930); English 3314–3332 (2012). Harvey, J. N. Organometallic reactivity: the role of
translation available in London, F. The general theory 16. Scott, N. M. & Nolan, S. P. Stabilization of metal–ligand bond energies from a computational
of molecular forces. Trans. Faraday Soc. 33, 8b–26 organometallic species achieved by the use of perspective. Dalton Trans. 40, 11184–11191 (2011).
(1937). N‑heterocyclic carbene (NHC) ligands. Eur. J. Inorg. 29. Grimme, S. Accurate description of van der Waals
3. Parsegian, V. A. in Van der Waals Forces: A Handbook Chem. 18, 5–1828 (2005). complexes by density functional theory including
for Biologists, Chemists, Engineers, and Physicists 17. Asay, M., Jones, C. & Driess, M. N‑Heterocyclic empirical corrections. J. Comput. Chem. 25,
(Cambridge Univ. Press, 2005). carbene analogues with low-valent group 13 and 1463–1473 (2004).
4. Eisenschitz, R. & London, F. Über das Verhältnis der group 14 elements: syntheses, structures, and 30. Grimme, S. Semiempirical GGA-type density
van der Waalsschen Kräfte zu den homöopolaren reactivities of a new generation of multitalented functional constructed with a long-range dispersion
Bindungskräften. Z. Phys. 60, 491–527 (in German) ligands. Chem. Rev. 111, 354–396 (2011). correction. J. Comput. Chem. 27, 1787–1799
(1930). 18. Jones, C. Bulky guanidinates for the stabilization of (2006).
5. Pace, N. C., Scholtz, J. M. & Grimsley, G. R. Forces low oxidation state metallacycles. Coord. Chem. Rev. 31. Alrichs, R., Penco, R. & Scoles, G. Intermolecular
stabilizing proteins. FEBS Lett. 588, 2177–2184 254, 1273–1289 (2010). forces in simple systems. Chem. Phys. 19, 119–130
(2014). 19. Edelmann, F. T. in Advances in Organometallic (1977).
6. Biedermann, F. & Schneider, H.‑J. Experimental Chemistry Vol. 57 (eds hill, F. E. & Fink, M. J.) 32. Becke, A. D. & Johnson, E. R. A density-functional
binding energies in supramolecular complexes. Chem. 183–352 (Academic Press, 2008). model of the dispersion interaction. J. Chem. Phys.
Rev. 116, 5216–5300 (2016). 20. Mindiola, D. J., Holland, P. L. & Warren, T. H. in 123, 154101 (2005).
7. Mosher, H. S. & Tidwell, T. T. Frank C. Whitmore and Inorganic Syntheses (ed. Rauchfuss, T. B.) (Wiley, 33. Jurečka, P., Černy, J., Hobza, P. & Salahub, D. R.
steric hindrance: a duo of centennials. J. Chem. Educ. 2010). Density functional theory augmented with an empirical
67, 9–14 (1990). 21. Bourget-Merle, L., Lappert, M. F. & Severin, J. R. The dispersion term. Interaction energies and geometries
8. Newman, M. S. Steric Effects in Organic Chemistry chemistry of β‑diketiminatometal complexes. Chem. of 80 noncovalent complexes compared with ab initio
(Wiley, 1956). Rev. 102, 3031–3066 (2002). quantum mechanics calculations. J. Comput. Chem.
9. Power, P. P. Some highlights from the development 22. Schreiner, P. R. et al. Overcoming lability of extremely 28, 555–569 (2007).
and use of bulky monodentate ligands. J. Organomet. long alkane carbon–carbon bonds through dispersion 34. Zhao, Y. & Truhlar, D. G. Density functionals with
Chem. 689, 3904–3919 (2004). forces. Nature 477, 308–311 (2011). broad applicability in chemistry. Acc. Chem. Res. 41,
10. Clyburne, J. A. C. & McMullen, N. Unusual structures 23. Fokin, A. A. et al. Stable alkanes containing very long 157–167 (2008).
of main group organometallic compounds containing carbon–carbon bonds. J. Am. Chem. Soc. 134, 35. Zhao, Y. & Truhlar, D. G. Applications and validations
m‑terphenyl ligands. Coord. Chem. Rev. 210, 73–99 13641–13650 (2012). of the Minnesota density functionals. Chem. Phys.
