Inorganic Chemistry

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 9

Different Types of Ligands

A ligand is an ion or molecule which donates a pair of electrons to the central metal atom or
ion to form a coordination complex. The word ligand is from Latin, which means “tie or
bind”. Ligands can be anions, cations, and neutral molecules. Ligands act as Lewis bases
(donate electron pairs), and central metal atoms are viewed as Lewis acids (electron pair
acceptors). The nature of bonding from metal to ligand varies from covalent bond to ionic
bond.

Occasionally ligands can be cations (NO+, N2H5+) and electron-pair acceptors. Examples for
anionic ligands are F–, Cl–, Br–, I–, S2–, CN–, NCS–, OH–, NH2– and neutral ligands are NH3,
H2O, NO, CO.

A ligand is an ion or molecule which binds to the central metal atom to form a coordination
entity or complex compounds. Classification of ligands is on the basis of the number of
binding sites with the central metal atom, charge and size.

Monodentate Ligands
Monodentate ligands are also called “one-toothed“ because they bite the metal atom only in
one place.

Br– , F– , H2O , CN–,

Bidentate Ligands
Lewis bases, which donate two lone pairs of electrons to the central metal atom, are known as
bidentate ligands. They are often referred to as “chelating ligands”. The complex which
contains chelating ligands is called “Chelates”.

Ethylenediamine (en)

Acetylacetonate ion (acac)

Oxalate ion (ox)


Tridentate Ligands and Polydentate Ligands
Tridentate ligands have three lone pairs of electrons to the central metal atom or ion.
Molecules with four donor atoms are called tetradentate, five donor atoms are called
pentadentate, and six donor atoms are called hexadentate. They are generally mentioned as
polydentate ligands

Chelate Effect
The chelate effect explains the enhanced affinity of chelating ligands for central metal ions or
atoms compared to the affinity of nonchelating monodentate ligands for the same metal.

Ambidentate Ligands
Ligands with more than one potential donor atom are known as ambidentate ligands. For
example, thiocyanate ion(NCS–) which can bind to the central metal atom or ion with either
nitrogen or sulfur atoms.

Bridging Ligands
A bridging ligand is one which is bound to more than one metal atom.

The bridge ligand is preceded by “mu,\eta”, which relates to hapticity. The list of bridging
ligands is as follows:

N3- : [Ir3N(SO4)6(H2O)3]4- , NH2- : HgNH2Cl

Isomerism in Coordination Complexes


The existence of coordination compounds with the same formula but different arrangements
of the ligands was crucial in the development of coordination chemistry. Two or more
compounds with the same formula but different arrangements of the atoms are
called isomers. Because isomers usually have different physical and chemical properties, it is
important to know which isomer we are dealing with if more than one isomer is possible. As
we will see, coordination compounds exhibit the same types of isomers as organic
compounds, as well as several kinds of isomers that are unique. Isomers are compounds with
the same molecular formula but different structural formulas and do not necessarily share
similar properties. There are many different classes of isomers, like stereoisomers,
enantiomers, and geometrical isomers. There are two main forms of isomerism: structural
isomerism and stereoisomerism (spatial isomerism).

Class I: Structural Isomers

Isomers that contain the same number of atoms of each kind but differ in which atoms are
bonded to one another are called structural isomers, which differ in structure or bond type.
For inorganic complexes, there are three types of structural
isomers: ionization, coordination and linkage. Structural isomers, as their name implies,
differ in their structure or bonding, which are separate from stereoisomers that differ in the
spatial arrangement of the ligands are attached, but still have the bonding properties. The
different chemical formulas in structural isomers are caused either by a difference in what
ligands are bonded to the central atoms or how the individual ligands are bonded to the
central atoms. When determining a structural isomer, you look at (1) the ligands that are
bonded to the central metal and (2) which atom of the ligands attach to the central metal.

Ionization Isomerism

Ionization isomers occur when a ligand that is bound to the metal center exchanges places
with an anion or neutral molecule that was originally outside the coordination complex. The
geometry of the central metal ion and the identity of other ligands are identical. For example,
an octahedral isomer will have five ligands that are identical, but the sixth will differ. The
non-matching ligand in one compound will be outside of the coordination sphere of the other
compound. Because the anion or molecule outside the coordination sphere is different, the
chemical properties of these isomers is different. A hydrate isomer is a specific kind of
ionization isomer where a water molecule is one of the molecules that exchanges places.

