You are on page 1of 9

CH2207 – Coordination Chemistry and Ligand Types

General Trends in Transition Metal Chemistry


The most important characteristic of a transition metal in determining a large amount of their
chemistry is the oxidation state. This is formalism and doesn’t strictly define the electronic structure
of the metal, however it does provide a useful method for distinguishing between different forms
and different metals and explaining their chemical reactions. Metals will always seek ways to
increase or decrease their electron-density through combination with the appropriate donor types
(Hard Soft Acid Base theory/metal-ligand complementation).
As the transition metals are followed from left to right on the periodic table there is an increase in
the number of available oxidation states, scandium is found exclusively in the +3 oxidation state,
titanium in the +3 and +4 states and so forth, manganese is found in any oxidation state between +2
and +7, post-manganese the number of available oxidation states (as well as the magnitude)
decreases until zinc which exists only in the +1 oxidation state. Early transition metals are generally
low electronegativity, higher oxidation state, electron-poor ‘hard’ metals (oxophilic). The late
transition metals are then higher electronegativity, lower oxidation state, electron rich ‘soft’ metals
(sulphilic). The electronegativities of the transition elements tend to range from 1.1 (Lanthanum) to
2.4 (Gold) (using the Pauling scale) and increase down the early group triads while increasing down
the latter group triads.

Another important characteristic of transition metal complexes is the 18 electron rule, or the
effective atomic number (EAN). A stable organometallic complex will result when the sum of metal
valence electrons plus the total number of electrons donated by the ligands totals 18. This is the d-
block equivalent of the octet rule (and essentially is the octet rule taking into account the 5 d
orbitals, thus 10 additional electrons, available to the transition metals). There are 3 types of
complex regarding the 18 electron rule, each described by a separate molecular orbital diagram;

Class A – weak σ-donors (small Δo); these contain 6 low-lying molecular orbitals and 6 medium
energy molecular orbitals. These are 12-22 electron species;

Class B – strong σ-donors (large Δo); these contain 6 low-lying molecular orbitals and 3 medium
energy molecular orbitals. These are 12-18 electron species due to the now larger difference
between t2g and eg;

Class C – π-acceptors; these contain 9 low-lying molecular orbitals and are 18 electron species, they
are often homoleptic species containing the carbonyl (CO) ligand;

For early transition metals 16 electron and sub-16 electron configurations are common with
coordination numbers greater than 6. For middle transition metals 18 electron configurations are
common, with coordination numbers equal to 6. For late transition metals 16 electron and less than
16 electron configurations are common with coordination numbers less than 6.

There are several exceptions to the 18 electron rule, it acts only as a guideline, these exceptions
include;
-there is a small value of Δo (from first to second row Δo increases ~40% and a further 10% from
second to third row)
-there is a low metal electron count (early transition metals)
-there are stearic interactions, large ligands may prevent a metal achieving 18 electrons
-electron rich d8 metal complexes, these are square-planar and possess 8 low-lying molecular
orbitals, these abide by a 16 electron rule.

Any species with less than 18 valence electrons can be considered coordinatively unsaturated,
these will seek to achieve 18 valence electrons through addition of more ligands. Those with 18
valence electrons are therefore deemed coordinatively saturated.

Electron-Counting in Transition Metal Complexes


Electron-counting is, very simply, the process of determining the number of valence electrons
associated within any given transition metal complex.
The ionic method treats each ligand as a closed shell entity, as atoms or molecules that have
achieved an octet of electrons. The charge on the ligand is therefore determined by the nature of
the species – how many electrons must be added to achieve the octet.

There are several types of ligand;


-neutral 2 electron donors, i.e. CO, PR3
-anionic 2 electron donors, i.e. X-, -CR3
-anionic 4 electron donors, i.e. ƞ3-C3H5- (allyl)
-anionic 6 electron donors, i.e. ƞ5-C5H5- (cyclopentadienyl)
-anionic donors that can show variable electron donation through use of available lone pairs, -OR
(alkoxides, 2/4/6 electron), -NR2(amide, 2/4 electron).

There are also briding ligands that differ from those listed above as they donate to more than one
metal atom. These are discussed further on.

To determine the electron-count using the ionic method the oxidation state of the metal (and hence
the d-electron count) for the metal must be known, by considering the charge of the ligands and the
overall charge of the species this is easily calculated. The total number of electrons on the metal and
donated by the ligands is then the electron-count for that compound.

Bridging Ligands
Bridging in ligands is denoted by mu, μx, the x denoting how many metal atoms the ligand bridges.
Halides are able to donate upto 3 lone pairs, maintaining the single negative charge in each case,
allowing for 3 metal atoms to be bridged by any one halide. Carbon monoxide may also act as a
bridging ligand however in doing so it only donates one electron to each metal atom, it also
undergoes a change in bond order (from 3 to 2 formally) which can be proven with spectroscopic
data which shows a weakening in the CO bond by the resulting shift in wavelength of absorption.