(2000). 24. Grimme, S. & Schreiner, P. R. Steric crowding can Lett. 502, 1–13 (2011).
11. Twamley, B., Haubrich, S. T. & Power, P. P. in Advances stabilize a labile molecule: solving the 36. Johnson, E. R. & Becke, A. D. Van der Waals
in Organometallic Chemistry Vol. 44 1–65 (Academic hexaphenylethane riddle. Angew. Chem. Int. Ed. 50, interactions from the exchange hole dipole moment:
Press, 1999). 12639–12642 (2011). application to bio-organic benchmark systems.
12. Ni, C. & Power, P. P. in Metal–Metal Bonding Vol. 136 25. Echeverría, J., Aullón, G., Danovich, D., Shaik, S. & J. Chem. Phys. Lett. 432, 600–603 (2006).
(ed. Parkin, G.) 59–111 (Springer, 2010). Alvarez, S. Dihydrogen contacts in alkanes are subtle 37. Becke, A. D. & Johnson, E. R. A unified density-
13. Arduengo, A. J. III Looking for stable carbenes: the but not faint. Nat. Chem. 3, 323–330 (2011). functional treatment of dynamical, nondynamical, and
difficulty in starting anew. Acc. Chem. Res. 32, 26. Pyykkö, P. Strong closed-shell interactions in inorganic dispersion correlations. J. Chem. Phys. 127, 124108
913–921 (1999). chemistry. Chem. Rev. 97, 597–636 (1997). (2007).

10 | ARTICLE NUMBER 0004 | VOLUME 1 www.nature.com/natrevchem


REVIEWS

38. Ruszinsky, A., Perdew, J. P. & Csonka, G. J. A simple 63. Schäfer, A., Weidenbruch, M., Peters, K. & von 86. Seidu, I., Seth, M. & Ziegler, T. Role played by
but fully nonlocal correction to the random phase Schnering, H. Hexa-tert-butylcyclotrisilane, a strained isopropyl substituents in stabilizing the putative triple
approximation. J. Chem. Phys. 134, 114110 (2011). molecule with unusually long Si–Si and Si–C bonds. bond in ArʹEEArʹ[E = Si, Ge, Sn; Arʹ = C6H3‑2,6-
39. Eshuis, H., Yarkony, J. & Furche, F. Fast computation of Angew. Chem. Int. Ed. 23, 302–303 (1984). (C6H3‑2,6‑Pri2)2] and Ar*PbPbAr* [Ar* = C6H3‑2,6-
molecular random phase approximation correlation 64. Wiberg, N., Schuster, A., Simon, A. & Peters, K. (C6H2‑2,4,6‑Pri3)2]. Inorg. Chem. 52, 8378–8388
energies using resolution of the identity and imaginary Hexa-tert-butyldisilane — the molecule with the (2013).
frequency integration. J. Chem. Phys. 132, 234114 longest Si–Si bond. Angew. Chem. Int. Ed. 25, 79–80 87. Stender, M., Phillips, A. D., Wright, R. J. & Power, P. P.
(2010). (1986). Synthesis and characterization of a digermanium
40. Furche, F. & Perdew, J. P. The performance of 65. Pyykkö, P. & Atsumi, M. Molecular single-bond analogue of an alkyne. Angew. Chem. Int. Ed. 41,
semilocal and hybrid density functionals in 3d covalent radii for elements 1–118. Chem. Eur. J. 15, 1785–1787 (2002).
transition-metal chemistry. J. Chem. Phys. 124, 186–197 (2008). 88. Phillips, A. D., Wright, R. J., Olmstead, M. M. &
044103 (2006). 66. Pauling, L. Nature of the Chemical Bond 239 (Cornell Power, P. P. Synthesis and characterization of
41. Jiménez-Hoyos, C. A., Janesko, B. G. & Scuseria, G. E. Univ. Press, 1960). 2,6‑Dipp2‑H3C6SnSnC6H3‑2,6‑Dipp2
Evaluation of range-separated hybrid and other 67. Paolini, J. P. The bond order–bond length (Dipp = C6H3‑2,6‑Pri2): a tin analogue of an alkyne.
