0% found this document useful (0 votes)
6 views

Inorganic Study Guide

The document is a study guide for the Inorganic Chemistry courses CHE311-W and CHE441-A at the University of South Africa, authored by Prof. Fikru Tafesse. It outlines the course objectives, key chapters from the prescribed textbook, and essential topics such as spectroscopy, coordination chemistry, and organometallic chemistry. The guide emphasizes the importance of understanding various instrumental techniques and provides a detailed overview of vibrational and electronic spectroscopy.

Uploaded by

shillah mogane
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
6 views

Inorganic Study Guide

The document is a study guide for the Inorganic Chemistry courses CHE311-W and CHE441-A at the University of South Africa, authored by Prof. Fikru Tafesse. It outlines the course objectives, key chapters from the prescribed textbook, and essential topics such as spectroscopy, coordination chemistry, and organometallic chemistry. The guide emphasizes the importance of understanding various instrumental techniques and provides a detailed overview of vibrational and electronic spectroscopy.

Uploaded by

shillah mogane
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 67

University of South Africa (UNISA)

DEPARTMENT OF CHEMISTRY

Inorganic chemistry study guide for CHE311-W and 441-A

Prof. Fikru Tafesse


Pretoria, 2006.

A study guide to accompany the prescribed text, Inorganic chemistry 4th edition, Atkins, Overton, (ISBN
0199264635), 2006, OUP
2
Preface

In writing this study guide, my aim is to provide additional information and resource on the
chapters covered in the prescribed text book for students that pursue inorganic chemistry
courses at third and honours levels. The study guide is anticipated to serve simultaneously both
CHE311-W and CHE441-A. The prescribed textbook for both courses is Atkins, Overton
Inorganic chemistry, 4th edition, 2006 (ISBN 0199264635), Oxford university press. Students
must acquire a copy of the prescribed textbook. The page numbers and chapters referred to in
this study guide correlate with the 4th edition of the prescribed textbook. Details of the course
contents for the two courses are given below:

CHE311-W (third level inorganic chemistry)

The following chapters from the prescribed textbook will be covered.

Chapter 6 (Physical techniques in inorganic chemistry, pages 169 – 194) and chapter 8 (an
introduction to coordination compounds are prerequisites. The concepts covered in those
chapters are revisions from previous years. The following three chapters form the core of the
course:

Chapter 19. d-metal complexes: electronic structure and spectra (pages 459-490

Chapter 20. Coordination chemistry: reaction of complexes (pages 491-524)

Chapter 21. d-metal organometallic chemistry (pages 525-571)

CHE441-A (honours level inorganic chemistry)

Chapter 7. Molecular symmetry is recommended as essential.

A review of chapters 19, 20 and 21 ( pages 459-571) is an absolute requirement. Honours


students are expected to have a clear understanding of the concepts outlined in those chapters.
The following additional three chapters comprise the syllabus for the course.

Chapter 22. the f-block metals (pages 572-585)

Chapter 25. Catalysis (pages 680-710)

Chapter 26. Biological inorganic chemistry (pages 711-768)


3

Honours level students are expected to solve all of the exercises and problems provided in the
chapters. Additional reading from the cited references is recommended for deeper insight into
the concepts outlined.

It is very difficult for me to cite all the references and resources I have utilised in compiling this
study guide. I have extensively used web based educational sites from the internet. However, I
would like to acknowledge the following three web sites among others.
http://wwwchem.uwimona.edu.jm/ courses; http://www.merlot.org; http://www.chem.ox.ac.uk .
I have also included my own lecture notes I utilised for teaching inorganic chemistry for a
number of years in different institutions. It is my sincere wish that students will find help in the
pages that follow.

Prof. Fikru Tafesse, chemistry department UNISA, 2006.


4
Aims and objectives of CHE311-W and 441-A

These courses begin with a review of vibrational and electronic spectroscopy. They then
canvass simple crystal field theory, highlighting both strengths and weaknesses of the model.
The model is then extended to the more general ligand field theory and the concept of
interelectronic repulsion is introduced followed by a description of Russell-Saunders terms.

Electronic spectra of high-spin octahedral complexes are interpreted using Orgel diagrams and
tetrahedral complexes are treated in this way as well. The treatment is then extended to the use
of Tanabe-Sugano diagrams, especially to cover the interpretation of the more difficult d2 - d7
octahedral cases as well as low-spin complexes.

It is expected that students should be able to interpret ALL simple electronic spectra and predict
both position and intensity based on Orgel/Tanabe-Sugano diagrams and the appropriate
selection rules and appreciate the relationships between the electron configuration, co-
ordination geometry, and spectroscopic properties of transition metal complexes.

Magnetic properties of high-spin octahedral and tetrahedral complexes are simplified in terms of
systems having A and E ground terms or T ground terms. It is expected that students should be
able to calculate the spin-only magnetic moment for any high-spin complex and predict the
nature of any variations found experimentally.

An introduction to inorganic reaction mechanisms, a subject that has rapidly grown in


importance during the last thirty years, will be presented to you. This introduction is necessarily
short and selective, since the course is primarily concerned with coordination compounds of first
row transition metal ions. The primary objective of this part of the course is to make you familiar
with the following topics:

• The current classification of inorganic reaction mechanisms;


• The mechanism of substitution at square-planar complexes;
• The mechanism of substitution at octahedral complexes;
• The mechanism of electron transfer reactions.

A short course on organometallic chemistry involving only transition metals is also presented.
The material covered in your prescribed textbook on the subject needs no additional guide.
Organometallic chemistry has developed enormously during the last 50 years. Currently
approximately half of the world's research publications in chemistry relate to organometallic
5

compounds. Over and above, organometallics play a vital role in the economy of the developed
nations. The aim of the course is to introduce to you:

• the IUPAC nomenclature of organometallic compounds;


• classification based on the 18-electron rule;
• preparations and structures of some metal carbonyls and metallocenes;
• vibrational spectra of metal carbonyls;
• Catalysis - hydrogenation and carbonylation.

The course content of CHE441-A is expanded to include the f-block metals, Catalysis and
bioinorganic chemistry. The treatment of the topics in CHE441-A is much wider and deeper
than those of CHE311-W.

Instrumental techniques used in inorganic chemistry

Technique Nature of the effect Applied to Information obtained

X-ray diffraction Scattering, mainly by Crystalline Complete structure,


electrons, followed by substances including accurate
interference (λ = 0.01- internuclear distances
1nm). and bond angles. The
location of light atoms
e.g. H and the
distinction of atoms with
similar scattering factors,
in the presence of heavy
atoms, is difficult.

Neutron Scattering, mainly by Crystalline As for X-ray diffraction,


diffraction nuclei, followed by substances but is especially useful in
interference (λ = 0.1 nm) locating H atoms.

Electron Diffraction (atom or Usually gases, As for X-ray diffraction,


diffraction molecule), mainly by sometimes but limited to species
nuclei but also by liquids or with small numbers of
electrons (λ = 0.01-0.1 solids structural parameters.
nm)

Microwave Absorption of radiation Gases having Interatomic distances


spectroscopy due to a change in dipole and dipole moments –
dipole moment during a moments limited to small
rotation (λ = 0.1-3.0 cm). molecules.
6

Infrared (IR) Absorption of radiation, Gases, liquids For small molecules –


spectroscopy due to a change in or solids interatomic distances,
dipole moment during a bond force constants
vibration (λ = 10-1 -10-4 and molecular
cm). geometries. For large
molecules –
identification and
characterization.

Raman Scattering of radiation Gases, liquids As for IR but also


spectroscopy with changed frequency or solids applicable to vibrational
due to a change in transitions not observed
polarizability during a using IR.
vibration.

Absorption of radiation, Gases, liquids, Assignment of electronic


Electronic due to a change in solids or transitions. Identification
spectroscopy dipole moment during an solutions and characterization.
electronic transition (λ = Particularly useful for
102 – 103 nm). transition metal
complexes.

Photoelectron Ejection of electrons Gases (UV) or Binding energies of


spectroscopy from molecular orbitals solids (X-rays) valence electrons (UV)
using UV radiation or X- or core electrons (X-
rays. rays).

Extended X-ray Back-scattering, off Especially Oxidation state and


absorption fine ligands, of useful for chemical environments
structure photoelectrons. metallo- of atoms, bond lengths.
(EXAFS) proteins and
heterogeneous
catalysts

Nuclear Interaction of radiation Usually Chemical environments


magnetic with a nuclear transition solutions of of atoms, structures of
resonance in a magnetic field (λ = diamagnetic groups of atoms
spectroscopy 102-107 cm; 3 KHz-300 compounds
MHz). Nuclie having
magnetic moments (e.g.
1
H, 13C, 14N, 19F, 31P) are
required.

Nuclear Interaction of radiation in Solids Degree of ionic


quadrupole the radio-frequency character in a bond.
resonance region with nuclei having
spectroscopy quadrupole moment.

Mossbauer Recoil-free emission and Solids Spin and ionization state


spectroscopy resonance absorption of of a nucleus.
low energy λ-radiation.
7
Electron spin Interaction of microwave Gases, liquids Oxidation and spin
resonance radiation with unpaired or solids states.
spectroscopy electrons in
paramagnetic species.

Mass Detection of fragments Gases or Fragmentation patterns


spectrometry by mass/charge ratio. volatile for molecules, masses of
species molecules and
fragments.

Unit 1

Raman and Infrared spectroscopy

Theoretical consideration

Introduction

Matter is made from charged building blocks (electrons, nuclei, positive and negative ions). A
specific kind of matter, such as a specific molecule, a solid, or an atom has a specific charge
distribution. These charges are not only isolated charges (monopoles), there are also positive
and negative charges forming electric dipoles (and higher multipoles). There is also a way that
the electric field of the light wave itself can shift charges and in this way, to induce a dipole
moment. The electric (and, in much extend also the magnetic)field of a light wave can interact
with the permanent existing and induced dipole moments. In a light wave the fields change
about 1015 times per second, that means the charged piece of matter move (oscillate) very fast.

On the other hand changes (oscillations) of charge distribution, such as an oscillating dipole can
result in the generation of electromagnetic fields. These processes between light and matter are
the basic processes of each spectroscopy. Since the charge distributions are very specific for a
specific kind of matter, such as for a specific molecule, the exchange of energy between the
electromagnetic wave and the matter is also very specific and provides unique information on
the matter. In the quantum picture, theses processes are described by the annihilation of a
photon, which is accompanied by the transition of the system in a higher energy state or vice
8

versa, by generating a photon when the system goes from a higher energy state to a lower one.
There is a specific probability that photons (or light waves) interact with matter. This can be
described by the so called “cross section” of a spectroscopic effect.

