Anda di halaman 1dari 40

EPR NOTES

BY
PROF. P. SAMBASIVA RAO
DEPARTMENT OF
CHEMISTRY
PONDICHERRY UNIVERSITRY
PONDICHERRY – 605 014

INTRODUCTION
Electron Paramagnetic Resonance spectroscopy has been matured into a powerful,

versatile, non-destructive and non-intrusive analytical method. Its history began in the year of

1945 , when discovered by E.Zavoisky [1]. EPR spectroscopy is also called Electron Magnetic

Resonance spectroscopy (EMR) or Electron Spin Resonance spectroscopy (ESR). The basic

principles of EPR have been well developed and treated by several authors [2-13]. The essential

condition required to study EPR is presence of at least one stable unpaired electron. These

electrons can be free radicals, transition metal complexes with unpaired electron in their d or f

orbital. They can be generated either by chemically, electrochemically or photo chemically.


This method is an ideal complementary technique used in a wide range of studies, i.e., in

chemistry, physics, biology, medicine, etc. This technique is also used for dating of fossils [14],

and in tool imaging in recent years.

Some of the information available from EPR study is:

 Spin state of paramagnetic centers.

 Magnitude of hyperfine interaction.

 Zero-field splitting of S1/2 state.

 Possible identity of paramagnetic centers (metals in metal clusters, free radicals).

 Possible identity of legating atoms.

 Changes in the oxidation state of the metal.

Theory

Electrons in molecular systems, by virtue of their spin and orbital motion, have spin

magnetic moment s and orbital magnetic moment l given by

µs = -(2.0023 eh/ 4me) S ……... (1)

l = - ( eh / 4me) L .…….. (2)

Here, S is spin angular momentum, L is orbital angular momentum; e is the magnitude of the

electronic charge, h is the Planck’s constant, me is the mass of the electron. The factor 2.0023,

known as free-electron gyromagnetic ratio, arises out of the ‘anomalous’ Zeeman effect and can

be derived from Dirac's relativistic wave mechanics and includes a correction of +0.0023 due to

relativistic mass variation [15]. The total magnetic moment (e), therefore, can be expressed as

e = S + L = (eh / 4me) [2.0023 S + L] ...……(3)

Here e = eh/4me is known as the Bohr magnetron. For a free electron, the total magnetic

moment reduces to
e = (2.0023 eh/ 4me) S ……… (4)

For a free electron system, the spin momenta are quantized along the applied

field axis with Z component SZ equal to ±1/2, i.e., the time independent components of the

précessing spin vector lie either parallel or antiparallel to the direction of the applied magnetic

field B0. Thus only two energy states are possible for a free electron about the unperturbed

energy level at ± BoβSZ. When the system is subjected to an oscillating electromagnetic field, the

magnetic moment vector precesses about the direction of the field with an angular frequency ω,

known as Larmor frequency and is given by

ω = (2.0023/4me)ehBo ………(5)

The Hamiltonian represents the interaction of the magnetic moment with the applied field

Bo and is given by

Η = e·Bo ……… (6)

H = (2.0023 / 4me)eh S·Bo ...……(7)

Here S is the effective spin including the orbital contribution and is also known as

‘fictitious spin’. A system designating with the fictitious spin of S’ splits into (2S’+1) levels and

transitions are allowed between these according to the selection rules. The only difference

between the true spin and the fictitious spin is that the latter defines the effective spin angular

momentum endowed due to any orbital contribution. A small rotating magnetic field B 1 rotating

in the same sense as the net magnetization vector causes exchange of energy between itself and

the system of spins causing transitions to occur between the Zeeman states, i.e., at resonance W B1

= |e|Bo/2me. Experimentally microwave field replaces the rotating field B1. This alternating field

can be shown to be equivalent to the rotating field as far as EPR and NMR are concerned. The

requirement that the oscillating field should be applied perpendicular to the static external field
can be proved using quantum mechanics and this will in turn leads to selection rules for magnetic

dipole transitions.

According to the well-known result of the time dependent perturbation theory [16], the

time independent transition probability is proportional to the square of the absolute value of the

transition moment P is given by

P =  i| V |f 2 ……… (8)

Here i| and |f are the initial and final state vectors and V is the time independent part of the

perturbation that connects these states. In EPR, i and f corresponds to MS values. If B1 =

2B1cosωt and V = g B1.S, then

P =  i | V |f 2 = <−1/2 |g B1.S| +1/2>2 ……… (9)

P = g[ 1/2| BzSz|+1/2  +  1/2| BxSx|+1/2 +  1/2| BySy|+1/2]2 …......(10)

Since first term in square brackets is zero, B 1 parallel to axis of quantization will be

ineffective. Also when B1 is perpendicular to the axis of quantization, the operators S x and Sy

lead to

f = i ±1 ...…… (11)

And hence the selection rule ΔΜS = ±1. However, in triplet spin ground states, it can be shown

that transitions can occur at low fields with B1 parallel to the axis of quantization. In general,

unpaired electrons in transition metal complexes do possess orbital angular momentum. The spin

and orbital angular momenta may couple to give a resultant momenta J, with (2J+1) degeneracy.

The J values themselves range form |L+S| to |L−S|. The corresponding g-factor, known as the

Lande’s splitting factor [17] is given by g=

J (J+1) – L (L1) + S(S+1)

1+ …….. (12)
2J (J+1)

J = |L−S| for less than half filled d-shell and J = |L+S| for more than half filled d-shell.

The resonance condition is realized when a circularly polarized electromagnetic radiation

having proper sense of precession with the same frequency as that of the precessing spin is

applied. If this frequency of radiation matches with the energy difference between the two states,

then the resonance condition is

hν = gβeBo ………(13)

This is represented schematically in Fig.1. The resonance condition can be satisfied either

keeping the field constant or varying the frequency or vice versa. In general, EPR experiments

are carried out at a fixed frequency. The two common frequencies are (i) X-band frequency

(range about 9-9.5 GHz), the field strength is about 330 mT

(ii) Q-band frequency (range about 35 GHz), the field strength is about 1350 mT.
Fig.1 Representation of the resonance condition in the frequency mode

In common, X-band is used. This frequency corresponds to 3 cm microwave region and most of

the EPR work is done in this region.

In the presence of applied magnetic field, the Boltzmann distribution is established in

such a way that the population of the upper level, i.e., ΜS = +1/2 is lower than the ΜS = −1/2. The

ratio of population is

N+/N- = exp(-gβeBo/kT) ......... (14)

Here N+ and N− are populations of ΜS = +1/2 and ΜS = −1/2 spin levels respectively, k =

Boltzmann constant and T = absolute temperature in Kelvin. At room temperature hν < kT


condition is satisfied for spin systems leading to a population difference between the two levels

of about sixteen in one million spins.

