Sei sulla pagina 1di 66

Inorganic Spectra Dr. Khalil K.

Abid

INORGANIC SPECTRA
By Dr. Khalil K. Abid
Department of Chemistry , College of Science
University of Mustansiriyah, Baghdad – IRAQ

Email: altameemikhalil@gmail.com; abidk@uomustansiriyah.edu.iq

1
Inorganic Spectra Dr. Khalil K. Abid

Infrared spectra
Introduction
Spectroscopy is the study of the interaction of electromagnetic radiation with matter. In a chemical
context, spectroscopy is used to study energy transitions in atoms and molecules. The transitions are
interpreted and can serve to identify the molecule or give clues about the molecular structure. The
technique used differ from each other due to the kind of excitation and therefore information that they
produce: in the radio wave regime, Nuclear Magnetic Resonance (NMR) and Electron Spin Resonance (ESR)
give information on the change of spin. Microwaves give information on the change or orientation. IR
spectroscopy gives information on the change of configuration. UV–VIS and X–rays give information on the
change of electron distribution and on electron transitions and γ–rays give information on the change of
nuclear configuration. When a molecule interacts with electromagnetic radiation, energy is absorbed and
the molecule is promoted, or is said to undergo a transition, to a higher energy state (excited state). In order
for absorption to occur, the energy of the radiation must match the energy difference between the quantized
energy levels of the molecule.
For example, E1 and E2 are the quantized energy levels and ΔE is the energy difference (ΔE = E2 - E1) that
must match the energy of the incident radiation.

E2 incident radiation
E ΔE ( a photon were ΔE = hν) ΔE = hν = hcλ = hcῡ
h = Planck's constant, 6.626 × 10 – 34 J·s
E1 c = speed of light, 3.00 × 108 m·s – 1

Radiation can be characterized by its frequency (ν), its wavelength (λ), or its wavenumber (ῡ ).
Although the wavenumber (cm-1) is not an S.I. unit, it is conventionally used to describe the transitions in
infrared (IR) spectroscopy, which we shall discuss in a moment. The unit of frequency, s -1 ("per second"), is
known as a hertz (Hz). This unit is sometimes convenient for very low energy transitions, such as in nuclear
magnetic resonance (NMR) spectroscopy. In general, an absorption spectrum is obtained by recording the
amount of radiation absorbed by the sample as a function of the frequency or wavelength of the incident
radiation. Each type of spectroscopy focuses upon a specific region of the electromagnetic spectrum
(Figure below). We will be primarily interested with infrared (IR) (4000 - 200 cm-1) and nuclear magnetic
resonance (NMR) (10 - 900 MHz) spectroscopies.

2
Inorganic Spectra Dr. Khalil K. Abid

The Origins of The Infrared Spectrum: Absorption of energy in the infrared region (ῡ = 4000 - 200 cm-1)
arises from changes in the vibrational energy of the molecules. There are two types of vibrations that cause
absorptions in an IR spectrum. Stretching involves rhythmical displacement along the bond axis such that
the interatomic distance alternately increases and decreases. Bending involves a change in bond angles
between two bonds and an atom common to both. In the most basic terms, the infrared spectrum is formed
as a consequence of the absorption of electromagnetic radiation at frequencies that correlate to the
vibration of specific sets of chemical bonds from within a molecule. First, it is important to reflect on the
distribution of energy possessed by a molecule at any given moment, defined as the sum of the contributing
energy terms:
E total= E electronic + E vibrational + E rotational + E translational
The translational energy relates to the displacement of molecules in space as a function of the normal
thermal motions of matter. Rotational energy, which gives rise to its own form of spectroscopy, is observed
as the tumbling motion of a molecule, which is the result of the absorption of energy within the microwave
region. The vibrational energy component is a higher energy term and corresponds to the absorption of
energy by a molecule as the component atoms vibrate about the mean center of their chemical bonds. The
electronic component is linked to the energy transitions of electrons as they are distributed throughout the
molecule, either localized within specific bonds, or delocalized over structures, such as an aromatic ring. In
order to observe such electronic transitions, it is necessary to apply energy in the form of visible and
ultraviolet radiation. The fundamental requirement for infrared activity, leading to absorption of infrared
radiation, is that there must be a net change in dipole moment during the vibration for the molecule or the
functional group under study.
While it was stated that the fundamental infrared absorption frequencies are not the only component to be
evaluated in a spectral interpretation, they are the essence and foundation of the art. For the most part, the
basic model of the simple harmonic oscillator and its modification to account for an harmonicity suffice to

3
Inorganic Spectra Dr. Khalil K. Abid

explain the origin of many of the characteristic frequencies that can be assigned to particular combinations
of atoms within a molecule. From a simple statement of Hooke’s law we can express the fundamental
vibrational frequency of a molecular ensemble according to Equation :
ν = 1/ 2πc ĸ/μ
where ν = fundamental vibration frequency, k = force constant, and μ = reduced mass. The reduced
mass, μ = m1m2/(m1 + m2), where m1 and m2 are the component masses for the chemical bond under
consideration. This simple equation provides a link between the strength (or springiness) of the covalent
bond between two atoms (or molecular fragments), the mass of the interacting atoms (molecular fragments)
and the frequency of vibration. Although simple in concept, there is a reasonably good fit between the bond
stretching vibrations predicted and the values observed for the fundamentals. It can be shown that the
number of normal modes of vibration for a given molecule can be determined as:
number of normal modes = 3N – 6 (nonlinear) = 3N – 5 (linear)
where N is the number of component atoms in the molecule. In practice, apart from the simplest of
compounds, most molecules have nonlinear structures, except where a specific functional group or groups
generate a predominant linear arrangement to the component atoms. If we calculate the number of modes
for a simple hydrocarbon, such as methane (nonlinear, tetrahedral structure), a value of nine are obtained.
This would imply that nine sets of absorption frequencies would be observed in the spectrum of methane
gas. In reality, the number observed is far less, corresponding to the asymmetric and symmetric stretching
and bending of the C – H bonds about the central carbon atom. The reason for the smaller than expected
number is that several of the vibrations are redundant or degenerate, that is, the same amount of energy is
required for these vibrations.

Simple Inorganics: Characterization of compounds via infrared spectroscopy is not limited to organic
compounds. Any inorganic compound that forms bonds of a covalent nature within a molecular ion
fragment, cation or anion, will produce a characteristic absorption spectrum, with associated group
frequencies. In a manner, certain aspects have already been covered for inorganic compounds in the form
of salts of carboxylic acids and amino and ammonium compounds, which can be extended to metal
complexes, chemical fragments associated with heterooxy groups (nitrates, sulfates, phosphates, silicates,
etc.), and transition metal carbonyl compounds. All ionic compounds (containing more than one atom) and
coordination compounds produce characteristic spectra. Many of the associated group frequencies can be
used diagnostically for characterization. The structure and orientation of the ion or complex, both as an

4
Inorganic Spectra Dr. Khalil K. Abid

isolated entity or within a crystal lattice, are important factors that affect the appearance and nature of the
infrared spectrum. Hydration of compounds (water of crystallization) also has a large effect on the
spectrum, and often adds a lot of complexity, in the form of additional absorption bands and structure to
existing bands. A few example for common inorganic ions are included below.
Group frequency (cm-1) Functional group/assignment
1490–1410/880–860 Carbonate ion
1130–1080/680–610 Sulfate ion
1380–1350/840–815 Nitrate ion
1100–1000 Phosphate ion
1100–900 Silicate ion
3300–3030/1430–1390 Ammonium ion
2200–2000 Cyanide ion, thiocyanate ion, and related ions
The practical situation– obtaining the spectrum and interpretation the results:
1. The sample (or spectrum) is a ‘‘total unknown’’ and an identification is required – examples include
forensic samples, environmental waste samples, or new discovery samples, where a new material has been
synthesized or discovered.
2. The sample (or spectrum) is an unknown and it needs to be characterized or classified – examples include
commercial applications where new additives or components are included in a material to provide a specific
property; in such cases this could be considered the basis of competitive product analysis.
3. The sample generally is known but the existence of a specific chemical class needs to be determined;
example include contaminant analysis, analysis for toxicology or environmental reasons, material additives
4. The sample is a complete known and the interpretation is required to confirm the material composition
and/or quality – examples include product quality control and the confirmation of a structure or functionality
of a newly synthesized material. A Quick Diagnostic Assessment of an Infrared Spectrum
Step 1: Overall Spectrum Appearance
Step 2: Testing for Organics and Hydrocarbons – Absorptions in the Region 3200–2700 cm-1
Step 3: Testing for Hydroxy or Amino Groups – Absorptions in the Region 3650–3250 cm-1
Step 4: Testing for Carbonyl Compounds – Absorptions in the Region 1850–1650 cm-1
Step 5: Testing for Unsaturation – Weak to Moderate Absorption in the Region 1670–1620 cm-1
Step 6: Testing for Aromatics – Well-defined Absorptions in the Region 1615–1495 cm-1
Step 7: Testing for Multiple Bonding (often with a bond order of 2 or higher), absorption in the region
2300–1990 cm-1.

5
Inorganic Spectra Dr. Khalil K. Abid

In general some changes noticed in the infrared spectra of any ligand in the coordinated compounds
when compared with the spectra of the free ligand, which could be summarized in the following :
– Some bands shifted to lower or upper region( 5 – 40 cm-1). Why?
– Some sharp bands could be changed to a broad ones. Why ?
– Some bands could be split to many bands. Why ?
– A new bands will be recorded in the region( 600 – 200 cm -1) not founds in the original spectra. Why ?

1) Ammine & it`s derivatives: The simplest complexes , hexaammine geometry such as [Co(NH3)6]Cl2
showed an infrared spectra ( Fig. 2) quite close as 1:1 (metal:ligand) and that spectra attributed to ;
antisymmetric and symmetric NH3 , stretching, degenerate deformation, symmetric deformation and rocking
vibration appears in the regions of ; 3400 – 3000, 1650 – 1550, 1370 – 1000, 950 – 600 cm-1 respectively. The
NH3 stretching frequencies of the complexes are lower than those of the free NH3 molecule for two reasons;
One is the effect of coordination , upon coordination the N – H bond is weakened and the NH3 frequencies
are lowered , the stronger the M – N bond the weaker the N – H bond . Thus the NH3 stretching frequencies
may be used as a rough measure of M – N bond strength. The other reason is the effect of the counter ion
( Cl-), due to the formation of hydrogen bonding N – H ….. Cl. The intensity of the M – N stretching mode
increase as the M – N bond becomes more ionic and as the M – N stretching frequency become more lower.
So the vibrations changes as follow:
ν( M+4 – N ) > ν( M+3 – N ) > ν( M+2– N ) > ν( M+1 – N )
For a series of divalent metals, the order is found to be parallel to Irving – Williams series :
Zn+2 < Cu+2 > Ni+2 > Co+2 > Fe+2 > Mn+2
When NH3 groups of hexammine complexes are partly replaced with other groups, the degenerate are split
because of lowering of symmetry and new bands belonging to other groups are appears as the complex
[Co(NH3)4X2]. The stretching force constant for Co – N bond found to be 1.05 UBF, while it is 0.99 , 0.91 and
0.62 for X= F, Cl and I respectively.

Fig. 2 infrared spectra of [Co(NH3)6]Cl2

6
Inorganic Spectra Dr. Khalil K. Abid

2 – Complexes of ethylenediamine and it`s derivatives: The conformation of ethylenediamine(en) in metal


complexes is of particular interest in coordination chemistry. It was found that the CH2 rocking mode
couples strongly with the NH2 rocking and CN stretching modes that appear in the same frequency region.
X – ray of the trans – [[Co(en)2Cl2]Cl indicate that the chelating ethylenediamine takes the gauche
conformation (Fig –3a ). The M – N stretching vibrations have been assigned in the region (410 –380 cm-1 )
for [M(en)3]+2 ion (M= Zn, Cd, Fe, Ni, Co and Mn) and in the region ( 550 – 400 cm-1 ) for M = Cu, Pd and Pt.
For [M(en)2X2 ], the stretching band shifted to 400 –260 cm-1 for the first group and to 420 – 360 cm-1 for the
second group. To distinguish between cis and trans isomers we should compare the spectra in the region
(1700 – 1500 cm-1 ) (NH2 bending ), 950 – 850 cm-1 ( CH2 rocking) and 610 – 500 cm-1 ( M – N stretching).
Ethylenediamine take the trans conformation when it function as a bridging group between two metal atoms
(Fig. – 3b), the infrared spectra found to be quite simple when compared with other conformations. The Infra
– red spectra of diethylenetriamine (dien) complexes have been reported for [Pd(dien)X]X and
[Co(dien)(en)X]X2 . The later exist in four isomeric forms (Fig. – 4a). Their infrared spectra revealed that two
of these isomers contain dien in the mer – configurations, while the other two take the fac – configuration.
The infra – red spectra of ( trien ) complexes as [M(trien)X2 ]X , M = Cr, Co and Rh, have been measured
and their infra – red spectra were reported. These compounds give three isomers ( Fig. – 4b) which can be
distinguished, by CH2 rocking vibration in the region 920 – 870 cm-1. For Co complex cis – α – isomer exhibit
two strong bands at 905, 870 cm-1 , and cis – β – isomer show four bands at 918, 898, 868 and 862 cm -1 ,
while trans isomer give only one band at 874 cm-1 with a weak band at 912 cm-1

Fig. – 3 a) Rotational isomers of 1,2 – disubstituted ethane b) Structure of bridging en

7
Inorganic Spectra Dr. Khalil K. Abid

Fig. – 4 a) Structure of [Co(dien)(en)X] b) Structure of [M(trien)X2]X

3 – Complexes of Pyridine and it`s derivatives: In pyridine metal complexes, the vibrational frequencies of
M – N and M – X are very useful in elucidating the stereochemistry of the complexes . For cyanopyridine
complexes, the infrared spectra were used to determine whether the coordination occur through the nitrile
or the pyridine nitrogen. It was found that 3 – and 4 – cyanopyridine complexes coordinate to the metal
through the pyridine nitrogen, while the 2 – cyanopyridine coordinate to the metal via the nitrile nitrogen.
Imidazole and it`s derivatives forms a complexes with many transition metals. Imidazole is biologicaly
important since imidazole nitrogen of histidyl residue coordinate to metal ion in many metalloprotiens, thus
the identification of the v( M – N ) in biological systems provides a valuable information about the structure
of the active side of a metalloprotiens. Infrared spectra of 2,2 – bipyridine and it`s analogues showed that
the high frequencies bands does not affected in the complexes since they are not metal sensitive, while the
low frequency bands v ( M – N ) were affected depending on the oxidation state and the type of the metal
ion, which can be summarize as follows :
1 – In terms of MO theory, Cr(III), Cr(II), Cr(I), Cr(0), V(II), V(0), Ti(I), Ti(0), Fe(III), Fe(II) and Co(III) have
filled or partly filled t2g ( bonding) and empty eg(antibonding) orbitals. The v( M – N) for these metal
complexes are in the region 300 – 390 cm-1
2 – On the other hand, Co(II), Co(I), Co(0), Mn(II), Mn(0), Mn(-1), Ni(II), Cu(II) and Zn(II) have filled or partly
filled eg orbitals The v( M – N) for these metal complexes are in the region 180 – 290 cm-1 .
3 – No change in frequencies recorded for Cr(II) – Cr(0) and Co(II) – Co(0), while a dramatic decrease were
recorded for Co(III) – Co(II).
4 – The former series in 1 and 2 indicate that the strength of the bond M – N not changed.
Depending upon the nature of alkyl group( R ) an alkylsubstituted α – diimine ( R – N =CH – CH =N – R )
coordinates to the metal [Pt(II) or Pd(II)] as mono or bidentate (chelating) ligand. The stretching band for

8
Inorganic Spectra Dr. Khalil K. Abid

C = N was observed at 1615 – 1624 cm-1 for monodentate coordination and at 1590 – 1604 cm-1 for
bidentate coordination. The metal isotop technique has been used to assign the M – N vibration of metal
complexes with many other ligand . For example the complex [M(DAPD)] , it was found that the stretching
vibration for Ni – N2 was higher than Ni – N1 (why?). For the complex [M(DMNAPY], the M – N stretching
frequencies are lower than those of the corresponding tris – bipy complexes due to the weakening of the
M – N bond as a result to the strain of four membered chelate ring of the [M(DMNAPY] complexes.

