Gibb
Gibb
Gibb
General Editor
A.D. Buckingham F.R.S., Professor of Chemistry,
University of Cambridge
Series Foreword
The field of science known as 'Chemical Physics' has
greatly expanded in recent years. It is an essential part
of both physics and chemistry and now impinges on
biology, crystallography, the science of materials and
even on astronomy. The aim of this series is to present
short, authoritative and readable books on different
topics in chemical physics at a level that is appreciated
by the non-specialist and yet is of prime interest to the
expert-in fact, the type of book that we all welcome
and enjoy.
I was grateful to be given the opportunity to help
plan this series, and warmly thank the authors and
publishers whose efforts have brought it into being.
A.D. Buckingham,
University Chemical Laboratory,
Cambridge, U.K.
T.e. GIBB
Department of Inorganic and Structural Chemistry,
University of Leeds
Preface
2 Hyperfine Interactions 22
2.1 The chemical isomer shift 23
2.2 Magnetic hyperfine interactions 28
2.3 Electric quadrupole interactions 30
2.4 Combined magnetic and quadrupole interactions 36
2.5 Relative line intensities 38
References 43
3 Molecular Structure 45
3.1 Iron carbonyls and derivatives 46
3.2 Geometrical isomerism in Fe and Sn compounds 54
3.3 Linkage isomerism in cyano-complexes of Fe 56
3.4 Conformations in organometallic compounds of Fe 60
Contents
3.5 Stereochemistry in tin compounds 63
3.6 Molecular iodine compounds 67
Appendix Quadrupole splitting in cis- and trans-isomers 69
References 71
Bibliography 245
Observed Mossbauer Resonances 247
Index 248
CHAPTER ONE
0.1 )
E +lMV2
2
= E 'Y +lM(V
2
+ v)2 (1.3)
4 Principles of M6ssbauer Spectroscopy
Table 1.1 Typical energies of nuclear and chemical interactions
E'Y =E -1..Mv
2
2- Mv V
0.4)
(1.6)
and depends on the mass of the nucleus and the energy of the 'Y-
ray. However, the Doppler energy ED is dependent on the thermal
motion of the nucleus, and will therefore have a distribution of
values which is temperature dependent. A mean value, ED, can be
defined which is related to the mean kinetic energy per translation-
The Mossbauer Effect 5
al degree of freedom, Ek =tkT, by
- _/~
ED = 2v (EkE R) = E'Y
j2Ek
Mc 2 0.7)
(a)
-Ey
-Ey
Fig. 1.1 The statistical distribution in the -y-ray energy for (a) emission,
(b) absorption, and (c) the resonant overlap for successive emission and ab-
sorption.
ER = (1 - f)tLw
or
(1.8)
From the form of the equation it can be seen that f is large when
80 is large (a strong lattice) and when the temperature T is small.
This decrease in f with rise in temperature follows in general from
equation (1.9), in that thermal energy will increase the amplitude
of vibration of the nucleus, but the precise form of the tempera-
ture dependence can only be predicted using a model for the lat-
tice vibrations. This problem is discussed in more detail in
Chapter 6.
It ·should be noted that although the recoilless fraction can be
increased by lowering the temperature there is a limiting value at
absolute zero which is still less than unity. For instance, from
equation (1.1 0) when T = 0
(1.11)
Transmission
as
recorded Radiation
incident
on
absorber
_ _ NonIsonan~
Radiation { - - - --
lost in Non -resonant
absorber by
non-resonant
- - - - - ---t---
Resonant
absorption
o +
Relative Ooppler shift E" fEy
N(E) dE =-
Ir dE
(1.12)
21T (E -Ky + (r(2)2
where N(E) is the probability of the energy being between E and
E + de, f is the recoilless fraction of the source and r is the Heisen-
berg width. This function which is usually referred to as the
Lorentzian distribution has a maximum value at E = E'"(" The cross-
section for resonant absorption, aCe), is similarly expressed [6, 7]
as
(r/2)2
aCe) = ao (1.13)
(E - Ky + (r/2)2
where ao is a nuclear constant called the absorption cross-section
given by
(1.14)
The Mossbauer Effect 11
where Ie and Ig are the nuclear spin quantum numbers of the excit-
ed and ground states and ex is the internal conversion coefficient
of the -y-ray [Note: -y-emission does not always lead to an observ-
able -y-ray. In a proportion of events an atomic s-electron is eject-
ed instead in a process known as internal conversion. ex is defined
as the ratio of the number of conversion electrons to the number
of -y-ray photons emitted. Values for ao are often quoted in units
of barns (1 bam = 10-24 cm 2 = 10- 28 m 2)]. For ao to be large,
both E'Y and ex should be small.
In any real absorber of finite thickness the resonant absorption
is in direct competition with other non-resonant processes such as
Compton scattering. A general evaluation of the integrated absorp-
tion intensity or 'transmission integral' is extremely difficult, but
to a first approximation the basic behaviour is comparatively
simple [8]. The recorded absorption spectrum for infinitely thin
sources and absorbers has a Lorentzian distribution with a width
at half-height of rr = 2r. For an absorber of finite thickness con-
taining t a resonant atoms per unit area of cross-section and with a
recoilless fraction of faJ an 'effective thickness' is usually defined
by the dimensionless parameter Ta = faaot a. The absorption shape
approximates very closely to Lorentzian but with a broadened
wid th r r given by
( 1.15)
provided that Ta < 5; i.e. the line broadens with an initially linear
dependence on thickness. Self-resonance in the source matrix
results in an additional line broadening which is approximately
independent of absorber thickness.
The probability that any given photon will be resonantly absorb-
ed can be enhanced by increasing the thickness Ta. However, the
intensity of the incident -y-ray beam is also strongly attenuated by
conventional non-resonant absorption processes, and this places
an upper limit on the value of Ta which can be used effectively.
The absorption may be defined in terms of the transmitted inten-
sity at the resonant maximum (10) and the transmitted intensity
at a large Doppler velocity where the absorption is zero (I ..) by
(1.16)
12 Principles of Mossbauer Spectroscopy
It has been evaluated [8] to be
(1.17)
1·0.--------------------------,
0·8
N
i::
~0·6
...,
.
N
i:::
I
I
";;" 0·4
....
.......
~
4 6 8 10
Thickness (Tl
Fig. 1.3 The function A If = 1 - e- TI2 J oUTI2) showing how the absorption,
A. shows a saturation behaviour with increasing thickness, T.
The Mossbauer Effect 13
1.4 The Mossbauer spectrometer
The Mossbauer spectrum is a record of the transmission of reso-
nant r-rays through an absorber as a function of the Doppler velo-
city with respect to the source. It is therefore quite simply a
record of transmission as a function of the energy of the incident
radiation. The major difference from other forms of transmission
spectroscopy is that the hazard to health in using high-energy radia-
tion, combined with the high cost of producing the radioisotopes,
necessitates the use of relatively low photon flux-densities. In con-
sequence, much longer counting times are required to achieve a
given resolution (hours rather than minutes). The statistics of the
random counting predicts that the standard deviation in N register-
ed events is VN: thus the standard deviation in 10 000 counts is
1% and in 1 000000 counts is 0.1 % of the total count. A prospec-
tive gain in resolution must be balanced against the longer experi-
mental time involved, which brings with it the problem of long-
term reproducibility in the measuring equipment. To double the
resolution requires four times the counting time, and makes it
imperative to obtain as large a percentage absorption as possible.
The measurement of a Mossbauer spectrum is almost exclusively
carried out by repetitively scanning the whole velocity range re-
quired, thereby accumulating the whole spectrum simultaneously,
and allowing continuous monitoring of the resolution. This can
be achieved electromechanically, and a typical modem Mossbauer
spectrometer is illustrated schematically in Fig.lA. The following
description is a brief outline of the major principles involved. For
a more detailed account of instrumentation and experimental
methods the reader is referred to a recent review by Kalvius and
Kankeleit [9] .
The major component is a device known as a multichannel
analyser which can store an accumulated total of r-counts, using
binary memory storage like a computer, in one of several hundred
individual registers known as channels. Each channel is held open
in tum for a short time interval of fixed length, which is derived
from a very stable constant-frequency clock device. Any r-counts
registered by the detection system during that time interval are
added to the accumulated total already stored in the channel. The
sequential accessing of the channels is completed in about 1/20
of a second, and is repeated ad infinitum.
14 Principles of Mossbauer Spectroscopy
r-------,
'---r---,;---:-r---' :s,,~ ,
I AbsorberI
I I
I I
I Cryostat I
' - _ _ _ _ _ .J Pulse
Waveform
generator amplifier
Constant Multichannel
frequency Discriminator
clock Address analyser
control
Data output to
1
C.R.T. display
Typewriter
Tape punch
Magnetic tape
x-v plotter
The timing pulses from the clock may also be used to synchro-
nize a voltage waveform, which is used as a command signal to the
servo-amplifier controlling an electromechanical vibrator. The
latter moves the source relative to the absorber. A waveform, with
a voltage increasing linearly with time, imparts a motion with a
constant acceleration in which the drive shaft and the source
spend equal time intervals at each velocity increment. The multi-
channel analyser and the drive are synchronized so that the velo-
city changes linearly from -v to +v with increasing channel num-
ber. In this way the source is always moving at the same velocity
when a given channel is open.
The 'Y-ray detection system is conventional. A scintillation
counter, gas proportional counter, or lithium-drifted germanium
detector may be used according to circumstances [10]. The pulses
from the detector are amplified, passed through a discriminator
which rejects most of the non-resonant background radiation, and
finally are fed to the open channel address.
The geometric arrangement of the source, absorber and detec-
The Mossbauer Effect 15
tor is important. A 'Y-ray emitted along the axis of motion of the
source has a Doppler shift relative to the absorber of E"{v/ c, but
any 'Y-ray which travels to the detector along a path at an angle 0
to this axis has an effective Doppler shift of only E"{ v cosO/c. This
'cosine-effect' of solid-angle will therefore cause a spread in the
apparent Doppler energy of the 'Y-rays and hence line broadening
[11]. The effect can be reduced by maintaining an adequate
separation between the source and detector, or by collimation.
The velocity range required to completely encompass the absorp-
tion is usually less than ± 1a mm s-l, so that the overall displace-
ment of the source during a scan is barely discernible. Absolute
calibration of the drive is not easy, but can be achieved by mount-
ing a diffraction grating or mirror on the shaft as part of an inter-
ferometer. For most purposes it is more convenient to rely on an
internal standard. As will be seen in the next chapter, it is possible
for an absorber to show more than one resolved absorption line.
If such a material is chemically reproducible and stable, and the
lines are measured on an instrument with absolute calibration, it
can then be used in other laboratories to establish the velocity
amplitude and linearity of un calibrated instruments. For example,
it is common practice to use a magnetic a-iron foil which has six
lines as a calibrant for 57Fe work, and to quote all measurements
of velocity relative to the centroid of this reference spectrum.
Very few Mossbauer resonances are easy to record at room tem-
perature, the main exceptions being 57Fe and 119Sn, and in many
instances it is necessary to cool at least the absorber, and some-
times the source as well, to increase the recoilless fractions.
Because the source is moving, this is not without its difficulties.
There is also the possibility that some of the hyperfine interactions
(to be described in Chapter 2) are temperature dependent. For
this reason alone, low-temperature measurements have become
popular [9]. Numerous commercial cryostats are now available,
the commonly used refrigerants being liquid nitrogen for tempera-
tures down to 78 K and liquid helium down to 4.2 K. Cryostats
with a fully-variable temperature control are also used, particularly
in the study of phase transitions. In some applications it is desir-
able to apply a large external magnetic field to the absorber, and
superconducting magnet installations are available which can pro-
duce magnetic flux densities of up to lOT.
Although this discussion has been weighted in favour of using a
16 Principles o! Mossbauer Spectroscopy
transmission experiment to observe resonant absorption, it is also
possible to detect resonant fluorescence by a scattering experiment.