The effectiveness of the coupling of the spin and orbital motion, however, depends of the

exact nature of the environment of the paramagnetic ion. Surrounding ionic charges, bonded

ligands etc., produce a strong electrostatic field and the spin-orbit coupling will breakdown due

to the ‘quenching’ of the orbital angular momentum. Thus EPR systems can be divided into three

main categories, i.e., weak-field, intermediate field and strong field cases.

EPR is sensitive to the changes in symmetry of the environment. Apart from this, a

number of interactions shift and split the energy levels of the unpaired electron. The former may

cause a shift in the center of gravity of the spectrum and the latter influences the fine structure. In

favorable circumstances, a detailed analysis of the EPR spectrum leads to a very accurate

determination of bonding parameters and the electronic structure of the ground state, not easily

possible in many other types of spectroscopy. A general consideration of the interaction involved

in the case of a paramagnetic species (with particular emphasis on transition metal ions) in a

crystal field is formulated in terms of generalized Hamiltonian. A consequence of this, involving

only spin-operators, is the spin-Hamiltonian, which is an artificial, though practical concept is

described subsequently.

The spin Hamiltonian consists of various terms, which have been arisen due to different

types of interactions between electron spin with either the applied magnetic field or nuclear spin

or another electron spin and the nuclear spin either with the applied magnetic field or another

nuclear spin. Below, the origin and importance of these interactions are mentioned.

Spin Hamiltonian
Most EPR data can be described in terms of “Spin-Hamiltonian’ involving smaller

number of terms by use of an effective fictitious spin without a detailed knowledge of spin-orbit

coupling, the magnitude of crystal field splitting etc.

Spin angular momentum,

S = [S (S+1)] 1/2 h/2 ……… (15)

Spin magnetic moment,

µ = −g βe S ……… (16)

Energy in a magnetic field

E=−µB ……… (17)

Replacing µ by the corresponding operator βeg.S, the corresponding energy operator is

HS = gβeSB ……… (18)

Since this operates only on the spin part of the wave function, it is called ‘spin Hamiltonian’.

This Hamiltonian gives the electron Zeeman interaction energy only and so, it is otherwise called

Zeeman interaction Hamiltonian. Basically, it represents all possible interactions that affect the

spin system and give rise to its energy states. For the most general hypothetical case with

electronic spin S>1/2, and nuclear spin I≥1, the effective spin Hamiltonian is made up of the

following terms;

HS = Hez + Hss + Hhfs + Hq - Hnz ……… (19)

where, Hez = Electron Zeeman interaction

Hss = Electron Spin-Spin interaction

Hhfs = Electron Spin-Nuclear Spin Hyperfine interaction (I≠0)

Hnz = Nuclear Zeeman interaction

Hq = Nuclear Quadrupole interaction


In general, the hyperfine terms, which are almost always much smaller than the leading

term Hez splits the energy levels into (2I+1) sublevels, leading to (2I+1) lines in the EPR

Spectrum.

All the interactions mentioned above could be written in terms of the interaction matrices

involving S and I in the following form:

HS = gβeSB + IAS + SDS + IQI − gNBβNI ……… (20)

Here, βN and gN are the nuclear magneton and nuclear g-factor respectively.

D = Dipolar interaction matrix, or zero field term

A = Hyperfine interaction matrix.

Q = Nuclear quadrupole interaction matrix

The concept of fictitious spin will be discussed in a more detailed way. Just like in the

case of a quantum state described by J splits it into (2J+1) levels in an external field, a system

designated with the fictitious spin ‘S’ splits into (2S+1) levels and transitions are allowed

between these according to the selection rules. The only difference between the true spin S and

fictitious S' is that the latter defines the effective spin-angular momentum endowed due to any

orbital contribution. Whereas the spin-only g-factor would deviate from this due to admixture of

higher lying states into the ground state via spin-orbit coupling and hence effectively takes into

account of the effect of crystal field terms and spin-orbit term of the generalized Hamiltonian.

Since the isotropic g-factor of a free electron is modified into a tensor when orbital momentum is

not completely quenched in the principal axis system of the g-tensor, the Hamiltonian is written

as,

H =  (BxgxxSx + BygyySy + BzgzzSz) ...…… (21)

If the tensor is cylindrically symmetric, gxx = gyy = gzz = g and Bx = By = Bz, then
H = [BzgS] ..……. (22)

And if the tensor is axially symmetric,

H = [Bzg//Sz + g(BxSx + BySy)] ...…... (23)

Likewise, the hyperfine terms which consist of a dipolar part and isotropic part is written

generally as:

H = Axx Ix Sx + Ayy Iy Sy +Azz Iz Sz …….. (24)

Isotropic hyperfine interaction arises due to (a) direct unpaired spin-density in an s-orbital

or in a molecular orbital (M.O.) with s-orbital contribution, (b) spin-polarization, due to isotropic

hyperfine coupling in an isolated paramagnetic atom or ion, where the electron is in a p or d

orbital arises via polarization of core-s-electrons, (c) also configuration interaction between a

ground state M.O. orbital with no s-orbital contribution and states with finite s-orbital

contribution. In systems with more than one unpaired electron, the spin-degeneracy is removed

even in the absence of external magnetic field by second spin-orbital coupling known as the

zero-field interaction.

Zeeman Interaction: HZee

This arises due to the interaction of the external magnetic field with the spin and orbital

magnetic moment of the electrons and nuclear spin magnetic moment and is given by

HZee =  (L + 2S) B – gn n B ∑I Ii. ……… (25)

Where n is the nuclear magneton and Ii is the various magnetic nuclear spins. The first term is

known as the electron Zeeman interaction. The second term, which is the nuclear Zeeman

interaction, causes to a high degree of approximation, merely a constant shift of all the energy
levels and is of minor importance in EPR, except in cases where the hyperfine coupling is much

smaller than the nuclear Zeeman interaction.

Hyperfine interaction:Hhyf

The magnetic moments of electrons and nuclei are coupled via the so-called contact

interaction. This interaction first introduced by Fermi to account for hyperfine structure in

atomic spectra. It represents the energy of the nuclear moment in the magnetic field produced at

the nucleus by electric currents, which are associated with the spinning electron. It has the form

H = a I·S = a (IXSX+IYSY+IZSZ) ……… (26)

The coupling constant 'a' is proportional to the squared amplitude of the electronic wave function

at the nucleus. If 'a' has the dimensions of energy,

a = (8/3)gβgNβN|ψ(0)|2 ……… (27)

Here |ψ(0)|2 is the squared amplitude of the unpaired electron density at the nucleus. Isotropic

hyperfine interactions arises due to,

(a) Direct unpaired spin density in an s orbital or in a molecular orbital with s orbital

interaction.