4 – Lattice water and Aquo Complexes : Water in inorganic salt may be classified as lattice or coordinated
water. There is ,however, no definite borderline between the two. The former term denote water molecules
trapped in the crystalline lattice, either by weak hydrogen bonds to the anion or by weak ionic bonds to the
metal, or by both ; where as the latter denotes water molecules bonded to the metal through partially
covalent bonds. Spectra of water molecules are highly sensitive to their surroundings. In general lattice
water absorbs at 3350 – 3200 cm-1 ( antisymmetric and symmetric OH stretching) and at 1630 – 1600 cm-1
(HOH bending). Vibrational spectroscopy is very important in elucidating of structures of aqua complexes.
For example, TiCl3.6H2 O should be formulated as trans – [Ti(H2O)2Cl2]Cl.2H2O since it exhibit only one TiO
stretching( about 500 cm-1 ) and one TiCl stretching ( 336 cm-1 )
Complexes of [M(acac)2X2] may take the cis – or trans - structure. Although the steric and electronic
consideration will favor the trans – isomer, the grate stability of cis – isomer is expected in terms of metal –
ligand π - bonding. This is the case of [Ti(acac)2F2 ], which take the cis - configuration with two v(TiF) at
633 and 618 cm-1 . In case of [Re(acac)2Cl2 ], however, both forms can be isolated , the trans – isomer exhibit
one v(ReO) and one v(ReCl) at 464 and 309 cm-1 respectively, while each one of these bands split into two
in the cis – isomer { 472 , 460 cm-1 for v(ReO) and 346, 333 cm-1 for v(ReCl) respectively }.

9
Inorganic Spectra Dr. Khalil K. Abid

For [VO(acac)2L], were L is a substituted pyridine, both cis – and trans – isomer can be expected. They
can be distinguished by their infra – red spectra, since v(V – O ) and v(V = O) of the cis – isomer are lower
than those for trans – isomer. In addition to that , in cis – isomer v(V – O ) split into two bands.

5 – Hydroxo ( OH ) complexes: The spectra of hydroxo complexes are expected to be similar to those of
metal hydroxide. The hydroxo group can be distinguished from the aqua group since the former lack the
HOH bending mode near 1600 cm-1 . Furthermore the hydroxo complexes exhibit the MOH bending mode
below 1200 cm-1 . For example, this mode is at 1150 cm-1 for the complex [Sn(OH)6] -2 ion and at 1165 cm-1
for [Pt(OH)6 ] -2. The OH group can also form bridging group between two metals which give a bending

mode at 955 cm-1 .

6 – Aldehydes , Ketones and Esters: In aldehydes , ketones and easters the v (C=O) shifted to a lower region
, while the v (C – O) shifted to a higher region. The amount of shift depending on metal ions , which seem to
obey the Irving – Williams order. β – diketones form a metal chelate ring with the metal ion. Acetylacetonate
is the most common example for metal complexes , which can form a square planar geometry with 1 : 2
(M:L) and octahedral with 1 : 3 (M:L). The spectra showed two strong bands at 1577 and 1529 cm-1 assigned
for C….C coupled with C …O and C …O coupled with C…C. The vibration of M – O are most interesting,
since they give us a direct information about the strength of M – O bond.

10
Inorganic Spectra Dr. Khalil K. Abid

Electronic Spectra
Types of spectra: Spectra are broadly classified into two groups (i) emission spectra(ii) absorption spectra
i. Emission spectra are of three kinds (a) continuous spectra,(b) band spectra and (c) line spectra.
Continuous spectra: Solids like iron or carbon emit continuous spectra when they are heated until they
glow. Continuous spectrum is due to the thermal excitation of the molecules of the substance.
Band spectra: The band spectrum consists of a number of bands of different colours separated by dark
regions. The bands are sharply defined at one edge called the head of the band and shade off gradually at
the other edge. Band spectrum is emitted by substances in the molecular state when the thermal excitement
of the substance is not quite sufficient to break the molecules into continuous atoms.
Line spectra: A line spectrum consists of bright lines in different regions of the visible spectrum against a
dark background. All the lines do not have the same intensity. The number of lines, their nature and
arrangement depends on the nature of the substance excited. Line spectra are emitted by vapours of
elements. No two elements do ever produce similar line spectra.
ii. Absorption spectra: When a substance is placed between a light source and a spectrometer, the
substance absorbs certain part of the spectrum. This spectrum is called the absorption spectrum
of the substance.Electronic absorption spectrum is of two types. d-d spectrum and charge transfer
spectrum. d-d spectrum deals with the electronic transitions within the d-orbitals. In the charge – transfer
spectrum, electronic transitions occur from metal to ligand or vice-versa.
The ultraviolet and visible spectra of coordination compounds of transition metals involve transitions
between the d orbitals of the metals. Therefore, we will need to look closely at the energies of these orbitals
and at the possible ways in which electrons can be raised from lower to higher energy levels.The electronic
absorption spectrum provides a convenient method for determining the magnitude of the effect of ligands
on the d orbitals of the metal. Although in principle we can study this effect for coordination compounds of
any geometry. The absorption spectrum can be used to determine the magnitude of the octahedral ligand
field parameter for a variety of complexes. In explaining the colors of coordination compounds, we are
dealing with the phenomenon of complementary colors: if a compound absorbs light of one color, we see
the complement of that color. For example, when white light passes through a substance that absorbs red
light, the color observed is green. Green is the complement of red, so green predominates visually when red
light is subtracted from white.Complementary colors can conveniently be remembered as the color pairs on
opposite sides of the color wheel shown in the margin.

11
Inorganic Spectra Dr. Khalil K. Abid

The approximate wavelengths and complementary colors to the principal colors of the visible
spectrum are given in table. Absorption of light results in the excitation of electrons from lower
to higher energy states; because such states are quantized, we observe absorption in "bands‘ , with
the energy of each band corresponding to the difference in energy between the initial and final states.
To gain insight into these states and the energy transitions between them, we first need to consider how
electrons in atoms can interact with each other.
An example from coordination chemistry is the deep blue color of aqueous solution of Copper(II)
compounds, containing the ion [Cu(H2O)6]+2 a blue color is a consequence of the absorption of light
600 – 1000 nm (maximum near 800 nm; Figure – 5 ), in the yellow to infrared region of the spectrum.
The color observed(blue) is the average complementary color of the light absorbed. It is not always possible
to make a simple prediction of color directly from the absorption spectrum, in large part because many
coordination compounds contain two or more absorption bands of different energies and intensities. The
net color observed is the color predominating after the various absorptions are removed from white light.

Fig. – 5 Absorption Spectra of [Cu(H2O)6]+2

12
Inorganic Spectra Dr. Khalil K. Abid

Electronic spectra of transitions metal complexes


Electronic absorption spectroscopy requires consideration of the following principles:
a. Franck-Condon Principle: Electronic transitions occur in a very short time (about 10 -15 sec.) and hence the
atoms in a molecule do not have time to change position appreciably during electronic transition .So the
molecule will find itself with the same molecular configuration and hence the vibrational kinetic energy in
the exited state remains the same as it had in the ground state at the moment of absorption.
b. Electronic transitions between vibrational states: Frequently, transitions occur from the ground
vibrational level of the ground electronic state to many different vibrational levels of particular excited
electronic states. Such transitions may give rise to vibrational fine structure in the main peak of the
electronic transition. All the molecules are present in the ground vibrational level, nearly all transitions that
give rise to a peak in the absorption spectrum will arise from the ground electronic state. If the different
excited vibrational levels are represented as υ1, υ2, etc., and the ground state as υ0, the fine structure in the
main peak of the spectrum is assigned to υ0 → υ0 , υ0 → υ1, υ0 → υ2 etc., vibrational states. The υ0 → υ0
transition is the lowest energy ( the long wavelength ) transition.
c. Symmetry requirement: This requirement is to be satisfied for the transitions discussed above.
Electronic transitions occur between split ‘d’ levels of the central atom giving rise to so called d-d or
ligand field spectra. The spectral region where these occur spans the near infrared, visible and U.V. region.
Ultraviolet UV Visible Vis Near infrared NIR
50,000 – 26300 26300 -12800 12800 -5000 cm-1
200 - 380 380 -780 780 - 2000 nm
Russel-Saunders or L-S coupling scheme : An orbiting electronic charge produces magnetic field
perpendicular to the plane of the orbit. Hence the orbital angular momentum and spin angular momentum
have corresponding magnetic vectors. As a result, both of these momenta couple magnetically to give rise
to total orbital angular momentum. There are two schemes of coupling: Russel- Saunders or L-S coupling
and j-j coupling.
a. The individual spin angular momenta of the electrons, si, each of which has a value of ± ½, combine to
give a resultant spin angular momentum (individual spin angular momentum is represented by a lower case
symbol whereas the total resultant value is given by a upper case symbol): Ʃsi =S; Two spins of each ± ½
could give a resultant value of S =1 or S= 0; similarly a resultant of three electrons is 1 ½ or ½ .The resultant
is expressed in units of h/2 π . The spin multiplicity is given by (2S+1). Hence, If n is the number of unpaired
electrons, spin multiplicity is given by n + 1.

13
Inorganic Spectra Dr. Khalil K. Abid

b. The individual orbital angular momenta of electrons, li, each of which may be 0, 1 ,2, 3 , 4 ….. in units of
h/2π for s, p, d, f, g, …..orbitals respectively, combine to give a resultant orbital angular momentum, L in
units of h/2π . Σ li = L. The resultant L may be once again 0, 1, 2, 3, 4…. which are referred to as S, P, D, F
G,… respectively in units of h/2π.The orbital multiplicity is given by (2L+1).
0 1 2 3 4 5
S P D F G H
c. Now the resultant S and L couple to give a total angular momentum, J. Hence, it is not surprising that J is
also quantized in units of h/2π.The possible values of J quantum number are given as
J = ( L + S ) , ( L + S - 1 ), ( L + S – 2 ) , ( L + S – 3 ), ….. L - S
The symbol | | indicates that the absolute value (L – S) is employed, i.e., no regard is paid to ± sign. Thus
for L = 2 and S = 1, the possible J states are 3, 2 and 1 in units of h/2π. The individual spin angular
momentum, si and the individual orbital angular momentum, ji, couple to give total individual angular
momentum, ji. This scheme of coupling is known as spin-orbit coupling or j -j coupling.

Term symbols:
1. Spectroscopic terms for free ion ground states: The rules governing the term symbol for the
ground state according to L-S coupling scheme are given below:
a. The spin multiplicity is maximized i.e., the electrons occupy degenerate orbitals so as to retain
parallel spins as long as possible (Hund’s rule).
b. The orbital angular momentum is also maximized i.e., the orbitals are filled with highest positive m
values first.
c. If the sub-shell is less than half-filled, J = L– S and if the sub-shell is more than half –filled, J = L +S
The term symbol is given by 2S+1 LJ. The left-hand superscript of the term is the spin multiplicity, given by
2S+1 and the right- hand subscript is given by J. It should be noted that S is used to represent two things;
(a) total spin angular momentum and (b) total angular momentum when L = 0. The above rules are illustrated
with examples.
For d4 configuration: L = 2 i.e., D; S = 2; 2S+1 = 5; and J = L- S = 0; Term symbol = 5D0
For d9 configuration: L = 2 i.e., D ; S = 1 /2 ; 2S+1 = 2 ; and J = L+ S = 3/2 ; Term symbol = 2D5/2
Spin multiplicity indicates the number of orientations in the external field. If the spin multiplicity is three,
there will be three orientations in the magnetic field.- parallel, perpendicular and opposed. The spectro-
scopic term symbols for dn configurations are given in the Table below. The terms are read as follows: The
left-hand superscript of the term symbol is read as singlet, doublet, triplet, quartet, quintet, sextet, septet,

14
Inorganic Spectra Dr. Khalil K. Abid

octet, etc., for spin multiplicity values of 1, 2, 3, 4, 5, 6,7, 8, etc., respectively. 1S0 (singlet S ); 2S1/2 (doublet S
one–half); 3P2 (triplet P two ); 5I8 (quintet I eight). It is seen from the Table that dn and d10 - n have same term
symbols, if we ignore J values. Here n stands for the number of electrons in dn configuration.
dn Term dn Term
d 0 1 S0 d 10 1 S0

d1 1D3/2 d9 2D3/2

d2 3F2 d8 3F4

d4 5D0 d6 5D4

d5 6S5/2

The empty sub -shell configurations such as p0, d 0, f 0 and full filled subshell configurations such as p6,
d10, f14, etc., have always the term symbol 1S0 since the resultant spin and angular momenta are equal to
zero. All the inert gases have term symbols for their ground state 1S0 . Similarly all alkali metals reduce to
one electron problems since closed shell core contributes nothing to L , S and J; their ground state term
symbol is given by 2S1/2. Hence d electrons are only of importance in deciding term symbols of transition
metals.