The resonant absorption event may be recorded by detecting the
scattered 'Y-ray on de-excitation (alternatively, if the internal con-
version coefficient IX is large, the conversion-electron or X-ray
produced may be registered). This has the advantage that non-
resonant radiation from the source is not recorded, and the major
remaining contributions to the radiation background are scattered
higher-energy 'Y-rays and non-resonant Rayleigh scattering. The
main disadvantages are that the intensity of the scattered radiation
is very weak, and the large solid geometry requirement for a high
count rate results in broadening of the Mossbauer line. A further
problem, which may arise if scattered 'Y-rays are counted, is the
possibility of anomalous effects dependent on the scattering angle.
These are due to interference between resonant Mossbauer scatter-
ing and the non-resonant Rayleigh scattering [12] and to diffrac-
tion [13].
The principal application of scattering experiments is in the stu-
dy of surface phenomena, and examples of this are given in Chap-
ter 9.
Natural
rrl abund-
Isotope E.y/keV (mm s-1) Ig Ie a ance (%) Nuclear decay*
.---~~-
~iCo
~- ---r--270d
EC (99.84 % )
119mSn
50
250d 89.54
IT
2H7keV
t-'-t---L-...--- 14'41 keY
t- ----''---1..__ 0
, 0
t + -----"'---
119s
50 n
129mTe 193 05
33 d _ _52~_~. 76
70m-:-=-:!--.-
" - - . - - - 138·9
- - i - - - 80·2 IT 12d
'---rrrH-If- 487
'--t--..-- 73-0 keY
----..,,.-LH-+t+- 278
1'-r''+--':1:±Ll1- 2H keY
'-"-L......l~r- 0 '--''--....IL..--O
~-
1-7 x 107y
Fig. 1.5 The nuclear decay schemes for the Mossbauer resonances in
57Fe, 119S n , 1291 and 1931r.
(1.18)
References
[1] Mossbauer,R.L.(1958)Z.Physik, 151, 124.
[2] Kuhn, W. (1929) Phil. Mag., 8,625.
[3] Metzger, F. R. (1959) Progr. Nuclear Phys., 7,54.
[4] Lipkin, H. J. (1960) Ann. Phys., 9,332.
[5] Petzold, J. (1961) Z. Physik, 163, 71; Wisscher, W. M. (1960) Ann.
Phys., 9, 194.
[6] Frauenfelder, H. (I962) The Mossbauer Effect, W. A. Benjamin Inc.,
N.Y.
[7] Jackson, J. D. (1955) Can. J. Phys., 33, 575.
[8] Margulies, S. and Ehrman, J. R. (1961) Nuclear Instr. Methods, 12, 131.
[9] Kalvius, G. M. and Kankeleit, E. (1972) Recent improvements in in-
strumentation and methods of Mossbauer spectroscopy. In Mossbauer
Spectroscopy and its Applications, International Atomic Energy
Agency, Vienna, p.9.
[10] (1965) Alpha-, beta-, and gamma-ray spectroscopy VoLl, ed. K. Sieg-
bahn, North-Holland, Amsterdam.
[11] Riesenman, R., Steger, J. and Kostiner, E. (1969) Nuclear Instr.
Methods, 72, 109.
[12] Black, P. J., Longworth, G. and O'Connor, D. A. (1964) Proc. Phys.
Soc., 83,925.
[13] Black, P. J. and Duerdoth, I. P. (1965) Proc. Phys. Soc., 84,169.
[14] Mulholland, H. and Jones, C. R. (1968) Fundamentals of Statistics,
Butterworths, London.
[15] Beyer, W. H. (editor), (1968) Handbook of Tables for Probability and
Statistics, 2nd edition. The Chemical Rl.!1.Jber Co., Cleveland.
CHAPTER TWO
Hyperjine Interactions
Ze 2
Eo = - - -
foo'1'2-
dr
(2.1)
41TEO 0 r
(2.2)
(2.3)
(2.5)
1·00
0·95
50·90
;;;
'"E
'"g 1·00
L.
~
Source: Zn1Z7m Te
0·98
Absorber : No 3HZ 1Z7106
Fig. 2.1 The 1271 and 1291 spectra of Na3H2I06 showing the smaller line-
width and larger chemical isomer shift of opposite sign for the latter ([ 41 ,
Fig. 1)
motion the mean value (v) is zero so that only the second-order term
in (v2) can influence the Mossbauer resonance. Hence for v ~ c
(V2»)
II =vo (1 +-- (2.7)
2 2c
DE vo - v <v 2 )
(2.8)
(2.10)
where mzis the magnetic quantum number and can take the values
I, 1 - 1, ... , - I. In effect, the magnetic field splits the energy
level into 21 + 1 non-degenerate equi-spaced sublevels with a separa-
tion of JiB/I.
In a Mossbauer experiment there may be a transition from a
ground state with a spin quantum number Ig and magnetic moment
Jig to an excited state with spin Ie and magnetic moment Jie- In a
magnetic field, both states will be split according to equations
(2.9) and (2.10). Transitions can take place between sub-levels
provided that the selection rule llm z = 0, ± 1 is obeyed [this is cal-
led a magnetic dipole (M I) transition; for other selection rules see
Section 2.5 on line intensities]. The resultant Mossbauer spectrum
contains a number of resonance lines, but is nevertheless symmetri-
cal about the centroid.
HyperJine Interactions 29
A typical example of magnetic hyperfine splitting is illustrated
schematically in Fig.2.2, which is drawn to a scale appropriate to
+,
119Sn. For this isotope Ig = Ie =T, Ilg = -1.041IlN and
Ile = +0.67IlN· The change in sign of the magnetic moment results
in a relative inversion of the multiplets. The six lines are the allow-
ed t:..m z = 0, ± I transitions, and the resultant spectrum is indicated
in the stick diagram. The lines are not of equal intensity, but the
3:2: I: I: 2:3 ratio shown here is often found for example in the
57Fe and 119S n resonances in randomly oriented polycrystalline
samples. A more detailed account of relative line intensities is
given later in the chapter.
_1
I
I
2
I 1
I.e=l2 I /
(,./ -2
\'\ ' +12
\
\ +12
+1
123 2
/
/
I
I
Ig=~ /
\
\
\
\
\
\ 1
456 -2
Velocity
-I' 0, +1'
3
I
4
I
2
I
5
I I6
Fig. 2.2 The energy-level scheme and resultant spectrum for magnetic hyper-
t
fine &p!itting of an. I g =} ~ Ie = transition. The relative splittings are
scaled In accord WIth the magnetIc moments of 119Sn ; Ilg = -1.041 IlN and
Ile = +0.67IlN- The line intensity ratios of 3:2: I: 1:2:3 are appropriate to a
poly crystalline absorber_
30 Principles of M6ssbauer Spectroscopy
The magnetic field may be applied by an external magnet, or it
may be intrinsic to the compound in which case it is usually refer-
red to as an 'internal' field. An unpaired electron in the atomic
environment can induce an imbalance in electron spin-density at
the nucleus, and thereby generate a local magnetic flux density
which can be as large as 100 T. Examples of this may be found
in Chapter 5.
The sign of an internal magnetic flux density, B, can sometimes
be obtained by applying an additional externally generated mag-
netic flux density, B o, with a large magnet. If the resultant flux
density in the Mossbauer spectrum has increased to B + Bo by
parallel alignment of B with Bo (i.e. the internal field is parallel
to the magnetization) then the sign is positive, whereas a field of
B - Bo signifies antiparallel alignment and a negative sign. This
method fails if the magnetic anisotropy of the matrix is large
enough to prevent rotation of the internal field into the direction
of the applied field.
Time-dependent effects such as relaxation are discussed separ-
ately in Chapter 6.
1
:Jf' = - - eQ • 'VE (2.12)
6
a2 v
'ViEi = - = -Vii (2.13)
ax/-'Oxi
(Xi, Xi = X, y, z)
There are thus nine values of Vii to express 'VE in a Cartesian axis
system, X, y, z. A 'principal' axis system may always be defined
*"
such that all the Vii terms with i j are zero (the matrix of the
tensor is diagonal), leaving the three finite 'principal' values Vxx ,
Vyy , and Vzz . Furthermore, ~E is a traceless tensor, so that
(2.14)
(2.15)
32 Principles of M6ssbauer Spectroscopy
such that I Vzz I > I Vyy I ~ I Vxx I and 0 .;;;; 7'/ .;;;; 1. The z axis is then
referred to as the 'major' axis, and x is the 'minor' axis. The cor-
rect assignment of these axes is often obvious from the molecular
geometry if the overall symmetry is high; e.g. in a molecule with
axial symmetry, the z axis is the symmetry axis. (Note: in an arbi-
trary axis sytem it is necessary to specify five parameters which
maybe Vxx, VX)', Vzz , Vyy,and Vyz,oraltemative1ytheprincipal
values Vzz and 7'/ with three angles to specify their orientation with
respect to the arbitrary x, y, z axes.)
The Hamiltonian in equation (2.12) is more often written as
2
= e qQ [31; _ 1(1 + 1) + 1/(/; -1;)] (2.16)
41(21 - 1)
with energy levels at ±(e 2qQ/4)(1 + 7]2/3)1/2. Values for higher spin
states must be calculated numerically.
Hyperfine Interactions 33
A Mossbauer transition between states can take place according
to the selection rule f1mz = 0, ± 1 for Ml dipole radiation. The
most common example is the Ig = } -+ Ie = f- transition in which
the ground state with Iz = ±} is unsplit, and the excited state has
two levels with I z = ±f- and ±t separated by (e 2qQ/2) (1 + T/ 2/3 )1/2.
The resultant Mossbauer spectrum is a doublet with a separation
called the quadrupole splitting of f1 = (e 2qQ/2)(1 +7]2/3)1/2. When
7] = 0, f1 = e 2qQ/2 or half the quadrupole coupling constant defin-
ed in nuclear quadrupole resonance spectroscopy. Typical energy
level schemes for} -+ %and T-+ i (scaled to 1291) transitions and
the resultant spectra with relative line intensities for randomly
oriented polycrystalline samples are illustrated in Fig.2.3.
The quadrupole spectrum is symmetrical only in the case of a
} -+ %transition. In this particular instance there is no direct means
of establishing the magnitude of T/ or the sign of Vzz . However,
as detailed later in the chapter, it is possible to determine these
parameters, either from the angular dependence of the line inten-
sities in single crystals, or from the spectrum observed when a
large external magnetic field is applied.
Once e 2qQ and T/ have been measured, they may be related to
the chemical environment of the resonant nucleus. The coupling
constant e 2qQ is the product of a nuclear constant (eQ) and the
maximum value of Vii (eq). Considerable confusion is often caused
by referring to the sign of Vzz = eq without specifying whether e
is taken to be the charge of the electron or the proton. For this
reason it is preferable to refer to the sign of e 2qQ or of q.
°
If Ig = or} so that Qg = 0, then the magnitude of Qe has to be
established empirically by comparing the observed quadrupole split-
ting values with estima ted values of Vzz. Fortunately, use of the
quadrupole splitting is not entirely dependent on these intrinsi-
cally difficult calculations. The relative order of magnitude of
e2qQ can be related comparatively easily to the valence orbitals
of the atom. The numerical value of Vzz due to an occupied elec-
tron orbital expressed by the wave function lJf is given by the
integral
-
Vzz = q = - (1 -R) JlJf * (3COS 28-1) lJf dr (2.19)
e 41T€Q r3
I \" 345 +1
I
I
\ -2
I \
I \
I \ 12 +1
+1 I -2
-2
+1
-2
I
I
/
19 =12 Ig =1 I
1.---
I
+~
-2
<1:,
",-..