density functional approaches on test sets relevant for relationship. J. Comput. Chem. 11, 1160–1163 J. Am. Chem. Soc. 124, 5930–5931 (2002).
transition metal-based homogeneous catalysts. (1990). 89. Pu, L., Twamley, B. & Power, P. P. Synthesis and
J. Phys. Chem. A. 113, 11742–11749 (2009). 68. Bock, H., Meuret, J. & Ruppert, K. Sterically characterization of 2,6‑Trip2H3C6PbPbC6H3‑2,6‑Trip2
42. Ryde, U., Mata, R. A. & Grimme, S. Does DFT‑D overcrowded or charge perturbed molecules: XXIII. (Trip = C6H2‑2,4,6‑i-Pr3): a stable heavier group 14
estimate accurate energies for the binding of ligands Hexakis(trimethylsilyl)disilane: structure and element analogue of an alkyne. J. Am. Chem. Soc.
to metal complexes? Dalton Trans. 40, 11176–11183 photoelectron spectrum of a sterically overcrowded 122, 3524–3525 (2000).
(2011). molecule. J. Organomet. Chem. 445, 19–28 90. Mitoraj, M., Michalak, A. & Ziegler, T. A. Combined
43. Swart, M., Solá, M. & Bickelhaupt, F. M. Inter- and (1993). charge and energy decomposition scheme for bond
intramolecular dispersion interactions. J. Comput. 69. Weidenbruch, M. et al. Hexa‑t‑butyldigerman und analysis. J. Chem. Theor. Comput. 5, 962–975 (2009).
Chem. 32, 1117–1127 (2011). Hexa‑t‑butylcyclotrigerman: moleküle mit den derzeit 91. Wu, L.‑C., Jones, C., Stasch, A., Platts, J. A. &
44. Yang, L., Adam, C., Nichol, G. S. & Cockroft, S. L. How längsten Ge–Ge− und Ge–C‑Bindungen. Overgaard, J. Non-nuclear attractor in a molecular
much do van der Waals dispersion forces contribute to J. Organomet. Chem. 341, 335–343 (in German) compound under external pressure. Eur. J. Inorg.
molecular recognition in solution? Nat. Chem. 5, (1988). Chem. 32, 5536–5540 (2014).
1006–1010 (2013). 70. Puff, H. et al. Bindungsabstände zwischen 92. Wagner, J. P. & Schreiner, P. R. London dispersion
45. Hansen, A. et al. The thermochemistry of london organylsubstituierten Zinnatomen: III. Offenkettige decisively contributes to the thermodynamic stability
dispersion-driven transition metal reactions: getting Verbindungen. J. Organomet. Chem. 363, 265–280 of bulky NHC-coordinated main group compounds.
the ‘right answer for the right reason’. ChemistryOpen (in German) (1989). J. Chem. Theor. Comp. 12, 231–237 (2016).
3, 177–189 (2014). 71. Wiberg, N. et al. Tetrasupersilyl-tristannaallene and 93. Hänninen, M., Pal, K., Day, B. M., Pugh, T. &
46. Kronik, L. & Tkatchenko, A. Understanding molecular -tristannacyclopropene (tBu3Si)4Sn3 — isomers with Layfield, R. A three-coordinate iron–silylene complex
crystals with dispersion-inclusive density functional the shortest S=Sn double bonds to date. Eur. J. Inorg. stabilized by ligand–ligand dispersion forces. Dalton
theory: pairwise corrections and beyond. Acc. Chem. Chem. 1999, 1211–1218 (1999). Trans. 45, 11301–11305 (2016).
Res. 47, 3208–3216 (2014). 72. Peng, Y. et al. Substituent effects in ditetrel alkyne 94. Albers, L., Rathjen, S., Baumgartner, J., Marschner, C.