Total degeneracy: We have seen that the degeneracy with regard to spin is its multiplicity which is given by
(2S+1). The total spin multiplicity is denoted by Ms running from S to -S. Similarly orbital degeneracy, ML, is
given by ( 2L+1 ) running from L to – L. For example, L= 2 for D state and so the orbital degeneracy is
(2x2+1) =5 fold. Similarly, for F state, the orbital degeneracy is seven fold. Since there are (2L+1) values of
ML, and (2S+1) values of Ms in each term, the total degeneracy of the term is given by: 2(L+1)(2S+1). Each
value of ML occurs (2S+1) times and each value of Ms occurs (2L+1) times in the term. For 3F state, the total
degeneracy is 3x7 =21 fold and for the terms 3P, 1G, 1D, 1S, the total degeneracy is 9,9,5,1 fold respectively.
Each fold of degeneracy represents one microstate.

Number of microstates: The electrons may be filled in orbitals by different arrangements since the orbitals
have different ml values and electrons may also occupy singly or get paired. Each different type of
electronic arrangement gives rise to a microstate. Thus each electronic configuration will have a fixed
number of microstates. The numbers of microstates for p2 are given below (for excited and ground states).

No. of microstates for p2. Each vertical column is one micro state. Thus for p2 configuration, there are 15 microstates

15
Inorganic Spectra Dr. Khalil K. Abid

Selection rules:
1. La Porte selection rule: This rule says that transitions between the orbitals of the same sub shell are
forbidden. In other words, the for total orbital angular momentum is ΔL = ± 1. This is La Porte allowed
transitions. Thus transition such as 1S→ 1P and 2D→ 2P are allowed but transition such as 3D→ 3S is
forbidden. Hence transitions from gerade to ungerade (g to u) or vice versa are allowed, i.e., u → g or
g → u but not u → u or g → g. In the case of p sub shell, both ground and excited states are odd and in the
case of d sub shell both ground and excited states are even. As a rule transition should be from even to odd
or vice versa. Electronic transitions within the same p or d sub-shell are forbidden.
2. Spin selection rule: The selection Rule for Spin Angular Momentum is Δ S = 0. Thus transitions such
as 2S → 2P and 3D → 3P are allowed, but transition such as 1S→ 3P is forbidden. The same rule is also
stated in the form of a statement, Electronic Transitions between the different states of spin multiplicity are
forbidden. The selection Rule for total angular momentum, J, is Δ J = 0 or ± 1. The transitions such as 2P1/2
→ 2D3/2 and 2P3/2 → 2D3/2 are allowed, but transition such as 2P1/2 → 2D5/2 is forbidden since Δ J= 2. There is
no selection rule governing the change in the value of n, the principal quantum number. Thus in hydrogen,
transitions such as 1s → 2p, 1s → 3p, 1s → 4p are allowed. Usually, electronic absorption is indicated by
reverse arrow, ← , and emission is indicated by the forward arrow, → , though this rule is not strictly
obeyed. These rules would seem to rule out most electronic transitions for transition metal complexes.
However, many such complexes are vividly colored, a consequence of various mechanisms by which these
rules can be relaxed. Some of the most important of these mechanisms are as follows:
1. The bonds in transition metal complexes are not rigid but undergo vibrations that may temporarily
change the symmetry. Octahedral complexes, for example, vibrate in ways in which the center of
symmetry is temporarily lost; this phenomenon, called vibronic coupling, provides a way to relax the
first selection rule. As a consequence, d-d transitions having molar absorptivities in the range of
approximately 10 to 50 L mol -1 cm -1 commonly occur (and are often responsible for the bright colors of
many of these complexes).
2. Tetrahedral complexes often absorb more strongly than octahedral complexes of the same metal in the
same oxidation state. Metal-ligand sigma bonding in transition metal complexes of Td symmetry can be
described as involving a combination of sp 3 and sd 3 hybridization of the metal orbitals; both types of
hybridization are consistent with the symmetry. The mixing of p – orbital character (of u symmetry) with
d-orbital character provides a second way of relaxing the first selection rule.

16
Inorganic Spectra Dr. Khalil K. Abid

Mechanism of breakdown of selection rules:


1. Spin-orbit coupling: For electronic transition to take place, Δ S = 0 and Δ L= ± 1 in the absence of spin-
orbit coupling. However, spin and orbital motions are coupled. Even, if they are coupled very weakly, a little
of each spin state mixes with the other in the ground and excited states by an amount dependent upon the
energy difference in the orbital states and magnitude of spin –orbit coupling constant. Therefore electronic
transitions occur between different states of spin multiplicity and also between states in which Δ L is not
equal to ± 1. For example, if the ground state were 99% singlet and 1% triplet (due to spin– orbit coupling)
and the excited state were 1% singlet and 99 % triplet, then the intensity would derive from the triplet –triplet
and singlet-singlet interactions. Spin-orbit coupling provides small energy differences between degenerate
state. To this point in the discussion of multielectron atoms, the spin and orbital angular momenta have
been treated separately. In addition, the spin and orbital angular momenta couple with each other, a
phenomenon known as spin-orbit coupling. In multielectron atoms, the S and L quantum numbers combine
into the total angular momentum quantum number J. The quantum number J may have the following
values: J = L + S , L + S - 1 , L + S - 2 ,..., [ L – S ]
Spin-orbit coupling acts to split free-ion terms into states of different energies. The 3P term therefore
splits into states of three different energies, and the total energy level diagram for the carbon atom as an
example can be shown as :
1S ----------------- 1S2
1D ----------------- 1D2 3P2

3P 3P1

3P0

The final step in this procedure is to determine which term has the lowest energy. This can be done by
using two of Hund's rules:
1 – The ground term (term of lowest energy) has the highest spin multiplicity
2 – If two or more terms share the maximum spin multiplicity, the ground term is the one having the
highest value of L.
3 – For subshells (such as P2) that are less than half-filled, the state having the lowest J value has the
lowest energy (3P0 above); for subshells that are more than half filled, the state having the highest J value
has the lowest energy. Half-filled subshells have only one possible J value. These three rules give us that, in
3d2 the configuration 3F is lower in energy than 3P, so for Ti+2 the ground term is 3F. Hund`s rules are

17
Inorganic Spectra Dr. Khalil K. Abid

reasonably reliable for predicting which term has lower energy( the ground state), but are not very reliable
for predicting the order in which terms of higher energy lie. Thus, for Ti+2 the rules predict the order
3F < 3P < 1G < 1D < 1S but the observed order is 3F < 1D < 3P < 1G < 1S
The spin multiplicity rule is fairly reliable for predicting the ordering of n terms, but the greatest L rule is
reliable only for predicting the ground terms, there is generally little correlation of L with the order of the
higher terms. The procedure for predicting the ground term of an atom or ion may be summarized as
follows:
1- Identify the microstate that has the highest value of multiplicity.
2- Identify the highest permitted value of L for that multiplicity.
2. La Porte selection rule: Physically 3d (even) and 4p (odd) wave functions may be mixed, if centre of
inversion (i) is removed. There are two processes by which i is removed.
a. The central metal ion is placed in a distorted field (tetrahedral field, Tetragonal distortions, etc.,) The most
important case of distorted or asymmetric field is the case of a tetrahedral complex. Tetrahedron has no
inversion centre and so d-p mixing takes place. So electronic transitions in tetrahedral complexes are much
more intense, often by a factor 100, than in a analogous octahedral complexes. Trans isomer of [Co(en) 2Cl2] +
in aqueous solution is three to four times less intense than the cis isomer because the former is centro-
symmetric. Other types of distortion include Jahn –Teller distortions.
b. Odd vibrations of the surrounding ligands create the distorted field for a time that is long enough
compared to the time necessary for the electronic transition to occur (Franck Condon Principle). Certain
vibrations will remove the centre of symmetry. Mathematically this implies coupling of vibrational and
electronic wave functions. Breaking down of La Porte rule by vibrionic coupling has been termed as
“Intensity Stealing”. If the forbidden excited term lies energetically nearby a fully allowed transition, it would
produce a very intense band. Intensity Stealing by this mechanism decreases in magnitude with increasing
energy separation between the excited term and the allowed level.

Splitting of energy states: The symbols A(or a) and B (or b) with any suffixes indicate wave functions which
are singly degenerate. Similarly E (or e) indicates double degeneracy and T (or t) indicates triple
degeneracy. Lower case symbols, a1g, a2g, eg, etc., are used to indicate electron wave functions(orbitals) and
upper case symbols are used to describe electronic energy levels. Thus 2T2g means an energy level which is
triply degenerate with respect to orbital state and also doubly degenerate with respect to its spin state.
Upper case symbols are also used without any spin multiplicity term and they then refer to symmetry (ex.,

18
Inorganic Spectra Dr. Khalil K. Abid

A1g symmetry). The subscripts g and u indicate gerade (even) and ungerade (odd). d orbitals split into two
sets - t2g orbitals and eg orbitals under the influence crystal field. These have T2g and Eg symmetry
respectively.

Splitting of Free – ion Term in Octahedral Symmetry

Crystal field components of the ground states of Lanthanide ions f n (n=1 to 14) configuration

19
Inorganic Spectra Dr. Khalil K. Abid

Energy level diagram


Energy Level Diagrams are described by two independent schemes - Orgel Diagrams which are
applicable to weak field complexes and Tanabe –Sugano (or simply T-S) Diagrams which are applicable to
both weak field and strong field complexes.
Racah parameters: The Racah parameters are A, B and C. The Racah parameter A corresponds to the partial
shift of all terms of a given electronic configuration. Hence in the optical transition
considerations, it is not taken into account. The parameter, B measures the inter electronic repulsion among
the electrons in the d-orbitals. The decrease in the value of the interelectronic repulsion parameter, B leads
to formation of partially covalent bonding. Racah redefined the empirical Condon –Shortley parameters so
that the separation between states having the maximum multiplicity (for example, difference between is a
function of 3F and 3P or 4F and 4P is a function of a single parameter, B. However, separations between terms
of different multiplicity involve both B and C.
Tanabe –Sugano diagrams: Tanabe-Sugano diagrams are special correlation diagrams that are particularly
useful in the interpretation of electronic spectra of coordination compounds. In Tanabe-Sugano diagrams,
the lowest-energy state is plotted along the horizontal axis; consequently, the vertical distance above this
axis is a measure of the energy of the excited state above the ground state.
It is an exact solutions for the excited sate energy levels in terms of Dq, B and C are obtained from
Tanabe-Sugano matrices. However, these are very large (10 x 10) matrices and calculations are not feasible.
For this reason Tanabe-Sugano have drawn energy level diagrams known as T-S diagrams or energy level
diagrams. The T-S diagrams are valid only if the value of B, C and Dq are lower for a complex than for the
free ion value. Quantitative interpretation of electronic absorption spectra is possible by using Tanabe –
Sugano diagrams or simply T-S diagrams. These diagrams are widely employed to correlate and interpret
spectra for ions of all types, from d2 to d8. Orgel diagrams are useful only qualitatively for high spin
complexes whereas T-S diagrams are useful both for high spin and low spin complexes. The x-axis in T-S
diagrams represent the ground state term. Further, in T-S diagrams, the axes are divided by B, the
interelectronic repulsion parameter or Racah Parameter. The x-axis represents the crystal field strength in
terms of Dq/ B or Δ / B and the Y-axis represents the energy in terms of E/B. The energies of the various
electronic states are given in the T-S diagrams on the vertical axis and the ligand field strength increases
from left to right on the horizontal axis. The symbols in the diagram omit the subscript, g, with the
understanding that all states are gerade states. Also, in T.S. diagrams, the zero of energy for any particular

20
Inorganic Spectra Dr. Khalil K. Abid

d n ion is taken to be the energy of the ground state. Regardless of the ligand field strength, then, the
horizontal axis represents the energy of the ground state because the vertical axis is in units of E/B and
x-axis is also in units of Δ /B. Thus, the unit of energy in T-S diagram is B, Racah Parameter. For example,
for the d 2 configuration, the lowest-energy state is described by the line in the correlation diagram joining
the 3T1g state arising from the 3F free-ion term with the 3T1g arising from the strong-field term, t2g2. In the
Tanabe-Sugano diagram (Figure – 6 ), this line is made horizontal; it is labeled 3T1g ( F ) and is shown to
arise from the 3F term in the free-ion limit (left side of diagram).
The Tanabe-Sugano diagram also shows excited states. In the d 2 diagram, the excited states of the
same spin multiplicity as the ground state are the 3T2g, 3T1g (P), and the 3A2g . These are the same triplet
excited states shown in the d 2 correlation diagram. Excited states of other spin multiplicities are also
shown but, they are generally not as important in the interpretation of spectra. A good example of the utility
of T-S diagrams in explaining electronic spectra is provided by the d 2 complex [V(H2O)6]+3 . The ground
state is 3T1g ( F ) under ordinary conditions this is the only electronic state that is appreciably occupied.
Absorption of light should occur primarily to excited states also having a spin multiplicity of 3. There are
three of these, 3T1g, 3T1g ( F ) and 3A2g. Therefore, these allowed transitions are expected, as shown in Figure
– 7. Consequently, we expect three absorption bands for [V(H2O)6]+3 one, corresponding to each allowed
transition. Two bands are observed at 17,800 and 25,700 cm -1 . A third band, at approximately 38,000 cm -1,
is apparently obscured in aqueous solution by charge transfer bands nearby.