,, +~
-2
+1
-2
li
Velocity Velocity
-V 0 +V
4
! ! 8
:5
Z
6
Fig. 2.3 The energy-level schemes and resultant spectra for quadrupole
t t
splitting of Ig = -+ Ie = ~ and Ig = -+ Ie = (scaled to 1291) transitions. f
I
11=-
q
f IJI* (3 sin e cos 2r!J) IJIdr
2
r3
(2.20)
Hyper/ine Interactions 35
Table 2.1 The magnitude of q and 1/ for the p- and d-orbitals
Orbital 41TEoq/(i - R) 1/
_~(r-3) 0
Pz 5
Px +~(r-3) -3
5
Py +~(r-3) +3
5
d X 2_y2 +~(r-3) 0
7
dz 2 -!(r- 3) 0
7
dxy +~(r-3) 0
7
dxz -~(r-3) +3
7
dyz -~(r-3) -3
7
where e is the angle between the magnetic axis and the major axis
of the electric field gradient tensor. All the magnetic hyperfine
lines are shifted by a quantity
,
I
I
I .l
/
I ........ /
/--,
, 3 5
_e_2_ _ ~""
,,
,, ..... _- " '
........
\
' ....
,, Z 4
,, - - /.... /
1
-- -
......-
,,_____ --L....L.L.++-+-_ +1
" 2
1=1 //
/ "
---<.
9 2 "
"" ,
" ,
,-------....&-.. . . . - _12
Magnetic Magnetic
+
quadrupole
I
Velocity
----'1
-V 0 +V
I~ ---,----'-,-3
' -----'---51
6
1/=0
1/-0,4
c:
o
';;;
'e... ~--
...
oil
c:
I-
-2 o 2 4
Velocity /( mm 5- 1)
(2.24)
ml m 2
L {<hl - ml m II2m 2)2e(l, m) - I (2.26)
Magnetic spectra (M 1)
C C2 e (J, m)
m2 -m1 m (1) (2) (2)
+!. I
+~
2 2
+1 1 4 ~O + cos 2 8)
+!.2 +!.
2
0 vi I
6" ~2 sin 2 8
I
-"2 +!. -1 V~
I
io + cos 2 8)
- .,.
2 12
-z3 +!.
2
I
0 0
+~ -"2 +2 0 0
2
+!.2 2
I
+1 V} I
12 ~(l + cos 2 8)
I
-2"
I
-2 0 vi- I
6" ~
2
sin 2 8
-2
3 I
-2" -1 1 I
"4 io + cos 2 8)
+!. +!. 1
-2' -2 "2 !.
2
+ ~4 sin 2 8
+~ +!. I
~(1 + cos 2 8)
-2' -2 "2
L
mlm2
C 2 e(J, m) = 1
n
42 Principles of Mossbauer Spectroscopy
Random
r
I I
m2 +~
2
+12 -21 +12 -21 -23
+12 +12 +12 1 1
m1 -2 -~ -2
Fig. 2.6 The effect of orientation upon the relative line intensities of a
t
magnetic hyperfine splitting and a quadrupole splitting of a } -+ transition
in an oriented absorber with a unique principal axis sytem.
References
[1] Pound, R. V. and Rebka, G. A. (1959) Phys. Rev. Letters, 3, 554.
[2] Kistner, O. C. and Sunyar, A. W. (1960) Phys. Rev. Letters, 4, 412.
[3] Breit, G. (1958) Rev. Mod. Phys., 30,507.
[4] Reddy, K. R.,Barros, F. de S. and DeBenedetti, S. (1966)Phys.
Letters, 20, 297.
[5] Pound, R. V. and Rebka, G. A. (1960) Phys. Rev. Letters, 4, 274.
44 Principles of M6ssbauer Spectroscopy
[6] Abragam, A. (1961) The Principles of Nuclear Magnetism. Clarendon
Press, Oxford.
[7] Kubo, M. and Nakamura, D. (1966) Adv. Inarg. Chern. and Radiachem.,
8,257.
[8] Ingallis, R. (1962) Phys. Rev., 128, 1155.
[9] Rose, M. E. (1955) Multipale Fields. Chapman and Hall, London.
[10] Condon, E. U. and Shortley, G. H. (1935) The Theory of Atomic
Spectra. Cambridge Dniv. Press.
[11] Greenwood, N. N. and Gibb, T. C. (1971) Mossbauer Spectroscopy,
Chapman and Hall, London.
[12] Housley, R. M., Grant, R. W. and Gonser, U. (1969)Phys. Rev., 178,
514.
CHAPTER THREE
Molecular Structure
45
46 Principles of M6ssbauer Spectroscopy
ticipating in covalent bonding to more than one atom. The unpair-
ed electrons in the valence shells of paramagnetic compounds pro-
duce characteristic effects which are discussed in Chapter 5.
The resonant atom in a diamagnetic compound gives a simple
spectrum, usually with a quadrupole splitting. In the event that
there are several non-equivalent resonant atoms in the structure,
then each should give a different contribution to the spectrum;
indeed the observation of two quadrupole spectra from a com-
pound with two resonant atoms is proof that these are not in the
same environment. The converse, that one quadrupole spectrum
indicates identity of environment, is not necessarily true, since
any differences may be less than the resolution of the measurement.
.,
-1 o 2 3 4
Velocity /(mm 5- 1)
g
\,8
OC - -Fe-H -
I
Fe - -CO
1
o
/\
3
48 Principles of M6ssbauer Spectroscopy
a detectable difference in parameters. In a situation like this it
is often advantageous to consider the Mossbauer data in combina-
tion with other spectroscopic measurements. In the event, the pre-
sence of bridging carbonyl bands in the infra-red spectrum elimi-
nates structure (3), and the total number of bands corresponds to
prediction from the symmetry of structure (I). As may be seen
from Table 3.1, the Mossbauer parameters for [Fe2(CO)gH]- are
similar to those for Fe2(CO)9. The replacement of a bridging car-
bonyl in Fe2(CO)9 by hydrogen causes a reduction in chemical
isomer shift, [) , and an increase in the quadrupole splitting, D.,
exactly analagous to the observed changes in Fe3(COh2 and
[Fe3(CO)11H]-. As a result, the Mossbauer data in conjunction
with the infra-red evidence constitute satisfactory proof that struc-
ture (I) is correct.
The chemical isomer shift is not a direct function of stereo-
chemistry because it involves the electron wavefunctions with
spherical symmetry. Nevertheless, a significant difference in shift
between two related compounds is usually evidence of either a
change in co-ordination number, or in the co-ordinated ligands, or
both. Such a change may have a large effect on the p- and d-
electron shielding or on the bond hybridization. On the other hand,
the quadrupole splitting derives from asymmetric occupation of p-,
d- and f-electron orbitals and is therefore related directly to the
geometry of the compound. A study of these parameters for a
large number of related compounds reveals systematic relationships
* Values in rom s-1. Ii is given relative to iron metal at room temperature. ffars = 1,2-
bis( dimethy larsino )tetrafluorocyclo butene.
Molecular Structure 49
with such features as co-ordination number, site symmetry, and
the nature of the bonding groups. Diagnostic interpretation of the
spectra from new uncharacterized compounds can then become
possible.
As an example, the Mossbauer parameters for iron carbonyls are
not very sensitive to substitution, but the quadrupole splitting is
often indicative of geometry [3]. An iron atom in 6-coordination
tends to give only a small quadrupole splitting as seen in for
example Fe2(CO)9, [Fe2(CO)sH]- and (OChFe(PMe2hFe(COh-
The latter (structure 4) has an Fe-Fe bond to complete the 6-co-
ordination (with a quadrupole splitting of .l = 0.68 mm s-l). In
2
ac,,1
"Fe-CO
if- I
C
o
4 5
o
C
I 2
Oc I 2
~"le-p~e2 I,,-CO
I
Oc ........ 1 ___ PMe,_I __ CO
OC .:--.. ~Fe __ Fe
Me,P-Fe ..... OC- -PMer , .......CO
8 I Co
(, I
C
o
6 7
3090
~.......
-.
A..~
...\ :
~
.
.....
C
: ......".
::I HFe3(CO)10CNMez
<32'30 \
:'
:"'\
..
,',
......
NO
-1 ,0 -0·5 o +0·5 HO
Veloci ty I( mm s-') relat ive to iron
oc co
oc co
0
(j
/Me
OC
I"
Me Ph
/Ph
°0
jP--Me
Me
10
Molecular Structure 53
oc
11
oc co
12
and in one instance the difference in sign between the cis- and
trans-isomers has been verified by the applied field method [13] :
This general rule for distinguishing isomers will remain valid pro-
vided that the geometry is close to octahedral and provided that
the electron distribution in one M-A bond is not significantly influ-
enced by its geometrical relationship to the other, i.e. there is no
large 'trans-effect' (note that a gross distortion of the geometry
may lead to a large value of the asymmetry parameter and/or a
change in the principal axis system of the electric field gradient
tensor).
Similar results have been found in 6-coordinate organotin com-
pounds using the 23.87-keV Ig = ~ -+ Ie = %resonance of 119Sn. An
empirical study of observed quadrupole splittings reveals a charac-
teristic distinction between on the one hand groups bonding
through carbon such as alkyl and aryl groups, and on the other
Molecular Structure 55
hand halides and nitrogen and oxygen bonded donor ligands. In
consequence, a compound such as Ph2SnCl2.phen behaves as if it
were of the type MA2B4 where both the chlorine atoms and the
1,IO-phenanthroline are classed as B ligands. This rather peculiar
effect will be mentioned again later in the chapter. Such com-
plexes also show the 1:2 ratio for cis- and trans-structures as may
be seen from Table 3.3, and it is therefore possible to derive un-
known stereochemistries [14]. The earlier assignment of trans-
phenyl groups to Ph 2Sn(acach is evidently incorrect. The cis-
isomers in chelated tin compounds are usually distorted from an
octahedral geometry. As described on p.67 this may result in a
change in the sign of the quadrupole splitting so that it corresponds
to that of a trans-isomer; however the magnitude of the splitting
still remains approximately the same as expected for a more regu-
lar geometry, and the method is still diagnostic of configuration if
this problem is borne in mind.
As a cautionary note, mention must be made of the syn- and
anti-isomers of the iron cyclopentadienyl derivative Cp(OC)Fe-
(PMe2hFe(CO)Cp (structure 13) which have the cyc10pentadienyl
ring on the same and opposite sides of the Fe2P2 plane respectively.
In this case the 57Fe spectrum is the same in both compounds.
The immediate Fe environment is identical, and is not influenced
significantly by geometrical changes at a much greater distance [3].
Table 3.3 The relationship between the quadrupole splitting and cis-/trans-
isomerism in organotin compounds
Stereochemistry Stereochemistry
Compound by Mossbauer by other methods
13
OM
Fell FeW Fe 3+
-....-
•..,....... 0#\ .............,.., •• ,.#.,••• -
~ "~'-', :'
".
'.
-
(0)
. ".:
II I I ,'.:
-4 -3 -2 -1 0 1 2 3
Velocity relative to Fe/(mm 5- 1 )
o ,.......... . ..
;." .~...~..... :.:....:....,.,...
'\,.
~ •••••0:...., ...;',.........;':.-::•••
". -,
-
,
".
.~
'" .......".
~c
.' -
o .' -
~5- (b) .'
-
..,~
...
-
. .
..... -
.....: I .
.J 1 I
-2 -1 0
Velocity relative to Fe/(mm s-1)
A Crill Fe 2+ 0.6 Fe 2+
B 0.4 Crill, 0.6 Fell Fe 2+ 0.6 Cr 3+
C [Fell (0.6 Fe 3+, Cr3+) ?]
D Fell Cr3+ 0.6 Fe 2+
14
biplanar (structure 14) [18]. It can be shown that the room tem-
perature N.M.R. spectrum results from a fast averaging effect by
cooling the solutions to about 120 K, where the spectrum becomes
more complicated as the molecular motion slows down. Clearly
the molecule can adopt more than one conformation but with
only a very small energy barrier between them which cannot pre-
vent their interconversion at room temperature. However, three
different interpretations of the nature of the low-temperature
'static' form of the complex have been given: a biplanar COT ring
acting as a I ,3-diene as in the solid state [19] ; a tub-shaped COT
ring acting as a I ,S-diene [20]; and a fast isomerization in which
the Fe(COh moiety moves between two equivalent structures in
the I ,3-diene and S,7-diene positions of a tub-shaped COT ring [21].