47. Berland, K. et al. I. van der Waals forces in density analogues: multiple versus single bonded isomers. & Müller, T. Dispersion-energy-driven Wagner–
functional theory: a review of the vdW‑DF method. Chem. Sci. 1, 461–468 (2010). Meerwein rearrangements in oligosilanes. J. Am.
Rep. Prog. Phys. 78, 066501 (2015). 73. Wiberg, N., Amelunxen, K., Blank, T., Nöth, H. & Chem. Soc. 138, 6886–6892 (2016).
48. Grimme, S. in The Chemical Bond: Chemical Bonding Knizek, J. Tetrasupersilyldialuminum [(t‑Bu)3Si]2Al– 95. Andersen, R. A. et al. The molecular structures of
Across the Periodic Table (eds Frenking, G. & Shaik, S.) Al[Si(t‑Bu)3]2: the dialane(4) with the longest Al–Al bis(pentamethylcyclopentadienyl)-calcium and
477–500 (Wiley, 2014). bond to date. J. Organometallics 17, 5431–5433 -ytterbium in the gas phase; two bent metallocenes.
49. Grimme, S., Hansen, A., Brandenburg, J. G. & (1998). J. Organomet. Chem. 312, C49–C52 (1986).
Bannwarth, C. Dispersion-corrected mean-field 74. Uhl, W. Tetrakis[bis(trimethylsilyl)methyl]dialan(4), 96. Andersen, R. A., Blom, R., Boncella, J. M., Burns, C. J.
electronic structure methods. Chem. Rev. 116, eine Verbindung mit Aluminium–Aluminium-Bindung. & Volden, H. V. The thermal average molecular
5105–5154 (2016). Z. Naturforsch. B 43, 1113–1118 (in German) structures of bis(pentamethylcyclopentadienyl)
50. Bondi, A. Van der Waals volumes and radii. J. Phys. (1988). magnesium(ii), -calcium(ii) and -ytterbium(ii) in the
Chem. 68, 441–451 (1964). 75. Wehmschulte, R. J. et al. Reduction of a gas phase. Acta Chem. Scand. 41A, 24–35 (1987).
51. Gomberg, M. Triphenylmethyl, ein Fall von tetraaryldialane to generate Al–Al π-bonding. Inorg. 97. Andersen, R. A., Blom, R., Burns, C. J. & Volden, H. V.
dreiwerthigem Kohlenstoff. Ber. Dtsch. Chem. Ges. 33, Chem. 32, 2983–2984 (1993). Synthesis and thermal average gas phase molecular
3150–3163 (in German) (1900). 76. Wiberg, N. et al. Ditrielanes (R3Si)2E–E(SiR3)2 and structures of bis(pentamethylcyclopentadienyl)-
52. Gomberg, M. An instance of trivalent carbon: heterocubanes (R3Si)4E4Y4 (R3Si = tBu3Si, tBu2PhSi; strontium and -barium; the first organo-strontium and
triphenylmethyl. J. Am. Chem. Soc. 22, 757–771 E = Al, Ga, In, Tl; Y = O, Se). Eur. J. Inorg. Chem. -barium structures. J. Chem. Soc., Chem. Commun.
(1900). 341–350 (2002). 768–769 (1987).
53. Lankamp, H., Nauta, W. Th. & MacLean, C. A new 77. Wiberg, N. et al. Tris(tri-tert-butylsilyl)digallanyl 98. Blom, R., Faegri, K. Jr & Volden, H. V. Molecular
interpretation of the monomer-dimer equilibrium of (tBu3Si)3Ga2: a new type of compound for a heavy structures of alkaline earth-metal metallocenes:
triphenylmethyl- and alkylsubstituted-diphenyl methyl- group 13 element. Angew. Chem. Int. Ed. 36, electron diffraction and ab initio investigations.
radicals in solution. Tetrahedron Lett. 9, 249–254 1213–1215 (1997). Organometallics 9, 372–379 (1990).
(1968). 78. Power, P. P. π‑Bonding and the lone pair effect in 99. Williams, R. A., Hanusa, T. P. & Huffman, J. C.