(b) Spin polarization: isotropic hyperfine coupling in an isolated paramagnetic atom or

ion, where the electron is in a p or d type orbital arises via polarization of core s-

electrons.

(c) Configuration interaction between a ground state molecular orbital with no s-orbital

contribution with states having finite s-orbital spin density.

Dipolar coupling arises out of point dipole-dipole interaction between p or d orbitals with

the nucleus and follows a (3cos2 1) variation given by

Hl = gβgNβN(3cos2 1) 1/<r3> …….. (28)


An alternative notation is to use the Dirac delta function δ(r) and represent H l by the

operator

Hl = (8/3) gβgNβN δ(r) I·S ……… (29)

The signs of the experimental values of the hyperfine tensor cannot be inferred from EPR

spectra. The ratio of unpaired p and s densities in simple free depends on the hybridization and

hence leads to an estimation of bond angle [8]. Often hyperfine coupling to ligand-magnetic

nuclei in transition metal complexes, when the unpaired electron formally occupies metal and

orbital can lead to an estimation of the covalency of the metal ligand bonds [18].

Spin-orbit interaction

This represents the coupling between the magnetic moments arising from the spin and

orbital motion of the unpaired electrons. The energy of the electron spin moment in a magnetic

field B is SB. So the Hamiltonian for the spin orbit coupling is

HL·S = ξ(r)L·S …….. (30)

HL·S = λ L·S …….. (31)

λ = ± ξ(r)/2S …….. (32)

Here λ is the spin-orbit coupling constant of the ion and is a function of the effective nuclear

charge. For more than half filled shell λ is negative and for less than half filled shells, it is

positive. In the spin Hamiltonian formalism, the effective of spin-orbit coupling is assimilated in

to the fictitious spin concept (vide infra).

Spin-Spin interaction

There is also a magnetic coupling between the magnetic moments of the electron and

nucleus, which is entirely analogous to the classical dipolar coupling between two bar magnets.

The classical interaction energy E between two magnetic moments e and N is given by
E = (e·N/r3) −3 (e·ř) (N· ř)/r5 …….. (33)

Here ř is radius vector from e to N and r is the distance between the two moments.

The quantum mechanical version of this is obtained by substituting

e = −gβS ……… (34)

N = gNβNI …….. (35)

Then, the dipolar interaction Hamiltonian

Hd = −gβgNβN{I·S/r3−3(I·ř)(S·ř)/r5} ……… (36)

When more then one unpaired electron is involved in the system, with a ground state

triplet or higher spin multiplicity, direct dipole-dipole interactions among these leads to the

splitting of the spin- states via the spin-spin interactions. When the external magnetic field is

much stronger than the magnitude of spin-spin coupling constant, the above vector dot products

can be expanded to give

HSS=[(gβ)2(3cos21)/r3]SJ·SK …….. (37)

Here θ is the angle between the external magnetic field and the vector joining S J and SK. When

the higher symmetry is in the ion site, this spin-spin interaction contributes a constant energy to

all Zeeman levels, which is not detected by transitions between the levels and hence do not

appear in the EPR spectrum.

For sites of lower symmetry, the electron distribution is polarized and the spin-spin

interaction becomes dependent on MS2 rather than MS because the spin-spin interaction detects

only relative orientations of electron spins, and not absolute orientations.

When the dipoles are always oriented parallel to an external field, the zero field splitting

tensor D is given by

H=S·D·S ……… (38)


Nuclear Quadrupole Interaction

This interaction is relevant only to systems having nuclei with spin I ≥ 1 and arises as a

result of the interaction of the nuclear electric quadrupole moment with the electric field gradient

at the nucleus due to the surrounding electrons. It is expressed as

HQ = QXXIX2 + QYYIY2 +QZZIZ2 ……… (39)

HQ = Q’ [ IZ2 − 1/3 I(I+1)] + Q’’[IX2 −IY2] .…….. (40)

Here Q's are rather the principle values of the quadrupole coupling constant tensor. The

quadrupolar interaction affects the EPR in two ways. Firstly, since there will be a competition

between the electric field gradient at the nucleus and the hyperfine field to quantize nuclear spin-

angular momentum about their respective axis, the |mI >'s are no longer good quantum numbers;

hence the selection rule Δms = ± 1, ΔmI = 0, breaks down and ‘forbidden’ transitions with ΔmI = ±

1, ± 2, become allowed. Secondly quadrupolar effects cause the intensities of normal transitions

and the hyperfine spacing to become unequal, the later showing a progressive increase or

decrease from the ends towards the center [18]. In single crystals, especially when the external

field is perpendicular to the symmetry axis, the analysis becomes difficult due to the presence,

sometimes, of intense forbidden transitions [19].

Thus a proper choice of the spin Hamiltonian based on the symmetry of the paramagnetic

species can explain the observed results, which often can lead to a correct solution of the spin

Hamiltonian parameters such as g, A, Q, etc. It is also possible to theoretically derive these

parameters from knowledge of the ground state molecular orbital and optical spectroscopic data

with the use of perturbation theory [20, 21]. The method of obtaining M.O. co-efficients from

the observed magnetic resonance parameters has been discussed. It is possible, in favorable

cases, therefore to obtain bonding parameters for transition metal complexes from EPR data.
Crystal field parameters for d electrons

Both d1 and d9 systems have a single unpaired electron in the outermost d-orbital and

hence, give rise to 2D in the free ion state. The energy level splitting for d 1 and d9 systems is

identical, except the ordering of the energy levels is inverted for both ions in any symmetry like

octahedral, tetrahedral, square planar etc. In octahedral symmetry, six ligands are arranged

octahedrally around the central d-metal ion and it is clear that the repulsive forces exerted by the

ligands would be strongest along the direction of the X, Y and Z- axes, because in O h field all

ligands are aligned along X, Y and Z- axes. The dx2─y2 and dz2 orbitals (known as eg set) of the d-

metal ion are aligned along the X, Y and Z- axes respectively and the remaining d xy, dxz and dyz

orbital (known as t2g set) are directed in between the X, Y and Z- axes. It means that in O h field,

the repulsion effects on eg orbital will be more than on the t 2g set. Due to this, the 2D term splits

into eg and t2g levels in Oh field. Fig.2 and Fig.3 show the splitting of d-orbitals under various low