Raman and infrared (IR) spectroscopy provide information about the vibrational and vibrational-
rotational modes of molecules. Vibrational rotational bands are generally observed when the
samples are in gaseous state, where the molecules are able to rotate freely. In condensed
(liquid or solid) phases, only vibrational frequencies of the sample can be observed. IR and
Raman spectroscopies are complementary techniques, since transitions allowed in Raman may
be forbidden in IR or vice-versa. This depends on symmetry considerations. An IR active mode
is one in which a particular vibration causes a change in the dipole moment of the molecule,
while only those vibrations which change the molecular polarizability lead to Raman scattering.
Therefore the activity of a certain vibrational mode depends highly on its symmetry and the
symmetry of the molecule. One simple symmetry rule is the so called mutual exclusion rule for
molecules with a center of symmetry. In such molecules no normal mode may be active in both
the infrared and the Raman spectrum. In summary, the number and kind (Raman or IR) of
active vibrations gives information about the molecular symmetry providing insights into the
structure of the molecule.

The simplest model of a vibrating diatomic molecule is a harmonic oscillator, for which the
potential energy varies with the internuclear distance and the quantized vibrational energies are:

E v = hv (v + 1
2 ) v = 0, 1, 2, L

Where v is the vibrational quantum number, ν is the vibrational frequency, and h is Planck’s
constant. For a harmonic oscillator the vibration frequency is a function of both, the strength of
the bond (i.e. the bond force constant k) and the reduced mass μ.

1 k
v=
2 µ
The vibrational energy levels are equally spaced in the harmonic oscillator model and the
separation between adjacent energy levels E is:
k
∆E = hv = h
µ
9

The number of possible vibrational modes that a molecule has can be determined from the
number of atoms N in the molecule and the molecular shape. The Cartesian coordinates of
each atom specify the molecular geometry. The molecule, therefore, has 3N degrees of
freedom. Three of these degrees are used in specifying the transitional motion of the molecule
in space, and either two (linear molecules) or three (non linear molecules) degrees are needed
to specify its rotational motion. The remaining degrees of freedom correspond to the number of
vibrational modes: 3N-5 for linear molecules and 3N-6 for non-linear molecules

The term "infra red" covers the range of the electromagnetic spectrum between 0.78 and 1000
μm. In the context of infrared spectroscopy, wavelength is measured in "wave numbers", which
have the units cm-1.

Wave number = 1 / wavelength in centimeters

It is useful to divide the infrared region into three sections; near, mid and far infrared;

Region Wavelength range (µm) Wave number range (cm-1)


Near 0.78 - 2.5 12800 - 4000
Middle 2.5 - 50 4000 - 200
Far 50 -1000 200 - 10
The most useful I.R. region lies between 4000 - 670cm-1.

Theory of infra red absorption

IR radiation does not have enough energy to induce electronic transitions as seen with UV.
Absorption of IR is restricted to compounds with small energy differences in the possible
vibrational and rotational states.
For a molecule to absorb IR, the vibrations or rotations within a molecule must cause a net
change in the dipole moment of the molecule. The alternating electrical field of the radiation
(remember that electromagnetic radiation consists of an oscillating electrical field and an
oscillating magnetic field, perpendicular to each other) interacts with fluctuations in the dipole
moment of the molecule. If the frequency of the radiation matches the vibrational frequency of
the molecule then radiation will be absorbed, causing a change in the amplitude of molecular
vibration.
10

Molecular rotations
Rotational transitions are of little use to the spectroscopist. Rotational levels are quantized, and
absorption of IR by gases yields line spectra. However, in liquids or solids, these lines broaden
into a continuum due to molecular collisions and other interactions.

Molecular vibrations
The positions of atoms in molecules are not fixed; they are subject to a number of different
vibrations. Vibrations fall into the two main categories of stretching and bending.
Stretching: Change in inter-atomic distance along bond axis

Bending: Change in angle between two bonds. There are four types of bend:

• Rocking
• Scissoring
• Wagging
• Twisting

Vibrational coupling
In addition to the vibrations mentioned above, interaction between vibrations can occur
(coupling) if the vibrating bonds are joined to a single, central atom. Vibrational coupling is
influenced by a number of factors;

• Strong coupling of stretching vibrations occurs when there is a common atom between the two
vibrating bonds
• Coupling of bending vibrations occurs when there is a common bond between vibrating groups
• Coupling between a stretching vibration and a bending vibration occurs if the stretching bond is
one side of an angle varied by bending vibration
• Coupling is greatest when the coupled groups have approximately equal energies
• No coupling is seen between groups separated by two or more bonds
11

Diatomic molecules

Harmonic vibrations

IR spectroscopy involves transitions between vibrational states. In order to describe the


transitions between vibrational states the vibration of a harmonic oscillator can be used as a
model. Consider a diatomic molecule, for example carbon monoxide (CO), as two balls
connected by a spring and lying on a frictionless table (see Figure 4). If the spring is stretched
and the balls then released the system will vibrate with a constant frequency.

The CO molecule as a harmonic oscillator

The frequency of vibration depends on the strength of the spring and the masses of the balls
according to the equation

1 k
V =  
2π µ

Where ν is the frequency of vibration


k is the force constant
and µ is the reduced mass of the system (excluding the spring).

The reduced mass is given by the equation:

mC mO
µ=
mC + mO

Where mC and mO are the masses of carbon and oxygen, respectively. However for a diatomic
molecule such as CO the vibrational energy is quantized. The equation for the allowed
vibrational energies is:

Evib = hν (v + ½) v = 0,1,2,……. Where Evib is the vibrational energy, ν is the frequency


of the vibration and v is the vibrational quantum number.
Combining the equations above, the following is attained:

k
E vib = h

  (v + 1
) v = 0, 1, 2, K
µ
2

For transitions between vibrational energy levels the vibrational quantum number, v, can only
change by 1, i.e. ∆v = ±1. At room temperature most molecules are in the vibrational ground
state (v = 0). So the only important transition is between the ground state and the first excited
state state (v = 1). Thus ∆E = E1 – E0 = hν [ (1 + ½) - ( 0 + ½)] = hν.
12

k
Hence ∆E = h

 
µ

The transition v = 0 to v = 1 is known as the fundamental vibrational absorption.

Exercise1: Calculate the frequency (in S-1) and the quantum energy (in J) corresponding to the
wave number of 2143 cm –1 of the radiation absorbed in the fundamental vibrational transition of
12 16
C O. Calculate the reduced mass of 12C16O and the force constant for the C-O bond in
12 16
C O.

( answer: v = 6.425 X 10 13 S-1 , E = 1.856 X 10 3 kg S –2 (i.e. 1.856 X 10 3 Nm –1, µ=


1.139 X 10 –26 kg)

Exercise 2: Calculate the force constant for the H-Cl bond in the H35Cl molecule. The wave
number of the radiation absorbed in the fundamental vibrational transition is 2886 cm –1.

( Answer : 4.770 X 10 2 Nm –1 )

Mechanism of absorption of IR radiation

The conditions required for molecules to absorb IR radiation cam be summarized by two
selection rules:

First selection rule: In order for molecules to absorb, there must be a change in dipole
moment during the vibration.

Second selection rule: Only transitions for which Δv = + 1 can occur. Most transitions will
occur from the ground state (v=0) to the first excited state (v=1). This is the fundamental
vibrational absorption. Since molecules are not perfect harmonic oscillators (some
anharmonicity is present), the second selection rule breaks down and transitions from the
ground state to the second and third excited states ( v=2 and v=3) do occur. The transition
from c=0 to v=2, which is called the first overtone, occurs at a frequency about twice that of
the fundamental and the transition from v=0 to v=3, the second overtone occurs at a
frequency about three times that of the fundamental. The intensity of the first overtone is
about ten times lower than the intensity of the fundamental and the intensity of the second
overtone about 100 times less than that of the fundamental.
13

A linear molecule has 3n-5 fundamental modes of vibration, where n is the number of atoms
in the molecule. These are three directions of motion for each atom, minus three translational
and two rotational motions for the molecule as a whole. A non-linear molecule has 3n-6
fundamental modes of vibration. Theses are three directions of motion for each atom, minus
three translational and three rotational motions for the molecule as a whole.

Examples of infrared active and inactive absorption bands in CO2

There is no change in dipole moment during the symmetric stretch vibration and the 1340
cm-1 band is not observed in the infrared absorption spectrum (the symmetric stretch is called
infrared inactive). There is a change in dipole moment during the asymmetric stretch and the
2350 cm-1 band does absorb infrared radiation (the asymmetric stretch in infrared active). A
related vibrational spectroscopic method is Raman spectroscopy, which has a different
mechanism and therefore provides complementary information to infrared absorption. Raman
spectroscopy is the measurement of the wavelength and intensity of inelastically scattered
light from molecules. The Raman scattered light occurs at wavelengths that are shifted from
the incident light by the energies of molecular vibrations. The mechanism of Raman
scattering is different from that of infrared absorption, and Raman and IR spectra provide
complementary information. Typical applications are in structure determination,
multicomponent qualitative analysis, and quantitative analysis.

Referring back to the absorption bands of carbon dioxide, polarizability depends on how
tightly the electrons are bound to the nuclei. In the symmetric stretch the strength of electron
binding is different between the minimum and maximum internuclear distances. Therefore the
14

polarizability changes during the vibration and this vibrational mode scatters Raman light (the
vibration is Raman active). In the asymmetric stretch the electrons are more easily polarized
in the bond that expands but are less easily polarized in the bond that compresses. There is
no overall change in polarizability and the asymmetric stretch is Raman inactive.

Infrared Absorption spectrometers .


Dispersive and Fourier-transform spectrometers are used in infrared absorption
spectroscopy. Common light sources are tungsten lamps, Nernst glowers, or glowbars. IR
detectors consist of a semiconductor such as PbS or liquid-nitrogen-cooled HgCdTe (also
called MCT).

Dispersive infrared spectrometer

Dispersive IR spectrometers use a diffraction grating in a monochromator to disperse the


different wavelengths of light. Dispersive IR spectrometers have largely been replaced with
FTIR instruments. They find some use in specific applications, such as monitoring a single IR
wavelength to measure the kinetics of a fast reaction.