54. Stein, M., Winter, W. & Rieker, A. Hexakis(2,6‑di‑tert- multiple bonds between heavier main group elements. Structures of ionic decamethylmetallocenes:
butyl‑4‑biphenylyl)ethane — the first unbridged Chem. Rev. 99, 3463–3503 (1999). crystallographic characterization of
hexaarylethane. Angew. Chem. Int. Ed. Engl. 17, 79. Fischer, R. C. & Power, P. P. π‑Bonding and the lone bis(pentamethylcyclopentadienyl)calcium and -barium
692–694 (1978). pair effect in multiple bonds involving heavier main and a comparison with related organolanthanide
55. Kahr, B., van Engen, D. & Mislow, K. Length of the group elements: developments in the new millennium. species. Organometallics 9, 1128–1134 (1990).
ethane bond in hexaphenylethane and its derivatives. Chem. Rev. 110, 3877–3923 (2010). 100. Hollis, T. K., Burdett, J. K. & Bosnich, B. Why are
J. Am. Chem. Soc. 108, 8305–8307 (1986). 80. Arp, H., Baumgartner, J., Marschner, C., Zark, P. & bis(pentamethylcyclopentadienyl) complexes, [MCp2*],
56. Wagner, J. P. & Schreiner, P. R. London dispersion in Müller, T. Dispersion energy enforced dimerization of a of calcium, strontium, barium, samarium, europium,
molecular chemistry — reconsidering steric effects. cyclic disilylated plumbylene. J. Am. Chem. Soc. 134, and ytterbium bent? Organometallics 12,
Angew. Chem. Int. Ed. 54, 12274–12296 (2016). 6409–6415 (2012). 3385–3386 (1993).
57. Schwertfeger, H., Fokin, A. A. & Schreiner, P. R. 81. Weidenbruch, M., Kilian, H., Peters, K., von 101. Timofeeva, T. V., Lii, J.‑H. & Allinger, N. L. Molecular
Diamonds are a chemist’s best friend: diamondoid Schnering, H. G. & Marsmann, H. Compounds of mechanics explanation of the metallocene bent
chemistry beyond adamantane. Angew. Chem. Int. Ed. germanium and tin, 16. A tetraaryldistannene with a sandwich structure. J. Am. Chem. Soc. 117,
47, 1022–1036 (2008). long tin–tin multiple bond and differing environments 7452–7459 (1995).
58. Maier, G., Pfriem, S., Schäfer, R. & Mausch, R. Tetra- at the tin atoms. Chem. Ber. 128, 983–985 (1995). 102. Rekken, B.‑D. et al. Dispersion forces and
tert-butyltetrahedrane. Angew. Chem. Int. Ed. 17, 82. Guo, J.‑D., Liptrot, D. J., Nagase, S. & Power, P. P. The counterintuitive steric effects in main group molecules:
520–521 (1978). multiple bonding in heavier group 14 element alkene heavier group 14 (Si–Pb) dichalcogenolate carbene
59. Balci, M., McKee, M. & Schleyer, P. v. R. Theoretical analogues is stabilized mainly by dispersion force analogues with sub‑90° interligand bond angles.
study of tetramethyl- and tetra-tert-butyl-substituted effects. Chem. Sci. 6, 6235–6244 (2015). J. Am. Chem. Soc. 135, 10134 (2013).
cyclobutadiene and tetrahedrane. J. Phys. Chem. 104, 83. Lee, V. Ya. et al. (tBu2MeSi)2SnSn(SiMetBu2)2: a 103. Eaborn, C. & Smith, J. D. Organometallic compounds
1246–1255 (2000). distannene with a >Sn=Sn< double bond that is containing tris(trimethylsilyl)methyl or related ligands.
60. Monteiro, N. K. V., de Oliveira, J. F. & Firme, C. L. stable both in the solid state and in solution. J. Am. J. Chem. Soc., Dalton Trans. 1541–1552 (2001).