symmetry environments. In the case of d9 electron, where the t2g level is lower, then in the Oh

field, the two levels have the energies

 t2gVoctt2g  = - 4Dq

 egVocteg  = 6Dq

The energy difference between the above two levels is 10 Dq, where Dq is a measure of

the interaction of the ion with the crystal field. It is generally considered as a semi-empirical

parameter that can be calculated from the experimental data. Generally, the value of Dq in a

complex depends on the geometrical shape, the nature of the central metal ion and the type of

ligands. The Dq value for an octahedral


Fig. 2 Schematic representation of the splitting of d-orbitals under various crystal field

symmetries
Fig. 3:Crystal field splitting of the d-orbitals in (a) octahedral (b) octahedral with

tetragonal distortion (c) square planar and (d) distorted square planar

complex is greater than for a tetrahedral complex, within the same ligands at the same distance

from the central ion. This is because of the geometrical shape and also the number of ligands

present in the complex [22]. Dq also depends on the effective charge (Zeff) and the valence of the

central metal ion. This effect is probably due to the fact that the central metal ions with higher

charge will polarize the ligands more effectively. The ligand field effects contribute more to the

variation in Dq value for a given central ion and a given geometry. A ligand exerting a strong

field will give a low Dq value.

However, many complexes seem to possess symmetry lower than octahedral. Deviations

from octahedral symmetry are usually treated as perturbations on the high symmetry, which

causes a splitting in the degenerate levels as shown in Fig.I.3. In the present investigation in all

lattices, the symmetry of the metal ions is lower than the octahedral symmetry. In such a

situation, the ground state is not purely due to a single d-orbital, because, the orbital contribution

to the spin Hamiltonian is due to the mixing of the excited state wave function with the ground

state through spin-orbit coupling. Therefore, in the most general situation, the ground state is the

linear combination of all five orbitals. Without restricting to any symmetry considerations, the

coefficients of the d-orbitals and hence the ground state can be calculated using the procedure

developed by Swalen et al. [23]. In the method, the experimentally observed g-values have been

used to determine the five coefficients of the Kramers doublet. There are only three g-values and

a normalization condition to determine five coefficients. The fifth equation is obtained by

assuming the coefficients of the dxz and dyz orbitals to be equal.


Relaxation [24]

In the spin Hamiltonian formalism outlined above, the paramagnetic entities are

considered as if they are in isolation. Such ideal system of ‘non-interacting’ spins is seldom

attainable in practice. Also the system of spins is energetically coupled to the lattice via spin-

orbit-lattice interaction, which affects the lifetime of the excited state by transferring the energy

to lattice via radiation less processes. In an exact solution of the dynamics of the energy transfer

between the microwave field and the magnetic dipoles, these two processes have to be taken into

account [25].

The spin lattice relaxation, or the longitudinal relaxation, (T 1) measures the efficiency

with which the spins can transfer their energy to the surrounding medium. The energy of the

magnetic dipoles is not conserved in this process and leads to the establishment of spin

populations governed by Boltzmann distribution.

The spin-spin relaxation or transverse relaxation time (T2) is a measure of the rate at

which the assembly of spins comes to internal equilibrium at a given temperature and bears no

immediate relationship to the lattice temperature. The energy of the system is conserved in this

process. The spin-spin relaxation controls the natural width of resonance when complication

from ‘saturation’ does not occur.

In ideal paramagnetic systems, the T2 process leads to very broad resonance and often

these may be even beyond detection. Usually, paramagnetic compounds are doped into

isomorphous diamagnetic host-lattices to reduce the dipolar broadening. The spin-lattice

relaxation is more of a property of the individual system and can be altered only be temperature

variation.
There are two theorems, which are very important for EPR spectroscopy. They are Jahn-

Teller and Kramer’s theorems.

Jahn -Teller Theorem


Jahn-Teller theorem [26] states that for a non-linear molecule or ion in an electronically

unstable degenerate state, distortion must occur to lower the symmetry, to remove the

degeneracy and have the lower energy. Cu(II), d9 is a best example, which undergoes Jahn-Teller

distortion when it is in octahedral configuration. If the two ligands along the z-axis in an

octahedral complex are moved either towards or away from the metal ion, the resulting complex

is said to be distorted tetragonal. Generally, such distortions are not favored, since they only

result in a loss of binding energy.

Kramer’s Theorem
Kramer’s theorem [27] states that for any molecule with an odd number of
electrons, the ground state will always be degenerate and the degeneracy can be lifted
only by an external field. This means that a separation between Kramer’s doublets is a
function of D. In other words, if a system has an odd number of electrons, it can show
EPR spectrum even at room temperature. For non-Kramer’s ion, EPR can be
observable only if D is very small. Hence, by using this theorem, one can predict the
observability of EPR.
Magnetically and chemically inequivalent sites

A paramagnetic system with anisotropic g and A tensors will give rise to EPR resonance

depending on the orientation of the magnetic field B with respect to the tensor axes. In single

crystals, depending on the space group and the number of molecules per unit cell (Z), there will

be several different spatial orientations of the paramagnetic sites. Species that are chemically

identical (i.e., they are described by identical spin Hamiltonian parameters) but are spatially

oriented differently are referred to as magnetically distinct sites. It is also possible that due to

charge compensation process [28] in the lattice, depending upon different relative configurations
of the ‘radical vacancy’ directions, there is exist in many different sets of spin Hamiltonian

parameters (although these may differ only slightly). The species themselves would be expected

to be identical when the charge-compensating vacancies are not taken into account. Such sites

are referred to as chemically distinct sites. These chemically distinct sites are necessarily

magnetically distinct, whereas the converse need not necessarily hold.

Line Width

There are four possible factors that can contribute to the width of EPR absorption lines

from solid samples.

(a) Anisotropy of g factors and hyperfine interactions or spin-spin interactions.

(b) Interactions with magnetic dipoles of neighboring electronic and nuclear spin.

(c) Exchange interactions with neighboring unpaired electrons.

(d) Spin lattice and spin-spin relaxation times (T1 and T2).

One or more of these factors may be the major contribution to the width and

shape of the absorption line.

The first factor is important only for powder samples in which the small crystals have a

random orientation. If the anisotropy in g and hyperfine interaction or the presences of a zero-

field splitting are the main factors in determining the line width. The powder will have a width

and shape determined by these parameters. In this case, narrower lines can be obtained by using

single crystals.

The second factor can be the significant one since the magnetic field produced by an

electron at a distance of 0.4 nm is approximately 60 mT. Since the field or a magnetic dipole

depends on the third power of the distance, a simple way to reduce the broadening is to study the

ion in an isomorphic crystal, which is diamagnetic.