Schematic diagram of a dispersive IR absorption spectrometer

Fourier-transform infrared (FTIR) spectrometer

Most modern IR absorption instruments use Fourier-transform techniques with a Michelson


interferometer. To obtain an IR absorption spectrum, one mirror of the interferometer moves to
generate interference in the radiation reaching the detector. Since all wavelengths are passing
through the interferometer, the interferogram is a complex pattern. The absorption spectrum as
a function of wave number (cm-1) is obtained from the Fourier transform of the interferogram,
which is a function of mirror movement (cm). This design does not have the reference cell of a
dispersive instrument, so a reference spectrum is recorded and stored in memory to subtract
from the sample spectrum.
15

Unit 2 UV-Visible spectroscopy

Introduction

Many compounds absorb ultraviolet (UV) or visible (Vis.) light. The diagram below shows a
beam of monochromatic radiation of radiant power P0, directed at a sample solution. Absorption
takes place and the beam of radiation leaving the sample has radiant power P.

The amount of radiation absorbed may be


measured in a number of ways:

Transmittance, T = P / P0
% Transmittance, %T = 100 T

Absorbance,
A = log10 P0 / P
A = log10 1 / T
A = log10
100 / %T
A = 2 - log10 %T

The last equation, A = 2 - log10 %T, is worth remembering because it allows you to easily
calculate absorbance from percentage transmittance data.The relationship between absorbance
and transmittance is illustrated in the following diagram:
16

So, if all the light passes through a solution without any absorption, then absorbance is zero,
and percent transmittance is 100%. If all the light is absorbed, then percent transmittance is
zero, and absorption is infinite.

The Beer-Lambert Law

Now let us look at the Beer-Lambert law and explore its significance. This is important because
people who use the law often don't understand it - even though the equation representing the
law is so straightforward:
A=εbc
Where A is absorbance (no units, since A = log10 P0 / P)
ε is the molar absorbtivity with units of L mol-1 cm-1
b is the path length of the sample - that is, the path length of the cuvette in which the sample is
contained. We will express this measurement in centimetres.
c is the concentration of the compound in solution, expressed in moles per liter
The reason why we prefer to express the law with this equation is because absorbance is
directly proportional to the other parameters, as long as the law is obeyed. We are not going to
deal with deviations from the law.
Let's have a look at a few questions...
Question : Why do we prefer to express the Beer-Lambert law using absorbance as a measure
of the absorption rather than %T ?
Answer : To begin, let's think about the equations...
A=εbc
%T = 100 P/P0 = e -εbc
Now, suppose we have a solution of copper sulphate (which appears blue because it has an
absorption maximum at 600 nm). We look at the way in which the intensity of the light (radiant
power) changes as it passes through the solution in a 1 cm cuvette. We will look at the
reduction every 0.2 cm as shown in the diagram below. The Law says that the fraction of the
light absorbed by each layer of solution is the same. For our illustration, we will suppose that
this fraction is 0.5 for each 0.2 cm "layer" and calculate the following data:

Path length / cm 0 0.2 0.4 0.6 0.8 1.0


%T 100 50 25 12.5 6.25 3.125
Absorbance 0 0.3 0.6 0.9 1.2 1.5
17

A = εbc tells us that absorbance depends on the total quantity of the absorbing compound in the
light path through the cuvette. If we plot absorbance against concentration, we get a straight line
passing through the origin (0,0).

Note that the Law is not obeyed at


high concentrations. This deviation
from the Law is not dealt with here.

The linear relationship between concentration and absorbance is both simple and
straightforward, which is why we prefer to express the Beer-Lambert law using absorbance as a
measure of the absorption rather than %T.
Question : What is the significance of the molar absorbtivity, ε ?
Answer : To begin we will rearrange the equation A = εbc :
ε = A / bc
In words, this relationship can be stated as "ε is a measure of the amount of light absorbed per
unit concentration".
Molar absorbtivity is a constant for a particular substance, so if the concentration of the solution
is halved so is the absorbance, which is exactly what you would expect.
18

Let us take a compound with a very high value of molar absorbtivity, say 100,000 L mol-1 cm-1,
which is in a solution in a 1 cm pathlength cuvette and gives an absorbance of 1.
ε=1/1×c
Therefore, c = 1 / 100,000 = 1 × 10-5 mol L-1

Now let us take a compound with a very low value of ε, say 20 L mol-1 cm-1 which is in solution in
a 1 cm pathlength cuvette and gives an absorbance of 1.
ε=1/1×c
Therefore, c = 1 / 20 = 0.05 mol L-1
The answer is now obvious - a compound with a high molar absorbtivity is very effective at
absorbing light (of the appropriate wavelength), and hence low concentrations of a compound
with a high molar absorbtivity can be easily detected.

Question : What is the molar absorbtivity of Cu2+ ions in an aqueous solution of CuSO4 ? It is
either 20 or 100,000 L mol-1 cm-1
Answer : I am guessing that you think the higher value is correct, because copper sulphate
solutions you have seen are usually a beautiful bright blue colour. However, the actual molar
absorbtivity value is 20 L mol-1 cm-1 ! The bright blue colour is seen because the concentration of
the solution is very high.
β-carotene is an organic compound found in vegetables and is responsible for the colour of
carrots. It is found at exceedingly low concentrations. You may not be surprised to learn that the
molar absorbtivity of β-carotene is 100,000 L mol-1 cm-1 !

Electronic transitions

The absorption of UV or visible radiation corresponds to the excitation of outer electrons. There are three
types of electronic transition which can be considered;

1. Transitions involving π, σ, and n electrons


2. Transitions involving charge-transfer electrons
3. Transitions involving d and f electrons (not covered in this Unit)

When an atom or molecule absorbs energy, electrons are promoted from their ground state to
an excited state. In a molecule, the atoms can rotate and vibrate with respect to each other.
19

These vibrations and rotations also have discrete energy levels, which can be considered as
being packed on top of each electronic level.

Absorbing species containing π, σ, and n electrons

Absorption of ultraviolet and visible radiation in organic molecules is restricted to certain functional
groups (chromophores) that contain valence electrons of low excitation energy. The spectrum of a
molecule containing these chromophores is complex. This is because the superposition of rotational and
vibrational transitions on the electronic transitions gives a combination of overlapping lines. This appears
as a continuous absorption band.

Possible electronic transitions of π, σ, and n electrons are;

σ → σ* Transitions

An electron in a bonding σ orbital is excited to the corresponding antibonding orbital. The


energy required is large. For example, methane (which has only C-H bonds, and can only
20

undergo σ → σ* transitions) shows an absorbance maximum at 125 nm. Absorption maxima due
to σ → σ* transitions are not seen in typical UV-Vis. spectra (200 - 700 nm)

n → σ* Transitions

Saturated compounds containing atoms with lone pairs (non-bonding electrons) are capable of
n → σ* transitions. These transitions usually need less energy than σ → σ * transitions. They can
be initiated by light whose wavelength is in the range 150 - 250 nm. The number of organic
functional groups with n → σ* peaks in the UV region is small.

n → π* and π → π* Transitions

Most absorption spectroscopy of organic compounds is based on transitions of n or π electrons


to the π* excited state. This is because the absorption peaks for these transitions fall in an
experimentally convenient region of the spectrum (200 - 700 nm). These transitions need an
unsaturated group in the molecule to provide the π electrons.
Molar absorbtivities from n → π* transitions are relatively low, and range from 10 to100 L mol-1
cm-1 . π → π* transitions normally give molar absorbtivities between 1000 and 10,000 L mol-1
cm-1 .
The solvent in which the absorbing species is dissolved also has an effect on the spectrum of
the species. Peaks resulting from n → π* transitions are shifted to shorter wavelengths (blue
shift) with increasing solvent polarity. This arises from increased solvation of the lone pair,
which lowers the energy of the n orbital. Often (but not always), the reverse (i.e. red shift) is
seen for π → π* transitions. This is caused by attractive polarisation forces between the solvent
and the absorber, which lower the energy levels of both the excited and unexcited states. This
effect is greater for the excited state, and so the energy difference between the excited and
unexcited states is slightly reduced - resulting in a small red shift. This effect also influences
n → π* transitions but is overshadowed by the blue shift resulting from solvation of lone pairs.

Charge - Transfer Absorption

Many inorganic species show charge-transfer absorption and are called charge-transfer
complexes. For a complex to demonstrate charge-transfer behaviour, one of its components
must have electron donating properties and another component must be able to accept
21

electrons. Absorption of radiation then involves the transfer of an electron from the donor to an
orbital associated with the acceptor.
Molar absorbtivities from charge-transfer absorption are large (greater that 10,000 L mol-1 cm-1).

Unit 3 The electronic spectra of complexes

The Russell Saunders Coupling Scheme

Quantum Numbers: Electrons in an atom reside in shells characterised by a particular value of n, the
Principal Quantum Number. Within each shell an electron can occupy an orbital which is further
characterised by an Orbital Quantum Number, l, where l can take all values in the range:

l = 0, 1, 2, 3, ... , (n-1),

traditionally termed s, p, d, f, etc. orbitals.

Each orbital has a characteristic shape reflecting the motion of the electron in that particular orbital, this
motion being characterised by an angular momentum that reflects the angular velocity of the electron
moving in its orbital.

A quantum mechanics approach to determining the energy of electrons in an element or ion is based on
the results obtained by solving the Schrödinger Wave Equation for the H-atom. The various solutions for
the different energy states are characterised by the three quantum numbers, n, l and ml.

ml is a subset of l, where the allowable values are: ml = l, l-1, l-2, ..... 1, 0, -1, ....... , -(l-2), -(l-1), -l.

There are thus (2l +1) values of ml for each l value,


i.e. one s orbital (l = 0), three p orbitals (l = 1), five d orbitals (l = 2), etc.

There is a fourth quantum number, ms, that identifies the orientation of the spin of one electron relative to
those of other electrons in the system. A single electron in free space has a fundamental property
associated with it called spin, arising from the spinning of an asymmetrical charge distribution about its
own axis. Like an electron moving in its orbital around a nucleus, the electron spinning about its axis has
associated with its motion a well defined angular momentum. The value of ms is either + ½ or - ½.