Stability and electronic structures of substituted Chem. Soc. 128, 11643–11651 (2006). 104. Eaborn, C., Hitchcock, P. B., Smith, J. D. &
tetrahedranes, silicon and germanium parents — a 84. Guo, J.‑D., Nagase, S. & Power, P. P. Dispersion force Sullivan, A. C. Crystal structure of the tetrahydrofuran
DFT, ADMP, QTAIM and GVB study. New. J. Chem. 38, effects on the dissociation of ‘Jack‑in‑the-box’ adduct of tris(trimethylsilyl)-methyl-lithium, [Li(thf)4]
5892–5904 (2014). diphosphanes and diarsanes. Organometallics 34, [Li{C(SiMe3)3}2], an ate derivative of lithium. J. Chem.
61. Nemirowski, A., Reisenauer, H. P. & Schreiner, P. R. 2028–2033 (2015). Soc., Chem. Commun. 827–828 (1983).
Tetrahedrane — dossier of an unknown. Chem. Eur. J. 85. Hinchley, S. L. et al. Spontaneous generation of stable 105. Buttrus, N. H. et al. The crystal structure of [(pmdeta)
12, 7411–7420 (2006). pnictinyl radicals from ‘Jack‑in‑the-box’ dipnictines: a Li(μ-Cl)Li(pmdeta)][Li{C(SiMe3)3}2]
62. Wiberg, N. Sterically overloaded supersilylated main solid-state, gas-phase, and theoretical investigation of [pmdeta = Me2N(CH2)2NMe(CH2)2NMe2]. A novel
group elements and main group element clusters. the origins of steric Stabilization. J. Am. Chem. Soc. linear chlorine-centred cation. J. Chem. Soc., Chem.
Coord. Chem. Rev. 163, 217–252 (1997). 123, 9045–9053 (2001). Commun. 969–970 (1986).

NATURE REVIEWS | CHEMISTRY VOLUME 1 | ARTICLE NUMBER 0004 | 11


REVIEWS

106. Al‑Juaid, S. S. et al. Metalation of tris(trimethylsilyl)- 117. Evans, W. J., Hughes, L. A. & Hanusa, T. P. Synthesis extended transition state method. Inorg. Chem. 51,
and tris(dimethylphenylsilyl)methane with and crystallographic characterization of an unsolvated, 7794–7800 (2012).
methylsodium: the first dialkylsodate. Angew. Chem. monomeric samarium bis(pentamethyl­cyclopenta­ 129. Nguyen, T. et al. Synthesis of a stable compound with
Int. Ed. 33, 1268–1270 (1994). dienyl) organolanthanide complex, (C5Me5)2Sm. J. Am. fivefold bonding between two chromium(i) centers.
107. Al‑Juaid, S. S. et al. Crystal structures of organometallic Chem. Soc. 106, 4270–4272 (1984). Science 310, 844–861 (2005).
compounds of lithium and magnesium containing the 118. Evans, W. J., Forrestal, K. J. & Ziller, J. W. Reaction 130. Power, P. P. Stable two-coordinate, open-shell (d1–d9)
bulky ligands C(SiMe3)2(SiMe2X) X = Me, Ph, NMe2, or chemistry of sterically crowded tris(pentamethylcyclo­ transition metal complexes. Chem. Rev. 112,
C5H4N-2. J. Organomet. Chem. 631, 76–86 (2001). pentadienyl)samarium. J. Am. Chem. Soc. 120, 3482–3507 (2012).
108. Eaborn, C., Hitchcock, P. B., Smith, J. D. & Sullivan, A. C. 9273–9282 (1998). 131. Wagner, C. L. et al. Dispersion-force-assisted
A novel monomeric alkyl–lithium compound. Crystal 119. Evans, W. J., Hughes, L. A. & Hanusa, T. P. Synthesis disproportionation: a stable two-coordinate copper(ii)
structure of [Li{C(SiMe2Ph)3}(tetrahydrofuran)]. and X‑ray crystal structure of complex. Angew. Chem. Int. Ed. 55, 10444–10447