Fig. 6 Tanabe – Sugano Diagram For d 2 in Octahedral Ligand Field

21
Inorganic Spectra Dr. Khalil K. Abid

Fig. – 7 Absorption Spectra of [V(H2O)6]+3

The molar absorptivities ( ε ) for most bands of [M(H2O)6]+2 are similar (1 – 20 L mol -1 cm -1 ) except for
the spectrum of [Mn(H2O)6]+2which has much weaker bands. Solutions of [Mn(H2O)6]+2 are an extremely pale
pink, much more weakly colored than solutions of the other ions shown. Why is absorption by [Mn(H 2O)6]+2
so weak? To answer this question, it is useful to examine the corresponding Tanabe-Sugano diagram, in
this case for a d 5 configuration. We expect [Mn(H2O)6]+2 be a highspin complex, because H20 is a rather
weak-field ligand. The ground state for weak-field d 5 is the 6A1g . There are no excited states of the same
spin multiplicity, and consequently there can be no spin-allowed absorptions. That [Mn(H2O)6]+2 is colored
at all is a consequence of very weak forbidden transitions to excited states of spin multiplicity other than 6
(there are many such excited states, hence the rather complicated spectrum).
Examples of spectra illustrating the selection rules and the ways in which they may be relaxed are
given in the following sections of this chapter. Our first example will be a metal complex having a d 2
configuration and octahedral geometry, [V(H2O)6]+3 . In discussing spectra, it will be particularly useful to be
able to relate the electronic spectra of transition metal complexes to the ligand field splitting, for octahedral
complexes. To do this it will be necessary to introduce two special types of diagrams, correlation diagrams
and Tanabe-Sugano diagrams. Figure – 8 is an example of a correlation diagram for the configuration d 2.
These diagrams make use of two extremes:
1. Free ions (no ligand field). The terms 3F , 3P, 1G, 1D and 1S were obtained for a d 2 configuration,
with the 3F term having the lowest energy.These terms describe the energy levels of a " free " d 2 ion
( example, a V+3 ion) in the absence of any interactions with ligands. In correlation diagrams.
2. Strong ligand field. There are three possible configurations for two d electrons in an octahedral
ligand field.

22
Inorganic Spectra Dr. Khalil K. Abid

In our example, the possible electron configurations of V+3 in an extremely strong ligand field t2g2 would
be the ground state; the others would be excited states. In correlation diagrams, we will show these states
on the far right, as the "strong field limit." Here, the effect of the ligands is so strong that it completely
overrides the effects of LS coupling. The way of how to analyze the electronic spectra of complexes improve
our understanding of their bonding. First of all we shall start with the electronic spectra of atoms:
1. The free-ion states (terms arising from LS coupling) are shown on the far left.
2. The extremely strong-field states are shown on the far right.
3. Both the free-ion and strong-field states can be reduced to irreducible representations.
Each free-ion irreducible representation is matched with (correlates with) a strong-field irreducible
representation having the same symmetry . Transitions to excited states having the same spin multiplicity
as the ground state are more likely than transitions to states of different spin multiplicity. To emphasize this,
the ground state and states of the same spin multiplicity as the ground state are shown as heavy lines, and
states having other spin multiplicities are shown as dashed lines.
In the correlation diagram the states are shown in order of energy. A non crossing rule is observed: lines
connecting states of the same symmetry designation do not cross. Correlation diagrams are available for
other d-electron configurations. In actual coordination compounds, the situation is intermediate between
these extremes. At zero field, the ml and ms values of the individual electrons couple to form, for d 2 , the
five terms 3F , 3P, 1G, 1D and 1S representing five atomic states with different energies. At a very high ligand
field, the t2g2 ,t2g eg, and eg2 configurations predominate. The correlation diagram shows the full range of
in-between cases in which both factors are important. The aspect of this problem that is important to us is
that free-ion terms (shown on the far left in the correlation diagrams) have symmetry characteristics that
enable them to be reduced to their constituent irreducible representations (in our example, these will be
irreducible representations in the Oh point group).
In an octahedral ligand field, the free-ion terms will be split into states corresponding to the irreducible
representations, as shown in previous lecture. Similarly, irreducible representations may be obtained for
the strong-field limit configurations t2g2 ,t2g eg, and eg2 The irreducible representations for the two limiting
situations must match; each irreducible representation for the free ion must match, or correlate with, a
representation for the strong-field limit( Figure – 8).Tanabe-Sugano diagrams for d 2 through d 8 are shown
in Figures 9 and 10. The cases of d 1 and d 9 configurations will be discussed later. The diagrams for d 4, d 5,

23
Inorganic Spectra Dr. Khalil K. Abid

d 6, and d 7 have apparent discontinuities, marked by vertical lines near the center. These are configurations
for which low spin and high spin are both possible. For example, consider the configuration d 4 :
S=3 S=5

In the weak-field part of the Tanabe-Sugano diagram, the ground state is 5Eg , having the expected spin
multiplicity of 5. On the right (strong-field) side of the diagram, the ground state is 3T1g (correlating with the
3H term in the free-ion limit), having the required spin multiplicity of 3. The vertical line is thus a dividing line
between weak- and strong-field cases: high-spin (weak-field) complexes are to the left of this line and low-
spin (strong-field) complexes are to the right. At the dividing line, the ground state changes from 5Eg to
3T1g .The spin multiplicity changes from 5 to 3 to reflect the change in the number of unpaired electrons.

Fig. – 8 Correlation Diagram for d 2 in Octahedral Ligand Field

24
Inorganic Spectra Dr. Khalil K. Abid

Fig. – 4 Spin-allowed :. 'Transitions for d2 Configuration

Fig. – 10 Simplified Tanabe-Sugano Diagrams of d2-d8 Electron Configurations in Oct. Ligand Fields.

25
Inorganic Spectra Dr. Khalil K. Abid

Jahn – Teller distortions and spectra


To discussed the spectra of d 1 and d 9 complexes, we might expect each to exhibit one absorption
band corresponding to excitation of an electron from the t2g to the eg levels. However, this view must
be at least a modest oversimplification, because examination of the spectra of [Ti(H2O)6]+2 ,( d 1 ) and
[Cu(H2O)6]+2 (d 9 ) shows these coordination compounds to exhibit two closely overlapping absorption bands
rather than a single band. To account for the apparent splitting of bands in these examples, it is necessary
to recall that some configurations can cause complexes to be distorted. For example, a d 9 metal in an
octahedral complex has the electron configuration t2 g6eg3 ;according to the Jahn-Teller theorem, such a
complex should distort. If the distortion takes the form of an elongation along the z axis (the most common
distortion observed experimentally), the t2 g and eg orbitals are affected. Distortion from Oh to D4h symmetry
results in stabilization of the molecule: the eg orbital is split into a lower a1 g level and a higher b1 g level.
When degenerate orbitals are asymmetrically occupied, Jahn-Teller distortions are likely . For example,
the first two configurations below should give distortions, but the third and fourth should not. In practice,
the only electron configurations for Oh symmetry that give rise to measurable Jahn-Teller distortions are
those that have asymmetrically occupied eg orbitals, such as the high-spin d 4 configuration. The Jahn-
Teller theorem does not predict what the distortion will be; by far, the most common distortion observed is
elongation along the z axis. Although the Jahn-Teller theorem predicts that configurations having
asymmetrically occupied t2 g orbitals, such as the low-spin d 5 configuration, should also be distorted, such
distortions are too small to be measured in most cases.

The Jahn-Teller effect on spectra can easily be seen from the example of [Cu(H2O)6]+2 a d 9 ,complex.

Symmetry labels for configurations: Electron confjgurations have symmetry labels degeneracies, as:
T designates a triply degenerate
asymmetrically occupied state.
E designates a doubly degenerate
asymmetrically occupied state
A or B designate a nondegenerate state.

Each set of levels in an A or B state is symmetrically occupied.

26
Inorganic Spectra Dr. Khalil K. Abid

When a 2D term for d 9 is split by an oct. ligand field, two configurations result:

eg
t2 g

The lower energy configuration is doubly degenerate in the eg orbitals and has the designation 2Eg ; the
higher energy configuration is triply degenerate in the t2 g levels and has the designation 2T2 g . Thus, the
lower energy configuration is the 2Eg and the higher energy configuration is the 2T2 g as, in the diagram
below. This is the opposite of the order of energies of the orbitals (t2 g lower than eg )
Eg

2T2 g

2D B2g
A1g
2Eg

B1g
Free ion term Effect of Effect of
O4h field D4h field
In summary, the 2D free-ion term is split into 2Eg and 2T2 g by a field of Oh symmetry, and further split on
distortion to D4h symmetry. The labels of the states resulting from the free-ion term are in reverse order to
the labels on the orbitals; for example, the b1g atomic orbital is of highest energy, whereas the B1g state
originating from the 2D free-ion term is of lowest energy.
For a d 9 configuration, the ground state in octahedral symmetry is a 2Eg term and the excited state is a
2T2 g term. On distortion to D4h geometry, these terms split, as shown before. In an octahedral d 9 complex,
we would expect excitation from the 2Eg state to the 2T2 g state and a single absorption band. Distortion of
the complex to D4h geometry splits the 2T2 g into two levels, the 2Eg and the B2g. Excitation can now occur
from the ground state (now the B1g state) to the A1g , the Eg , or the B2g. The B1g A1g transition is too low
in energy to be observed in the visible spectrum. If the distortion is strong enough, therefore, two separate
absorption bands may be observed in the visible region, to the Eg or the B2g levels (or a broadened or

27
Inorganic Spectra Dr. Khalil K. Abid

narrowly split peak is found , as in [Cu(H2O)6]+2 . For a d 1 complex, a single absorption band, corresponding
to excitation of a t2 g electron to an eg orbital, might be expected:

Ground State 2T2 g Excited State 2Eg

However, the spectrum of [Ti(H2O)6]+3 ,example of a d 9 ' complex, shows two apparently overlapping
bands rather than a single band. How is this possible? One explanation commonly used is that the excited
state can undergo Jahn-Teller distortion, the eg orbitals can split these orbitals into two of slightly different
energy (of A1g and B1g symmetry). Excitation can now occur from the t2 g level to either of these orbitals.
Therefore, as in the case of the d 9 configuration, there are now two excited states of slightly different
energy. The consequence may be a broadening of a spectrum into a two-humped peak, as in [Ti(H2O)6]+3 ,or
in some cases into two more clearly defined separate peaks. There are several useful relationships between the
diagrams for d1-d9.
1) Electron-Hole Inversion: One "hole" in a d shell (i.e. d9) = d1, leading to the same free ion terms of
maximum spin multiplicity. d1 and d9 2D; d2 and d8 3F and 3P
d3 and d7 4F and 4P d4 and d6 5D

But for the same ligand field symmetry, splittings of each free ion term are reversed on going dn to d10-n.

e.g. in Oh field
2) Octahedral vs Tetrahedral Fields: The splitting of d orbitals is inverted on changing from an Oh to a Tet.
field. So too, although the order of free ion terms remains the same, the splittings of these terms in a weak
field will be inverted as will the configurations of the strong field limits (d5 unchanged).

e.g. d1 case

28
Inorganic Spectra Dr. Khalil K. Abid

Charge Transfer Transitions: There is another important class of transition in which the electron moves
from a molecular orbital centered mainly on the ligand to one centered mainly on atom, or vice versa . In
these the charge distribution is considerably different in ground and excited states, and so they are called
Charge Transfer Transitions. The absorption band, can be very intense; it is responsible for the vivid colors
of some of the halogens in donor solvents. The metal on charge transfer (CT) gives rise to intense
absorptions, whereas ‘d–d’ bands are much weaker. In some spectra, CT absorptions mask bands due to
‘d–d’ transitions, although CT absorptions (as well as ligand-centred n – π* and π – π* bands) often occur
at higher energies than ‘d–d’ absorptions.
There are two classes of these bands ; ligand to metal ( L M ) and metal to ligand ( M L ). In
general, most of CT are of the first class. Charge transfer transitions usually lie at the extreme blue end of
the visible spectrum, or in the ultraviolet region. Also, nearly all observed CT transitions are fully allowed,
hence the CT bands are strong, and the extinction coefficient are typically 10 3 to 104 , or more . There are of
course many forbidden CT transition that give rise to weak bands; these are seldom observed because they
are covered up by the strong CT bands, which lie between 400nm (25,000 cm-1); 200nm(50,000 cm-1 ).
Such absorption bands involve the transfer of electrons from molecular orbitals that are primarily ligand
in character to orbitals that are primarily metal in character (or vice versa). For example, consider an
octahedral d 6 complex with σ-donor ligands. The ligand electron pairs are stabilized, as shown in Figure –
11 . The possibility exists that electrons can be excited, not only from the t2g level to the eg but also from the
σ – orbitals originating from the ligands to the eg. The latter excitation results in a charge-transfer transition;
it may be designated as charge transfer to metal (CTTM) or ligand to metal charge transfer (LMCT). This type
of transition results in formal reduction of the metal. A CTTM excitation involving a cobalt (III) complex, for
example. Examples of charge-transfer absorptions are numerous. For example, the octahedral complexes
[IrBr6]-2 , (d 5) and [IrBr6]-3 , (d 6) both show charge-transfer bands. For [IrBr6]-2 , two bands appear, near 600
nm and near 270 nm; the former is attributed to transitions to the t2g levels and the latter to the eg . In [IrBr6]-3
the t2g levels are filled, and the only possible CTTM absorption is therefore to the e g. Consequently, no low
energy absorptions in the 600nm range are observed, but strong absorption is seen near 250 nm,
corresponding to charge transfer to eg. A common example of tetrahedral geometry is the permanganate
ion, MnO4 - , which is intensely purple because of a strong absorption involving charge transfer from orbitals
derived primarily from the filled oxygen p orbitals to empty orbitals derived primarily from the
manganese(VII).

29
Inorganic Spectra Dr. Khalil K. Abid

Figure – 11a Charge Transfer to Metal. b – Charge Transfer to Ligand.

Similarly, it is possible for there to be charge transfer to ligand (CTTL), also known as metal to ligand
charge transfer (MLCT), transitions in coordination compounds having π -acceptor ligands. In these cases,
empty π* orbitals on the ligands become the acceptor orbitals on absorption of light. Figure – 11b
illustrates this phenomenon for a d 5 complex. CTTL results in oxidation of the metal; a CTTL excitation of
an iron(III )complex would give an iron(IV) excited state. CTTL most commonly occurs with ligands having
empty π* orbitals, such as CO, CN -, SCN -, bipyridine, and dithiocarbamate (S2CNR2-). In complexes such as
[Cr(CO)6] which have both σ-donor and π –acceptor orbitals, both types of charge transfer are possible. It is
not always easy to determine the type of charge transfer in a given coordination compound. Many ligands
give highly colored complexes that have a series of overlapping absorption bands in the ultraviolet part of
the spectrum as well as the visible. In such cases, the d-d transitions may be completely overwhelmed and
essentially impossible to observe.
Finally, the ligand itself may have a chromophore and still another type of absorption band, an
intraligand band, may be observed. These bands may sometimes be identified by comparing the spectra of
complexes with the spectra of free ligands. However, coordination of a ligand to a metal may significantly
alter the energies of the ligand orbitals, and such comparisons may be difficult, especially if charge-transfer
bands overlap the intraligand bands.