In summary then, each electron in an orbital is characterised by four quantum numbers:

Quantum Numbers
Principal Quantum Number - largely governs size of orbital and
n
its energy
Azimuthal/Orbital Quantum Number - largely determines shape
l
of orbital
ml Magnetic Quantum Number
ms Spin Quantum Number - either + ½ or - ½ for single electron
22

Russell Saunders coupling

The ways in which the angular momenta associated with the orbital and spin motions in many-electron-
atoms can be combined together are many and varied. In spite of this seeming complexity, the results are
frequently readily determined for simple atom systems and are used to characterise the electronic states of
atoms.

The interactions that can occur are of three types.

• spin-spin coupling
• orbit-orbit coupling
• spin-orbit coupling

There are two principal coupling schemes used:

• Russell-Saunders (or L - S) coupling


• and j - j coupling.

In the Russell Saunders scheme it is assumed that spin-spin coupling > orbit-orbit coupling > spin-orbit
coupling.

This is found to give a good approximation for first row transition series where J coupling is ignored,
however for elements with atomic number greater than thirty, spin-orbit coupling becomes more
significant and the j-j coupling scheme is used.

Spin-Spin coupling
S - the resultant spin quantum number for a system of electrons. The overall spin S arises from adding the
individual ms together and is as a result of coupling of spin quantum numbers for the separate electrons.

Orbit-Orbit coupling
L - the total orbital angular momentum quantum number defines the energy state for a system of
electrons. These states or term letters are represented as follows:

Total Orbital Momentum


L 0 1 2 3 4 5
S P D F G H

Spin-Orbit coupling

Coupling occurs between the resultant spin and orbital momenta of an electron which gives rise to J the
total angular momentum quantum number. Multiplicity occurs when several levels are close together and
is given by the formula (2S+1).
23

The Russell Saunders term symbol that results from these considerations is given by:
(2S+1)
L

As an example, for a d1 configuration:

S= + ½, hence (2S+1) = 2
L=2
and the Ground Term is written as 2D

The Russell Saunders term symbols for the other free ion configurations are given in the Table below.

Terms for 3dn free ion configurations


# of quantum # of energy Ground
Configuration Excited Terms
states levels Term
d1,d9 10 1 2
D -
2 8 3
d ,d 45 5 3
F P, 1G,1D,1S
d3,d7 120 8 4
F 4
P, 2H, 2G, 2F, 2 x 2D, 2P
3
H, 3G, 2 x 3F, 3D, 2 x 3P,
d4,d6 210 16 5
D 1
I, 2 x 1G, 1F, 2 x 1D, 2 x 1S
4
G, 4F, 4D, 4P, 2I, 2H, 2 x
d5 252 16 6
S 2
G, 2 x 2F, 3 x 2D, 2P, 2S

Note that dn gives the same terms as d10-n

Hund's Rules
The Ground Terms are deduced by using Hund's Rules.
The two rules are:
1) The Ground Term will have the maximum multiplicity
2) If there is more than 1 Term with maximum multipicity, then the Ground Term will have the largest
value of L. A simple graphical method for determining just the ground term alone for the free-ions uses a
"fill in the boxes" arrangement.

Ground
dn 2 1 0 -1 -2 L S
Term
d1 ↑ 2 1/2 2
D
2 3
d ↑ ↑ 3 1 F
3 4
d ↑ ↑ ↑ 3 3/2 F
4 5
d ↑ ↑ ↑ ↑ 2 2 D
5 6
d ↑ ↑ ↑ ↑ ↑ 0 5/2 S
6 5
d ↑↓ ↑ ↑ ↑ ↑ 2 2 D
7 4
d ↑↓ ↑↓ ↑ ↑ ↑ 3 3/2 F
8 3
d ↑↓ ↑↓ ↑↓ ↑ ↑ 3 1 F
9 2
d ↑↓ ↑↓ ↑↓ ↑↓ ↑ 2 1/2 D
24

To calculate S, simply sum the unpaired electrons using a value of ½ for each.
To calculate L, use the labels for each column to determine the value of L for that box, then add all the
individual box values together.

For a d7 configuration, then:


in the +2 box are 2 electrons, so L for that box is 2*2= 4
in the +1 box are 2 electrons, so L for that box is 1*2= 2

in the 0 box is 1 electron, L is 0


in the -1 box is 1 electron, L is -1*1= -1
in the -2 box is 1 electron, L is -2*1= -2

Total value of L is therefore +4 +2 +0 -1 -2 or L=3.

Note that for 5 electrons with 1 electron in each box then the total value of L is 0.
This is why L for a d1 configuration is the same as for a d6.

The other thing to note is the idea of the "hole" approach.

A d1 configuration can be treated as similar to a d9 configuration. In the first case there is 1 electron and
in the latter there is an absence of an electron ie a hole.

The overall result shown in the Table above is that:


4 configurations (d1, d4, d6, d9) give rise to D ground terms,
4 configurations (d2, d3, d7, d8) give rise to F ground terms
and the d5 configuration gives an S ground term.

The Crystal Field Splitting of Russell-Saunders terms

The effect of a crystal field on the different orbitals (s, p, d, etc.) will result in splitting into subsets of
different energies, depending on whether they are in an octahedral or tetrahedral environment. The
magnitude of the d orbital splitting is generally represented as a fraction of ∆oct or 10Dq.

The ground term energies for free ions are also affected by the influence of a crystal field and an analogy
is made between orbitals and ground terms that are related due to the angular parts of their electron
distribution. The effect of a crystal field on different orbitals in an octahedral field environment will cause
the d orbitals to split to give t2g and eg subsets and the D ground term states into T2g and Eg, (where upper
case is used to denote states and lower case orbitals). f orbitals are split to give subsets known as t1g, t2g
and a2g. By analogy, the F ground term when split by a crystal field will give states known as T1g, T2g, and
A2g.

Note that it is important to recognise that the F ground term here refers to states arising from d orbitals
and not f orbitals and depending on whether it is in an octahedral or tetrahedral environment the lowest
term can be either A2g or T1g.
25

The Crystal Field Splitting of Russell-Saunders terms


in high spin octahedral crystal fields.
Russell-Saunders Terms Crystal Field Components
S A1g
P T1g
D Eg , T2g
F A2g , T1g , T2g
G A1g , Eg , T1g , T2g
H Eg , 2 x T1g , T2g
I A1g , A2g , Eg , T1g , T2g

Note that, for simplicity, spin multiplicities are not included in the table since they remain the same for
each term. The table above shows that the Mulliken symmetry labels, developed for atomic and
molecular orbitals, have been applied to these states but for this purpose they are written in CAPITAL
LETTERS.

Mulliken Symbols
Mulliken Symbol
for atomic and molecular Explanation
orbitals
Non-degenerate orbital; symmetric to principal
a
Cn
Non-degenerate orbital; unsymmetric to principal
b
Cn
e Doubly degenerate orbital
t Triply degenerate orbital
(subscript) g Symmetric with respect to center of inversion
(subscript) u Unsymmetric with respect to center of inversion
Symmetric with respect to C2 perp. to principal
(subscript) 1
Cn
Unsymmetric with respect to C2 perp. to
(subscript) 2
principal Cn
(superscript) ' Symmetric with respect to sh
(superscript) " Unsymmetric with respect to sh
For splitting in a tetrahedral crystal field the components are similar, except that the symmetry
label g (gerade) is absent.
The ground term for first-row transition metal ions is either D, F or S which in high spin octahedral fields
gives rise to A, E or T states. This means that the states are either non-degenerate, doubly degenerate or
triply degenerate.

We are now ready to consider how spectra can be interpreted in terms of energy transitions between these
various levels.
26

Spectroscopy of First row Transition Metal Complexes

Crystal Field Theory copes reasonably well for d1 (d9) systems but not for multi-electron systems, which
are the more common. To deal with these systems we need to introduce a new concept, that of the
electronic state. Electronic configurations refer to the way in which the electrons occupy the d orbitals, so
for Ti(III) we write an electronic configuration of [Ar] 3d1 and in an octahedral crystal field the lowest
energy configuration would be written as t2g1 eg0. The electronic state refers to energy levels available to
a group of electrons. This is much more complex than the single electron case since not only is it
necessary to consider the crystal field effects of the repulsion of the metal electrons by the ligand
electrons, but it is necessary to include the repulsions between the electrons themselves.

When describing electronic configurations, lower case letters are used, thus t2g1 etc.
For electronic states, upper case (CAPITAL) letters are used and by analogy, a T state is triply
degenerate. Subscripts 1 and 2 are used to distinguish states of like degeneracy and g and u subscripts
indicate the presence of a centre of symmetry eg. T1g, T2g, T1u and T2u.
These symbols are further modified to show the spin multiplicity of the electronic state using the Russell-
Saunders Notation.
If we consider the Ti(III) case, the electronic configuration is d1. In an octahedral crystal field this would
give rise to a t2g1 arrangement and the excitation of the low lying electron to the higher level would then
give an eg1 arrangement.
Only one transition is expected and this roughly corresponds to what is observed for Ti(III) complexes,
although it is somewhat more complicated due to Jahn-Teller considerations.
27

One approach taken when we consider the Russell-Saunders scheme with the various electronic states
makes use of what are called Orgel diagrams. The relevant Orgel diagram for the D ground state is given
below:
28

For a Fe(II) high spin octahedral complex we would write the free ion electronic configuration as d6 and
in the octahedral crystal field it would be described as t2g4 eg2.
The Russell-Saunders scheme that takes into account the electron-electron interactions would be
described by a free ion ground state of 5D. In octahedral and tetrahedral crystal fields, this D state is split
into E(g) and T2(g) terms. To decide which side of the D Orgel diagram should be applied to the
interpretation can be quickly determined by looking at the electronic configuration and noting that the
ground state is triply degenerate and the excited state is doubly degenerate (i.e. we must use the right-
hand-side).
It is expected then that there should be 1 absorption band found in the electronic spectrum and that the
energy of the transition corresponds directly to ∆. The transition is written in a notation that is read from
right to left, which in this case is 5Eg <-- 5T2g.
29

Spectroscopy of First Row Transition Metal Complexes - F ground states

We have already looked at Transition metal ion complexes with D ground states which accounts for the
d1, d4, d6 and d9 configurations.
We turn now to d2, d3, d7 and d8 configurations for which the ground state of the free ions are given by
Russell-Saunders F terms.