J. Chem. Soc., Chem. Commun. 1390–1391 (1983). bis(pentamethylcyclopenta­dienyl) complexes of (2016).
109. Eaborn, C., Hitchcock, P. B., Smith, J. D. & samarium and europium: (C5Me5)2Sm and (C5Me5)2Eu. 132. Boynton, J. N. et al. Linear and nonlinear two-
Sullivan, A. C. Preparation and crystal structure of the Organometallics 5, 1285–1288 (1986). coordinate vanadium complexes: synthesis,
tetrahydrofuran adduct of lithium bis 120. Evans, W. J., Gonzales, S. L. & Ziller, J. W. Synthesis characterization, and magnetic properties of V(ii)
[tris(trimethylsilyl)methyl]cuprate, [Li(THF)4] and X‑ray crystal structure of the first amides. J. Am. Chem. Soc. 135, 10720–10728 (2013).
[Cu{C(SiMe3)3}2]. The first structural characterization tris(pentamethylcyclopentadienyl)metal complex: 133. Lin, C.‑Y. et al. Dispersion force stabilized two-
of a Gilman reagent. J. Organomet. Chem. 263, (η5‑C5Me5)3Sm. J. Am. Chem. Soc. 113, 7423–7424 coordinate transition metal–amido complexes of the
c23–c25 (1984). (1991). –N(SiMe3)Dipp (Dipp = C6H3‑2,6‑Pri2) ligand:
110. Al‑Juaid, S. S., Eaborn, C., Hitchcock, P. B., 121. Ahlquist, M. S. G. & Norrby, P.‑O. Dispersion and structural, spectroscopic, magnetic, and computational
McGeary, C. A. & Smith, J. D. The crystal structure of back-donation gives tetracoordinate [Pd(PPh3)4]. studies. Inorg. Chem. 52, 13584–13593 (2013).
bis{tris(trimethylsilyl)methyl}magnesium: an example Angew. Chem. Int. Ed. 50, 11794–11797 (2011). 134. Faust, M. et al. The instability of Ni{N(SiMe3)2}2: a
of two‑co‑ordinate magnesium in the solid state. 122. Lyngvi, E., Sanhueza, I. A. & Schoenebeck, F. fifty year old transition metal silylamide mystery.
J. Chem. Soc., Chem. Commun. 1989, 273–274 Dispersion makes the difference: bisligated transition Angew. Chem. Int. Ed. 54, 12914–12917 (2015).
(1989). states found for the oxidative addition of Pd(PtBu3)2 to 135. Bower, B. K. & Tennent, H. G. Transition metal
111. Al‑Juaid, S. S. et al. Preparation, crystal structure, and Ar‑OSO2R and dispersion-controlled chemoselectivity bicyclo[2.2.1]hept‑1‑yls. J. Am. Chem. Soc. 94,
reactivity of bis {tris(trimethylsilyl) methyl} in reactions with Pd[P(iPr)(tBu2)]2. Organometallics 34, 2512–2518 (1972).
magnesium. J. Organomet. Chem. 480, 199–203 805–812 (2015). 136. Liptrot, D. J., Guo, J.‑D., Nagase, S. & Power, P. P.
(1994). 123. Maseras, F. & Eisenstein, O. Opposing steric and Dispersion forces, disproportionation and stable high-
112. Eaborn, C. & Hitchcock, P. B. The first structurally electronic contributions in OsCl2H2(PPr3i)2. A valent late transition metal alkyls. Angew. Chem. Int.
characterised solvent-free ς‑bonded diorganocalcium, theoretical study of an unusual structure. New Ed. 55, 13655–13659 (2016).
Ca[C(SiMe3)3]2. Chem. Commun. 1961–1962 (1997). J. Chem. 22, 5–9 (1998). 137. Lewis, R. A. et al. Reactivity and Mössbauer
113. Westerhausen, M., Rademacher, B. & Poll, W. 124. Minenkov, Y., Occhipinti, G., Heyndrickx, W. & spectroscopic characterization of an Fe(iv) ketimide
Trimethylsilyl-substituierte Derivate des Dimethylzinks Jensen, V. R. The nature of the barrier to phosphane complex and reinvestigation of an Fe(iv) norbornyl
— Synthese, spektroskopische Charakterisierung und dissociation from grubbs olefin metathesis catalysts. complex. Inorg. Chem. 52, 8218–8227 (2013).