Ligand to meta ( L M ): Most of metal complexes had this type of transition which can be expected to be
divided to four types of transitions in octahedral configuration. Fig. 12, shows a partial MO diagram for
such complexes, and each of transitions shown is a group of transitions, since the excited orbital
configuration gives rise to several different of similar but not identical energies. Transition of the ν1 type will
obviously be of lowest energy. Second since the π and π* orbitals involved are both approximately non
bonding , they will not vary steeply with M – L distance as the ligand vibrate. The bands for these transitions
should be relatively narrow.

30
Inorganic Spectra Dr. Khalil K. Abid

A third factor that should assist in identifying the ν1 . Set of bands is that they will be missing whenever
the π*(t2g ) orbitals are filled (d 6 complexes). The energies for ν1 will be decrease in the sequence MCl6,
MBr6 and MI6 which is the order of decreasing the ionization potentials ( easier oxidizability ) of the halogen
atoms. As the oxidation state of the metal increase (easier oxidizability ) like RuCl 6-3, RuCl6-2 ,its orbitals
should be deeper , thus the transition should go to lower energy.
Transition of the ν2 type should give the lowest energy CT bands in t2g6 complexes like those for PtX6-2
complexes. Since the transition is from a mainly the nonbonding level to a distinctly antibonding one, the
bands should be fairly broad. The transition assigned to the ν2 sets all have half – widths of 2000 cm_1 to
4000 cm_1. The shift of energy in these bands with change of halogen and change of metal oxidation state
are again as expected for L M transitions. Transition of the ν3 set are all expected to be broad and weak
and are not observed. The ν4 transitions have been observed in a few cases, but in many cases they must
lie beyond the range of observation.

Fig. 12 Partial MO diagram for octahedral complex

In tetrahedral complexes like NiX4-2, CoX4-2 and MnX4-2 a strong L M spectra can be observed and
assigned in much the same way as for octahedral complexes. In the spectrum of [Cr(NH 3)6]+3 ,when one NH3
ligand replaced by one weaker ligand (Cl - ) moves the lower energy band to lower energy than that for
[CrCl(NH3)5]+2 . That arises because the Cl- ligands have π lone pair electrons that are not directly involved
in bonding. The band is an example of an LMCT transition in which a lone pair electron of Cl - is promoted
into a predominantly metal orbital. The LMCT character of similar bands in is [CrX(NH 3)5]+2 confirmed by the
decrease in energy in steps equivalent to about 8000 cm-1 as X is varied from Cl to Br to I .

31
Inorganic Spectra Dr. Khalil K. Abid

Metal to Ligand( M L ): A transfer of charge of electrons from metal to ligands is most commonly
observed in complexes with ligands that have low lying π* orbitals, especially carbonyl (CO ), cyanide
( CN ) and aromatic ligands (diimine, phenanthroline and dithiolene). If the metal ion have a low oxidation
number, in which case the d orbitals will be relatively high in energy, the transition will occur at low energy.
A diimine ligand may also be easily substituted into a complex with other ligands that favor a low oxidation
state. Two example are, [ W(CO)4(phen)] and [ Fe(CO)3(bipy))]. Octahedral metal carbonyl such as Cr(CO) 6

and Mo(CO)6 pair of intense bands which can be assigned to transition from bonding (mainly metals) to the
antibonding( mainly ligands)components of the metal ligands π bonding interactions.

Ligand Field Spectra: So far we have been dealing with the energy levels of isolated atoms or ions. Now we
need to consider how these are affected by the presence of surrounding groups such as anions, ligands or
solvent molecules. Since we know that orbitals have their energies changed by crystal field, with a tendency
to split degeneracies, it is likely that the energy states derived from these will be changed.
If we go back to the simple case of one electron in d orbital set the only term arsing is that of 2D. We shall
ignore spin-orbit coupling for the present. Just as an octahedral field splits the d-orbitals into t2g and eg
subsets, the microstates contributing to 2D are no longer all the same in energy - they are also split into the
same two groups.The states thus arising are called 2T2g and 2Eg. 2Eg will be at higher energy than 2T2g

Tanabe-Sugano Diagrams
and the difference will be a measure of Δ. Note that the spin degeneracy is unaffected. Obviously when you
consider the cases where you have more than one d electron the result is less trivial. It is important to
introduce forces in order of their magnitude, the largest first. The point at which ligand field effects are
introduced depends on its magnitude. With the assumption of RS coupling for the free ion we can imagine
three cases for the magnitude of the crystal field.
a) Term separation > spin-orbit coupling > crystal field
b) Term separation > crystal field > spin-orbit coupling
c) Crystal field > term separation > spin-orbit coupling

32
Inorganic Spectra Dr. Khalil K. Abid

Type a) met in the lanthanides where we are considering transitions between states derived from an
incompletely filled 4f shell. Not going into deep detail of Ln spectra, but note the main points and compare
these with the TM complexes:
1) The crystal field in lanthanide complexes is weak (~ 100 cm-1) compared with the term separation (1000 -
30000 cm-1) and spin-orbit coupling (~1000 cm-1). Therefore the spectra of lanthanide complexes are rather
like the free ion spectra - just small perturbations by the crystal field.
2) Because the crystal filed is weak, i.e. metal ion is not much perturbed by the ligands, differences between
the effects of different ligands are also small. Therefore the spectrum does not change much on changing
the ligand. All compounds of a given ion give rather similar spectra. This does not mean that the changes
that do occur are unimportant, but they are not large.
3) Δ does not change much with small changes in M-L bond length i.e. with bond vibrations. Spectral band
energies do not differ much during vibrations and therefore the bands are sharp.
Types b) and c) are important for ordinary transition elements and they correspond to Weak and Strong
field cases respectively.

Weak Field Approach: This is essentially a perturbation method. Take the energies and types of terms
known from atomic spectra and permit them to be perturbed by the crystal field.

For example : in an octahedral complex (Oh symmetry)


S is unsplit and gives A1g
P is unsplit and gives T1g
D is split and gives T2g + Eg
F is split and gives A2g + T1g + T2g

33
Inorganic Spectra Dr. Khalil K. Abid

Knowing the symmetry of the ligand field enables one to predict the degeneracies of the derived states
but gives no information about their relative energies.
Use of Tanabe-Sugano Diagrams: Used mainly for interpreting electronic spectra, but also of value for
magnetic properties including EPR of compounds of dn ions. In interpreting spectra we must consider
a) Number of bands; b) Band energies; c) Intensities – i) solutions - εmolar; ii) solids – relative
intensities; d) Breadth and Shape
1) Determination of Stereochemistry and Δ Values: e.g. For a complex ML6n+ it should be possible to find a
value of D which will explain all the observed d-d bands. Moreover, the value of Δ should be "reasonable"
when compared with values for related cases.

Octahedral Ni2+ is d8.

Bands are relatively low in intensity. The Tanabe Sugano Diagram for a d8 complex in an octahedral field
is below. The band assignments are shown on the figure as arrows. Δ for en ~ 11000 cm-1 , while Δ for H2O
~ 9000 cm-1.
Note ν2 for Ni(H2O)62+ involves contribution from 1Eg . For comparison, solid KNiCl3. What is its structure?
It has a similar spectrum to these two compounds with bands at 6700, 12700 and 22000 cm-1. The bands are
not very strong. This suggests that the Ni2+ is actually surrounded by 6Cl- ligands as in NiCl2 for which
bands at 6900, 12900 and 22100 cm-1. The Δ values form these are reasonable: Cl < H2O < en ~6900, ~9000,
~11000 agree with spectrochemical series.

34
Inorganic Spectra Dr. Khalil K. Abid

2) Can use spectra to distinguish between various possible donor atom sets: e.g. NO2. can coordinate via N
or O. Δ values differ with N > O. Electronic Spectra of Ni(en)2(NO2)2 and Ni(4-methylpyridine))4(ONO)2 Curve
(A) is for Ni(N-ethylethylenediamine)2(NO2)2 and curve (B) is for Ni(4-methylpyridine)4(ONO)2 Both show
3A2g → 3T2g and 3A2g → 3T1g (F) transitions. [In each case there is a further spin-allowed band to 3T1g(P)
level above 25000 cm-1 not shown in the diagram].

35
Inorganic Spectra Dr. Khalil K. Abid

3) Information about the nature of bonding: For complexes of a known geometry the energy level diagrams
we have constructed using free ion terms in Racah parameters do not fit all the bands. e.g. for tetrahedral
Co(II). Expect three spin-allowed bands. One at highest energy is: 4A2 → 4T1(P) for CoCl4 2- observed at 15000
cm-1 for CoBr4 2- observed at 13500 cm-1 for CoI4 2- observed at 12500 cm-1 BUT for free ion (i.e. Δ = 0)
4F – 4P = 15B = 14540 cm-1 .To reduce B, i.e. allow for covalency , energy of 4T1(P) increases with Δ. For
CoCl4 2- lower two transitions can be fitted with Δ ~ 3100 cm-1. With this Δ 4T1(P) predicted at ~16200 cm-1
observed at 15000 (if we try to fit this ν3 band alone, we find Δ ~ 400 cm-1 this is unreasonable).
The situation is worse for CoBr4 2- and CoI4 2- as ν3 at 13700 and 12500 cm-1 respectively, i.e. < free ion 4P !
This is a problem. We need to take into account covalency in M-L bond. If covalent contributions are
present, the metal d electrons will be partly shared by the ligands i.e. will spend < 100% of their time on the
metal. If they are spread out more onto the ligands their interelectronic repulsions will be less and the 4P - 4F
term separation will decrease (thus we modify the diagram to account for this). The extent to which this is
necessary gives an experimental measure of bond covalency. Measured by the required reduction in the
Racah parameter B as 4F - 4P = 15B and B’< B (expressed as β = B’/B).
For the above three compounds: CoCl4 2- β ~ 0.74
CoBr4 2- β ~ 0.72
CoI4 2- β ~ 0.69. Thus they are still fairly ionic.
4) Distortions from regular symmetry: If the point symmetry at the metal ion is lower than O h or Td some of
the spectral transitions may be split. We can determine the effects of changing the symmetry upon the
orbital degeneracies by using the table of relationships of irreducible representations.

36
Inorganic Spectra Dr. Khalil K. Abid

In a 6-coordinate Ni(II) metal ion all the orbital-triplet states are split by a tetragonal distortion thus in D4h
symmetry 3T2g → 3B2g + 3Eg ; 3T1g → 3A2g + 3Eg. Similarly for the related Fe(II) complexes (see below)

In Oh expect 1 band: 5T2g → 5Eg. In D4h 5Eg becomes 5A1g + 5B1g and we therefore expect 2 bands. Here
splitting increases on changing X = Cl, Br, I. Ground state also split (into 5B2g + 5Eg) but transitions at too low
energy to be seen. [Can use Mossbauer studies to determine splittings less than 1000 cm -1.]

Electronic Spectra of Complexes: The region for the electronic spectra of the complexes usually applied at
the range 200 – 800nm. For transition metal complexes we can identify a maximum of four types of
transitions as follows:
1- Absorption bands attributed to π – π* and n – π* transitions for the organic part of the complexes
(ligands) , located at the ultraviolet region.
2- Absorption bands attributed to charge transfer transition ( metal to ligand or ligand to metal), located
between last part of ultraviolet and first part of visible region.
3- Absorption bands attributed to d – d transition of the metal ion located at the visible region.
4- Absorption bands attributed to the counter ion ( if presents), located at the same region of the charge
transfer absorption bands.
Absorption bands for 1 and 3 seems to be very strong while it is medium for 4 and weak for 3, why?. Any
complex should had one absorption bands ( 2 ) while some complexes could had the others( two or one of
them), why? . Let us take the simplest possible case with a d1 configuration, [Ti(H2O)6]+3 the d electron will
occupy a t2g orbital. On irradiation with light of frequency c, equal to ∆o/h, where h is Plank`s constant. The
d electron captured the quantum of radiation and excited from t2g to the eg orbital. The absorption bands
found to be in the visible spectrum of complex and is responsible for the violet color. Three features of this
absorption band are of importance, the position, the intensity and the breadth. Fig. 13. The position of the
band is related to the splitting of the d orbitals, the spectrum tell us that ∆o in the complex is 20.000cm -1.
37
Inorganic Spectra Dr. Khalil K. Abid

Since there are 83.7cm-1 per kilo joule, that means the splitting energy is about 250kJ per mole. The
intensity of the absorption band is extremely weak compared to the other bands because the transition
involved in this band is in the same orbital (d – d) and according to the Laporta rule it is forbidden ( in
tetrahedral complexes the absorption band is greater with the factor 10* , why?).
The absorption band seems to be a broad band due to the vibration of the complex. Complexes with a d 9
configuration ( CuII) will have a same spectrum as d1 although it is not simple as Ti( III) complex, several
nearly superposed bands, due to the distortion of the octahedral complex ( Jahn – Teller effect). For
complexes having more than one but less than nine electrons, we must employ an energy level diagram
based upon the Russell – Saunders states of the relevant dn configuration in the free ( uncomplexed )
ion.It can be shown that just as the set of five d orbitals is split apart by the electrostatic field of surrounding
ligands to give two or more sets of lower degeneracy, so also are the various Russell – Saunders states of a
dn configuration. The number and types of the components into which an octahedral field will split a state of
given L is the same regardless of the dn configuration from which it arises .
Although the states into which a given free ion state is split are the same in number and type in both
octahedral and tetrahedral field, the pattern of energies is inversed in one case relative to the other, why.
The electronic spectra of the complex [Cr(NH3)6]+3 in aqueous solution showed two types of absorption
bands, Fig. 14,The bands at the lowest energy attributed to d – d transition, while the very intense band
attribute to charge transfer transition . A closer analysis shows that there is also a third transition hidden
under the very intense CT band. The ground configuration for this complex is t2g3 so the excited state is
t2g2eg1 t2g3 ; because there are three t2g orbitals and two eg orbitals, there are in fact six possible
transitions because any of the three t2g electrons can migrate to either of the two eg orbitals. In the absence
of interelectron repulsions, all six transitions occur at the same energy. However, because there are
interelectron repulsions, the transition energies depend on which orbitals are specially involved in the
transition. It may be helpful to see qualitatively from the viewpoint of simple MOT why there are two bands.
The dz2 dxy transition promotes an electron from the xy plane into the already electron rich z direction ( it
is electron rich because both dyz and dxz are occupied). However, the dz2 dxz transition merely relocates
an electron that is already largely concentrated along the z axis , Fig. 15. The repulsion between electrons
in these two cases are not the same, and as a result the two eg t2g transitions lie at different energies. All the
other transitions resemble one or other of these two cases, and three transitions fall into one group and the
other three fall into the second group.