Once again, one approach taken to aid in the interpretation of these spectra is to use an Orgel diagram.
The relevant Orgel diagram for the F ground state is given below:

It should be noted that whenever the ground state is an F term there will be a P term found at higher
energy with the same spin multiplicity. The separation of these terms in the free ion is measured in terms
of the B Racah parameter and is equivalent to 15B (where B is usually around 1000 cm-1).
In octahedral and tetrahedral crystal fields, the F state is split into A2(g), T1(g) and T2(g) terms, while the P
term generates an additional T1(g) term.
The lines showing the A2 and T2 terms are linear and depend solely on ∆. The "non-crossing rule" results
in the lines for the two T1 terms being curved to avoid each other and as a result this introduces a
"configuration interaction" in the transition energy equations.
That is, whenever there are lines with equivalent Russell-Saunders terms they are not allowed to cross on
the Orgel diagram and they in fact diverge. All other lines are expected to be linear.
30

For Cr(III) complexes, we would start by writing the free ion electronic configuration as d3 and in an
octahedral crystal field this would be described as t2g3 eg0.
The Russell-Saunders scheme that takes into account the electron-electron interactions would be
described by a free ion ground state of 4F. To decide which side of the F Orgel diagram should be applied
to the interpretation can be quickly determined by looking at the electronic configuration and noting that
the ground state is singly degenerate (i.e. we must use the left-hand-side where the lowest term is A2).
It is expected then that there should be 3 absorption bands found in the electronic spectrum. The energy of
the first transition corresponds directly to ∆ and the transition is written as 4T2g <-- 4A2g.

If we now consider the d2 electronic configuration, then the Russell-Saunders free ion ground term state is
a 3F. For an octahedral complex, the lowest energy state is a triplet which tells us that we need to be using
the right-hand-side of the F Orgel diagram.
The first transition is written as 3T2g <-- 3T1g and the energy of this transition does NOT correspond
directly to ∆. Interpretation of d2 spectra are made even more complicated since as we increase the size of
∆ we note that at some point the 3T1g(P) and 3A2g lines cross. To determine the sequence of the
transitions (that is whether we are on the left-hand-side or right-hand-side of the intersection) normally
requires that all 3 absorption bands can be observed. This is often not the case and an alternative approach
to interpretation of these spectra is needed. This makes use of what are known as Tanabe-Sugano
diagrams.
31

Interpretation of the spectra of first-row transition metal


complexes using Tanabe-Sugano diagrams

For ALL octahedral complexes, except high spin d5, simple CFT would predict that only 1 band
should appear in the electronic spectrum corresponding to the absorption of energy equivalent

to Δ. If we ignore spin-forbidden lines, this applies to those ions with D ground terms e.g. d1, d9
as well as to d4, d6.

The observation of 2 or 3 peaks in the electronic spectra of d2, d3, d7 and d8 high spin
octahedral complexes requires further treatment involving electron-electron interactions. Using
the Russell-Saunders (LS) coupling scheme, these free ion configurations give rise to F ground
states which in octahedral and tetrahedral fields are split into terms designated by the symbols
A2(g), T2(g) and T1(g).
To derive the energies of these terms and the transition energies between them is beyond the
needs of introductory level courses and is not covered in general textbooks[10,11]. A listing of
some of them is given here as an Appendix. What is necessary is an understanding of how to
use the diagrams, created to display the energy levels, in the interpretation of spectra.
In the laboratory component of the course we will measure the absorption spectra of some
typical chromium(III) complexes and calculate the spectrochemical splitting factor, Δ. This
corresponds to the energy found from the first transition below and as shown in Table 1 is
generally between 15,000 cm-1 (for weak field complexes) and 27,000 cm-1 (for strong field
complexes).

For the d3 octahedral case, 3 peaks can be predicted which would correspond to the following
transitions and energies:

4
1. T2g ← 4A2g transition energy = ∆
4
2. T1g(F) ← 4A2g transition energy = 9/5 * ∆ - C.I.
4
3. T1g(P) ← 4A2g transition energy = 6/5 * ∆ + 15B' + C.I.

where C.I. is the configuration interaction


32

Complex ν1 ν2 ν3 ν2/ ν1 ν1/ ν2 ∆/B

Cr3+ in emerald 16260 23700 37740 1.46 0.686 20.4

K2NaCrF6 16050 23260 35460 1.45 0.690 21.4

[Cr(H2O)6]3+ 17000 24000 37500 1.41 0.708 24.5

Chrome alum 17400 24500 37800 1.36 0.710 29.2

[Cr(C2O4)3]3- 17544 23866 ? 1.37 0.735 28.0

[Cr(NCS)6]3- 17800 23800 ? 1.34 0.748 31.1

[Cr(acac)3] 17860 23800 ? 1.33 0.752 31.5

[Cr(NH3)6]3+ 21550 28500 ? 1.32 0.756 32.6

[Cr(en)3]3+ 21600 28500 ? 1.32 0.758 33.0

[Cr(CN)6]3- 26700 32200 ? 1.21 0.829 52.4

For octahedral Ni(II) complexes the transitions would be:

3
1. T2g ← 3A2g transition energy = ∆
3
2. T1g(F) ← 3A2g transition energy = 9/5 * ∆ - C.I.
3
3. T1g(P) ← 3A2g transition energy = 6/5 * ∆ + 15B' + C.I.

where C.I. again is the configuration interaction and as before the first transition corresponds
exactly to Δ.

For M(II) the size of Δ is much less than for M(III) and typical values for Ni(II) are 6500 to 13000
cm-1 as shown in the following Table .
33

Complex ν1 ν2 ν3 ν1 ν1/ ν2 ∆/B

NiBr2 6800 11800 20600 1.74 0.576 5

[Ni(H2O)6]2+ 8500 13800 25300 1.62 0.616 11.6

[Ni(gly)3]- 10100 16600 27600 1.64 0.608 10.6

[Ni(NH3)6]2+ 10750 17500 28200 1.63 0.614 11.2

[Ni(en)3]2+ 11200 18350 29000 1.64 0.610 10.6

[Ni(bipy)3]2+ 12650 19200 ? 1.52 0.659 17

For d2 octahedral complexes, few examples have been published. One such is V3+ doped in
Al2O3 where the vanadium ion is generally regarded as octahedral.

Complex ν1 ν2 ν3 ν2/ ν1 ν1/ν2 ∆/B

V3+ in Al2O3 17400 25200 34500 1.448 0.6906 30.90

[VCl3(MeCN)3] 14400 21400 ? 1.486 0.6729 28.68

K3[VCl6] 14800 23250 ? 1.571 0.6365 24.78

Interpretation of the spectrum highlights the difficulty of using the right-hand side of the Orgel
diagram above for many d2 cases where none of the transitions correspond exactly to Δ and
often only 2 of the 3 transitions are clearly observed.
The first transition can be unambiguously assigned as:
3
T2g ← 3T1g transition energy = 4/5 * ∆ + C.I.
But, depending on the size of the ligand field ( Δ) the second transition may be due to:
3
A2g ← 3T1g transition energy = 9/5 * ∆ + C.I.
for a weak field or
3
T1g(P) ← 3T1g transition energy = 3/5 * ∆ + 15B' + 2 * C.I.
for a strong field.
34

The transition energies of these terms are clearly different and it is often necessary to calculate
(or estimate) values of B, Δ and C.I. for both arrangements and then evaluate the answers to
see which fits better.

The difference between the 3A2g and the 3T2g (F) lines should give ∆. In this case ∆ is equal to either:
25200 - 17400 = 7800 cm-1
or 34500 - 17400 = 17100 cm-1.
Given that we expect ∆ to be greater than 15000 cm-1 then we must interpret the second transition as to
the 3T2g(P) and the third to 3A2g. Further evaluation of the expressions then gives C.I. as 3720 cm-1 and B'
as 567 cm-1.

Solving the equations like this for the three unknowns can ONLY be done if the three transitions
are observed. When only two transitions are observed, a series of equations[14] have been
determined that can be used to calculate both B and Δ. This approach still requires some
evaluation of the numbers to ensure a valid fit. For this reason, Tanabe-Sugano diagrams
become a better method for interpreting spectra of d2 octahedral complexes.

Using Tanabe-Sugano diagrams

The use of Orgel diagrams allows a qualitative description of the spin-allowed electronic transitions
expected for states derived from D and F ground terms. Only 2 diagrams are needed for high spin d2-d9
and both tetrahedral and octahedral ions are covered.

Tanabe-Sugano diagrams were developed in the 1950's to give a semi-quantitative approach and include
both high and low spin ions and not only the spin-allowed transitions but the spin-forbidden transitions as
well.

At first glance they can appear quite daunting, but in practice they are much easier to use for interpreting
spectra and provide much more information. The obvious differences are the presence of the additional
lines and that the ground state is shown as the base line along the X axis rather than as a straight line or
curve originating from the Y axis.

On the X axis ∆/B' is plotted while on the Y axis E/B' is plotted, where B' is the modified Racah B
parameter that exists in the complex.

A separate diagram is needed for each electronic configuration d2-d9 and for the d4-d7 cases both the
high spin and low spin electronic configurations are shown. The high spin is on the left-hand-side of the
vertical line on the diagram.

For the d2 case where it is difficult to use an Orgel diagram, the TS diagram is shown below. The ground
state is 3T1g which is plotted along the base line.
35

Note that the transitions that occur are dependent on the sizes of ∆ and B and the A2g term may be either

higher or lower than the T1g (P) term (depending on whether ∆/B' is greater than about 15).

For the V(III) aqua ion, transitions are observed at 17,200 and 25,600 cm-1 which are assigned to the 3T2g

← 3T1g and 3T1g(P) ← 3T1g (F) respectively.


Interpretation requires taking the ratio of these frequencies and then finding the position on the diagram

where the height of the 3T1g(P) / 3T2g exactly matches that ratio.

For a ratio of 1.49, this is found on the diagram at ∆/B' of 28.0. (see next page)

Reading off the position on the Y axis for the three spin-allowed lines gives E/B' values of 25.9, 38.6 and
52 (3T2g, 3T1g and 3A2g)

To determine the value of ∆ and B' is now relatively straightforward since from the first transition energy
of 17,200 cm-1 and the value of E/B' of 25.9 we can equate B' as:
B' = 17,200/ 25.9 or B' = 665 cm-1

The value of ∆ can then be determined from the ∆/B' ratio of 28.0 and the value just calculated for B' of
665 cm-1.
This gives ∆ as 28.0 x 665 = 18,600 cm-1

The transition 3A2g ← 3T1g would be predicted to occur at 52 x 665 that is 34,580 cm-1 which is in the UV
region and not observed. (Possibly obscured by charge transfer bands).