Although the Mbssbauer effect cannot be observed in a solute
molecule in the liquid phase, it is comparatively easy to freeze a
solution into a glassy matrix where the individual molecules are
not subject to the same nearest neighbour interactions as in the
pure solid. It is imperative of course that the solute should not
precipitate during freezing, and that the solvent should be both
chemically inert and form a glass rather than a well-defined crystal-
line matrix.
Freezing (COT)Fe(COh in several solvents is seen to have a
negligible effect on the Mbssbauer parameters (Table 3.S, [22]).
By way of contrast, the Fe(COh complex with cycloocta-l ,S-
diene and the 1,4-diene with norbornadiene show substantially
different parameters (Table 3 .S), and it is clear that the conjugated
1,3-diene structure found in the solid is retained in the low-
temperature solution.
There are also examples where conformational changes are seen
in solution. The compound (tr-CsHs)Fe(COhSnCI3 shows the
62 Principles of Mossbauer Spectroscopy
Table 3.5 S7Fe Mossbauer parameters in complexes with cyclooctatetraene
Compound Ll* 8*
same S7Pe and 1l9Sn spectra in both the solid state and dispersed
in a glassy polymethylmethacrylate matrix. The Mossbauer para-
meters are not influenced by the nature of the matrix. The S7Pe
spectra of [(1T-CsHs)Fe(COhhSnC12 are also uninfluenced by the
matrix, but the 119Sn spectrum shows four lines in the polymethyl-
methacrylate rna trix as against only two in the solid state [231.
This is illustrated in Pig.3.5. Assignment of the 4-line spectrum
into two quadrupole doublets can be made in more than one way,
but the ambiguity does not affect the prime result, that the 119 Sn
can have two different environments. Conformational effects have
also been reported in solutions of [(1T-C sHs)Fe(COhl SiC1 2CH 3 and
[(1T-CsHs )Fe(COhl SnC1 2 CH3 , and it would appear that in the pre-
sent instance two rotational conformers exist related by a 1200
rotation about an Sn-Pe bond (structures 15 -16). Structure 16
is that found in the solid state.
CI
15 16
Molecular Structure 63
100
99
98
97
100
98
3·0 4·0
(0)
.."
c:
",;;
"e
on
c:
"
L.
I-
(b)
-6 -4 -2 0 2 4 6
Velocity I (mm 5- 1)
Sign of Ll Stereo-
Compound e 2qQ /(mm s-l) chemistry Ref.
r
Me2SnF2 +
Me2SnCI2
CS2[Me2SnCl41
+
+
3.4
465
4.28
~ octahedral
28
26
K2[Me2SnF41 + 4.12 trans-alkyl 26
Me2SnCl2(pyOh + 4.1 29
Pr2SnCI2.2J}picoline + 3.99 27
Me3SnNCS 3.77 ,25
Me3SnOH 2.91 25
Ph3SnF 3.62 25
Ph3SnCI 2.51 trigonal 25
(Me4N)[Me3SnCI21 3.31 bipyramidal 26
(Me4N) [Ph3SnCI21 3.02 26
Et3SnCN 3.17 26
Ph3SnCl.piperidine 2.95 27
Me3SnC6H5 1.39 } { 26
tetrahedral
r
Ph3SnC6F5 0.97 26
Me2Sn(oxinh +
2.06
1.67 }
Ph2Sn(oxinh + octahedral 26
Ph2Sn(NCS)2.1, I O-phenanthro- cis-alkyl
line + 2.36 26
Pr~SnCI2.2 morpholine + 2.41 27
Molecular Structure 67
stereochemistry of new compounds [27]. Ph3SnCl.piperidine
for example has e2qQ negative in sign as predicted for a trigonal
bipyramidal structure with equatorial phenyl groups, and contrary
to prediction for equatorial Cl and piperidine ligands. Pr~SnCI2.2{3-
picoline evidently has trans-propyl groups. However, although
Pr~Sna2.2 morpholine has a quadrupole splitting approximately
half that of the {3-picoline compound, leading one to suspect a cis-
geometry, the sign of e2 qQ is positive in both cases and contrary
to the arguments given earlier in the chapter regarding cis-/trans-
isomerism. Moreover, other cis-complexes such as Me2Sn(oxin)z
for which an X-ray structure is available also show a positive value
of e 2qQ. The problem here is that chelating ligands cause substan-
tial distortion of the octahedral bond angles. The result is that
the principal direction of the electric field gradient tensor has to
be redefined and the value of 11 may be quite large. In this example
it is therefore partly fortuitous, although convenient, that the 2: 1
trans/cis ratio described earlier still holds.
17
CI
CI
CI
"'1/ "'1/
/'"CI / "Br
Br Br
CI
'"
/1", /
/ /1",
CI",
CI
Br
CI
(0) (b)
CI",
/
CI
/1", /1",
"/ Br CI",
"/
/1", /1"
/
Br CI
Br CI CI CI Br CI
(C) (d)
CI Br CI CI"", Br
',/Br
'" '"
/
""'/ "'/ /1" / \
CI / CI / Br CI CI CI
(e) (I)
18
6·07.--...,---,.--;---.--,--.----,-,---,-.,.---,--.--...,---r-,--r-,-"'"1
5·84 , ............
,
f~ 5·61
0° "' ......
'"
~5-38 (0)
<3
5·15
4-92 '---:-8:--'----:6r---'----'4;---'----!2r--...L.----:-
O--'----::-2-J......~4-J......-6~--'----,8!:--'
3077 ...-, ,
.... ... . ~
.;;-
I
~
."
~H4
0
<..>
3069
-16 -12 -8 -4 0 4 8 12 16
Velocity f(mm 5- 1)
Fig. 3.7 The 1291 spectra of (a) the 1Ft cation and (b) the IF6" anion at
90 K. The single line in the former case is consistent with regular octahedral
symmetry, while the quadrupole splitting with a large asymmetry parameter
for IF6" indicates a stereo chemically active lone-pair of electrons. ([ 32],
Fig. O.
Molecular Structure 69
F
c~{>,
F
19
Appendix
VA=~
S R - z
thence
_d_V_~ = __q_A-c:-
dz (R - z)2
and
2
d VA
_ _ s = ___
2qA ::-
dz 2 (R - z)3
and as z -+ 0
A101ecular Structure 71
Vs = 2q A + 2qB + ~ + qB
(R2 + z2)1/2 (r2 + z2)1/2 (r + z) (r - z)
hence
. 2qA 2qB
V (ClS)=- - + - = - V (mono)
zz R3 r3 zz
References
[1] Gibb, T. C. (1970) In Spectroscopic Methods in Organometallic
Chemistry, Chapter 2, p.47, ed. W.O. George, Butterworths, London.
[2] Fannery, K., Kilner, M., Greatrex, R. and Greenwood, N. N. (1968)
Chen!. Comm., 593.
[3] Gibb, T. C., Greatrex, R., Greenwood, N. N. and Thompson, D. T.
(1967) J. Chem. Soc. (A), 1663.
[4] Clark, M. G., Cullen, W. R., Garrod, R. E. B., Maddock, A. G. and
Sams, J. R.(1973) Inorg. Chem., 12,1045.
[5] Herber, R. H., Kingston, W. R. and Wertheim, G. K. (1963) Inorg.
Chem., 2, 153.
[6] Greatrex, R., Greenwood, N. N., Rhee, I., Ryang, M. and Tsutsumi, S.
(1970) Chem. Comm., 1193.
[7] Dahl, L. F. and Rundle, R. E. (1957)J. Chem. Phys., 26,1751.
72 Principles of lIfossbauer Spectroscopy
[8] Erickson, N. E. and Fairhall,A. W.(1965)Ino~. Chern., 4, 1320.
[9] Wei, C. H. and Dahl, L. F. (1969) J. Arner. Chern. Soc., 91, l351.
[10] McDonald, W. S., Moss, J. R., Raper, G., Shaw, B. L., Greatrex, R. and
Greenwood, N. N. (1969) Chern. Cornrn., 1295.
[11] Cullen, W. R., Harbourne, D. A., Liengme, B. V. and Sams, J. R.
(1969) Inorg. Chern., 8,95.
[12] Berrett, R. R. and Fitzsimmons, B. W. (1966) Chern. Cornrn., 91;
(1967)1. Chern. Soc. (A), 525.
[13] Bancroft, G. M., Garrod, R. E. B., Maddock, A. G., Mays, M. J. and
Prater, B. E. (1970) Chern. Cornrn., 200.
[14] Fitzsimmons, B. W., Seeley, N. J. and Smith, A. W. (1969)J. Chern.
Soc. (A), 143.
[15] Maer, K., Beasley, M. L., Collins, R. L. and Milligan, W. O. (1968) J.
Arner. Chern. Soc., 90,3201.
[16] Kerler, W., Neuwirth, W., Fluck, E., Kuhn, P. and Zimmermann, B.
(1963) Z. Physik, 173, 321.
[17] Brown, D. B., Shriver, D. F. and Schwartz, L. H. (1968) Inorg. Chern.,
7,77.
[18] Dickens, B. and Lipscomb, W. N. (1962)J. Chern. Phys., 37, 2084.
[19] Kreiter, C. G., Maasbol, A., Anet, F. A. L., Kaesz, H. D. and Winstein,
S. (1966) J. Arner. Chern. Soc., 88,3444; Anet, F. L., Kaesz, H. D.,
Maasbol, A. and Winstein, S. (1967) 1. Arner. Chern. Soc., 89,2489.
[20] Cotton, F. A., Davidson, A. and Faller, J. W. (1966) J. Arner. Chern.
Soc., 88,4507.
[21] Keller, C. E., Shoulders, B. A. and Pettit, R. (1966) 1. Arner. Chern.
Soc., 88,4706.
[22] Grubbs, R., Breslow, R., Herber, R. H. and Lippard, S. J. (1967)
J. Arner. Chern. Soc., 89, 6864.
[23] Herber, R. H. and 9oscinney, Y. (1968)Inorg. Chern., 7, 1293.
[24] Gibb, T. C. (1970) J. Chern. Soc. (A), 2503.
[25] Goodman, B. A. and Greenwood, N. N. (1971)J. Chern. Soc. (A),
1862.
[26] Parish, R. V. and Johnson, C. E. (1971) J. Chern. Soc. (A), 1906.
[27] Goodman, B. A., Greenwood, N. N., Jaura, K. L. and Sharma, K. K.
(1971) J. Chern. Soc. (A), 1865.
[28] Goodman, B. A. and Greenwood, N. N. (1969) Chern. Cornrn., 1105.
[29] Fitzsimmons, B. W. (1970) 1. Chern. Soc. (A), 3235.
[30] Pasternak, M. and Sonnino, T. (1968)J. Chern. Phys., 48,1997.
[31] Wynter, C. I., Hill, J., Bledsoe, W., Shenoy, G. K. and Ruby, S. L.
(1969) 1. Chern. Phys., 50,3872.
[32] Bukshpan, S., Soriano, J. and Sharnir, J. (1969) Chern. Phys. Letters,
4,241.
CHAPTER FOUR
73
74 Principles of Mossbauer Spectroscopy
tion obtainable from the hyperfine interactions. The primary ob-
servable parameters become the chemical isomer shift and the quad-
rupole splitting, and these are both intimately linked to the covalent
bonding of the molecule.
Because of this distinction, it is convenient to discuss covalency
in diamagnetic substances in some depth before turning to some of
the more exotic phenomena associated with unpaired electrons. As
an introduction, it is instructive to consider, simply, the effects of
a change in the formal oxidation state of the resonant atom.
from Table 4.1, the largest effect is felt in the contribution from
11/I3sCO) 12. A progressive reduction in the number of 3d-electrons
causes a non-linear increase in the total s-e1ectron density.