If neighboring electron spins are close enough to have their orbital overlap appreciably,

an exchange interaction will occur between the spins. When this interaction is greater than kT,

the system exhibits the phenomena of ferromagnetism or antiferromagnetism. When the

interaction is smaller than kT, it primarily influences the line shape of the EPR spectrum. If the

exchange interaction is large, the effect is similar to what one would expect if the electron were

free to move throughout the crystal. The electron sees an average of all local sites in the crystal,

giving rise to a narrow line in which all interactions have been averaged out. Since the exchange

interaction is strongly dependent on the distance between magnetic ions, it is not present in

magnetically dilute systems.

For systems where the spins are strongly coupled to the vibration modes, the lifetime of a

given magnetic state is short, resulting in an uncertainty in the energy, which manifests itself as a

broad absorption line in the EPR spectrum. This particular broadening mechanism is strongly

dependent on the temperature so that lowering the temperature of the sample can sharpen lines

broadening in this manner. Extremely short relaxation times often occur when the ion has an

electronic excited state only a few hundred cm-1 above the ground state. In such cases, it is

necessary to go to liquid helium temperatures to be able to detect the EPR spectrum. The spin-

spin relaxation or transverse relaxation time (T2) is a measure of the rate at which the assembly

of spins comes to internal equilibrium at the given temperature.


Referneces:

1. E. Zavoisky, J.Phys, USSR, 9,211 and 245 (1945).

2. D. J. E. Ingram, ‘Spectroscopy at Radio and Microwave Frequencies’, Butterworth’s

Scientific Publications, London (1967).

3. D. J. E. Ingram, ’Free radical as studied by Electron Spin Resonance’, Academic Press,

New York (1958).

4. W. Low ‘Paramagnetic Resonance in Solids’, Solid State Physics, Suppl.2, Academic

Press, New York (1960).

5. G. E. Pake, ’Paramagnetic Resonance’, W.A. Benjamin, New York (1962).

6. C. P. Slichter, ‘Principles of Magnetic Resonance’, Harper and Row, New York (1963).

7. A. Carrington and A.D. McLachlan, ‘Introduction to Magnetic Resonance’, Harper and

Row, New York (1967).

8. P.W. Atkins and M.C.R. Symons, ‘The Structure of Inorganic Radicals’, Elesvier,

Amsterdam (1967).

9. P. B. Ayscough, ‘ESR in Chemistry’, Metheuen, London (1967).

10. B. Bleaney and K.W.H. Stevens, Rep.Progr.Phys., 16, 108 (1953).

11. J. A. Weil, J.R. Bolton and J.E. Wertz, ‘Electron Spin Resonance: Elementary Theory and

Practical Applications’, New York, (1994).

12. J. E. Freeman and R.B. Frankel, ‘Hyperfine Interactions’, Academic Press, New York

(1967).

13. B. R. McGarvey, ‘Transition Metal Chemistry’, (R.l. Carlin Ed.), Vol.3, p.89, Dekker,

New York (1968).


14. M. Ikeya, “New applications of Electron Spin Resonance: Dating, Dosimetry and

Microscopy”, World Scientific, Singapore, 1993.

15. R. McWeeny, ‘Spins in Chemistry’, Academic Press, New York (1970)*.

16. H. Eyring, J. Walter and G.E. Kimball, ‘Quantum Chemistry’, John Wiley and Sons, New

York, (1944).

17. E. U. Condon and G.H. Shortley, ‘Theory of Atomic Spectra, Cambridge University

Press, London, (1935).

18. A. Abragam and B. Bleaney, ‘Electron Paramagnetic Resonance of Transition Ions’,

Dower Publications Inc., New York (1970)

19. J. R. Byberg, S. J. K. Jensen and L. T. Muus, J. Chem. Phys., 46, 131 (1967).

20. D. Kivelson and R. Neiman, J. Chem. Phys., 35, 149 (1961).

21. H.R. Gersmann and J.D. Swalen, J. Chem. Phys., 36, 3221 (1962).

22. B.N.Figgis,”Introduction to Ligand Field Theory”, Interscience publishers,

Newyork(1967)

23. J. D. Swalen,B. Johnson,H. M. Gladney,J. Chem.Phys., 52 (1970) 4078

24. N. Bloembergen, E. M. Purcell and R. V. Pound, Phys.Rev., 73, 679 (1946).

25. F. Bloch, Phys.Rev., 70, 460 (1946).

26. H. A. Jahn and E. Teller, Proc.Roy.Soc.Lond, A161, 220 (1937).

27. H. A. Kramers, Proc.Amsterdam Acad.Sci.33, 959 (1930).

28. C. P. Poole, ‘Electron spin Resonance’, 2nd ed, Dover Publications, USA, (1996).
EXPERIMENTAL TECHNIQUES

In this section, a general description of the experimental techniques in EPR spectroscopy

is presented. Methods of extracting the spin-Hamiltonian parameters from EPR data are also

outlined. The resonance condition h = gB could be achieved either by keeping the microwave

frequency constant and sweeping the magnetic field or vice versa. Therefore, in two different

ways a spectrometer could be designed in order to observe the EPR absorption. However, in the

microwave region of the spectrum there are a number of experimental difficulties involved in

varying the frequency. Experimentally it is convenient to keep the frequency constant and vary

the field.

Although the technique of EPR spectroscopy is well suited to the study of solids, liquids

and gases, in the present work, we exclusively deal with the solid state only. The information

obtained from single crystals allows one to obtain both the magnitude and direction cosines of

the various magnetic tensors. The EPR spectrometers are classified into S, X, K, Q and W bands

depending upon the irradiating microwave frequency and their classifications given Table1.

Table-1: Classification of EPR spectrometers

Bands S X K Q W

Approximate frequency (in GHz) 3 9 24 35 70

Approximate wavelength (in nm) 90 30 12 8 4


Approximate field (mT) for g = 2 1.1 3.3 8.5 12.5 25

Some of the characteristics of the EPR spectrum are frequency dependent and more

information to be obtained by the spectra at different frequencies. Also the cavity dimensions

suitable for a specific sample may be chosen by varying the frequency. At higher frequencies,

the second order effects leading to inequivalent spacing of hyperfine splitting and forbidden

transitions will be less effective. The choice of radiation frequency is limited by several factors

such as sensitive and requirement of high magnetic fields homogeneous over the sample volume.

The working frequency of most commercial spectrometers is 9.5 GHz (X-band) and the next

commonly used frequency is 35 GHz (Q-band). Basic principles of EPR instrumentation and the

details of measurement techniques have been discussed by many authors [1-4]. The

spectrometer has to be operated under optimum conditions of microwave power, modulation

amplitude, and spectrometer gain, filter time constant, scan range and scan time.