The values of ∆ and B' can be compared to similar V(III) complexes and it should be noted that in general
for M(III) ions the D value is often about 3/2 times the value expected for M(II) ions.

The free ion value of B for a V(III) ion is 860 cm-1 and the reduction of this value noted for the observed
B' is a measure of what is described as the Nephelauxetic Effect.
A large reduction in B indicates a strong Nephelauxetic Effect. The Nephelauxetic Series is
given by:

F- > H2O > urea > NH3 > en ~ C2O42- > NCS- > Cl- ~ CN- > Br- > S2- ~ I-

Ionic ligands such as F- give a small reduction in B,


while covalently bonded ligands such as I- give a large reduction of B.
36

As an example of a Cr(III) complex, using the observed peaks found for [Cr(NH3)6]3+, Δ/B' is
found at 32.6. The E/B' for the first transition is given as 32.6 from which B' can be calculated as
661 cm-1. The third peak can then be predicted to occur at 69.64 * 661 = 46030 cm-1 or 217 nm
(well in the UV region and probably hidden by charge transfer or solvent bands).
It is important to remember that the width of many of these peaks is often 1-2000 cm-1. If it is
possible to assign peaks unambiguously, the techniques are valuable.
37

Calculations using Orgel diagrams

Orgel diagrams are useful for showing the energy levels of both high spin octahedral and tetrahedral
transition metal ions. They ONLY show the spin-allowed transitions.

For complexes with D ground terms only one electronic transition is expected and the transition energy
corresponds directly to ∆. Hence, the following high spin configurations are dealt with: d1, d4, d6 and d9.
38

D Orgel diagram

On the left hand side d1, d6 tetrahedral and d4, d9 octahedral complexes are covered and on the right hand
side d4, d9 tetrahedral and d1, d6 octahedral.
For simplicity, the g subscripts required for the octahedral complexes are not shown.

For complexes with F ground terms, three electronic transitions are expected and ∆ may not correspond
directly to a transition energy. The following configurations are dealt with: d2, d3, high spin d7 and d8.

F Orgel diagram
39

On the left hand side, d2, d7 tetrahedral and d3, d8 octahedral complexes are covered and on the right hand
side d3, d8 tetrahedral and d2 and high spin d7 octahedral.

Again for simplicity, the g subscripts required for the octahedral complexes are not shown.

On the left hand side, the first transition corresponds to ∆, the equation to calculate the second contains
expressions with both ∆ and C.I. (the configuration interaction from repulsion of like terms) and the third
has expressions which contain ∆, C.I. and the Racah parameter B.
4
1. T2g ← 4A2g transition energy = ∆
4
2. T1g(F) ← 4A2g transition energy = 9/5 *∆ - C.I.
4
3. T1g(P) ← 4A2g transition energy = 6/5 *∆ + 15B' + C.I.

On the right hand side,


The first transition can be unambiguously assigned as:
3
T2g ← 3T1g transition energy = 4/5 *∆ + C.I.
But, depending on the size of the ligand field (∆) the second transition may be due to:
3
A2g ← 3T1g transition energy = 9/5 *∆ + C.I.
for a weak field or
3
T1g(P) ← 3T1g transition energy = 3/5 *∆ + 15B' + 2 * C.I.
for a strong field.

TANABE-SUGANO DIAGRAMS
An alternative method is to use Tanabe Sugano diagrams, which are able to predict the transition energies
for both spin-allowed and spin-forbidden transitions, as well as for both strong field (low spin), and weak
field (high spin) complexes.

Note however that most textbooks only give Tanabe-Sugano diagrams for octahedral complexes and a
separate diagram is required for each configuration.

In this method the energy of the electronic states are given on the vertical axis and the ligand field
strength increases on the horizontal axis from left to right.

Linear lines are found when there are no other terms of the same type and curved lines are found when 2
or more terms are repeated. This is as a result of the "non-crossing rule".

The baseline in the Tanabe-Sugano diagram represents the lowest energy or ground term state.

The d2 case (not many examples documented).

The electronic spectrum of the V3+ ion, where V(III) is doped into alumina (Al2O3), shows three major
peaks with frequencies of: ν1=17400 cm-1, ν2=25400 cm-1 and ν3=34500 cm-1.
40

These have been assigned to the following spin-allowed transitions.


3 3
T2g <−−− T1g
3 3
T1g(P) <−−− T1g
3 3
A2g <−−− T1g

The ratio between the first two transitions is calculated as ν2 / ν1 which is equal to 25400 / 17400 =
1.448. In order to calculate the Racah parameter, B, the position on the horizontal axis where the ratio
between the lines representing ν2 and ν1 is equal to 1.448, has to be determined. On the diagram below,
this occurs at ∆/B=30.9. Having found this value, a vertical line is drawn at this position.

Tanabe-Sugano diagram for d2 octahedral complexes

On moving up the line from the ground term to where lines from the other terms cross it, we are able to
identify both the spin-forbidden and spin-allowed transition and hence the total number of transitions that
are possible in the electronic spectrum.
41

Next, find the values on the vertical axis that correspond to the spin-allowed transitions so as to determine
the values of ν1/B, ν2/B and ν3/B. From the diagram above these are 28.78, 41.67 and 59.68
respectively.

Knowing the values of ν1, ν2 and ν3, we can now calculate the value of B.

Since ν1/B=28.78 and ν1 is equal to 17,400 cm-1, then B=ν1/28.78 = 17400/28.78


or B=604.5cm-1. Then it is possible to calculate the value of ∆.

Since ∆/B=30.9, then: ∆=B*30.9 and hence: ∆ = 604.5 * 30.9 = 18680 cm-1

The d3 case

Calculate the value of B and ∆ for the Cr3+ ion in [Cr(H2O)6)]3+ if ν1=17000 cm-1, ν2=24000 cm-1 and
ν3=37000 cm-1.

SOLUTION.

These values have been assigned to the following spin-allowed transitions.

4 4
T2g <−−− A2g
4 4
T1g <−−− A2g
4 4
T1g(P) <−−− A2g

From the information given, the ratio ν2 / ν1 = 24000 / 17000 = 1.412

Using a Tanabe-Sugano diagram for a d3 system this ratio is found at ∆/B=24.00


42

Tanabe-Sugano diagram for d3 octahedral complexes

Interpolation of the graph to find the Y-axis values for the spin-allowed transitions gives:

ν1/B=24.00
ν2/B=33.90
ν3/B=53.11

Recall that ν1=17000 cm-1. Therefore for the first spin-allowed transition,
17000 /B =24.00 from which B can be obtained, B=17000 / 24.00 or B=708.3 cm-1.

This information is then used to calculate ∆.


Since ∆ / B=24.00 then ∆ = B*24.00 = 708.3 * 24.00 = 17000 cm-1.

It is observed that the value of Racah parameter B in the complex is 708.3 cm-1, while the value of B in
the free Cr3+ ion is 1030cm-1. This shows a 31% reduction in the Racah parameter indicating a strong
Nephelauxetic effect.
43

The Nephelauxetic Series is as follows:

F->H2O>urea>NH3 >en~C2O42- >NCS- >Cl-~CN->Br- >S2- ~I-.

Ionic ligands such as F-give small reduction in B, while covalently bonded ligands such as I- give
a large reduction in B.

• The method of finding the correct X-intercept is somewhat tedious and time-consuming. A plot of
the energy ratio ν2/ ν1 (from theTanabe Sugano diagram) vs ∆/B is plotted to determine exactly
where the ratio of the the observed transitions occur.

The Irving-Williams series

The general stability sequence of high spin octahedral metal complexes for the replacement of water by
other ligands is:

Mn(II) < Fe(II) < Co(II) < Ni(II) < Cu(II) > Zn(II)

This trend is essentially independent of the ligand.

In the case of 1,2-diaminoethane (en), the first step-wise stability constants (logK1) for M(II) ions are
shown below.

Notes
The sequence is generally quoted ONLY for Mn(II) to Zn(II) since there is little data available for the
other first row transition metal ions; their M(II) oxidation states are not very stable.
The position of Cu(II) is considered out-of-line with predictions based on Crystal Field Theory and is
probably a consequence of the fact that Cu(II) often forms distorted octahedral complexes.
One explanation: Crystal Field Theory is based on the idea that a purely electrostatic interaction exists
44

between the central metal ion and the ligands. This suggests that the stability of the complexes should be
related to the ionic potential; that is, the charge to radius ratio. In the Irving-Williams series, the trend is
based on high-spin M(II) ions, so what needs to be considered is how the ionic radii vary across the d-
block. For free metal ions in the gaseous phase it might be expected that the ionic radius of each ion on
progressing across the d-block should show a gradual decrease in size. This would come about due to the
incomplete screening of the additional positive charge by the additional electron, as is observed in the
Lanthanide Contraction.
For high-spin octahedral complexes it is essential to consider the effect of the removal of the degeneracy
of the d-orbitals by the crystal field. Here the d-electrons will initially add to the lower t2g orbitals
beforefilling the eg orbitals since for octahedral complexes, the t2g subset are directed in between the
incoming ligands whilst the eg subset are directed towards the incoming ligands and cause maximum
repulsion.

For d1-d3 (and d6-d8) the addition of the electrons to the t2g orbitals will mean that the screening of the
increasing attractive nuclear charge is not very effective and the radius should be smaller than for the free
ion. The position of d4 and d9 on the plot is difficult to ascertain with certainty since six-coordinate
complexes are expected to be distorted due to the Jahn-Teller Theorem. Cr(II) is not very stable so few
measurements are available. For Cu(II) however, many complexes are found to have either 4 short bonds
and 2 long bonds or 2 short and 4 long bonds. The radii are expected to show an increase over the d3 and
d8 situation since electrons are being added to the eg subset. The reported values have been found to lie on
both sides of the predicted value.

For d0, d5 and d10 the screening expected is essentially that of a spherical arrangement equivalent to the
absence of a crystal field. The plot above shows that these points return to the line drawn showing a
gradual decrease of the radius on moving across the d-block.

Once the decrease in radius with Z pattern is understood, it is a small step to move to a pattern for q/r
since this only involves taking the reciprocal of the radius and holding the charge constant. The radius
essentially decreases with increasing Z, therefore 1/r must increase with increasing Z.