Hapticity
Hapticity describes how many carbons, in a conjugated π-system, are bonded to the metal atom. It is
denoted by the symbol eta, ƞx, where x is the number of carbons. The x carbons are usually
conjoined and a ligand is referred to polyhapto if x is greater than 2. When the value of x is odd the
ligands are, almost exclusively, anionic and so donate x+1 electrons to the metal atom, when x is
even the ligands are, almost exclusively, neutral and so x electrons are donated. ƞ1, although
seemingly pointless, can be necessary in describing a ligand which can bond via several modes to the
metal atom when that ligand is bonded via only one carbon.
The Geometries of Coordinated Complexes

Coordination Number Geometries Examples


2 Linear Electron rich d10 transition metal
complexes, Ph3P-Au-Cl
3 Trigonal planar
4 Tetrahedral Usually obey 18 electron rule, i.e.
[Co(CO)3(NO)]
Square Planar Usually obey 16 electron rule,
[Ni(CN)4]2-
5 Trigonal bipyramid Fe(CO)5
Square pyramid
Note; the energy difference between
these two systems is low and so many 5
coordinate complexes switch between
the two
6 Octahedral (very common)
Trigonal prism (rare)
7-10 Rare for the first row transition metals,
only possible with 2nd/3rd row metals
and small ligands, more common in the
f-block

When considering polyhapto ligands (which could be seen either as one ligand or several) it is best to
define them by the number of electron pairs they donate, and so a cyclopentadienyl ligand is
considered 3 substituents, i.e. iron in [Fe(ƞ5-C5H5)2] is considered 6-coordinate and not 2.

π-bases
Any ligand that, upon coordination, retains one or more lone pairs may act as a π-base through
donation of the lone pair into an empty metal orbital, in doing so forming a metal-ligand double
bond.

 Halide donors, X-
In general Br-/Cl-/F- are poor π-bases due to their high electronegativity, however I- is able to
act as a stronger π-base.
 Carbon donors, R2C2-/RC3-
Carbenes (alkylidenes) contain a formal M=CR2 bond derived from R2C-, this is a 4-electron
donor that is essentially based on sp2 hybridized carbon. Carbynes possess M≡CR bonds and
are 6-electron donors based on linear sp hybridized carbon in which two orthogonal p-
orbitals are then able to donate into orthogonal orbitals on the metal.
 Oxygen donors, RO-/O2-
a) Alkoxides (alcoholates) of the form RO- can show three bonding modes.
1)sp3 oxygen as a 2-electron donor with 2 lone pairs forms a single
M-O bond with an M-O-R bond angle of around 107o
M O
R
2)sp2 oxygen as a 4-electron donor with 1 lone pair forms a
double M=O bond with an M-O-R bond angle of around 120o

M O
R
3)sp oxygen as a 6-electron donor with no lone pairs forms a
triple M≡O bond with an M-O-R bond angle of around 180o
M O R
When bound at a single metal alkoxides are versatile ligands that can show variable electron-
donation. They are also able to bridge 2 or 3 metals using the lone pairs of electrons available as
with halides. The multiple bonded species are more common with early transition metals and when
the R groups on the alkoxide are bulky.
The M-O bond length will decrease with higher π-contribution and so when performing an electron-
count on a species containing alkoxides it is often the bond length and the bond angle that provide
information on how many electrons each alkoxide is donating. Since the multiplicity is only a formal
application this method can often give confusingly large values for the number of valence electrons,
the reality is that many metal-oxygen bonds are in fact partially multiple.
b)Oxo ligands, O2- are akin to alkoxides but more likely to act as 4-electron donors, especially when
there are two or more oxo groups present. The M≡O unit is common in the absence of better π-
donors if only one oxo ligand is present. Oxo ligands can also doubly or triply bridge.

 Nitrogen donors, R2N-/RN2-/N3-


a) Amido ligands, R2N-, are very common 2/4-electron donors for early transition metals.
Dialkylamido ligands that act as 2-electron donors have a pyramidal geometry (sp3-like)
whereas amides acting as 4-electron donors have a trigonal planar geometry (sp2-like). The
lone pairs in R2N- are considerably higher in energy than those in RO-, thus amides are
stronger donors than the isoelectronic alkoxides and 4-electron donation is much more
prevalent than 2-electron donation for these donors. Amido ligands may also act as bridging
ligands in which case the sp2 form is seen, with 2-electron donation to each metal atom.
b) Alkylimido ligands, RN2-, are isoelectronic with the oxide ligand and bind in a similar
manner to single metal centres. They can donate 4/6-electrons terminally (with a bent
geometry and a linear geometry respectively), or doubly/triply bridge through donation of
4/6-electrons respectively.
c) Nitrido ligands, N3-, are powerful 6-electron donors forming strong M≡N triple bonds.
 Phosphides, R2P-, have a greater tendency to bond to electron-rich metals (as according to
HSAB theory) and are similar to the lighter amide analogues, but with a stronger tendency
towards both the pyramidal and the bridging forms. Again the planar and pyramidal 2/4-
electron forms can be differentiated by bond length and angle and in some cases both can
exist at once, interchanging via lone-pair resonance.