Instrumentation

The spectrometer used in this work is a JEOL JES TE100 ESR spectrometer working at a

X-band (9.5GHz). At this frequency, 100 KHz modulations are used for observing the first

derivative signal. The block diagram of a typical EPR spectrometer is shown in Fig. 4. For

single crystal X-band work, the crystal is mounted at the end of a Perspex rod with ‘quick fix’

and then introduction into the cavity. The other end of the Perspex rod is attached to a large

protractor, calibrated in degrees. This arrangement made it possible to do accurate rotations of

the crystal about preferred axis by rotating the Perspex rod. The error in mounting the crystal is

about  2 and that in orientations is about  . An accuracy of g = 2 in the


Fig.

4:

Block diagram of a typical X-band EPR spectrometer

orientation of the crystal with respect to the magnetic field could be achieved by this set up. The

axis of rotation is always perpendicular to the magnetic field. The organic free radical diphenyl

picryl hydrazyl (DPPH) with g = 2.0036 is used as an internal field maker

Crystal growth
Single crystal EPR analysis gives valuable information than that obtained from powders

and liquids. So, a brief discussion of the crystal growth is given in this chapter. The slow

evaporation method technique is employed for the crystal growth. Growing crystals by allowing

a saturated solution of a material to lose solvent by evaporation is one of the simplest methods

[5, 6]. Many interesting crystals are grown simply by evaporation of solvent or temperature

change. The evaporation of solvents makes the solution supersaturated so that, it attains the

equilibrium saturated state by eliminating the seed crystals to solution. But, if the solution

becomes too much supersaturated, crystals then would appear spontaneously throughout the

solution. The factors that control the crystal growth technique are (i) character of the solution (ii)

effect of additives and (iii) operating variable such as the degree of super saturation and the

temperature range.

The choice of solvent is an important factor that determines the growth of a crystal from

solution. Growth of a large crystal is almost impossible unless a solvent is found in which the

solute is appreciably soluble. The rate of growth depends on the temperature at which the

solution is maintained. At higher temperatures, the growth rate will be high. However, fine

crystals are obtained by slow evaporation at room temperature. All the crystals are generally

grown at room temperature.

Interpretation of EPR spectra

As one can measure EPR spectra from solution, powder and single crystal samples, the

procedure to obtain spin-Hamiltonian parameters from theses spectra must be identified. A brief

discussion is mentioned below. In order to calculate the g and A values, the following expression

has been used,


g = (gDPPH BDPPH)/B ……… (2.1)

where B is the magnetic field position at the EPR peak, B DPPH is the field position corresponding

to DPPH and gDPPH is the g-value of DPPH which is equal to 2.0036. The g-value is directly

calculated using the spectrometer frequency at which resonance occurs. The expression is as

follows

g = (h / B) ……… (2.2)

In this case,  is the resonance frequency.

The hyperfine (hf) coupling constant ‘A’ is given by the field separation between the

hyperfine components. If the spacing is unequal, an average of them is taken as the value of A.

For n number of hyperfine lines, the average hyperfine value is given by

A = (Bn-B1) / (n-1) ……… (2.3)

Here, Bn is the field position for the nth hyperfine line and B1 is the first hyperfine line field

position.

Spectra are measured both in single crystal and poly-crystalline forms. A brief outline of

the interpretation is given below.

Powders:

In powders, the paramagnetic systems are randomly oriented. Hence the spectrum is an

envelope of statistically weighed average of all these molecules. Concepts and interpretation of

powder line shapes have been described by Kneubuhl [7], Sands [8] and Ibers and Swalen [9]. A

brief pictorial summary of the evaluation of principal magnetic tensors from a few representative

examples is given in Fig.5. In favorable circumstances, therefore, the principal values can be

evaluated from powder data. However, powder line shapes become complicated, when more
than one type of species is present and or when hyperfine lines overlap, and especially so when

the tensors do not coincide. The first two complications can be circumvented to some extent by

measuring the spectra at two different frequencies, say X and Q–band

Fig 2. 2: Schematic diagram indicating the calculation of principle values of magnetic

tensors from powder data based on the delta function line shape
sorting out the field-dependent and field- independent terms in the Hamiltonian. Sometimes

power saturation techniques and temperature variation may also help in this respect.

The limitations in the interpretation of powder spectrum are:

Complication, if there are more than one chemically different paramagnetic species.

Low g-anisotropy of paramagnetic complex

Non-coincidence of g and A tensors

Low intensities and poor resolution may due to overlap of spectra from various micro

crystals of the sampl

Single crystal:

Many authors, for example, Schonland [10] Weil and Anderson [11], Pryce [12], Geusic

and Brown [13], Lund and Vanngard [14] and Waller and Rogers [15] have discussed in detail

the procedure for the evaluation of the principal values of magnetic tensors from single crystal

measurements. The method consists of measuring the variation of g2() for rotations about three

mutually perpendicular planes in the crystal, which may coincide with the crystallographic axes

or are related to the crystallographic axes by a simple transformation. From the maxima and

minima obtained in the three orthogonal planes, the matrix elements of the g 2 tensors can be

derived easily [10]. A Jacobi diagonalization of this matrix gives rise to the eigen values

corresponding to the principal values of the g tensor and the transformation matrix, which

diagonalises the experimental g2 matrix. This matrix provides the direction cosines of these

tensors with respect to the three orthogonal rotations. However, complications will arise, when

more than one magnetically distinct site per unit cell is present, because no apriority

predictability of the relations between sites and spectra in the three planes. This leads to several

possible permutations leading to many g2 tensors. For example, if a system has n sites, then
there are 2n3 tensors, including for not performing a proper rotation, i.e., clockwise or

anticlockwise. A careful examination, however, invariably leads to the proper combinations and

the corresponding direction cosines. In the case of hyperfine tensor, when g is not highly

anisotropic, the same procedure as above can be adopted. When this is not the case, Schonland

[10] has suggested that it is necessary to follow the variation of g 2A(θ) in the three principal

planes. The reason for this is as follows:

The Hamiltonian for a paramagnetic system, including only the electronic Zeeman and

hyperfine terms can be expressed as

H =  (g11B1S1 + g22B2S2 + g33B3S3) + (A11S1I1 + A22S2I2 + A33S3I3) ……… (2.4)

Let (n1, n2, n3) be the direction cosines of the magnetic field B with reference to the axes

of the g tensor and hyperfine tensor. Here, it is assumed that g and hyperfine tensors are

coincident. If M and m are the electron and nuclear spin quantum numbers, then the energy

levels are given by

EM,m = gBM + KMm ……… (2.5)

Here g and K are given by the equations

g = (g112n12 + g222n22 + g332n32)1/2 ……… (2.6)

and

K = 1/g [g112A112n12 + g222A222n22 + g332A332n32]1/2 ......... (2.7)

The magnetic field Bm where the transition |M, m>|M+1,m> occurs is given by

h = gBm + K ……… (2.8)

In other words,

Bm = h / g - (K/g)m ……… (2.9)

If A is hyperfine splitting and the lines are centered around (h/g), then
A = K/g …….. (2.10)

In order to obtain the matrix elements of the hyperfine tensor, the angular variation of (gK)2 is

considered, since (gK) has a linear angular dependence on g. Therefore,

(gK)2 = g4A2 …….. (2.11)

From this equation, the matrix elements of the hyperfine tensor matrix are evaluated using the

same procedure used to get g tensor matrix.