Unfortunately, the value of the fractional change in the nuclear
radius in equation (2.4), oR/R, cannot be determined independently
of the chemical environment. It is therefore necessary to construct
an empirical calibration of the chemical isomer shift scale using the
observed shifts of compounds whose electron configurations are
known. In this way the sign of oR/R and its approximate magni-
tude are obtained. However, no known compound of iron can be
assigned the free-ion configuration 3d5 or 3d 6 because all Fe 3+ and
Electronic Structure and Bonding: Diamagnetic Compounds 75
Table 4.1 Electron densities at r = 0 for different configurations of an Fe
atom [1,2]
Electrons per
cubic Bohr 3dB 3d7 3d6 3d5 3d 6 4s 2
radius Fe+ Fe 2+ Fe 3+ free atom
2 L
n
Il/Ins(O) 12 11878.9 11879.2 11879.8 11881.7 11885.8
W~ Fe (II) 5=2
D Fe(Il) 5=1
Fe(Jj) 5=0
Fe (III) 5= ~
Fe (III) 5 = ~
Fe(II1) 5=~
Fe (IV) 5=2
D Fe(IV) 5=1
D Fe (VI) 5=1
4.2 Iodine
The fortunate occurrence of suitable Mossbauer isotopes for the
contiguous series tin, antimony, tellurium, iodine and xenon has
greatly facilitated the study of chemical bonding in these elements.
Their compounds are diamagnetic, and the orbitals participating
most strongly in the bonding are the 5s- and Sp-orbitals, with some
contribution from 5d- in appropriate instances. The ability of
iodine to form compounds involving only one covalent bond to
another atom makes it possible to develop a quantitative interpreta-
tion of the Mossbauer parameters which is not always possible for
78 Principles of M6ssbauer Spectroscopy
other elements. In this discussion of covalency, it is instructive to
consider iodine first, before tackling the more difficult problems
which occur elsewhere.
Data can be obtained from both the 27.7-keV 1291 and 57.6-keV
1271 resonances, the former giving better spectra (see Fig.2.1) but
with the inconvenience of being an artificial radioactive isotope.
Both levels are populated by the ~-decay of a tellurium parent, so
that chemical isomer shift data are usually quoted with respect to
the customary source matrix, zinc telluride. The 1291 excited state
spin of Ie = ~ and ground state spin of Ig = ~ are interchanged in
127 1, but otherwise the analysis of the respective quadrupole spectra
are equivalent. The energy-level scheme appropriate to 129 1 is
illustrated in Fig.2.3.
The chemical isomer shift for 1271 or 1291 can be expressed in
terms of the deviation from the closed-shell electronic configura-
tion 5s 25p6 of the 1- anion [5,6]. The nephelauxetic effect of
covalent bonding causes an electron to become less effective in the
hyperfine interactions. In particular, a 5p-electron in a covalent
bond has a very much smaller value of (r- 3 ) so that it appears to
be a 'hole' in the c10sed-shell configuration. This concept has al-
ready been well proven in the interpretation of nuclear quadrupole
resonance (N.Q.R.) data. If we define the number of 5s- and 5p-
electron 'holes' in the 1- configuration to be hs and hp respectively,
the chemical isomer shift with respect to I - can be written as
This follows directly because if the major axis of the electric field
gradient tensor is defined as the z axis, then eq(pz) = -} eq(px, py).
The asymmetry parameter is given by
so that
where in both cases the reference zero is the source matrix ZnTe.
Note the inversion of sign caused by the opposing signs of the
oR/R values.
It can be seen from these equations that, in addition to permit-
ting a determina tion of 5p-im balance in the equivalent manner to
N.Q.R. spectroscopy, the M6ssbauer spectrum also allows a deter-
mination of the 5s-orbital occupation. The two figures in combina-
tion give a much more detailed description of the covalent bonding.
The 129 1 spectrum of solid iodine at 100 K is shown in Fig.4.2
and illustrates the good resolution which can be obtained with this
isotope [9]. The large number of lines in the quadrupole split
spectrum allows an unambiguous determination of both the value
and sign of e2qQ, and also of the asymmetry parameter 17. The
iodine atom in solid iodine has e2q 127Qg/h = -20S5 MHz, but sur-
prisingly has a non-zero asymmetry parameter of 17 = 0.16. This
is unexpected for a linear molecule and is contrary to the naive
assumption that the bonding is purely of p-character. The results
are consistent with admixture of, for example, an s2p 4d configura-
tion into the s2p 5 state as a result of intermolecular interactions.
However, frozen solutions of iodine in hexane, CC4, and solid
argon show a similar spectrum with e 2q 127Qg/h ~ -2250 MHz but
here the asymmetry parameter is zero [ 10]. Clearly in the latter
case a closer approach to a true isolated 12 molecule is found, and
the calculated hp value of ~0.99 is in good agreement with the
predicted value for a free molecule of 1.00.
Having established the calibration of the isomer shift scale
using 1- and the 12 molecule, it becomes feasible to study the
chemical bonding in other compounds. Sample experimental data
are collected in Table 4.2. Some of the numerical values for Up
and hp differ from their original source because all data have been
recalculated using the latest calibration figures.
The value of e2q127 Qg/h in IBr is -2892 MHz, and since this
~
~
(';
.....
d
;::
Velocity/(mm ,-" ;::;.
- 2'13 -1·17 -1'30 -0·88 -0'47 - 0·05 0 + 0'37 +1·20
1 . 80 ir----------r----------r----------r----------,----------r-r--------~--------_,----------._--,
Vl
~
(';
'"="'. ~~. .... <.~.,
2"
4 ~
~1 ' 76 I:::l
;::
2
" ~
1/2
§'" 1·74 5/2 ! 3/2 ttl
8 ex . s -r--- o;::
'0 \
.2: 1-72 512 ~
!j 26·8 k.v 2 3 ~ 5 678 ~'
z
HO 1 1/2
7/2 3/2 e:,
g.s, scum obs \ 5/2
7/2 IS'
1·68 ~
o so 400
Channel number ~
~
(SOchann.l. 4 ·175 mm
= ,-'I .....
;::;.
Fig.4.2 The 1291 resonance in solid 12912 at 100 K. The line numbered 1
9
on the decay scheme is off the scale of the spectrum. It is usually ignored ~o
because of its low intensity. ([ 91, Fig. 3). >::
;::
~
00
82 Principles of Mossbauer Spectroscopy
Table 4.2 Mossbauer data for the 1291 resonance
•
12
I", 1
E
--
E
~
c:
N
2
...
-
.~
a
~ 0
CO
-1~--~----------~----------~----------~----~
o 1000 2000 3000
e 2 q127Qg / MHz
Fig. 4.3 Values for the chemical isomer shift and quadrupole coupling con-
stant in iodine compounds plotted around the line calculated for pure 5p-
bonding. Note the substantial deviations in the iodomethanes due to an
s-character of less than 5%.
Compound S/(mm 8- 1)
relative to
121Sb in Sn02
4.4 Tin
Some of the difficulties encountered in using the 119 Sn resonance
to study molecular compounds were referred to in Chapter 3. In
particular, the failure of many unsymmetrical tin(lV) molecules to
show a detectable quadrupole splitting of the Ie = ~ excited state
is both unexpected and still unexplained quantitatively. The con-
Electronic Structure and Bonding: Diamagnetic Compounds 87
fusion caused by uncertainties in the stereochemistry of many
organotin compounds, and our poor knowledge regarding the ori-
gins of the electric field gradient at the tin nucleus have resulted
in the extensive use of empiricism. However, the problems are no
greater than with any of the other spectroscopic techniques, and
it is gratifying to observe the way in which Mossbauer spectroscopy
has drawn attention to many fundamental issues in the understand-
ing of the chemistry of tin, which in turn it is helping to solve.
Compounds in the tin (II) oxidation state are known to feature,
with few exceptions, a pyramidal bonding arrangement of three
ligands about the tin. Early theories regarding the origin of the
electric field gradient predicted that the sign of e 2qQ in these
compounds could be either positive or negative. However, recent
measurements on SnF 2, SnO, SnS, Sn3(P04h, SnC204, NaSnF3,
NaSn2FS, SnS04, Sn(HC02h, Sn(CH3C02h and K2Sn(C204h.H20
in applied magnetic fields [19, 20] have shown that e2 qQ is posi-
tive in every case. As Q is known to be negative, the sign of e2qQ
is consistent with an excess of electron density in the pz orbital
(for which q is negative, see p.35). Thus the lone-pair has a sig-
nificant pz character which becomes the dominant contribution
to the electric field gradient because of tighter binding to the nuc-
leus. However, neither simple covalent nor point-charge arguments
have been able to predict successfully the magnitude of the gradient.
A salutory warning as to the dangers inherent in the interpreta-
tion of small differences between related compounds is provided
by data for the trihalogeno complexes, [SnXYZ] - where X,Y,Z =
Cl, Br, I. The chemical isomer shifts for the [BulN] + salts of these
compounds are correlated in Fig.4.4 with the average Mulliken
electronegativity of the ligands, XM, and a linear relationship is
found [21]. A similar correlation exists between the chemical iso-
mer shift and the quadrupole splitting. The electric field gradient
is influenced by the polarity of the tin-ligand bonds and increases
with increasing polarization, presumably because the pz character
of the lone-pair becomes in greater imbalance with respect to
the covalent bonding orbitals. The trend in quadrupole splitting
values is MSnI 3 ) = 0.79 mm s-l < A(SnBr3) = 1.02 mm s-l
< A(SnCIJ) = 1.13 mm s-1. This does not necessarily imply that
the pz character has increased.
An unexpected and startling observation is the large difference
in both Mossbauer parameters between the [Bu~N]+ salts and cor-
88 Principles of Mossbauer Spectroscopy
3·0
Z·9
-;
..""
~
>
Z'8
Cl BrI
" Br~
Z'6
-6 -4 -2 0 2 4
Velocity I(mm s-\ )
Fe
Cl
y..
II
= -2[Fe] + 2 (3 sin 2 ~
2
- 1) [Cll
The angles are O! = 128.6 0 and {3 = 94.1 o. The direction and sign of
Vzz is a complex function of the ratio R = [Fe] /[Cl]. The quadru-
pole splitting in a single crystal of the compound shows unequal
line intensities (see p.4D) which establish that the principal axis, z,
is along k. In combination with the observed value for the asym-
metry parameter this leads to a unique value for R of 1 .2. It may
therefore be concluded that the more electronegative halogen
atoms withdraw a greater proportion of Sp-electron density from
the tin atom in a-bonding, and therefore make a smaller contribu-
Electronic Stmcture and Bonding: Diamagnetic Compounds 93
tion than the iron to the electric field gradient.
This compound is also of interest in being one of several for
which the crystal structure shows unusually short Sn-Fe and long
Sn-Cl bond distances. It was originally thought that this was the
result of intermolecular interactions or crystal-packing effects, but
the similarity of the Mossbauer spectra of the pure solid with those
of frozen solutions in inert rna trices confirms that the structure
reflects the internal bonding of the molecule itself.
In molecular orbital terms, one can say that the metal t'4: (3d)
levels and the 1T and 1T* ligand orbitals all have t2g symmetry and
combine to form molecular orbitals with the same symmetry. The
eg (3d) orbitals have the same symmetry as ligand a-orbitals, so
that a large part of n2 will be a-donation from the ligands.
Calculation of the relative importance of these effects is diffi-
cult, but one interesting example concerns a comparison of
[Fe(CN)6]4- and [Fe(CN)6j3- made by Shulman and Sugano [29].
They estimated many of the molecular orbital parameters from
electron spin resonance (E.S.R.) data, and found that the degree of
1T back-donation as typified by n3 is about one electron greater in
[Fe(CN)6]4- than in [Fe(CN)6j3-, while the term n2 remains
about the same. Therefore the value of neff is similar in both com-
pounds, and explains why they have almost the same chemical
isomer shift. Although the hexacyanoferrate(II) nominally has an
extra 3d-electron, its effect is largely nullified by an additional
delocalization of the t2gorbitals to the ligands. However, it is im-
portant to realize that the two configurations are distinct. The
full molecular orbital schemes still feature an unpaired-electron in
the hexacyanoferrate(III), whereas the hexacyanoferrate(II) is
diamagnetic.
The Mossbauer spectrum of a hexacyanoferrate(II) is only a
single line because of the regular octahedral geometry, but the same
is not true of substituted cyanides where a quadrupole splitting
will be seen. A simple example is the so-called nitroprusside ion,
[Fe(CN)s(NO)]2-, which shows a splitting of 1.705 mm s-l [30] .