In general the π-donating ability of a ligand decreases as the overall charge on the ligand decreases
and/or the electronegativity of the bonding atom increases and so;

RC3->N3->R2C2->RN2->O2->R2N->RO-
This sequence is dependent on the metal atom, for a given metal a combination of infrared and
Raman spectroscopy can give information on which donor is performing as a better π-donor.

π-Acceptors
π-acceptors are ligands that are able to accept electron-density from metals through π-type
(perpendicular) bonding. In order to act as a π-acid the ligand must have an empty orbital of
appropriate energy and symmetry.

-The carbonyl ligand, CO


Formally considered a triple C≡O bond (despite the resulting positive charge on the oxygen), the
carbonyl ligand contains an sσ* HOMO orbital that acts as a σ-donor, as with any ligand, and a pπ*
LUMO degenerate pair of orbitals that are able to accept electron density from the metal atom. This
process is called back-bonding and is synergic with the σ-donation from the carbon to the metal. CO
is a poor σ-donor, but a great π-acceptor, and so is only likely to form complexes with metals
capable of back-bonding. Back-bonding results in a bond order somewhere between 1 and 2, not
quite a double bond but stronger than a single bond, between the metal and the carbon, however
since electron density is entering the anti-bonding orbital of CO the C-O bond order decreases.
Conversely the σ-donation is removing electron density from the anti-bonding sσ orbital and so
increases the C-O bond order. The CO bond order is most sensitively measured using infrared
spectroscopy, in homoleptic complexes the CO strength decreases as the electron density on the
metal (and so back-bonding) increases.
Carbonyl ligands allow access to highly reduced metal species such as [V(CO) 5]3- as they accept the
electron density from the metals, stabilizing the formal charge.
Carbonyls may also display different modes of coordination;
Symmetric μ2 bridging in which 1-electron donation occurs to each metal (ν(CO)≈1800cm-1, C=O
bond formally)
Asymmetric μ2 bridging in which 2-electron donation occurs to each metal (ν(CO)≈1800cm-1, C≡O
bond formally)
μ3 bridging in which 1-electron donation occurs to each metal (ν(CO)≈1680cm-1, C-O bond formally)
Isocarbonyl in which both oxygen and carbon donate 2-electrons to separate metals.

-The nitrosyl ligand, .NO


Nitrogen monoxide, nitrosyl/.NO, is a stable free-radical that exists in equilibrium with N 2O2. The
bonding in the NO radical is similar to that of CO except the p π* orbital is singly occupied by the
additional electron and the NO bond order is reduced to 2.5 formally. The loss of this electron gives
NO+ which is isoelectronic with CO and binds to transition metals in the same manner but as a
cationic 2-electron donor. This is a stronger π-acceptor but a weaker σ-donor due to the
corresponding decrease in orbital energy due to the positive charge and increased electronegativity
of nitrogen compared to carbon. NO can also gain an electron to give NO-, an anionic 2-electron
donor with a formal N=O double bond.
As such M-N≡O (+NO) is a straight bond whereas M-N=O (-NO) is a bent linear bond. Again this
geometry change can be used to determine the form of NO present in a given complex.

Other π-acceptors include N2, CS and RNC.

Transition Metal Complexes with π-bonded Carbon Ligands – Alkenes and Alkynes as Ligands
There are two bonding components in the bonding of an alkene to a metal;
a)σ-donation of a π-bonding electron-
pair on the alkene to an empty σ-type
orbital on the metal
b)π-donation from a filled d-orbital on
the metal into an empty π*-orbital on
the alkene.

The interaction is, as with CO complexes, dominated by the back-bonding component and again
both components increase the M-C bond strength (they are synergic) however in this scenario both
weaken the C=C double bond.
For poorly π-basic metals such as Pd(II), Pt(II) and Cu(I) the C=C bond order is only slightly lowered,
in these cases bonding is almost entirely σ-type. In this case the double bond is retained and the sp2
geometry of each carbon remains intact.