Schonland has indicated the probable errors in the method described above to get the

principal values of g and hyperfine tensors. But, the errors are very small compared with the

experimental errors involved, such as mounting the crystal along the specific axis, measurement

of magnetic field etc.

Direction cosines of the Substitutional sites:

The single crystal X-ray analysis data provides the positional parameters p, q, r and the

unit cell dimensions a, b, c and , , . For crystal system with non-orthogonal crystal axes, the

positional parameters p, q, r of the various atoms can be changed over to an orthogonal

framework and the Cartesian co-ordinates x, y, z could be calculated using the relation

x a b cos  c cos  p

y = 0 b sin  (c/sin) (cos  - cos  cos ) q

z 0 0 d r

where, d = [ c2 – c2 cos2  - ( c2 / sin2 ) (cos  - cos  cos  )2 ]1/2


By setting the metal atom as the origin, the coordinates of the various atoms in the crystal

surrounding the metal are calculated. The normalized Cartesian co-ordinates of these atoms give

the direction cosines of the metal-ligand bond of the co-ordination polyhedron. The direction

cosines of these metal-ligand bonds can be compared with the direction cosines of the g and A-

tensors, obtained by the procedure described in the previous section. Sometimes, it is found that

the magnetic tensor directions coincide with some of the bond directions, which may not be so in

low symmetry cases.

SimFonia powder simulation:

The simulation of the powder spectrum is generally carried out to verify the experimental

spectrum with the theoretical one, obtained by using the spin Hamiltonian parameters calculated

from the experimental spectrum. The simulation of the powder spectrum is done using the

computer program SimFonia developed and supplied by Brucker company. The algorithm used

in the SimFonia program for powder simulation is based on perturbation theory, which is an

approximation. Previously, perturbation theory has been used in the interpretation of EPR

spectra because of the speed of calculation and the intuitiveness of the results. It is an

approximate technique for finding the energy eigen values and eigen vectors of the spin-

Hamiltonian. The assumption made is that there is a dominant interaction, which is much larger

than the other interactions. As the dominant interaction becomes larger when compared to the

other interactions, the approximation becomes better. The five interactions that are considered in

the SimFonia simulation program for the powder sample are:


Electronic Zeeman interaction: It is the interaction of the magnetic moment of the

electron with externally applied magnetic field i.e., the magnetic field from the spectrometer

magnet.

Zero-field splitting: It occurs in electronic systems in which the spin is greater than 1/2.

Nuclear hyperfine interaction: It is an interaction between the magnetic moment of the

electron with the magnetic moment of the nucleus.

Nuclear Quadrupole interaction: It is the interaction between the Quadrupole moment of

the nucleus with the local electric field gradients in the complex (for system having nuclear spin

greater than 1/2).

Nuclear Zeeman interaction: It is the interaction of the magnetic moment of the nucleus

with the externally applied magnetic field.

The assumption made in the simulations is that the electronic Zeeman interaction is the

largest, followed by the zero-field splitting, hyperfine interaction, nuclear quadrupole interaction

and the nuclear Zeeman term is the smallest. Perturbation theory works best when the ratio

between the successive interactions is at least ten. If the limits exceeded, perturbation theory still

gives a good picture of EPR spectrum; however, it may not be suitable for the quantitative

analysis. And if the EPR spectrum is to be simulated with larger hyperfine interactions, then

second order perturbation theory is selected to increase the accuracy of the simulation. The zero-

field splitting is always treated to second order because they do not produce a non-zero first

order term.

Only allowed EPR transitions are simulated, but under some circumstances forbidden

transitions can also appear. These corresponds to simultaneous flip of the nucleus and flop of the

electron and forbidden EPR lines occur between the allowed transitions or a Ms =  2
electronic transitions. These forbidden lines are not simulated because perturbation theory is not

the optimal method for calculating their positions and intensity. The SimFonia powder

simulation program simulates EPR spectra for spin 1/2 to spin 7/2 electronic systems. For spin

greater than 1/2, the zero-field splitting terms (D and E) are implemented. There are essentially

no restrictions on the spin of the nuclei. All the naturally occurring spins have been

programmed. The principal axes of the electronic Zeeman interaction and the zero-field splitting

are assumed to be coincident. SimFonia can simulate both types of line shapes i.e., Lorentzian

and Gaussian, as well as combination of the two. This technique is most efficient for many line-

complicated spectra. Detailed theory of the powder spectra simulation can be obtained from the

references [16, 17].

Simulation of single crystal EPR Spectra:

Once magnetic parameters are obtained from single crystal, it is important to check their

accuracy by simulating the EPR spectra at selected orientations or calculating the iso-frequency

plots at certain planes and then to compare these with experimental results. In case of organic

free radicals in solution where g is isotropic and hyperfine coupling constant are very much

lower than the electron Zeeman terms, it is enough to the first order perturbation theory (S z and Iz

are good quantum numbers) to generate the experimental spectrum by using Lorentzian line

shapes of appropriate width for the various transition. When it comes to simulation of EPR

spectrum of transition metal complexes in solids, the following additional points must be

remembered. The anisotropy in g and A tensors and the quadrupole coupling constant cannot be

ignored generally. Therefore any reasonable prediction to the resonance position can only be

achieved by carrying the perturbation at least up to second order in energy, the wave function
correct to first order. Using the second order perturbation equations to calculate line positions

and three-points quadrate formula of integration, EPR spectra of powder samples are simulated

using program EPR-NMR developed by Weil and his group[18].