For the sequence Mn(II) to Zn(II), the crystal field (q/r) trend expected would be:
45

Mn(II) < Fe(II) < Co(II) < Ni(II) > Cu(II) > Zn(II)

Apart from the position of Cu(II), this corresponds to the Irving-Williams series.

The discrepancy is once again accounted for by the fact that copper(II) complexes are often distorted or
not octahedral at all. When this is taken into consideration, it is seen that the Irving-Williams series can
be explained quite well using Crystal Field Theory.

References

1. Basic Inorganic Chemistry, F.A.Cotton, G.Wilkinson and P.L.Gaus, 3rd edition, John Wiley and
Sons, Inc. New York, 1995.
2. Physical Inorganic Chemistry, S.F.A.Kettle, Oxford University Press, New York, 1998.
3. Complexes and First-Row Transition Elements, D.Nicholls, Macmillan Press Ltd, London 1971.
4. The Chemistry of the Elements, N.N.Greenwood and A.Earnshaw, Pergamon Press, Oxford,
1984.
5. Concepts and Models of Inorganic Chemistry, B.E.Douglas, D.H.McDaniel and J.J.Alexander 2nd
edition, John Wiley & Sons, New York, 1983.
6. Inorganic Chemistry, J.A.Huheey, 3rd edition, Harper & Row, New York, 1983.
7. Inorganic Chemistry, G.L.Meissler and D.A.Tarr, 2nd edition, Prentice Hall, New Jersey, 1998.
8. Inorganic Chemistry, D.F.Shriver and P.W.Atkins, 3rd edition, W.H.Freeman, New York, 1999.
9. Basic Principles of Ligand Field Theory, H.L.Schlafer and G.Gliemann, Wiley-Interscience, New
York, 1969.
10. Y.Tanabe and S.Sugano, J. Phys. Soc. Japan, 9, 1954, 753 and 766.
11. (a). Inorganic Electronic Spectroscopy, A.B.P.Lever, 2nd Edition, Elsevier Publishing Co.,
Amsterdam, 1984.
11(b). A.B.P.Lever in Werner Centennial, Adv. in Chem Series, 62, 1967, Chapter 29, 430.
12. Introduction to Ligand Fields, B.N.Figgis, Wiley, New York, 1966.
13. E.Konig, Structure and Bonding, 9, 1971, 175.
14. Y. Dou, J. Chem. Educ, 67, 1990, 134.

Unit 4 : Introduction to Inorganic Reaction Mechanisms

Scope of the course

• Limited to solution reactions


• "classical inorganic" reactions which occur at metal centres in coordination complexes
46

Elementary Reaction Kinetics and Mechanism

Transition State Theory

Note: only applies when transition state in equilibrium with the reactants
does not apply in reactions which are at diffusion controlled limit
Therefore: fast reactions are favoured by

Note: since the rate is exponentially dependent on both S and Eact, small rate changes are "not
significant" for the interpretation of mechanism with these very crude theories.

Effects of Pressure on the Rate of Reactions

Measurements of Rates of Reactions

• "inert" species i.e. t1/2 > 1 min

We can use classical static techniques, e.g. light absorption, pH measurements


• "labile" species: i.e. t1/2 ca. 1 min-1ms
Use stop flow measurements, rapid mixing, fast spectroscopy
• "rapid"reactions: -relaxation techniques + fast spectrophotometry
47

General Comments:
We cannot conclude a mechanism from a rate law !
e.g.

(i) involvement of solvent (pseudo-first order behaviour)


(ii) complex reactions with only one rate limiting step.

Finally remember you can only disprove a mechanism you can never prove a mechanism.
A rate law can at best only be consistent with a mechanistic scheme, it can
never prove it.

Definitions:

(1) "Lability" and "Inert" are kinetic terms !!


(2) "Stable" and "Unstable" are thermodynamic statements
(3) "Intimate" Mechanism refers to the details of the mechanism on the molecular scale.

Substitution at Square Planar Metal Complexes

Examples of Square Planar Transition Metal Complexes:


Ni(II) (mainly d8) Rh(I) Pd(II) Ir(I) Pt(II) Au(III)

General Rate Law:

Factors Which Affect The Rate Of Substitution


1. Role of the Entering Group
2. The Role of The Leaving Group
3. The Nature of the Other Ligands in the Complex
4. Effect of the Metal Centre
48

An illustration of the importance of solvent on the substitution pathways for square planar reaction
centres

1. Role of the Entering Group

• rate is proportional to the nucleophilicity of entering group


• element specific
• nPt scale is the relative index of nucleophilicity obtained from the reaction of the standard complex
trans-[Pt(py)2Cl2] in methanol, nPt is defined according to Eqn:
snPt = Log(ky/ks)
49

• No correlation of nPt scale with other properties of the nucleophiles, such as basicity, redox potentials,
or pH.
E.g. strongly basic ligands such as OH- or OMe- do not react with Pt(II) substrates.
• For the reactivity series I- > Br- > Cl- the most important factor seems to be the polarizability or
"softness" resulting from the existence in low lying excited states.
• PR3 > AsR3 > SbR3 >> NR3
• -Notable exceptions to this series are NO2-, SeCN-, which are "biphilic"
• "Biphilic" ligands exhibit different reactivities depending on the charge on thecomplex.
Biphilic Ligands:
• Biphilic properties originate from -acceptor properties from non-bondingelectrons
• Availability of non-bonding electrons is dependent on ligand environment

2. The Role Of The Leaving Group

For the reaction


[Pt(dien)X]+ + py ----------> [Pt(dien)(py)]+ + X-
o
in H2O at 25 C the sequence of lability is;
NO3 > H2O > Cl >Br > I > N3 > SCN > NO2 > CN
with a spread of over 106 in rate across series.
• leaving group does not affect the nucleophilic discrimination factors only the intrinsic reactivity
• the series tend to parallel the strength of the Metal--L bond
50

3. The Nature of other Ligands in the Complex

• The Trans Ligands


Definition;
The trans effect is best defined as the effect of a coordinated ligand upon the rate of substitution
of ligands opposite to it.
For example, in the substitution reactions of Pt(II) square planar complexes, labilizing effect is in
the order;
H2O ~ OH- ~ NH3 ~ amines ~ Cl- < SCN- ~ I- < CH3- < Phosphines ~ H- < Olefins < CO ~ CN-
Note that the "labilizing effect" is used to emphasise the fact that this is a kinetic phenomenon.

This labilization may arise because of destabilisation (a thermodynamic term) of the ground
state and/or a stabilisation of the transition state.

The trans influence is purely a thermodynamic phenomenon. That is, ligands can influence the
ground state properties of groups to which they are trans. Such properties include;
(i) Metal-Ligand bond lengths
(ii) Vibration frequency or force constants
(iii) NMR coupling constants
The trans influence series based on structural data, has been given as;
R- ~ H- >= PR3 > CO ~ C=C ~ Cl- ~ NH3

* The cis ligands


In cases where a relatively poor nucleophile act as the entering group
51

Note: Compare with the trans series below, it acts in the same way.

4. Effect of the Metal Centre

The order of reactivity of a series of isovalent ions is;


Ni(II) > Pd(II) >> Pt(II)
• This order of reactivity is the same order as the tendency to form 5-coordinate complexes.
• More ready the formation of a 5-coordinate intermediate complex, the greater the stabilisation of the
transition state and so the greater the bimolecular rate enhancement.

M = Ni ky = 33 M-1 sec-1

Pd ky = 0.58 M-1 sec-1

Pt ky = 6.7 x 10-6 M-1 sec-1

The Intimate Mechanism for Substitution at Square Planar Complexes

The assiciative pathways utilized in the substitution of one ligand for another at a square planar
reaction centre
52

TYPICAL REACTION COORDINATES

The 1st order dependence of the rate of the reaction on:


• the concentration of the substrate
• the concentration entering reagent
Indicates that these complexes undergo substitution of ligands by a bimolecular mechanism.
Some important points concerning the "intimate" mechanism
53

Profile A applies when the energies of the reactants and products are almost equal, and in
which the energies of the transition states are almost equal.
Profile B is appropriate to a reaction for which the bond breaking is rate determining (d
mechanism from the intermediate)
Profile C depicts the situation when bond forming is rate determining, a situation in which the
entering group (Y) lies higher in the trans effect series than the leaving group.
In a few cases the intermediate is reasonable stable and so lies at lower energy than the
products.
For example;

A reaction profile for a reaction where a quasi-stable intermediate is formed before the transition
state: bond breaking in the five-coordinate intermediate is more important than bond making to
form the intermediate

General Substitution Reactions of Octahedral Complexes

Studies on octahedral complexes have largely been limited to two types of reaction:

• Replacement of coordinated solvent ( eg water). Perhaps the most thoroughly studies replacement
reactions of this type is the formation of a complex ion from a hydrated metal ion in solution.

• Anation: When the entering group is an ion the reaction is called anation.
54

• Solvolysis. Since the majority of such reactions have been carried out in aqueous solution,
hydrolysis is a more appropriate term. Hydrolysis reactions have been done under acidic or basic
conditions.

Octahedral Substitution: General Comments

[M(H2O)6]n+ + X- -----> [M(H2O)5X](n-x)+ + H2O

Observations Implications

• Rates quite similar to water exchange rates (H2O dissociation important factor)
• Rate increases with change on anion ("Outer Sphere" Complex)
• Rate unaffected by nucleophilicity, basicity, etc... (A and Ia paths thus unlikely)
• Rate strongly dependent on M (suggests D or Id mechanism)
• Kinetics 2nd order (does not eliminate any mechanism)
• Intermediates not detected (very few cases, e.g. [Co(CN)5H2O]2-)
• V +ve, generally (careful !!!!)
55

General Scheme for Substitution Reactions

Rates of Water Exchange


56

• The 1st Order rate constants for the substitution of inner sphere water in transition and non-
transition metal complexes fall over 10-orders of magnitude.
General Comments

1. The rate constants of a given ion are approximately constant, no matter what the nature of the
entering group. (evidence for dissociative activation)
2. The substitution rates divide the metal ions into 4-distinct groups

Class 1.
Exchange of water is very fast and is essentially diffusion controlled (k >= 108 sec-1) Ions include
Groups 1A , IIA (except Be2+, Mg2+), and IIB (except Zn2+), plus Cr2+, and Cu2+.
Class 2
Rate constants are in the range 104-108 sec-1. This includes most of the 1st row T.M divalent ions
(except V2+, Cr2+, Cu2+) and Mg2+ and the lanthanide M3+ ions.
Class 3.
Rate constants are in the range 1 - 104 sec-1. This includes Be2+, Al3+, V2+, and some 1st row T.M.
tri-valent ions.
Class 4.
Rate constants are in the range 10-3-10-6 sec-1. This includes Cr3+, Co3+, Rh3+, Ir3+, Pt2+

Factors which control the rate of H2O Substitution

1. For Class 1. ionic size, and ionic charge are clearly the important. As the ion becomes smaller the
substitution rate slows. Reflecting importance of orbital overlap between metal ion and departing
ligand. Group IIA, and IIB show similar trend, (Be2+ is anomalous due to complex hydrolysis
reactions).
2. For T.M. metals the correlation of rate with size is not obeyed, e.g. Cr2+, Ni2+, and Cu2+ have
identical radii.
3. Cu2+(d9) and Cr2+(d4) are structurally distorted by the Jahn-Teller effect, with bond to the axial
ligands longer and weaker than bonds to the equatorial groups. Therefore the ground state
structures are not far removed from the transition state structures.
57

To explain the behaviour of the other M2+ ions we need to look at the effects of d-electron
configuration.