The difference between the ionization isomers can be view within the context of the ions
generated when each are dissolved in solution. For example, when
pentaamminebromocobalt(II) chloride is dissolved in water, Cl− ions are generated:
CoBr(NH3)5Cl(s)→CoBr(NH3)+5Cl−(aq)

whereas when pentaamminechlorocobalt(II) bromide is dissolved, Br− ions are generated:

CoCl(NH3)5Br(s)→CoCl(NH3)+5Br−(aq)

Coordination Isomerism
Coordination isomerism occurs in compounds containing complex anionic and complex cationic
parts and can be viewed as an interchange of some ligands from the cation to the anion. Hence,
there are two complex compounds bound together, one with a negative charge and the other with a
positive charge. In coordination isomers, the anion and cation complexes of a coordination
compound exchange one or more ligands. For example, the [Zn(NH3)4][Cu(Cl4)][Zn(NH3)
4][Cu(Cl4)] and [Cu(NH3)4][Zn(Cl4)][Cu(NH3)4][Zn(Cl4)] compounds are coordination isomers

Linkage Isomerism

Linkage isomerism occurs with ambidentate ligands that are capable of coordinating in more
than one way. The best known cases involve the
monodentateligands: SCN−/NCS−SCN−/NCS− and NO−2/ONO−NO2−/ONO−. The only

diffrence is what atoms the molecular ligands bind to the central ion. The ligand(s) must
have more than one donor atom, but bind to ion in only one place. For example, the
(NO−2NO2−) ion is a ligand that can bind to the central atom through the nitrogen or the
oxygen atom, but cannot bind to the central atom with both oxygen and nitrogen at once, in
which case it would be called a polydentate ligandrather than an ambidentate ligand.

As with all structural isomers, the formula of the complex is unchanged for each isomer, but
the properties may differ. The names used to specify the changed ligands are changed as well.
For example, the NO−2 ion is called nitro when it binds with the N atom and is
called nitrito when it binds with the O atom.

Class 2: Geometric Isomers

The existence of coordination compounds with the same formula but different arrangements
of the ligands was crucial in the development of coordination chemistry. Two or more
compounds with the same formula but different arrangements of the atoms are
called isomers. Because isomers usually have different physical and chemical properties, it is
important to know which isomer we are dealing with if more than one isomer is possible.
Recall that in many cases more than one structure is possible for organic compounds with the
same molecular formula; examples discussed previously include n-butane versus isobutane
and cis-2-butene versus trans-2-butene. As we will see, coordination compounds exhibit the
same types of isomers as organic compounds, as well as several kinds of isomers that are
unique.

Planar Isomers

Metal complexes that differ only in which ligands are adjacent to one another (cis) or directly
across from one another (trans) in the coordination sphere of the metal are called geometrical
isomers. They are most important for square planar and octahedral complexes.

Because all vertices of a square are equivalent, it does not matter which vertex is occupied by
the ligand B in a square planar MA3B complex; hence only a single geometrical isomer is
possible in this case (and in the analogous MAB3 case). All four structures shown here are
chemically identical because they can be superimposed simply by rotating the complex in
space:
or an MA2B2 complex, there are two possible isomers: either the A ligands can be adjacent to
one another (cis), in which case the B ligands must also be cis, or the A ligands can be across
from one another (trans), in which case the B ligands must also be trans. Even though it is
possible to draw the cis isomer in four different ways and the trans isomer in two different
ways, all members of each set are chemically equivalent:

Because there is no way to convert the cis structure to the trans by rotating or flipping the
molecule in space, they are fundamentally different arrangements of atoms in space. Probably
the best-known examples of cis and trans isomers of an MA2B2 square planar complex are
cis-Pt(NH3)2Cl2, also known as cisplatin, and trans-Pt(NH3)2Cl2, which is actually toxic rather
than therapeutic.

Zeise's salt
Zeise's salt is the chemical compound with the formula K[PtCl3(C2H4)].H2O. The anion of this
air-stable, yellow, coordination complex contains an η2-ethylene ligand. The complex is
commonly prepared from K2[PtCl4] and ethylene in the presence of a catalytic amount
of SnCl2. The anion features a platinum atom with a square planar geometry.

Zeise's salt was one of the first organometallic compounds to be reported.[1] Its inventor W.
C. Zeise, a professor at the University of Copenhagen, prepared this compound in 1820s
while investiging the reaction of PtCl4 with boiling ethanol. He proposed that the resulting
compound contained ethylene. Justus von Liebig, an influential chemist of that era, often
criticised Zeise's proposal, but Zeise's theories were decisively supported in 1868 when
Birnbaum prepared the complex using ethylene.[2][3]
Zeise's salt received a great deal of attention during the second half of the 19th century
because chemists could not properly explain the molecular structure of the salt. This
question remained unanswered until the advent of x-ray diffraction in the 20th century.

Zeise's salt stimulated much scientific research in the field of organometallic chemistry, and
would be key in defining new concepts in chemistry such as "Hapticity".[1] The Dewar-Chatt-
Duncanson model explains how the metal is coordinated to the double bond

Related compounds

Zeise's dimer [{(η2-C2H4)PtCl2}2], derived from Zeise's salt by elimination of KCl followed
by dimerisation.