38
Inorganic Spectra Dr. Khalil K. Abid

Fig. 13 The visible absorption spectrum of [Ti(H2O)6]+3

Fig. 14 The spectrum of the d3 complex [Cr(NH3)6]+3

The spectrum of octahedral and tetrahedral complexes of Mn(II) in Mn(H2O)6]+2 and [MnBr4]-2 showed
for the first one, a very weak bands compared to other octahedral complexes and a large number of bands
and last a great variation in the width of the bands, with one extremely narrow indeed( Fig. 16). The
correspond to the 6S ground state of the free ion, which is not split by the Ligand field. All the excited states
have different spin multiplicity from the ground state, and transition to them are spin forbidden. Because of
weak spin orbit interaction, such transition are not totally absent, but they are very weak.
Tetrahedral complexes of Mn(II) are yellow – green, and the color is more intense than that of the
octahedral complexes. As shown in Fig. 16 the molar absorbance is 100 times than that of the octahedral

39
Inorganic Spectra Dr. Khalil K. Abid

due probably to the fact that in tetrahedral environment there is a mixing between p orbitals of the ligands
and d orbital of the metal ion, which is facilitated by overlap of metal d orbitals with ligand orbitals in the
tetrahedral complexes.

Fig. 15 The shift in electron density


a) relocate of electron density toward the ligands on z axis b) less relocation

Fig. 16 The d – d absorption spectra of the Mn(H2O)6]+2 ion (solid) and the [MnBr4]-2 ion ( dashed curve)

40
Inorganic Spectra Dr. Khalil K. Abid

Magnetic Properties of Coordination Compounds

Types of magnetism:
1. Diamagnetism: This arises due to paired electrons. When all the electrons in a molecule are paired, it is
called a diamagnetic compound. This compound will be slightly repelled by the external magnetic field. It
is not affected by temperature , negative in value.
2. Paramagnetism: This is due to unpaired electrons in a compound. The compound will be moderately
attracted by the external magnetic field. The dipoles will not be aligned uniformly but at random in the
absence of external field. Affected by temperature, positive in value .
3. Ferromagnetism: In this compound the magnetic dipoles are arranged in a parallel manner even in
the absence of magnetic field. Hence, these compound s will be magnetic eve in the absence of external
magnetic field. These compounds are strongly attracted by external magnetic field.
4. Antiferromagnetism: In his case, the magnetic dipoles are arranged anti parallel. These compounds
are weakly attracted by external field.
Variation of magnetic susceptibility with temperature: The variation of different types of magnetic
susceptibility with temperature is shown in Figure
Curie temperature (TC): At this temperature, ferromagnetism changes to paramagnetism.
Neel temperature (TN): Antiferromagnetism changes to paramagnetism at this temperature.
Components of paramagnetism: When unpaired electrons are present in a molecule, they have spin and
orbital motions. Paramagnetism results due to these spin and orbital angular motion.
The spin and orbital motion can couple in three ways, spin-spin, orbital-orbital, and spin-orbital.

Magnetic susceptibility and temperature

41
Inorganic Spectra Dr. Khalil K. Abid

Theoretical paramagnetic moment: This is the magnetic moment incorporating all the three types of
coupling, viz., spin-spin, orbital- orbital, and spin-orbital. This is given by Equation 1.

μ = g √ J (J +1) (1)
where μ is the magnetic moment, J is the total angular momentum quantum number and g is the Landé
splitting factor for the electron.
J = L + S , L + S - 1,.....,L – S (2)
L is the orbital-angular momentum and S is the spin-angular momentum. When the spin-orbit coupling
is negligible or absent in a complex but there is significant spin and Orbital contribution, the above
equation transforms as Eq. 3:

μ = √ [4S ( S + 1) + L ( L + 1)] (3)

Magnetic susceptibility and the spin-only formula: Paramagnetism arises from unpaired electrons.
Each electron has a magnetic moment with one component associated with the spin angular momentum
of the electron and a second component associated with the orbital angular momentum. For many
complexes of first row d-block metal ions we can ignore the second component and the magnetic
moment, , can be regarded as being determined by the number of unpaired electrons. The two equations
below are related because the total spin quantum number S = n/2 .

μeff = 2 √ S(S + 1) ………… (4)


μeff = √ n(n + 2) ……… (5)
The effective magnetic moment, μeff, can be obtained from the experimentally measured molar
Magnetic susceptibility, χm, and is expressed in Bohr magnetons (B.M.). In the laboratory, the continued
use of Gaussian units in magnetochemistry means that irrational susceptibility is the measured quantity
and the equation below is therefore usually applied.

μeff = 2.8 √ χm. T ……… (6)


Several methods can be used to measure χm : e.g. the Gouy balance (Figure – 17 ), the Faraday
balance (which operates in a similar manner to the Gouy balance). The Gouy method makes use of
the Interaction between unpaired electrons and a magnetic field; a diamagnetic material is repelled by a
magnetic field whereas a paramagnetic material is attracted into it.

42
Inorganic Spectra Dr. Khalil K. Abid

Fig. – 17 Schematic representation of a Gouy balance

The compound for study is placed in a glass tube, suspended from a balance on which the weight of
the sample is recorded. The tube is placed so that one end of the sample lies at the point of maximum
magnetic flux in an electromagnetic field while the other end is at a point of low flux. Initially the magnet is
switched off, but upon applying a magnetic field, paramagnetic compounds are drawn into it by an amount
that depends on the number of unpaired electrons. The change in weight caused by the movement of the
sample into the field is recorded, and from the associated force it is possible to calculate the magnetic
susceptibility of the compound. The effective magnetic moment is then derived using equation(4). For
metal complexes in which the spin quantum number S is the same as for the isolated gaseous metal ion,
the spin only formula can be applied to find the number of unpaired electrons and gives information
about, for example, oxidation state of the metal and whether the complex is low- or high-spin.

Spin-orbit coupling: Interaction of an electrons orbital angular momentum and its spin angular momentum
leading to a magnetic moment.
• For spin-orbit coupling to occur there MUST be a half-filled or empty orbital similar in energy (and
symmetry) to the orbital holding the unpaired electron(s).
• The nearby symmetry related orbital MUST NOT contain an e- with the same spin as the first electron
(Hund`s rule ).
• Electron in this orbital can make use of this nearby orbital vacancy (by moving into it) to circulate around
the centre of the complex and General orbital momentum (μL) → Additional magnetic moment. .
These conditions are fulfilled when any 1 or 2 of the three T2g orbitals (dxy, , dyz, , dxz) contain an odd
number of electrons. Therefore deviations from Spin-only values occur for: Low spin d5 complex ; i.e.
[Fe(CN)6]-3 ,High spin 3d6 complex i.e. Oct. Fe+2 High spin 3d7 complex i.e. Oct. Co+2; 3d1 complex i.e.

43
Inorganic Spectra Dr. Khalil K. Abid

Oct. Ti +3 ; 3d2 complex i.e. Oct. V+3 . In other 3d complexes the orbital angular momentum contributions
of the electrons cancel each other out and values near spin-only are observed. Details of the Russell–
Saunders coupling scheme to obtain the total angular momentum quantum number. For the d 2 config-
uration, for example, the 3F state in an octahedral field is split into states 3F2 , 3F3 and 3F4 . In a magnetic
field, each state with a different J value splits again; it is the very small energy differences between these
levels with which ESR (also called electron paramagnetic resonance, EPR) spectroscopy is concerned.
The overall splitting pattern for a d2 ion is shown in Figure – 18. The extent to which states of different J
values are populated at ambient temperature depends on how large their separation is compared with the
thermal energy available. It can be shown theoretically that if the separation of energy levels is large, the
magnetic moment is given by the previous equation. Strictly, this applies only to free-ion energy levels.
For d-block metal ions, equation (4) gives results that correlate poorly with experimental data (see table).
For many (but not all) first row metal ions, is very small and the spin and orbital angular momenta of
the electrons operate independently. For this case, the van Vleck formula , equation (4) has been derived;
strictly, equation (7) applies to free ions but, in a complex ion, the crystal field partly or fully quenches the
orbital angular momentum. Data in Table reveal a poor fit between observed values of μeff and those
calculated from equation (7). μeff = 2 √ S(S + 1) + L(L + 1) ……. (7)
If there is no contribution from orbital motion, then equation 7 reduces to equation 4 , which is the
spin-only formula we met earlier. Any ion for which L = 0 (e.g. high-spin d5 Mn+2 or Fe+3 ) should obey
equation 4. However, some other complex ions also obey the spin-only formula. In order for an electron to
have orbital angular momentum, it must be possible to transform the orbital it occupies into an entirely
equivalent and degenerate orbital by rotation.

Fig. 18 Diagrammatic representation of spin and orbital contributions to μeff

44
Inorganic Spectra Dr. Khalil K. Abid

The electron is then effectively rotating about the axis used for the rotation of the orbital. In an
octahedral complex, for example, the three t2g orbitals can be interconverted by rotations through 90o ,
thus, an electron in a t2g orbital has orbital angular momentum. The eg orbitals, having different shapes,
cannot be interconverted and so electrons in eg orbitals never have angular momentum.
There is, however, another factor that needs to be taken into account: if all the t2g orbitals are singly
occupied, an electron in, say, the dxz orbital cannot be transferred into the dxy or dyz orbital because
these already contain an electron having the same spin quantum number as the incoming electron. If all
the t2g orbitals are doubly occupied, electron transfer is also impossible. It follows that in high-spin
octahedral complexes, orbital contributions to the magnetic moment are important only for the
configurations t2g1, t2g2 , t2g4 eg2 and t2g5 eg2 . For tetrahedral complexes, it is similarly shown that the
configurations that give rise to an orbital contribution are e 2t21 , e2t22, e4t24 and e4t25. These results lead us
to the conclusion that an octahedral high-spin d7 complex should have a magnetic moment greater than
the spin-only value but a tetrahedral d7 complex should not. However, the observed values of μeff for
[Co(H2O)6]+2 and [CoCl4]-2 are 5.0 and 4.4 B.M. respectively, i.e. both complexes have magnetic moments
greater than (spin-only). The third factor involved is spin– orbit coupling. Spin–orbit coupling is a
complicated subject and we can give only a brief mention here.
As a result of a mixing of states (which we have so far ignored), theory brings us to equation (8)this
modifies the (spin-only) formula to take into account spin–orbit coupling and, although dependent on
Δoct, applies also to tetrahedral complexes.
αλ
μeff = μ( spin only) ( 1 – ) ……. (8)
Δoct
where λ= spin–orbit coupling constant and α constant that depends on the ground term: α = 4 for
an A ground state, and α = 2 for an E ground state. The simple approach of equation (4) is not
applicable to ions with a T ground state. The values found to be is positive for less than half-filled
shells and negative for shells that are more than half-filled. Octahedral Ni(II) (d8) has a 3A2g ground state.
Equation needed:

μ( spin only) = √ n(n + 2) = √ 8 = 2.83

The value of Δoct = 11500cm-1 and the value of λ = – 315cm-1 , so the equation will be:
4x315
μeff = μ( spin only) ( 1 + ) = 3.14
11500

45
Inorganic Spectra Dr. Khalil K. Abid

The calculated value is significantly larger than (spin-only) as expected for a dn configuration with a
more than half-full shell; it agrees well with the experimental value. An important point is that spin–orbit
coupling is generally large for second and third row d-block metal ions and this leads to large
discrepancies between (spin-only) and observed values of μeff. The d1 complexes cis- [NbBr4(NCMe)2]
and cis – [TaCl4(NCMe)2] illustrate this clearly room temperature values of μeff are 1.27 and 0.45 B. M.
respectively compared with (spin-only) = 1:73 B.M.

Quenching of magnetic moments: The observed magnetic moments in complexes are somewhat less than
the expected values Equation 1 in the absence of spin-orbit coupling or when it is negligible: The reason
for this decrease in the value is that the actual orbital contribution is always less than the expected (ideal)
value. The orbital angular momentum is high in the free metal ion and it is reduced when ligands are
attached to it. When the orbital contribution to the magnetic moment is zero, we say that the orbital
contribution to the magnetic moment is quenched. When the ground state of a complex is either A or E,
the orbital contribution will be quenched. In other words, there is no orbital angular momentum for these
states.
‘A’ state is non-degenerate because it has only one level. Hence, rotation cannot change one orbital
into another equivalent orbital.
‘E’ state is doubly degenerate, which means that there are two energy levels with the same energy. The
orbitals giving rise to this are dx2 – y2 and dz2. One orbital cannot be changed into another by rotation
because their shapes are different. Hence, A and E terms do not have orbital angular momentum. In other
words, orbital angular momentum is quenched in these cases. Thus, complexes having A or E ground
states will have no orbital angular momentum contribution to the magnetic moment and hence, they will
have lower values than expected.
T state :This is a triply degenerate state caused by the t2g orbitals, dxy, dyz, and dzx. They have similar
shape and hence, one orbital can be transformed into another by simple rotation. Thus, if an electron is
present in a dxy orbital, it can occupy the dyz or dzx by simple rotation about the proper axis and the
electron can rotate in an orbit producing magnetic moment. Hence, the orbital angular momentum will not
be quenched and the orbital angular momentum contribution will add to the total magnetic moment of the
complex.