Struktur. J. Organomet. Chem. 421, 175–188 Eur. J. Inorg. Chem. 1507–1516 (2012). 138. Byrne, E. K. & Theopold, K. H. Redox chemistry of
(in German) (1991). 125. Minenkov, Y., Singstad, A., Occhipinti, G. & tetrakis(1-norbornyl)cobalt. Synthesis and
114. Eaborn, C., Jones, K. L., Smith, J. D. & Tavakkoli, K. Jensen, V. R. The accuracy of DFT-optimized characterization of a cobalt(v) alkyl and self-exchange
The remarkable thermal stability of geometries of functional transition metal compounds: rate of a Co(iii)/Co(iv) couple. J. Am. Chem. Soc. 193,
benzyl[tris(dimethylphenylsily)methyl]mercury. a validation study of catalysts for olefin metathesis 1282–1283 (1987).
How can a bulky ligand stabilize an organometallic and other reactions in the homogeneous phase. 139. Ruspic, C., Moss, J. R., Schürmann, M. & Harder, S.
compound towards unimolecular dissociation? Dalton Trans. 41, 5526–5541 (2012). Remarkable stability of metallocenes with superbulky
J. Chem. Soc., Chem. Commun. 1201–1202 (1989). 126. Wolters, L. P., Koekkoek, R. & Bickelhaupt, F. M. Role ligands: spontaneous reduction of Smiii to Smii. Angew.
115. Al‑Juaid, S. S., Eaborn, C., Lickiss, P. D., Smith, J. of steric attraction and bite-angle flexibility in metal- Chem. Int. Ed. 47, 2121–2126 (2008).
Davis, Tavakkoli, K. & Webb, A. D. Preparation, mediated C–H bond activation. ACS Catal. 5,
spectroscopic properties and thermal stabilities of 5766–5775 (2015). Acknowledgements
organomercury compounds containing the bulky 127. Wolstenholme, D. J., Dobson, J. L. & McGrady, G. S. The authors are grateful to the David Parkin Visiting
ligand (Me3Si)3C or (PhMe2Si)3C. J. Organomet. Chem. Homopolar dihydrogen bonding in main group Professorship at the University of Bath (P.P.P.), the English-
510, 143–151 (1996). hydrides: discovery, consequences, and applications. Speaking Union Lindemann Trust Fellowship (D.J.L.), the US
116. Ghotra, J. S., Hursthouse, M. B. & Welch, A. J. Dalton Trans. 44, 9718–9731 (2015). National Science Foundation (CHE‑1565501) and M. Hill for
Three‑co‑ordinate scandium(iii) and europium(iii); 128. Ndambuki, S. & Ziegler, T. Analysis of the putative his generosity, invaluable advice and support.
crystal and molecular structures of their Cr–Cr quintuple bond in ArʹCrCrArʹ
trishexamethyldisilylamides. J. Chem. Soc., Chem. (Arʹ = C6H3‑2,6(C6H3‑2,6‑Pri2)2 based on the Competing interests statement
Commun. 6, 9–670 (1973). combined natural orbitals for chemical valence and The authors declare no competing interests.

12 | ARTICLE NUMBER 0004 | VOLUME 1 www.nature.com/natrevchem


O N L I N E O N LY

Subject categories
Physical sciences / Chemistry / Inorganic chemistry / Chemical bond-
ing [URI /639/638/263/910]
Physical sciences / Chemistry / Inorganic chemistry / Organometallic
chemistry / Chemical bonding [URI /639/638/263/406/910]

ToC blurb

000 London dispersion forces in sterically


crowded inorganic and organometallic
molecules
David J. Liptrot and Philip P. Power
Despite its relative weakness, the London dispersion
force can strongly influence physical and chemical
properties of molecules. This Review highlights
how structure and reactivity of organometallic and
inorganic molecules are greatly affected by the
Carbon Hydrogen Silicon Lithium
cumulative effect of this force.

You might also like