"COD-platinum dichloride," (cyclooctadiene)PtCl2, derived from platinum(II)


chloride and 1,5-cyclooctadiene, is a common platinum(II) alkene complex.

Many other ethylene complexes have been prepared. For example,


ethylenebis(triphenylphosphine)platinum(0), [(C6H5)3P]2Pt(H2C=CH2), wherein the platinum
is three-coordinate and zero-valent (Zeise's salt is a derivative of platinum(II)).

Structure

In Zeise's salt and related compounds, the alkene rotates about the metal-alkene bond with
a modest activation energy. Analysis of the barrier heights indicates that the π-bonding
between most metals and the alkene is weaker than the σ-bonding. In Zeise's anion, this
rotational barrier cannot be assessed by NMR spectroscopy because all four protons are
equivalent. Lower symmetry complexes of ethylene, e.g. CpRh(C2H4)2, are, however, suitable
for analysis of the rotational barriers associated with the metal-ethylene bond.[5]

Vaska;'s complex
Vaska's complex is the trivial name for trans-
chlorocarbonylbis(triphenylphosphine)iridium(I) with the formula IrCl(CO)[P(C 6H5)3]2.
This square planar diamagnetic organometallic complex consists of a central iridium atom
bound to two mutually trans triphenylphosphine ligands as well as carbon monoxide, and
chloride. The complex was reported by Di Luzio and Vaska in 1961. Vaska's complex can
undergo oxidative additions and is notable for its ability to bind to O2 reversibly. It is a bright
yellow crystalline solid. Vaska's complex is the trivial name for trans-
chlorocarbonylbis(triphenylphosphine)iridium(I) with the formula IrCl(CO)[P(C 6H5)3]2.
This square planar diamagnetic organometallic complex consists of a central iridium atom
bound to two mutually trans triphenylphosphine ligands as well as carbon monoxide, and
chloride. The complex was reported by Di Luzio and Vaska in 1961. [1] Vaska's complex can
undergo oxidative additions and is notable for its ability to bind to O2 reversibly. It is a bright
yellow crystalline solid.
Preparation

The synthesis involves heating virtually any iridium chloride salt with P(C6H5)3 with a CO
source. The most popular method uses dimethylformamide (DMF) as a solvent.
Sometimes aniline is added to accelerate the reaction. Another popular solvent is 2-
methoxyethanol. The reaction typically is conducted under nitrogen. In the synthesis,
triphenylphosphine serves as both a ligand and a reductant, and the carbonyl ligand is
derived by decomposition of dimethylformamide probably via a deinsertion of an
intermediate Ir-C(O)H species. The following is a possible balanced equation for this
complicated reaction.[2]

H2IrCl6 + 3.5P(C6H5)3 + HCON(CH3)2 + 4C6H5NH2 + 1.5H2O → IrCl(CO)*P(C6H5)3]2 +


(CH3)2NH2+Cl- + 1.5 OP(C6H5)3

Typical sources of iridium used in this preparation are IrCl3.xH2O and H2IrCl6.

Reactions

Studies on Vaska's complex provided a conceptual framework for homogeneous catalysis.


Vaska's complex, with 16 valence electrons, is considered "unsaturated" and can thus bind
to one two-electron or two one-electron ligands to become electronically saturated with 18
valence electrons. The addition of two one-electron ligands is called oxidative addition.
Upon oxidative addition, the oxidation state of the iridium increases from Ir(I) to Ir(III). The
four-coordinated square planar arrangement in the starting complex converts to
an octahedral, six-coordinate product. Vaska's complex undergoes oxidative addition with
conventional oxidants such as halogens, strong acids such as HCl, and other molecules
known to react as electrophiles, such as iodomethane (CH3I).

An interesting characteristic of Vaska's complex is that it binds O2 reversibly.

IrCl(CO)[P(C6H5)3]2 + O2 ↔ IrCl(CO)*P(C6H5)3]2O2

The dioxygen ligand is bonded to Ir(I) via both oxygen atoms, so-called side-on bonding. In
myoglobin and hemoglobin, O2 binds "end-on," attaching to the metal via only one of the
two oxygen atoms. The oxygenation reaction is carried out simply by bubbling O2 through a
solution of Vaska's complex in toluene, which results in a colour change from yellow to
orange. The resulting dioxgen adduct reverts to the parent complex upon heating in boiling
benzene solution.

Spectroscopy

Infrared spectroscopy can be used to analyse the products of oxidative addition to Vaska's
complex because the reactions induce characteristic shifts of the stretching frequency of the
coordinated carbon monoxide.[3] These shifts are dependent on the amount of π-back
bonding allowed from the newly associated ligands. The CO stretching frequencies for
Vaska's complex and oxidatively added ligands have been documented in literature.

You might also like