The ground state configuration for the axial C4v symmetry is
Electronic Structure and Bonding: Diamagnetic Compounds 95
(dxz • dyz)4(dxy )2. The (dxz. dyz) orbitals are strongly delocalized
by 1T-donation to the 2p-orbitals of the nitrosyl group, and there is
also some 1T-donation from the dxy orbital to the equatorial
cyanides.
The electric field gradient may be presumed to originate mainly
from the unequal contributions of these 3d-orbitals. If, as with
the hexacyanoferrate(II), we define effective occupation numbers
nxy etc., it becomes possible to write the value of q in the quadru-
pole splitting e2 qQ in terms of the value for a single d-electron of
(4/7)(1 - R)(,-3)/(41TEO) as
q = (1
-41TEO
- - (r- 3) + -7 nxy
- R) [4 2
- - (n
7 xz
+n
yz
) ]
The value of nxy (incorporating the delocalization to the cyanides)
can be estimated to be approximately 1.74 via the E.S.R. orbital
reduction factor in [Fe(CN)6]3-. Using an estimated value for
the quadrupole splitting from a single d-electron of -4.0 mm s-l,
and the experimental value of 1.705 mm s-l, it is possible to der-
ive (nxz + nyz) = 2.63. If the delocalization to the nitrosyl is
indeed greater than to the cyanide, then this figure for the occu-
pation of dxz and dyz can be used to derive an approximate molecu-
lar orbital
The resultant occupation figure for the 1T* (NO) orbital of 34%
is an upper limit, and is in good agreement with independent esti-
mates of about 25%.
The nitroprusside ion has an unusually low chemical isomer shift
for an iron(II) cyanide complex (-0.258 mm s-l relative to iron
metal at 298 K). The effective 3d-popUlation, neff, is only -4.4,
and there is an unusually high 4s-population of -0.5. These two
factors combine to produce the exceptionally low shift observed.
The related complex ion [Fe(CN)s(NO)]3- has an extra electron
in an orbital of essentially d;- character, and in frozen solutions
its parameters are a quadrupole splitting of Ll = + 1.90 mm s-l and
a chemical isomer shift of 0 = +0.1 mm s-l [31 ]. In this case the
value of q is given by
96 Principles of Mossbauer Spectroscopy
(1 - R) -3 4 2 4
q= - - - ( r >[ +"7nxy -"7(n xz + n yz ) -"7nz2]
47T€Q
One would expect the value of nxy to be little changed from that in
the nitroprusside, and consideration of E.S.R. data leads to an esti-
mate for n z 2 of 0.6. The experimental value for the quadrupole
splitting can then be used to derive the values (n xz + nyz ) '='='- 1.6,
nxy ~ 1.9, nz 2 ~ 0.6, giving neff '='='- 4.1. Although these values are
approximate, the large decrease in (n;xz + nyz) compared to the nitro-
prusside reflects the large increase in 7T-donation to the NO+ ligand
as a result of a formal reduction from Fe(II) to Fe(l). This offsets
the increase in electron density about the central atom associated
with the addition of a non-bonding electron.
It will now be apparent that it is difficult to estimate the relative
effects of a- and 7T-bonding from the comparatively small amount
of information obtainable from a single compound. A more fruit-
ful approach is to compare data for large numbers of related com-
plexes. In Chapter 3 it was shown that cis- and trans-isomers of
low-spin iron(II) compounds may be distinguished by a 1 :2 ratio
of the quadrupole splittings. From data for large numbers of low-
spin iron(II) compounds, the chemical isomer shift has been found
to be an approximately additive property of the number and type
of the ligands [32, 33]. Any contributions to the shift from either
the second-order Doppler shift or zero-point motion are ignored
on the grounds that for comparative purposes they are self-
cancelling in similar compounds. Similarly, the quadrupole split-
ting, ~, can be expressed in terms of partial quadrupole splittings
such that for the three stereochemistries FeAB s , cis-FeA2B4 and
trans-FeA 2B4
(a) based on stainless steel at 295 K as reference. Calculated values for /j can be con-
verted to iron metal as standard by subtracting 0.09 mm s-l.
(b) based on -0.30 mm s -1 for a-as an arbitrary value.
Wave
function
No. of (metal
elec- coeff.
Y net Ll*
14PO>
net Ll*
13dO>
trons
-
....
Cl)
u
~
1/I(e1u) = 0.591J.L(4P1) 4 0.3493 -0.693
+ 0.807 p(eiu)
1/I(a2u) = 0.4 71J.L( 4PO) 2 0.2218 +0.444
+ 0.882p(a2u)
1/I(a1g) = 0.633J.L(4s) + 2 0.4007
+ 0.774p(a1g)
-0.249 -0.811
* The contributions to the quadrupole splitting in units of that from a 14po> or 13do>
electron. Both quantities are defined as negative.
References
[1] Watson,R. E.(1960) Phys. Rev., 118,1036.
[2] Watson, R. E. (1960) Phys. Rev., 119,1934.
[3] Walker, L. R., Wertheim, G. K. and Jaccarino, V. (1961)Phys. Rev.
Letters, 6,98.
[4] Duff, K. J. (1974) Phys. Rev. B, 9, 66.
[5] Hafemeister, D. W., Pasquali, G. de and Waard, H. de (1964) Phys Rev.,
135, BI089.
[6] Perlow, G. J. and Perlow, M. R. (1966) J. Chern. Phys., 45, 2193.
[7] Ehrlich, B. S. and Kaplan, M. (1969) J. Chern. Phys., 50,204l.
[8] Das, T. P. and Hahn, E. L. (1958) Nuclear Quadrupole Resonance
Spectroscopy, Supplement I of Solid State Physics, Academic Press
Inc., New York.
[9] Pasternak, M., Simopoulos, A. and Hazony, Y. (1965)Phys. Rev., 140,
A1892.
[10] Bukshpan, S., Goldstein, C. and Sonnino, R. (1968) J. Chern. Phys.,
49,5477.
[11] Pasternak, M. and Sonnino, T. (1968) J. Chern. Phys., 48, 1997.
[12] Pasternak, M. and Sonnino, T. (1968)J. Chern. Phys., 48, 2009.
[13] Bukshpan, S. and Sonnino, T. (1968) J. Chern. Phys., 48,4442.
[14] Bukshpan, S. and Herber, R. H. (1967) J. Chern. Phys., 46,3375.
[15] Bukshpan, S. (1968) J. Chern. Phys., 48,4242.
[16] Gibb, T. C., Greatrex, R., Greenwood, N. N. and Sarma, A. C. (1970)
J. Chern. Soc. (Aj, 212.
[17] Birchall, T., Della Valle, B., Martineau, E. and Milne, J. B. (1971)
J. Chern. Soc. (Aj, 1855.
[18] Long, G. G., Stevens, J. G., Tullbone, R. J. and Bowen, L. H. (1970)
J. Arner. Chern. Soc., 92,4230.
Electronic Structure and Bonding: Diamagnetic Compounds 101
(19] Gibb, T. C., Goodman, B. A. and Greenwood, N. N. (1970) Chern.
Cornrn., p.774.
[20] Donaldson, J.D., Fihnore, E. J. and Tricker, M. J. (1971) J. Chern. Soc.
(A), 1109.
[21) Goldstein, M. and Tok, G. C. (1971) J. Chern. Soc. (A), 2303.
[22] Herber, R. H. and Cheng, Hwan Sheng (1969) Inorg. Chern., 8,2145.
[23] Clausen, C. A. and Good, M. L. (1970) Inorg. Chern., 9,817.
[24] Carter, H. A., Qureshi, A. M., Sams, J. R. and Aubke, F. (1970)
Canad. J. Chern., 48,2853.
[25] Greene, P. T. and Bryan, R. F. (1971)J. Chern. Soc. (A), p.2549.
[26] Herber, R. H. (1973) J.Inorg. Nucl. Chern., 35,67.
[27] Davies, A. G., Milledge, H. J., Pux1ey, D. C. and Smith, P. J. (1970)
J. Chern. Soc. (A), 2862.
[28] Gibb, T. C., Greatrex, R. and Greenwood, N. N. (1972) J. Chern. Soc.
(A), p.238.
[29] Shulman, R. G. and Sugano, S. (1965)J. Chern. Phys., 42,39.
[30] Danon, J. and Iannarella, L. (1967) J. Chern. Phys., 47, 382.
[31] Raynor,J. B. (1971)J.Inorg. Nucl. Chern., 33,735.
[32] Bancroft, G. M., Mays, M. J. and Prater, B. E. (1970)J. Chern. Soc.
(A), p.956.
[33] Bancroft, G. M., Garrod, R. E. B. and Maddock, A. G. (1971)J. Chern.
Soc. (A), 3165.
[34] Collins, R. L. (1965)J. Chern. Phys., 42, 1072.
CHAPTER FIVE
102
Electronic Structure and Bonding: Paramagnetic Compounds 103
spin iron(II) compounds. However, in many paramagnetic ions
there is a substantial degree of thermal excitation to electronic
levels which are only of the order of kT above the ground state.
The resulting temperature dependence of the orbital occupation
produces characteristic effects on the quadrupole splitting.
The most familiar example of this is the high-spin Fe 2+ ion. The
free-ion configuration 5 D (tigei) signifies a single 3d-electron out-
side a spherical half-filled shell. In a ligand field of cubic sym-
metry such as regular octahedral, the t2g levels remain degenerate
(as do the eg) and there is no finite electric field gradient. How-
ever, if the symmetry is lowered to trigonal or tetragonal, further
degeneracy is removed. The sixth electron is in the appropriate
lowest-lying unfilled state, and now generates an electric field
gradient at the iron nucleus. A ligand field of axial symmetry (i.e.
a tetragonal distortion of the octahedron) splits the t2g state into
a lower d xy singlet and an upper dxz,yz state. The sixth electron
will be in the d xy level and (from Table 2.1) will generate a quad-
rupole splitting in proportion to Vzz/e = q = +~(r-3}(l - R)/(4rre.o).
If the doublet lies lowest, then the electron is in dxz,yz and gives
q = _~(,-3) (1 - R)/( 4rre.o). If the distortion is trigonal, i.e. an
elonga tion along the (111) axis, the singlet ground state corresponds
to dz2 and gives q = -~(r-3}(l - R)/(4rre.o).
A singlet d xy ground state is found for example in
[Fe(H20)6](NH4S04h for which e2 qQ is positive, and a singlet
d z 2 state is found in the trigonally distorted [Fe(H20)61 SiF 6 for
which e 2qQ is negative. One of the few known examples of a
doublet ground state is in trigonally distorted FeC03, and the low-
temperature quadrupole splitting value of +2.06 mm s-1 is about
half that of -3.61 mm s-1 for the fluorosilicate. All three com-
pounds have a distorted 6-coordination to oxygen. The quadru-
pole splitting originates not from asymmetric covalency with the
ligands, but from the sensitivity of the orbital state of an essen-
tially non-bonding electron to the geometrical environment. This
effect provides a powerful means of studying small distortions
from regular co-ordination.
At an infinitely high temperature, all the levels of the t 2g multi-
plet would be equally populated by thermal excitation, and the
electric field gradient would once again be zero. At intermediate
temperatures, a partial degree of thermal excitation will cause a
partial cancellation of the electric field gradient, so that the quad-
104 Principles of Mossbauer Spectroscopy
rupole splitting should decrease with rise in temperature. A full
mathematical treatment of the temperature dependence, (an out-
line of which is given below) was first given by Ingalls [I] .
The effects of the ligand field on the free-ion wavefunctions
are calculated using a perturbation Hamiltonian
where the successively smaller terms are Vo the axial field, VT the
tetragonal (or trigonal) distortion, VR the rhombic distortion, and
'!I.l.§ the spin-orbit coupling. This last term has a considerable
effect on the quadrupole splitting at low temperatures because
it causes admixture of excited states into the ground state. The
free-ion 5D state of Fe 2+ has S = 2 and L = 2 so that there are
(2S + 1)(2L + 1) = 25 energy states in the ligand field. Some typi-
cal energy level schemes from such calculations are illustrated in
Fig.5.1 for the lowest 15 levels only.