For strongly π-basic metals (low oxidations tates) and/or the presence of electron withdrawing
groups on the alkene can greatly increase the back-bonding, and so decreasing the strength of the
C=C double bond, consequently the carbons approach the sp3 geometry expected of singly bonded
carbons.
Aside from this change in geometry the dynamic processes by which the ligands move is affected by
this phenomenon. The σ-bonding aspect is unaffected by rotation, however the π-bonding aspect is
strongly affected by rotation; it is at a maximum when the interacting π-orbitals are co-planar and at
its weakest when they are perpendicular. When the metal-alkene bond has an appreciable π-
component rotation of the ligand about the metal is restricted and may be observed for some
systems using variable temperature NMR spectroscopy (VT-NMR). This barrier to rotation will effect
a stronger π-acid more strongly, i.e. C2F4 is more effected than C2H4.

Phosphines – PR3
Phosphines are neutral 2-electron donors that offer versatility since their donor properties, both
electronic and steric, can be readily modified by changes to the ‘R’ substituents. They also serve as a
spectroscopic handle, 31P-NMR can be used in the studies of transition metal complexes to
determine their properties. Phosphines are strong σ-donors that, depending on R, can also act as π-
acids, the extent of both the σ-donation and π-acceptance is dependent on R. Electron donating R
groups enhance the σ-donation whereas electron withdrawing R groups enhance the π-acidity.

The π-receptor type orbitals in phosphines are a combination of the phosphorus d-orbitals and a
degenerate set of two P-R σ*-orbitals.
This back-bonding mechanism is supported using evidence from redox-related pairs of complexes,
the M-P bond strengthens upon one or two-electron reduction of the metal centre, the bond is
strengthened as electron density is increased at the metal centre. The P-R bond is weakened by this
same process, suggesting evidence for the involvement of the σ*-orbitals in the back-bonding
process.

The relationship between σ-basicity and π-acidity can be shown by the following series;

This series was developed by Tolman who prepared several species of the form [Ni(CO)3(PR3)] and
compared the stretching frequencies of the CO using IR spectroscopy, a higher value would suggest a
stronger CO bond, which in turn meant less electron-density on the metal and so less donation from
the phosphine group, indicating a π-accepting R group. The opposite would then be true of weaker
CO stretching frequencies. Similar studies are also performed on rhodium complexes as they are less
toxic than their nickel analogues.

The steric properties of phosphines is of importance to catalysis and using the same nickel
complexes Tolman also defined the steric parameter, the ligand cone angle, in order to qualitatively
assess the bulk of a phosphine. The Tolman cone angle is defined as the angle of the cone of which
the point is located 228pm from the phosphorus atom and the edges run tangential to the van der
Waals radii of the R groups.
e.g. for R=CH3

or angles less than 140-160o steric effects are minimal. Bulky phosphines are able to stabilize low-
coordinate species and promote catalytically important metal-based reactions such as reductive
elimination, a bulky phosphine group will increase the rate of dissociative processes while
decreasing the rate of associative processes.

Mechanisms of Substitution in Transition Metal Complexes


Considering only octahedral, first-row transition metal complexes and the following reaction;

MLn+L’ ⇌ MLn-1L’+L

There are two potential extreme mechanisms, dissociative or associative.

The dissociative mechanism is akin to SN1 reactions in which the M-L bond first breaks before the M-
L’ bond forms, this reaction will follow first-order kinetics.

The associative mechanism is akin so SN2 in which an L’-M-L bond forms before the M-L bond breaks
to give the product, this reaction will follow second-order kinetics.

The third possible mechanism, the interchange reaction, is one in which both of the above processes
occur to an extent and the kinetics are dictated by both bond-breaking and bond-forming. If the
bond-breaking has a greater influence on the reaction rate then it can be considered dissociative
interchange (Id) and if the bond-forming has a greater influence on the reaction rate it can be
considered an associative interchange (Ia) reaction.

To assess the kinetics of a dissociative reaction consider;

Here k1<<k-1<<k2 under normal circumstances and so, since k-1[X]<<k2[Y];

[ ]
[ ]

The reaction is first order with respect to the substrate and so the rate of substitution ought to be
independent of [Y].
To assess the kinetics of an associative reaction consider;

By applying the steady state approximation to ML5XY the following rate law is obtained;

[ ] [ ][ ]
[ ][ ]

If the rate of a ligand substitution is shown to have no dependence on the concentration of the input
ligand, and only on the nature of that ligand, it suggests a dissociative mechanism, if the substitution
shows a dependence on the concentration of the input ligand then it suggests an associative
mechanism.

The nature of the metal is also important in determining the rate of ligand substitution. The larger
ratio of Z2/r the greater the electrostatic attraction between the metal and the ligand and so slower
dissociation processes occur.

There is also ligand field activation energy (LFAE), the difference in the ligand field stabilization
energies between the ground state (octahedral) and the transition state (often square pyramidal),
complexes with the highest CFSE (d3 and d6 low-spin in particular) in an octahedral ground rate tend
to have the slowest substitution rates (or the largest ΔG of transition, ΔGⱡ) whereas some are
atypically fast due to Jahn-Teller distortions, for example Cu(II) (d9).

You might also like