Orbital Occupation Effects on Substitution Reactions of Octahedral Complexes

The basic assumption is that there is a significant contribution to the activation energy in a substitution
reaction is the change in d-orbital energy on going from the ground state of the complex to the transition
state. Any loss in energy would contribute to the activation (and slow the rate)
Using this approach the [M(H2O)6]2+ ions show the following relative rates;
V2+ << Cr2+ >Mn2+ > Fe2+ > Co2+ > Ni2+ < Co2+ > Zn2+

General Comments

1. Series is generally in good agreement with experimental data.


2. The d3 ions are predicted to be substitutionally inert.
3. Except for Ni2+(d8), the high spin d4-d10 M2+ ions are predicted to be labile by either C4v or D3h
transition state pathways).
4. Some deviation that need further explanation; Cr2+(d4),and Cu2+(d9) complexes are especially
labile. This is due as mentioned earlier to the Jahn-Teller effect.
5. Mn2+, Fe2+, and Co2+ are predicted to have equal labilities, but we observe experimentally that
substitution rates decrease across the series. This is due to the increase in Zeff across the series
and increase in E(M--L).

Stereochemistry of Substitution of Octahedral Complexes

General Points about Stereochemistry

1. In general, trans-[Co(en)2ACl]2+ undergo stereochemical change on hydrolysis-whereas cis-


[Co(en)2ACl]2+react with retention of original chirality.
2. cis isomer generally react faster than trans isomers
3. Ligands A, that lead to stereochemical change in the trans series are those with p orbitals
4. If the reaction proceeds through a square pyramidal intermediate, retention of configuration will
be observed.
5. Proceeding through a trigonal bipyramid, has 75% chance of leading to stereochemical change.
58

Points 3 and 5

Stereochemistry of Substitution of Octahedral Complexes


59

Compound D or Id Mechanism

Tetragonal or Square Pyramid Trigonal Bipyramid


Intermediate Intermediate

cis product trans product cis product trans product

trans-ML4AXn+ 0 100 66.6 33.3

cis-ML4AXn+ 100 0 83.3 16.6

trans-M(L-
0 100 66.6 33.3
L)2AXn+

cis-M(L- = 58.3
100 0 16.7
L)2AXn+ = 25.0

Base-Catalysed Hydrolysis: The CB Mechanism (conjugate base


mechanism)

As mentioned earlier the rates of substitution of octahedral complexes are not sensitive to the nature of
the entering group-with one exception. In basic media Co(III) complexes having ligands of the type NH3,
RNH2, R2NH are sensitive to the nature of the entering group. The base catalysed reactions are generally
much more rapid than anation or hydrolyses in acid solution.

The agreed mechanism, involves the removal of a proton from the amine ligand. This step is generally
very fast, (105 faster), and represents rapid pre-equilibrium to the rate determining loss of leaving group.
60

If, however, K is quite small (it is in th erange 0.01-0.2 for Pt(IV) complexes) so that K[OH] << 1,, the
rate law would reduce to that experimental observed.
61

Electron Transfer Reactions

Outer Sphere Electron Transfer


Electron transfer reactions may occur by either of both of two mechanisms: outer or inner
sphere mechanisms. In principle all outer sphere mechanism involves electron transfer from
reductant to oxidant with the coordination shells or spheres of each staying intact. That is one
reactant becomes involved in the outer or second coordination sphere of the other reactant and
an electron flows from the reductant to oxidant. Such a mechanism is established when rapid
electron transfer occurs between two substitution-inert complexes.
62

Inner Sphere Electron Transfer


An inner sphere mechanism is one in which the reactant and oxidant share a ligand in their
inner or primary coordination spheres the electron being transferred across a bridging group.

Outer Sphere Electron-Transfer

Case 1

Electron-transfer will lead to change in the total optical rotation, thus if equal concentrations of reagents
are used:
Rate of Racemization [[equivalence]] Rate of Electron Transfer

Studied by :
* Isotopic labelling with 54Mn * 17O NMR studies

Outer Sphere Electron-Transfer - Mechanism


Electron transfer from the reductant to the oxidant, with the coordination shells or spheres of each staying
intact. (One reactant becomes involved with outer sphere or second coordination sphere of the other
reactant) Such reactions are observed in electron-transfer reactions of substitutionally inert
complexes
63

Elementary Steps in the Outer Sphere Mechanism


Formation of a precursor (cage) complex

Chemical activation of the precursor, electron transfer and


relaxation of the successor complex

Dissociation to the separated products

kobs = KAkel
Step 1. Formation of the Precursor Complex
This step is always considered to be fast, the rate is generally determined by kel.
Step 2. Chemical Activation and Electron Transfer
Important factors are:
(i) Solvent reorganisation
(ii) Electronic structure
(iii) M-L reorganisation small

Inner Sphere Electron Transfer

The reduction of hexaamminecobalt(3+) by hexaaquochromium(2+) occurs slowly (k = 10-3 M-1sec-1) by


an outer sphere mechanism.

However, if one ammonia ligand on Co(III) is substituted by Cl-, reaction now occurs with a substantially
greater rate (k = 6 x 105 M-1 sec-1).

For inner sphere electron transfer,The reductant to the oxidant share a ligand in their primary
coordination sphere, the electron being transfered across a bridging group.
64

Prerequisites for Inner Sphere Mechanism

• One reactant (usually the oxidant) possess at least one ligand capable of binding simultaneously to
two metal ions.
• The other reactant is substitutionally labile; i.e. one ligandmust be replaced by the bridging ligand.

Examples

Ligand transfer is not a requirement for inner sphere mechanism.

Taube's Conlcusion
Inner sphere mechanism can be unequivocally assigned when both oxidant and oxidised
reducing agent are substitution inert and when ligand transfer from oxidant to reductant has
accompanied electron transfer

The Elementary Steps and Rate Expression for the Inner Sphere Mechanism

Ox-X + Red(H2O) Ox--X--Red + H2O


Formation of a precursorcomplex
Ox--X--Red -Ox--X--Red+
65

Activation of the precursor, and electron transfer


-Ox--X--Red+ + H2O Ox(H2O)- + RedX+
Dissociation to the separated products
The overall reaction can be given by:

k1 is rate limiting, formation of the precursor complex (usually substitution of the bridging ligand for
H2O on the Red complex)
k3 is rate determining, electron transfer within the complex, or fission of the successor complex
Both extremes give 2nd order kinetics

Important Factors which Affect the Rate of Inner Sphere Reactions

1. Formation of the Precursor Complex

This step is closely related to the rates of substitution in Octahedral Complexes (see lecture 3),
* Octahedral d3 are relatively inert
* High spin d4 and d5 are labile

Thus the rates of electron transfer involving [V(H2O)6]2+ (d3) are comparable to the rate of
water substitution.

2. Electronic Structure of Oxidant and Reductant

Many electron transfer reactions are greatly enhanced when they proceed via an inner sphere
mechanism.
This can be explained by consideration of the symmetries of the reductant orbital from
which the electron is lost and the oxidant orbital into which the electron moves.
EXAMPLE
The *-> * electron transfer is the most accelerated, and is the route that requires
most chemical activation, i.e. bond rearrangement in the transition state. These
requirements seem to be largely over come by forming 3-centre bonds in the transition
state.
66

3. Nature of the Bridging Ligand

Inner sphere electron transfer reactions are very sensitive to the nature of the bridging ligand.
The bridging group has two roles;
1. To bring the metal centres together (Thermodynamic contribution), important
factor here is the stability of the intermediate, and M-L bond strengths.

2. The kinetic contribution is the transfer the electron, important factor is the
matching of the donor and acceptor MO's.
Bridging ligands can be organic or inorganic.

However the major difficulty is deciding whether the el-transfer is either outer or inner
sphere. A useful test has been developed, consider the reaction;

The azide ion can better form an intermediate, than thiocyanate.

Studies of inner sphere reactions involving organic bridging ligands, show that reduction rates can
be controlled by steric effects, electronic structure of the bridge, the point on attack on the bridge,
and its reducibility.

4. Fission of the Successor Complex

Occasionally, the rate determining step can be the fission of the binuclear complex, after el-
transfer has occured. For example

The reason of the stability of the binuclear complex in this case is that both Cr(III) (d3) and Ru(II)
(d6) are substitutionally inert.

Differentiation between Outer Sphere vs. Inner Sphere Electron Transfer

1. Outer sphere mechanism is open to all redox active systems, while inner sphere mechanism
requires substitutionally labile reactants and products.
2. Evidence from ligand transfer, labelling, etc..
67

3. V
4. Marcus Relationship (outer sphere reactions)
5. N3- vs. NCS- ratio (Outer Sphere ) ~ 1 ratio (Inner Sphere) >> 1

Relative Rates of Reduction of N3- and NCS- Complexes

Oxidant Reductant KN3-/kNCS- Reaction Type

[Co(NH3)5X]2+ Cr2+ 104 inner sphere

[Co(NH3)5X]2+ V2+ 27 not determined

[Co(NH3)5X]2+ Fe2+ >3 x 103 inner sphere

[Co(NH3)5X]2+ Cr(bipy)32+ 4 outer sphere

[Co(H2O)5X]2+ Cr2+ 4 x 104 inner sphere

You might also like