If the energy separation of a given excited level is Ej , its relative
population is given by Boltzmann's theory as e-Ej/kT for a ground
state population of unity. The thermal averaging takes place
within a time-scale less than the Mossbauer event (ca. 100 ns for
57Fe), so that the total resultant quadrupole splitting will be a
statistically averaged quantity. Each excited level has an electric
field gradient represented by V'z= eqj and f/j' and the quadrupole
splitting at any temperature T is given by
t:.T = ~e2Q
,,
,,...---2
,',,-----
dxr ,/,,-----
1''--'';';';;'''---''\ /"
,II , , - - - - -
2
1
I \\ /I~---
~'11/~
dlz, yz I
,
\ {,
,,,
,....--""-{
I,I,
\ 1
\ 1 \
,,
\ I
\ \ ___--'-_~I
dyz I
I,
( \\,'----
\
\
\
1------ 2
\
\
\
\
\
\
\
\
\
\
" ----2
\'
\\
"',1----
Cubic Axial Rhombic
field + distortion + distortion 1\
1\
2
'------1
Axial distortion
+ spin - orbit coupling
Fig. 5.1 The effect of axial (tetragonal), rhombic, and spin-orbit coupling
perturbations on the t2g energy levels of the Fe 2+ ion. The level scheme on
the right is a combined axial distortion + spin-orbit coupling, and the num-
bers indicate the degeneracy of the states.
(1 - e-Eo / kT )
I::. T = 1::.0 (1 + 2e-Eo / kT)
1500
Ground
state
d xy
900
600
300
oL__________~==========::==::====~;;~~1~50~ 60 0-
150
Ground
state
600
dxr,yr
-0·5L-------1.-------L------L-----------l co
o 200 400 bOO 800
T/K
Fig. 5.2 Calculated values for the quadrupole splitting flT/ flO = F in units
of the value for a 3d-electron. The curves are drawn for different values of
the splitting of the t2g levels (in cm- 1), and for a spin-orbit coupling para-
meter of A = -80 cm~1.
3.0.-----~--l
'", 2·0
-
E
E
~
c
Ia
~
1·0
o 50
Temperature I K
1000
so that
B=Bo+Bs+BL+BD
where
Bs(x) 2S +
=- 1 (2S + 1)x
- coth - - 1coth (1)
-- x
2S 2S 2S 2S
1 · 0 r - - - - - -_ __
c::
:3 0.6
:a'"...
c::
""
'"
E
t:::>
0·4
~
""
0·2
94
L = 1·0009
TN
88
. ..
82
100
...,<»
0
L-
96
...,""
c::
~ =0·9989
c::
:>
0
U
92
100
96
~ =0·9955
92
88L---~----~----~----~--~----~----~--~
-3 -2 -1 0 2 3
Fig. S.S(a) The Mossbauer spectrum of FeF3 in the vicinity of the Neel
tern perature. ([ 9 ], Fig. 2)
Electronic Structure and Bonding: Paramagnetic Compounds 115
88
84~~----~----~----~----~----~----~--~
-12 -8 -4 0 4 8 12
Doppler velocity / (m m !i-I)
100r:..~~"W-=-
97
94
91
7B·21 K
BB
85
94
94 70'23 K
91 88
Y
88 78·11 K
.
85 ~
~
~ 94
98
.
2-
~
0
67·6 K
~ 8S
96 .'=
~ 94
~
~ 92
Eo
~ 90
g' 88
""g1001-"'.~~~,:..,
u
98
96
94
92
90
n96K
88
100 ,....-_____ 4·2 K
98
BB
96 B
94
Fig.S.? Hyperfine structure in FeF2 when the
92
magnetic coupling is stronger than the quad-
90 rupole coupling. ([ 111, Fig.2).
8B
86
84_U4--L-_2~~-~0--L-~2-~~4-~-6~
Doppler velocity/(mm 5")
0-92 '. :
l'OO~lI.".~"
... ~ .:.... h-. 1""';'-'
-I-"~:
0'98
.
' , . : .-
~~ '::
0·96 : ~ ; i
245 K ': :,
0·94 '; :' ~
l'OOI~':"
0.98 ~ r ~ ..~~
; ... i ::
0'96 ' ; "
•• , l'
:'
.
0·94 236K:: :.;
.2
n~I~'" ~ ,'i ~
~ 0·98
.~
... ~~
".
~
f~·, a~
~:
~ ,'.t#. .•
~
:; 0·96 \
.:.~.
II" ,:
~
., 0·94
'ii
234K':
" p'
;
erc 0'92 ;
1'COI'~.;t:t*-
0-98
. '1 ..
.}I;,
',.'
,i!'"
.il.?
t ~ ~ .'
0·96 . ::
....
~.
.
o!) "."
0-94 230 K <
"
:
0·92 ~ .
l'OOl ~'!\"
0,98 \; . '
,-..~
.'
:;
" :: f
0·96 :;
194K '.
0·94 ::
1'001~
0·98 : .'--
:
0·96 : :
0'94 .:
147K '.
0·92 ~
! ,I ,
-4 . -2 0 4
V.loci\y/(mm s")
0·000
......
. ... .....
c
o
~ 0'020
5l
.0
00·040
":5
'il 0'060 4 Kbor
at
0·080
0'000 .. -..,
c 0'020
o
'z
e-
5l0'040 Low
.g spin
'0
:5 0'060
'z
"~
"- 0'080
65 Kbor
0'100
., .
0000
: ..... .... -.
:5 0'020
:;:;
Q,
g (}040
.0
o
'0
:5 0060
z
"o
.t 0{)80
Fig. 5.9 The 57Fe Mossbauer spectra of MnS2 containing 2 at.% 57Fe
impurity atoms under a pressure of 4,65 and 138 kbar. Note the change
from high-spin Fe 2+ to low-spin Fe 2+ with increase in pressure. ([ 19], Fig,l).
10
Iva 6
e KAu( CNIZi 2
e
:::::: KAu(CN)Z 8r2
'"
~4
KAu(CN)zCl z
Ci.
'"
~2
0.
..,::>
L
'"
::>
etO
-2
-4
-3 -2 -1 o 1 2 3 4 5 6
Chemical isomer shift / (mm 5- 1 )
Fig. 5.1 0 The chemical isomer shifts and quadrupole splittings of some
simple gold compounds of the type AuX, AuX 2, AuX3, AuX4" and AuX2Y2".
Note the linear correlations shown by gold(I) and gold(III) compounds.
+20
+10 NpBr4
NpCl"
Am(Thl NpOZ
0 t..::...::':::"::":::..:F=- NpF"
~pAlz ~...........- - KINpOz'C0 3
-10
(NpOZ'CZ04.2HzO
I", (NpOZ' OH. x HZO
-20
E
E
--
'0
-30
~~~- Rb(NpOZ)(N0 3' 3
NpFS
Na( NpOZ)(CZH30Z13
-40 (NH"IZNPz07' HZO
K3( NpO Z' Fs
-50
NpF6
-60 Ba3(NpOslz. x HZO
CQ3 (NpOS'Z. xHZO
[Coen3J NpOS. x HZO
-70 [Co(NH3'6J NpOS. x HZO
Lis Np 06
-80
-90
Fig. 5.11 A correlation of the 237N p chemical isomer shifts with oxidation
state. The shaded areas represent the predicted ranges. The more positive
shifts in those compounds with CO-<lrdination to oxygen rather than fluorine
are indicative of considerable covalent character in the former. (after [27)).
132 Principles of Mossbauer Spectroscopy
100,0
99·0
96'0 NpF,
(NpHI
97,0
100{)
99 ·0
li 96{)
~
~ 97'0
~.
i!
~ 100-0
co:
99·9
996
99-7
1000
99,5
99-0
Fig. 5.12 The S9.S-keV 237Np spectra of four neptunium compounds with
different oxidation states showing the large chemical isomer shifts, quadru-
pole splitting (N pF 3), and combined quadrupole/magnetic splitting
(KNp02C03 and K3Np02FS)' ([28], Fig.4).
References
[1] Ingalls, R. (1964)Phys. Rev., 133, A787.
[2] Gibb, T. C. (1968) J. Chern. Soc. (A), 1439.
[3] Ingalls, R., Ono, K. and Chandler, L. (1968) Phys. Rev., 172,295.
[4] Nozik, A. J. and Kaplan, M. (1967) Phys. Rev., 159,273.
[5] Campbell, M. J. M. (1972) Chern. Phys. Letters, 15,53.
[6] Dale, B. W. (1971) J. Phys. C, 4,2705.
[7] Edwalds, P. R., Johnson, C. E. and Williams, R. J. P. (1967) J. Chern.
Phys., 47, 2074.
[8] Watson, R. E. and Freeman, A. J. (1961) Phys. Rev., 123,2027.
[9] Wertheim, G. K., Guggenheim, H. J. and Buchanan, D. N. E. (1968)
Phys. Rev., 169,465.
[10] Wertheim, G. K. and Buchanan, D. N. E. (1967) Phys. Rev., 161,478.
[11] Johnson,C.E.(1967)Proc.Phys. Soc., 92,748.
[12] Johnson, C. E., Rickards, R. and Hill, H. A. O. (1969) J. Chern. Phys.,
50,2594.
[13] Konig, E. and Madeja, K. (1967) Inorg. Chern., 6,48.
[14] Jesson, J. P., Weiher, J. F. and Trofimenko, S. (1968)J. Chern. Phys.,
48,2058.
[15] Rickards, R., Johnson, C. E. and Hill, H. A. O. (1968) J. Chern. Phys.,
48,5231.
[16] Cox, M., Darken, J., Fitzsimmons, B. W., Smith, A. W., Larkworth,
L. F. and Rogers, K. A. (1972) J.es. Dalton, 1192.
[17] Drickamer, H. G. and Frank, C. W. (1973) Electronic Transitions and
the High Pressure Chernistry and Physics of Solids, Chapman and Hall,
London.
[18] Simanek, E. and Wong, A. Y. C. (1968) Phys. Rev., 160,348.
[19] Bargeron, C. B., Avinor, M. and Drickamer, H. G. (1971) Inorg. Chern.,
10,1338.
[20] Fisher, D. C. and Drickamer, H. G. (1971) J. Chern. Phys., 54,4825.
[21] Kaindl, G., Potzel, W., Wagner, F., Zahn, U. and Mossbauer, R. L.
(1969) Z. Physik, 226, 103.
134 Principles of Mossbauer Spectroscopy
[22] Potzel, W., Wagner, F. E., Zahn, U. and Mossbauer, R. L. (1970) Z.
Physik, 240, 306.
[23] Wagner, F. and Zahn, U. (1970)Z. Physik, 233, 1.
[24] Faltens, M. O. and Shirley, D. A. (1970) l. Chern. Phys., 53,4249.
[25] Gerth, G., Kienle, P. and Luchner, K. (1968) Phys. Letters, 27A, 577.
[26] Kalvius, G. M., Shenoy, G. K., Ehnholrn, G. J., Katila, T. E., Lounasmaa,
O. V. and Reivari, P. (1969) Phys. Rev., 187,1503.
[27] Frohlich, K., Gutlich, P. and Keller, C. (1972) l.C.S. Dalton, 971.
[28] Dunlap, B. D., Kalvius, G. M., Ruby, S. L., Brodsky, M. B. and Cohen, D.
(1968) Phys. Rev., 171,316.
CHAPTER SIX
Dynamic Effects
(6.1 )
135
136 Principles of Mossbauer Spectroscopy
in the direction of observation j is related to the recoilless fraction
by
(6.2)
Note that <xJ >is the mean value along direction j, so that it is
dependent on direction, whereas <v 2 > is averaged in all directions
and therefore can have only a single value.
Unfortunately the values of <v 2 >and <xJ> at any given tempera-
ture are a complex function of the dynamic motion of the atom.