46
Inorganic Spectra Dr. Khalil K. Abid

Configurations Ground-state term Orbital contribution


d1 2T2g Yes
d2 3T1g Yes
d3 4A2g No
d4 5Eg (high-spin) No
3T1g (low-spin) Yes
d5 6A1g (high-spin) No
2T2g (low-spin) Yes
d6 5T2g (high-spin) Yes
1A1g (low-spin) No
d7 4T1g (high-spin) Yes
2Eg (low-spin) No
d8 3A2g No
d9 2Eg No

Magnetic moment & structure: In the case of lanthanide complexes, all types of coupling should be
considered, spin-spin, spin-orbital, and orbital-orbital. Incorporating all these, we have equation 1, which
gives the theoretical magnetic moment for a complex. When the spin-orbit coupling and the orbital angular
momentum are zero, the magnetic moment is given by the equation 9.
μ = 2 √ S ( S + 1) (9)
The equation is known as the spin-only formula to get the magnetic moment. S = n/2, where ‘n’ is the
number of unpaired electrons. Substituting this value of S in the equation 10 we get Equation 11.
μ = 2 √ n/2 ( n/2 + 1) (10)
μ = √ n ( n + 2) (11)
This is known as the spin-only formula. As far as the first row transition metals (3d series) are
considered, spin-only formula works well showing that the orbital angular momentum does not contribute
significantly to the magnetic moment.

47
Inorganic Spectra Dr. Khalil K. Abid

Example 1: Iron(III) complex; It is a d5 system. The high-spin complex, whose electronic configuration
is t2g3eg2, has five unpaired electrons. Using the above equation, we can calculate the magnetic moment
using Equation 11.

μ = √ 5 ( 5 + 2) = 5.92 B. M.

However, the experimental value was found to be in the range 5.70 – 6.00 B.M. In the low-spin complex
the electronic configuration is t2g5eg0. There will be one unpaired electron. That is, n=1. Hence, the
calculated value
μ =√ n(n+2)] = √ 1 ( 1 + 2) = 1.73 BM. The experimental value is in the range 2.0 – 2.5 BM.

Lanthanides: Here, all the three kinds of couplings are significant and hence the spin-only formula will not
give the correct value for magnetic moment. The reason is that the d-orbitals are deeply buried and hence
are not perturbed by the ligands. Therefore, all the three components of coupling contribute appreciably to
the magnetic moment.

Spin-cross over region and effect of temperature: Magnetic measurements will tell us whether the complex
is a high-spin or low-spin complex as explained earlier. Many transition metal ions are able to form high-
spin and low-spin complexes depending up on the strength of the ligand field. When the
ligand is of intermediate field strength, both high-spin and low-spin complexes can coexist in equilibrium.

Example 2 : Fe+2 can form both high-spin, Fe(H2O)6+2, and low-spin, Fe(CN)6 -2, complexes. The high-spin
complex has electronic configuration, t2g4eg2 and has 4 unpaired electrons. Hence, S=4/2 =2 and is
paramagnetic. The electronic configuration of low-spin complex is t2g6eg0. S = 0 and is paramagnetic as
there are no unpaired electrons. In octahedral complexes, the d6 system will be having 5T2g as the ground
state in weak-field cases and 1A1g as the ground state in strong-field cases. This is shown
Near the cross-over point, the difference in energy between the two spin states is very small and hence,
they exist in equilibrium near this point. The actual population of the states depends on the temperature.
For the complexes of iron(II) mentioned above, the high-spin complex is predominant at high temperature
and the low-spin complex is predominant at low temperature.

48
Inorganic Spectra Dr. Khalil K. Abid

The effects of temperature on μeff : So far, we have ignored the effects of temperature on μeff . If a
Complex obeys the Curie Law (equation 12 ), then μeff is independent of temperature; this follows
from a combination of equations 3 and 5.
χ = C / T …….. (12)
where C = Curie constant; T = temperature in Kelvin. However, the Curie Law is rarely obeyed and
so it is essential to state the temperature at which a value of μeff has been measured. For second and
third row d-block metal ions in particular, quoting only a room temperature value of μeff is usually
meaningless; when spin–orbit coupling is large, μeff is highly dependent on T. For a given electronic
configuration, the influence of temperature on μeff can be seen from a Kotani plot of μeff against kT;
where k is the Boltzmann constant, T is the temperature in K, and is the spin– orbit coupling constant.
Remember that is small for first row metal ions, is large for a second row metal ion, and is even larger for
a third row ion. Figure – 19 shows a Kotani plot for a t2g4 configuration; four points are indicated on the
curve and correspond to typical values of μeff (298 K) for complexes of Cr(II) and Mn(III) from the first row,
and Ru(IV) and Os(IV) from the second and third rows respectively . Points to note from these data are:
. since the points corresponding to μeff (298 K) for the first row metal ions lie on the near-horizontal
part of the curve, changing the temperature has little effect on μeff ;
. since the points relating to μeff (298 K) for the heavier metal ions lie on parts of the curve with
steep gradients, μeff is sensitive to changes in temperature; this is especially true for Os(IV).

Spin crossover: The choice between a low- and high-spin configuration for d4, d5, d6 and d7 complexes is
not always unique and a spin crossover sometimes occurs; this may be initiated by a change in pressure
(e.g. a low- to high-spin crossover for [Fe(CN)5NH3]-3 at high pressure or temperature (e.g. octahedral
[Fe(phen)2(NCS – N)2] , octahedral [Fe(L1)2] and the square-based pyramidal complex L2 undergo low- to
high-spin crossovers at 175, 391 and 180K respectively). The change in the value of eff which
accompanies the spin crossover may be gradual or abrupt. (Figure – 19) .

49
Inorganic Spectra Dr. Khalil K. Abid

Fig. 19 – i Kotani plot for a t2g4 configuration; is the spin– orbit coupling constant. values of μeff(298 K) for Cr(II), Mn(III),
Ru(IV) and Os(IV) are indicated on the curve
ii – The dependence of the observed values of μeff on temperature for (a) [Fe(phen) 2(NCS – N)2] where low- to high-spin
crossover occurs abruptly at 175 K, and (b) [Fe(ptz)2(NCS – N)2] where low- to high-spin crossover occurs more gradually.

50
Inorganic Spectra Dr. Khalil K. Abid

X – Ray & ESR Spectroscopy

X–ray sources can be used to study scattering and diffraction phenomena of matter: X–ray Diffraction
(XRD), small angle X–ray scattering (SAXS) and wide angle X–ray scattering (WAXS). Especially XRD is a
conventional lab technique for the study of crystalline solids. The most common X–ray spectroscopy is the
absorption spectroscopy, X–ray Absorption Spectroscopy (XAS).

XAS: Theory: X-ray Absorption Spectroscopy, is a broadly used method to investigate atomic local structure
as well as electronic states. An X-ray strikes an atom and excites a core electron that can either be promoted
to an unoccupied level, or ejected from the atom. Both of these processes will create a core hole. If the
electron dissociates, this produces an excited ion as well as photoelectron and is studied by X-ray
Photoelectron Spectroscopy (XPS). The electrons that are excited are typically from the 1s or 2p shell, so the
energies are on the order of thousands of electron volts. XAS therefore requires high-energy X-ray excitation,
which occurs at synchrotron facilities. X-ray energy is about 104 eV (where "soft x-rays" are between 100 eV
– 3 keV and "hard x-rays" are above 3 keV) corresponding to wavelengths around 1 Angstrom. This
wavelength is on the same order of magnitude as atom-atom separation in molecular structures, so XAS is a
useful tool to deduce local structure of atoms. XAS is also utilized in analyzing materials based on their
characteristic X-ray absorption "fingerprints It is possible to deduce local atomic environments of each
separate type of atom in a compound. XAS is particularly convenient because it is a non-destructive method
to examine samples directly. Structures can be determined from samples that are both heterogeneous and
amorphous. When an X-ray strikes an atom, one of the core electrons is either excited to a higher energy
unoccupied state (a transition studied by XAS) or into an unbound state, called the continuum. When
electrons are ejected from an atom of a solid material, this is essentially the photoelectric effect and is
studied by X-ray photoelectron spectroscopy.

51
Inorganic Spectra Dr. Khalil K. Abid

Below is a diagram illustrating an atom absorbing an x-ray with the resultant ejection of a core electron
into the continuum. The level K refers to the n=1 level, L refers to the n=2 level, and M refers to the n=3 level.

The wave vector of a photoelectron Determining the wave vector of the electron dictates which subset of
XAS will be used ( EXAFS, NEXAFS, etc.). The kinetic energy Ek and threshold energy E0 (amount of energy
the photoelectron needed to promote the electron into the continuum) of the photoelectron can be
addressed, after considering the de Broglie equation: λ=h p where
p is the momentum of the electron, λ is the wavelength, and h is Planck’s constant.
Beer’s Law and the relation to XAS : Beer's Law can be applied to XAS for an x-ray beam that is both narrow
and monochromatic, striking the absorber at a 90 degree angle. The absorber has a uniform composition
and thickness. The important result of this derivation will be an expression describing the absorption
coefficient, ?M, of X-ray absorption.

Beer’s law can be written

where m and d refer to the mass and thickness of the sample, respectively. The terms k and kd are proportionality
constants whose units are dependent on the units of m or d, the wavelength, and also the elements comprising the
sample. The previous equation can be written in terms of natural logarithms, in which case

where the symbol ?M has been introduced, and is termed the


mass absorption coefficient. Units of μM are cm2/g, and amounts are not dependent on the physical or chemical state
of the element as X-ray absorption is primarily an atomic property.

52
Inorganic Spectra Dr. Khalil K. Abid

X‐Ray Crystallography: X-ray crystallography is a tool used for identifying the atomic and molecular
structure of a crystal, in which the crystalline atoms cause a beam of incident X – ray to diffract into many
specific directions. By measuring the angles and intensities of these diffracted beams, a crystallographer can
produce a three-dimensional picture of the density of electrons within the crystal. From this electron density,
the mean positions of the atoms in the crystal can be determined, as well as their chemical bonds, their
disorder and various other information.
Since many materials can form crystals—such as salts , metals, minerals, semiconductors, as well as
various inorganic, organic and biological molecules—X-ray crystallography has been fundamental in the
development of many scientific fields. In its first decades of use, this method determined the size of atoms,
the lengths and types of chemical bonds, and the atomic-scale differences among various materials,
especially minerals and alloys. The method also revealed the structure and function of many biological
molecules, including vitamins, drugs, proteins and nucleic acids such as DNA. X-ray crystallography is still
the chief method for characterizing the atomic structure of new materials and in discerning materials that
appear similar by other experiments. X-ray crystals structures can also account for unusual electronic or
elastic properties of a material, shed light on chemical interactions and processes, or serve as the basis for
designing pharmaceuticals against diseases.
– It is a type of analysis in which atomic positions are revealed through the scattering of x‐rays that
occurs because of electron clouds surrounding the atoms’ tendency to diffract x‐rays.
– It is useful for small molecule structures (up to only 100 atoms). And also especially for structural biology
that include: Nucleic acid structure, protein structure, enzyme/drug interactions
In a single-crystal X-ray diffraction measurement, a crystal is mounted on a goniometry . The goniometer is
used to position the crystal at selected orientations. The crystal is bombarded with a finely focused mono
chromatic beam of X-rays, producing a diffraction pattern of regularly spaced spots known as reflections.
The two-dimensional images taken at different rotations are converted into a three-dimensional model of the
density of electrons within the crystal using the mathematical method of Fourier transforms, combined with
chemical data known for the sample. Poor resolution (fuzziness) or even errors may result if the crystals are
too small, or not uniform enough in their internal makeup. X-ray crystallography is related to several other
methods for determining atomic structures. Similar diffraction patterns can be produced by scattering
electrons or neutrons, which are likewise interpreted as a Fourier transform. If single crystals of sufficient
size cannot be obtained, various other X-ray methods can be applied to obtain less detailed information;

53
Inorganic Spectra Dr. Khalil K. Abid

such methods include fiber diffraction, powder diffraction and small - angle e X- ray scattering(SAXS). If the
material under investigation is only available in the form of nano crystalline powders or suffers from poor
crystalline, the methods of electron crystallography can be applied for determining the atomic structure. For
all above mentioned X-ray diffraction methods, the scattering is elastic; the scattered X-rays have the same
wavelength as the incoming X-ray. By contrast, inelastic X-ray scattering methods are useful in studying
excitations of the sample, rather than the distribution of its atoms.
The Requirements : The molecule being observed must have a crystalline structure.To be eligible for X ‐Ray
Crystallography, Crystals must:
– Not have any cracks or flaws and be small enough to allow for photons from x ‐rays to interact with atoms
(less than 1mm)
– The crystal must be mounted ( secured) to prevent vibration of atoms which will skew reflections!
– It’s also useful to have a general idea of the structure of the crystal, X‐ray crystallography is useful for the
details of atom positioning in Space.
What is considered a ‘crystal’? – A regularly spaced array of atoms

Crystalline vs. Non‐crystalline X-ray crystallography can locate every


Structure of same molecule atom in a zeolite an aluminosilicate

Low Molecular Weight Molecules are easy to crystalize, but high molecular weight ones (Such as
Biological molecules) take a lot longer to make. The arrangement of the atoms within the crystal is what
determines the specific diffraction pattern. The dots projected can be light or dark depending on the amount
of photon intensity (how many photons per cubic area). Photon Intensity Depends on Wave Interactions (look
below). These varied photon intensities then help to find structure of molecule who’s atoms’ electron clouds
are diffracting the photons. For diffraction to occur it is required that photons being shot out towards
crystals have a wavelength that is approximately equal to the distance between atoms in the crystal.

54
Inorganic Spectra Dr. Khalil K. Abid

Wave interactions that affect photon intensity Each dot=reflection


Compilation of Reflections =Diffraction Pattern
Wave Interactions:
Constructive: Troughs and crests align=Amplitudes of waves add together
Destructive: Troughs and crests are not aligned and their amplitudes cancel each other out
Partial Interference:(most common) only slightly unaligned leads to complex pattern of high and low
amplitudes. Diffraction pattern = Electron Density Map = Solve Electron Density map = Structure

– Well once we collect one diffraction pattern, we must proceed to alter the angle at which Xrays hit the
crystal to get yet another diffraction pattern. You continue to alter the angles until you have a 360 degree
“view” of the crystal structure in terms of a ton of separate diffraction patterns.. Putting together all of these
diffraction patterns, then allows you to create 3D diffraction pattern (which is translated into an Electron
Density Map) to then discover structure. The higher the atomic number the larger the electron cloud!
**This is why Hydrogen’s (Very small atomic number and electron cloud) are hard to find in maps. From this
point you just look at the Electron Density Map and knowing a general idea of what general structure of the
crystal, you can assign each atom a specific element name according to amount of electron density. .. You
will get something like this for each cluster of ring groups. Next, just play “connect the dots” in the sense
that no bond rules are broken. You will end up with a crystal structure once you put everything together.