Single Crystal:

The procedure for the extraction of principal values of the magnetic tensors from single

crystal studied as formed extensive coverage and literature by different authors for example,

Schonland [10], Weil and Anderson [11], Pryce [12], Geusic and Brown [13], Vangurd and Lund

[14] and Waller and Rogers [15]. The basic theory of the general methods is described below:

Essentially, the method consists of measuring the variation of g2() for rotation about

three mutually perpendicular axes in the crystal, which may coincide with the crystallographic

axes by a simple transformation. From the maxima and minima obtained in the three orthogonal

planes, the matrix elements of the g2 tensor can be derived easily. A Jacobi diagnolization of this

matrix gives rise to the eigen values corresponding to the principal values of the g tensor and the

transformation matrix which diagonalizes the experimental g2 matrix gives the direction cosines

of these tensors with respect to the three orthogonal rotations. Complication arises, when more

then one magnetically distinct site per unit cell is present, since there is no apriority predictability

of the relations between sites and spectra in the three planes. This leads to several possible

permutations leading to many g2 tensors. A careful examination invariably leads to the

identification of the proper combinations and the corresponding direction cosines. In the case of

hyperfine tensor, when g is not highly anisotropic, the same procedure as above can be adopted.
When this is not the case, Schonland has suggested, that it is necessary to follow the variation

g2A() in the three principal planes. The reason for this is as follows

The Hamiltonian for the paramagnetic system, including only the electronic Zeeman and

hyperfine terms can be expressed as

H = (g11 B1S1 + g22 B2S2 + g33 B3S3) +A11S1I1+ A22S2I2 + A33S3I3

Let (n1, n2 , n3)be the direction cosines of the magnetic field B with respect to the axes of the g

tensor and hyperfine tensor. Here, it is assumed that g and hyperfine tensors are coincident. If

M and m are the electron and nuclear spin quantum numbers, then the energy levels are given by

EM,m = gBM + KMm

Here g and K are given by the equations

g = (g112n12 + g222n22 + g332n32)1/2

K= 1/g(g112 A112 n12 + g222 A222 n22 + g332 A332 n32)1/2

The magnetic field Bm where the transition |M,m>  |M+1,m> takes place is given by

hν = gβBm + βKm

If A is hyperfine splitting and the lines are centered around hν/gβ, then

A=K/g

Schonland has indicated the probable errors in the method described to get the principal

values of g and hyperfine tensors. But, the errors are very small compared with the experimental

errors involved, such as mounting the crystal along the specified axis.

Computer Program EPR-NMR [19]:

The program sets up spin-Hamiltonian (SH) matrices and determines their eigenvalues

(energies) using “exact” diagonalization. It is a versatile program, having many operating modes
tailored to a variety of applications. These modes can be grouped into four categories, in

increasing order of complexity as follows

1. Energy-level calculation,

2. Spectrum simulation,

3. Comparison with observed data,

4. Parameter optimization.

For each category, most of the operations of the lower categories remain available, so that

a good way to learn how to use the program effectively is to start at the lowest category and work

one’s way up.

Category I: In this category, the user provides the program with SH parameters, and directions

and magnitudes of applied magnetic fields.

Category II: In category II, the user also specifies an experiment, chosen from field-swept or

frequency-swept electron paramagnetic resonance (EPR) or nuclear magnetic resonance (NMR),

electron nucleus double resonance (ENDOR), or electron spin echo envelope modulation

(ESEEM). Also, the user must identify the transitions of interest. The “spectra” simulated

consist of sets of transition frequencies or magnetic field magnitudes, and possibly relative

transition probabilities. The program can also convolute these data with a line-shape function

(Lorentzian or Gaussian) to produce a plot.

Category III: For this category, the user also supplies appropriate observed single-crystal data,

with transition labels assigned, and the program determines the degree of consistency with data

calculated from the given SH parameters. This can include an error analysis on a user-selected

subset of SH parameters and /or magnetic-field directions.


Category IV: In the category IV, the user-selected subset of parameters may be optimized, so as

to give better agreement between observed and calculated transition frequencies. This uses a

non-linear least squares routine, which systematically varies the parameters so as to minimize

weighted differences between observed and calculated transition frequencies (or fields). In this

category, user-supplied SH parameters need only be estimates or outright guesses. Examples

dealing with organic and inorganic radicals are discussed. The discussion has been extended to

Transition metal ions. Term symbols, splitting of various states in ligand fields and the pattern

of electronic spectra for d1 to d9 has been highlighted in the lectures.

References

[1] D. J. E. Ingram, “Biological and Biochemical Applications of Electron Spin Resonance”,

Adam Hilder LTD, London (1969).

[2] C. P. Poole, “Electron Spin Resonance”, Dover Publications, United States of America

(1996).

[3] R. S. Alger, “Electron Paramagnetic Resonance Techniques and Application,”

Interscience, New York (1968).

[4] T. H. Wilmshurst, “Electron Spin Resonance Spectrometer”, Plenum, New York (1968).

[5] J. J. Gilman, “The Art and Science of Growing Crystals”, John Wiley & son Inc., New

York (1963).

[6] J. C. Brice, “The growth of Crystals From Liquids”, North-Holland Publishing Company,

London (1973).

[7] E. K. Kneubuhl, J. Chem. Phys., 33 (1960) 1074.

[8] R. H. Sands, Phys. Rev., 99 (1955) 1222.


[9] J. A. Ibers, J. D. Swalen, Phys. Rev., 127(1962) 1914

[10] D. S. Schonland, Proc. Phys. Soc., 73 (1959) 788.

[11] J. A. Weil, H. A. Anderson, J. Chem. Phys., 28 (1958) 864.

[12] M. H. L. Pryce, Proc. Phys. Soc., A63 (1950) 25.

[13)] E. Geusic, L. C. Brown, Phys. Rev., 112 (1958) 64.

[14)] Lund, T. Vanngard, J. Chem. Phys., 42 (1965) 2979.

[15] W. G. Waller, Max T. Rogers, J. Magn. Res., 9 (1973) 92.

[16] A. Abragam, B. Bleaney, “Electron Paramagnetic Resonance of Transition metal Ions”,

Clarendon Press, Oxford, (1970).

[17] J. R. Pilbrow, “Transition Ion Electron Paramagnetic Resonance”, Clarendon Press,

Oxford, (1990).

[18] Physical Chemistry, ATKINS 6th edition

[19] EPR-NMR Program developed by F. Clark, R. S. Dickson, D. B. Fulton, J. Isoya, A.Lent,

D. G. McGavin, M. J. Mombourquette, R. H. D. Nuttall, P. S. Rao, H. Rinneberg, W. C.

Tennant, J. A. Weil, University of Saskatchewan, Saskatoon, Canada (1996).

Anda mungkin juga menyukai