55
Inorganic Spectra Dr. Khalil K. Abid

Applications of X Ray Crystallography Knowledge:


– It will allow you to distinguish which is a proper reaction product structure (cis or trans)since crystals can
have same molecular formula, functional groups, and splitting patterns making NMR, IR and MS useless
– For Biological molecules it allows for the creation of drugs that utilize enzyme/substrate complex
functioning to alter behaviors and biological.

Synchrotron radiation: Synchrotron radiation takes its name from a specific type of particle accelerator, but
has become a general name for radiation from charged particles traveling at relativistic speeds in applied
magnetic fields. These magnetic fields force them to travel along curved paths. Synchrotron radiation is
produced in storage rings where electrons or positrons are kept circulating at constant energy. It radiation is
produced either in the bending magnets or in insertion devices such as wigglers or undulators in the straight
sections of the storage ring. These insertion devices force electrons to follow oscillating paths due to
alternating magnetic fields.
The main principle of the occurrence of the radiation is that the electrons are forced to move in circular
paths. At each bend induced by the magnetic field the electrons change their velocity and therefore have a
short period of acceleration during which these electrons radiate.

Developments in X – ray spectroscopies


Sample conditions: For studies in the hard X–ray regime, in–situ is quite easily applied, even reaching more
‘real–life’ conditions, as with hard X–rays it is not hard to measure solid, liquid and / or gaseous samples, but
for studies in the soft X–ray regime, the soft X–rays might lose a lot of intensity before reaching the actual
sample if in–situ measurements are performed. However, improvements in measurements in the soft X–ray
regime are being developed. For example, to measure in–situ soft X–ray (absorption and photoemission)
edges can be done by the construction of dedicated beam lines that allow for a mbar pressure in the
measurement chamber. Experiments under mbar pressure have given new insights in catalysis and surface
science, but also (mainly) in–situ XPS has given more information on interactions between water and
surfaces

Space–resolved: XAS can be combined with a number of X–ray microscopy techniques: transmission X–ray
microscopy (TXM) which exist in scanning (STXM) and full–field modes, fluorescence–yield X–ray
microscopy and X–ray induced photoemission electron microscopy (X–PEEM), which measures the electron
yield with a position–sensitive electron analyzer. In this subsection we will discuss all these microscopy
techniques in combination with some recent studies. X–PEEM in the soft X–ray regime leads to electron

56
Inorganic Spectra Dr. Khalil K. Abid

energies which are typically less than 10 eV, and therefore a small escape depth of about 2–5 nm making X-
PEEM an attractive surface–sensitive technique. The spatial resolution is typically limited to about 20 nm due
to spherical aberrations. X–PEEM is mainly used as a tool for studies of magnetic materials, since X–PEEM is
quite sensitive to magnetic fields. However, the technique can be used for all kinds of surfaces and interfaces
depending on the incident X–ray energy, but the current spatial resolution is still much better using another
X–ray microscopy discussed below (below 20 nm in the soft X–ray STXM), although the first results were
published on a double aberration corrected low–energy electron microscope (LEEM) with about 2 nm
resolution. Note that X–PEEM only works in high vacuum, thus samples cannot be tested under in–situ
conditions. It have shown the possibilities of TXM in the full–field mode for imaging of biomaterials such as
bone structures and cells and noted that besides the absorption contrast, their setup at the Stanford
Synchrotron Radiation Light source (SSRL) is capable of measuring resolutions of about 30–40 nm.
Time–resolved: Another trend in synchrotron–based X–ray science is the study of transitions in real time.
The first time–resolved XAS was carried out on biological systems and currently it is a major desire to study
biological matter with time–resolved techniques. The picosecond and femtosecond regime is quite routinely
applied for studies of atomic ions in solution. For example, a platinum photo-catalyst was studied with time–
resolved XAS in combination with FEFF calculations to study the photo–induced changes. Also femtosecond
time–resolved photo–electron events can be routinely studied. Attosecond pulses can only be realized in the
extreme ultraviolet and the X–ray regime and with attosecond pulses one could study the dynamics of
ultrafast electron transfer. An especially important development in time–resolved X–ray spectroscopy studies
is the construction of X–ray free electron lasers.

Energy–resolved: In the hard X–ray regime, the energy resolution is worse than in the soft X–ray regime and
therefore hard X–ray XAS may give less electronic structure information. A way to circumvent the loss of
information in hard X–ray XAS is to improve the resolution by doing High–Energy–Resolution–Fluorescence–
Detection(HERFD)–XAS. HERFD–XAS is also sometimes called lifetime–removed–XAS or lifetime–
suppressed–XAS. The XAS is then measured at constant emission energy. Fluorescence detected XAS was
originally developed in order to be able to measure dilute samples that do not absorb enough photons to
obtain a XAS spectrum in transmission mode. XAS in fluorescence mode is either detected without energy
resolution (e.g., with a photodiode) or with energy resolution (e.g., using a Ge detector). The better the energy
resolution, i.e., the smaller the bandwidth around the fluorescence line, the better will be the signal to back
ground ratio. The limit for improvement is set by the core hole lifetime.

57
Inorganic Spectra Dr. Khalil K. Abid

TGA & DSC Spectroscopy of inorganic compounds

Thermal Analysis; A class of techniques to measure chemical or physical properties of a substance as


function of temperature or time.

Thermo Gravimetric Analysis TGA:


I.Process: weight changes of a specimen as a function of temperature
II.General instrumentation
III.Output: plot of mass versus temperature
Calibration: Temperature calibration: use of a reference with a known Curie point Mass calibration: use of a
reference with exact known mass.

Applications of thermogravimetry:
1. Analysis of mixtures: Each part of a mixture behaves differently with temperature change
2. Oxidation studies: Study of the oxidation of metals at a constant temperature
3. Reduction studies: Study of the reduction of a solution/suspension of metal(s)
4. Exact chemical identification: Different TGA behaviours of identical samples with different history

CaC2O4.H2O → CaC2O4 + H2O T ~ 100 C


CaC2O4 → CaCO3 + CO T ~ 500 C
CaCO3 → CaO + CO2 T ~ 800 C

Analysis of mixtures

58
Inorganic Spectra Dr. Khalil K. Abid

Differential Scanning Calorimetry DSC: Difference in heat flow to a sample and to a reference is monitored
against time/temperature:
1. Power-compensated DSC:
– Separate heating of sample and reference
– Identical temperature difference,
– Measurement of electrical power
2. Heat flux DSC:
– In-concert heating of sample and reference,
– Measurement of temperature change
Theory of DSC
Heat :Energy transfer between two bodies due to thermal communication.
Calorimetry: Measuring the heat of physical/chemical interactions.
Heat capacity: The amount of heat needed to increase temperature of a substance.

Calibration:
1.Baseline: correction for the baseline slope of a blank run (no pan).
2.Heat flux and temperature: melting a reference (Indium) of known heat flux.
3.Heat capacity: use of a reference of known heat capacity (i.e. sapphire).

59
Inorganic Spectra Dr. Khalil K. Abid

Differential thermal analysis: Recording any temperature difference between sample and reference being
subjected to identical thermal cycles

Applications
DSC DTA
1. Glass transition temperature 1. Heat of reaction
2. Crystallinity and crystallization rate 2. Corrosion
3. Reaction kinetics 3. Identification of polymers
4. Heat of reactions 4. Ceramic and metalsindustry
5. Corrosion 5. Structural studies: decomposition temperature phase
transitions; melting-boiling crystallization point; thermal stability

60
Inorganic Spectra Dr. Khalil K. Abid

Thermal analysis is one of the most useful methods of analysis in collecting both physical and chemical
information. Probably the most used technique of thermal analysis is thermogravimetric analysis (TGA), it is
used in all types of applications, providing information about the bonding of components within the sample.
TGA can become an even stronger analytical technique when coupled with other thermal analysis techniques
such as differential scanning calorimetry (DSC) and spectroscopic techniques such as Fourier transform
infrared spectroscopy (FTIR) and mass spectrometry (MS). TGA measures the absolute amount and rate of
change in weight of a sample as either functions of time or temperature in a controlled environment.
TGA has a wide range of properties that can be measured such as thermal stability, oxidative stability,
effects of different atmospheres, moisture and volatile content, and sometimes the composition of multi-
component systems. TGA determines if and how different components within a material are bonded
differently. When TGA is coupled with DSC or differential thermal analysis (DTA), the mode of analysis is
called simultaneous DSC-TGA (or DTA) (SDT). SDT measures the amount and rate of change in weight, but
also measures the heat flow of the sample as a conventional DSC does. SDT measures the same properties
as TGA, but extends the list to include heats of reactions, melts, and boiling points. The three most important
signals that TGA collects while it is analyzing the sample are weight, rate of weight change—differential
thermogravimetry, and temperature. A differential thermogravimetry curve—DTG—is generated as the first
derivative of the weight with respect to temperature or time. The DTG curve can be used to provide both
qualitative and quantitative information about the sample. Qualitative modes of analysis include
fingerprinting a material and distinguishing between two or more overlapping reactions. Quantitative modes
include peak height and temperature at maximum weight loss measurements. The most important aspect of
TGA operation is the validity of measurements made. Confidence in data collection can be achieved through
regular calibration. For TGA, both mass and temperature calibrations must be performed. Most instrument
and software packages possess a relatively automated mass calibration procedure in which the user places
certified calibration weight onto the instruments sample platform.

Schematic principle of TGA measurement

61
Inorganic Spectra Dr. Khalil K. Abid

For example :
The TGA curve of two differently cured epoxy resins is shown in Figure and can be used to get first
information about the thermal stability of those systems.

Limitations :
TGA • Only provides meaningful data when a change in mass occurs.
• Some liquids can be measured, but this is generally very difficult to do.
• Very small samples are used, so non-homogeneous materials generally cannot be tested
DSC • Very sensitive to any change in the sample or crucible.
• Requires very good thermal contact with bottom of sample crucible
• Very sensitive to heating rate.
General Description: Thermogravimetric analysis (TGA) is an analytical technique used to determine a
material’s thermal stability and its fraction of volatile components by monitoring the weight change that
occurs as a specimen is heated. The measurement is normally carried out in air or in an inert atmosphere,
such as Helium or Argon, and the weight is recorded as a function of increasing temperature. Sometimes, the
measurement is performed in a lean oxygen atmosphere (1 to 5% O2 in N2 or He) to slow down oxidation. In
addition to weight changes, some instruments also record the temperature difference between the specimen
and one or more reference pans (differential thermal analysis, or DTA) or the heat flow into the specimen pan
compared to that of the reference pan (differential scanning calorimetry, or DSC). The latter can be used to
monitor the energy released or absorbed via chemical reactions during the heating process. In the particular
case of carbon nanotubes, the weight change in an air atmosphere is typically a superposition of the weight
loss due to oxidation of carbon into gaseous carbon dioxide and the weight gain due to oxidation of residual
metal catalyst into solid oxides.

62
Inorganic Spectra Dr. Khalil K. Abid

63
Inorganic Spectra Dr. Khalil K. Abid

DSC: The DSC can be used to obtain the thermal critical points like melting point, enthalpy specific heat or
glass transition temperature of substances. The sample and an empty reference crucible is heated at
constant heat flow. A difference of the temperature of both crucibles is caused by the thermal critical points
of the sample and can be detected.

Schematic principle of DSC measurement

For example: DSC curve of Poly(ethylenetherephthalate) (PET) is shown in Figure The glass transition
temperature is characterized by a dip in the curve. The point of recrystallization and the melting point result
in exothermal and endothermal peaks respectively.

DSC curve of PET TGA/DSC

A small lab jack can be used to adjust the magnet’s distance from the sample such that a 2-3% weight
gain or weight loss occurs once the magnet is positions above or below the sample. Figure 3 shows the
Curie point determination for nickel and alumenum. Note that the Curie point is denoted as the offset.

64
Inorganic Spectra Dr. Khalil K. Abid

SDT also has an alternate method for temperature calibration. The melting points of standard materials can
be determined by the onset of the endotherms and compared to the theoretical melt temperature. A good
exercise for TGA and SDT is to perform multiple analyses of calcium oxalate monohydrate. By performing
such an analysis the performance and precision of both you and the instrument can be measured. An overlay
of five calcium oxalate experiments is shown in Figure .

Performance testing using calcium oxalate monohydrate

Although calcium oxalate monohydrate is not typically a standard material, it does hold good utility in
intra-laboratory analysis. The weight change and peak temperature can be inputted into a spreadsheet
program to check your instrument and operators performance. The accuracy of the instrument can be used
to assess your instruments long-time performance, and help single out a damaged component of the
instrument. The baseline can also be quite usual in quantifying your instrument’s performance and
sensitivity. Small weight losses become increasingly difficult to measure if the instrument’s baseline is large
compared to that of the instrument. TGA is the foremost analysis technique in determining quantitative
properties of the original sample. A polyethylene (PE) sample filled with CaCO 3 was analyzed as shown the
following figure.

65
Inorganic Spectra Dr. Khalil K. Abid

TGA curve of polyethylene sample filled with calcium carbonate

By knowing the degradation reaction of CaCO3, the initial percentage of CaCO3 in the PE can be
calculated. At approximately 550oC the PE is completely decomposed; thus, the weight loss occurring at
approximately 650oC is due to the decomposition of CaCO3. The weight loss is a direct result of the evolution
of CO2 gas. The residue is the remaining CaO that fails to decompose. From the weight change and the
residue, the stoichiometric relationships can be used to determine a percentage of CaCO 3 that exists in the
original PE sample. Calculating the initial percentage of CaCO3 from the weight change is more accurate than
calculating it from the residue. Most polymers contain fillers; hence, the residue is a combination of CaO and
these fillers making this calculation less accurate.
TGA and SDT can also be used to demonstrate the important of reaction atmosphere. Calcium oxalate
monohydrate was analyzed under the same experimental conditions except the purge gas. The sample was
analyzed in air, CO2, and nitrogen of equal flow rates. Hi-Resolution TGA is useful to separate overlapping
weight losses. Hi-Resolution TGA exposes the sample to an isotherm once a weight loss is detected. The
isotherm allows the weight loss occurring at the lower temperature to complete before the second weight
loss begins.

66

Potrebbero piacerti anche