However, the temperature dependence of either quantity can be
compared with the predictions of a suitable theoretical analysis.
It is customary to consider the dynamics using the harmonic-
oscillator approximation. One such treatment by Housley and
Hess [I] calculates the mean-square velocity and mean-square dis-
placement of the particle in direction j to be given by
(6.3)
~3 Slope = m
3kT
..'"'"
c:
>.
~.'" 2
Intercepts - tm Wj (1)
4
kT 1
Slopes = In w2(-2)
J
Intercepts = 2m
" 1
Wj (-1)
2 kT
(Vj )T-+oo = -- (6.6)
m
138 Principles of Mossbauer Spectroscopy
kT
(6.7)
<xJ>r-> wJ( -2)
0 0 = _ _ __
m
00
(6.8)
(6.9)
(6.10)
2 1'1 1
(Xj )T->O = 2m -W-j(---l-) (6.11)
(6.12)
Dynamic Effects 139
(6.13)
Such an expression for (v 2 ) has been used for example to fit the
second-order Doppler shift of FeC12, the resulting value being
OE= 169 K [2].
Better agreement with experiment can usually be obtained using
a Debye model in which there is a distribution of frequencies. The
number of oscillations with frequency w isN(w), given by N(w) =
3w 2/wg where Wo is the maximum value of w, and 'hwo/k = OD
defines the 'Debye temperature' of the lattice. Thence
m
[i
= 9kO D + (~)4J8DIT x 3dx
OD 0 eX - 1
] (6.14)
2 -
(Xj ) -
311
wgm
IWo [1 + 1]
0 2" ehwlkT _ 1 wdw
9kBD 9kB D
<V 2 >D= - - (6.l7)
8m 16mc 2
N
E
u
co
I
o
Temperature! K
(6.20)
1_3_/2_ =
f +
'If
o
sin 2 e) sin ede
These expressions are strictly valid only for axial symmetry, and
become more complex for a non-axial field gradient because of
state mixing. However, the final result in the more general analy-
sis is still a value of unity.
The above argument is only rigorous if the probability for recoil-
less absorption (or emission) is the same in all directions. If it is
not isotropic, then the intensity of the absorption will be weighted
in favour of a particular orientation of the crystallites. Writing the
recoilless fraction f as a function of e, we have for a thin absorber
144 Principles of Mossbauer Spectroscopy
"\.' .
V ." '" \ f ..'.
~:~':'~':"':..
t=~'5<1O:S ... •
.",\
.!
1'\ 1\ 1foo \.1'J
K
.
•••••,~ .' . ~ \\ \'7/
f
.'
:.. :-~.. 'tJ .;i'~
t=3·5<10··s ...'::-..•.., ~:.:...... \
'" /.: \.. I
"'-'. .~...
.\ 'ti 213K
:t.:...
:'~:e:'. " .......
.' \. '/' ."""'
...-:' /'
§ t=17x10··s
:::
~•
. ~~'..
":..
,'. '\
\U.
'.,. /" ••
...... '-'.-.. .
. j ~. .~ ,"
...
228 K
c. •• • • \
'. . .......~""
....
~. ~ ~.
t=85<10·'Os \ .1'
~ "\ :/'250K
"1--·
. -........e._,TL,..-~~.~.
..... :
\
r-::'"
. .
. . ...~ \\""
. ~.
Source: lS1Sm,O, .~ 1
\ t
,
Absorber: Eu,S,
-
• Experimental spectra
Calculated spectra
~ I
f
~
Statistical error ~/
-20 -10 o 10
Velocity / I mm 5-1)
Fig. 6.3 The 151Eu spectra of EU3S4 between 83 K and 325 K. The weaker
line at 85 K is due to Eu 2+, and the stronger line to Eu 3+. The fast electron
hopping which occurs at higher temperatures results in an averaged spectrum,
and the solid curves are computed with a relaxation model whose time con-
stant is T. ([ I 31, Fig. l).
Dynamic Effects 147
Fe 2+ and 50% of the Fe 3+ cations combine to give an averaged
contribution [14, 15].
The phenomenon of atomic diffusion is more difficult to
study. If the nuclei jump instantaneously between lattice sites,
then the effect should be manifested as a broadening of the reso-
nance line without a change in recoilless fraction. If 70 is the mean
resting time between jumps, then the fractional increase in the
width of the energy distribution at half-height is given by
Ar = 212/70 (6.23)
(6.24)
...... .
.- ....
100 .' ... ~.II',.':
.... ..:\:..,.. .."r'~.
..-. :.··;:.:......
,
99 300 K
98
. '.~.,
r."ot' I I - . : ....
... ....
.,...~•• ; - .·I~"·
•.." ...".....,..,...... ...;:.
100 ... ..... , ..." ,
.,.... .. ___........ "" ••-~
~~
,... -.:.
."..
78K
99
...-.......:•~.
f.ro::,1'01;;...•·t,..... .;.: .;:........•••• ';0'-.:","':" .
..
100 ~ •...,~.... '. "~'" :r\ -•
AI._ .. .." - . . " ••-
99 54 K
, ,
...'~V"A.:·.. ._.,;" -:.".' ..,"~
...
....
'" ..."":
.. ..-
.to .. .,
...
.. .. , .. -
.: .... "
.' .. ~
99 : '. :
20 K
..
.-~
"" _'_..'.' e.
. .' .
•• ..1 ••
~
'.
99
4·2 K
98 L-------~
20-0-------1~0~
0 ------0~----~10~O~--~2~O~O----~
Voloc lty/ Imm ,-1)
r.':"~~:~;.~..
' ..
'
,:,..
c: ., "'
o
'.~ ;",,"'..;. / ~ :
: '. !,
'E
II>
c:
C
.....L. (0) (b) ',' (c)
-2 -1 0 1 2 -2 -1
.
..
0 1 2 -2 -1 0 1 2
Velocity/( mm S-1) Velocity/( mm s- '} Velocity/( mm S-I}
1
If
2.1l
.Il
~ Itt>
I
±S/2
I
I
, ,
... '
,,
±3/2 I
o ....•.
.' I I I' I'
· . >~S~. ±1/2
\
....
~ ~.
c:
....0-
.Q
I-
oof)
.J:J
«
-10 -5 0 5 10
Velocity/(mm 5-')
Fig.6.7 The 57Fe spectrum at 77 K of 0.14 at.% Fe 3+in A1203. The pre-
dicted lines for the three Kramers' doublets are shown as bar diagrams, but
the I±~> contributions are not resolved. ([301, Fig. 2).
156 Principles of Mdssbauer Spectroscopy
times and causes collapse of the magnetic splittings, and this pro-
cess is largely complete at 1.4% Fe.
Of the other common iron electronic configurations, low-spin
iron(lII) shows magnetic effects only if magnetically dilute «5%)
to prevent spin-spin relaxation and at low temperature to slow the
spin-lattice relaxation. Such conditions have been achieved by
doping K 3Fe(CN)6 into K 3Co(CN)6 [31] .
6.S Superparamagnetism
Having considered the case of paramagnetic ions which can show
magnetic hyperfine interactions, it is appropriate finally to con-
sider the inverse situation of a magnetically ordered material which
appears to be paramagnetic. Using a naive description, one can say
that a magnetically ordered ion with Sz :::: ~ will still fluctuate bet-
ween the Sz :::: +~ and Sz :::: -~ states as for a paramagnetic ion,
but that the spin will statistically spend more time in one state
than the other. In magnetically ordered solids the spin-exchange
interactions are strong. The spin-wave description of the magnetism
results in an averaged value (Sz) which is responsible for the magne-
tic hyperfine splitting. This value of (Sz) is defined along the
direction of magnetization and has a long relaxation time because
the co-operative interaction is long-range. Thus a magnetically
ordered solid usually shows a static hyperfine splitting at all tem-
peratures below the ordering temperature.
The value of (Sz) is not, however, a static quantity, and within
each magnetic domain fluctuates with a relaxation time 7
:::: 70 exp (KV/kT) where KV is effectively the energy barrier to be
overcome in changing the direction of the magnetization, and kT
is the thermal energy. Since V is the volume of the domain, K
is the energy per unit volume. Thus the relaxation time can be
decreased by decreasing the volume of the domain or by raising
the temperature. Provided that the relaxation rate is slower than
the Mossbauer lifetime then the full hyperfine splitting is seen,
but if the two are comparable then a relaxation narrowing can be
expected.
This effect has been clearly demonstrated in several iron oxides
such as a-Fe203 [32] and a-FeOOH [33] by decreasing the par-
ticle size and thus the magnetic domain size to the order of <200
A (20 nm). Although the magnetic hyperfine splitting is still
Dynamic Effects 157
present at low temperatures, it is seen to collapse with increase in
temperature to give the spectrum characteristic of the paramagne-
tic phase well before the Neel temperature is reached. The finer
details of the relaxation are often obscured because most methods
of generating small particles tend to give a distribution in particle
size with a consequent distribution in the relaxation time.
Probably the most common type of magnetic relaxation occurs
in non-stoichiometric compounds where the ordered cations on a
particular lattice site have been diluted with a non-magnetic sub-
stituent. This weakens the exchange interactions to the point
where they become localized to small clusters of cations. It is
thus analogous to the small particle-size effect, and a contributory
cause of the collapse of the magnetic hyperfine splitting. Several
examples will be found in Chapter 7.
References
[1] Housley, R. M. and Hess, F. (1966) Phys. Rev., 146,517.
[2] Johnson, D. P. and Dash, J. G. (1968)Phys. Rev., 172,983.
[3] Hazony, Y.(1966) 1. Chern. Phys., 45,2664.
[4] Lafleur, L. D. and Goodman, C. (1971) Phys. Rev. B, 4,2915.
[5] Housley, R. M. and Nussbaum, R. H. (1965),Phys. Rev., 138, A 753.
[6] Housley, R. M., Gonser, U. and Grant, R. W. (1968) Phys. Rev.
Letters, 20,1279.
[7] Housley, R. M., Grant, R. W. and Gonser, U. (1969) Phys. Rev., 178,
514.
[8] Grant, R. W., Housley, R. M. and Gonser, U. (1969) Phys. Rev., 178,
523.
[9] Goldanskii, V. 1., Gorodinskii, G. M., Karyagin, S. V., Korytko, L. A.,
Krizhanskii, L. M., Makarov, E. F., Suzdalev, 1. P. and Khrapov, V. V.
(1967) Doklady Akad. Nauk S.S.S.R., 147, 127.
[10] Karyagin, S. V. (1963) Doklady Akad. Nauk S.S.S.R., 148, 1102.
[11] Herber, R. H. and Chandra, S. (1970)J. Chern. Phys., 52,6045.
[12] Wickman, H. H. and Catalano, E. (1968) J. Appl. Phys., 39,1248.
[13] Berkooz, 0., Malamud, M. and Shtrikman, S. (1968) Solid State Cornrn.,
6,185.
[14] Banetjee, S. K., O'Reilly, W. and Johnson, C. E. (1967) 1. Appl. Phys.,
(1967) 38,1289.
[15] Hargrove, R. S. and Kundig, W. (1970) Solid State Cornrn., 8, 803.
[16] Knauer, R. C. and Mullen, J. G. (1968) Appl. Phys. Lett., 13,150.
[17] Knauer, R. C. (1971) Phys. Rev. B, 3,567.
[18] Cameron, J. A., Keszthelyi, L., Nagy, G. and Kacs6h, L. (1971) Chern.
Phys. Letters, 8,628.
[19] Dilorenzo, J. V. and Kaplan, M. (1968) Chern. Phys. Letters, 2, 509.
158 Principles of M6ssbauer Spectroscopy
[20] Wickman, H. H. and Wertheim, G. K. (1968) In Chemical Applications
of Mossbauer Spectroscopy, Chapter 11, ed. V. I. Goldanskii and
R. H. Herber, Academic Press, New York.
[21] Eibschutz, M. and van Uitert, L. G. (1969) Phys. Rev., 177,502.
[22] Wickman, H. H. and Nowik, I. (1967)J. Phys. Chern. Solids, 28,2099.
[23] Ofer, S. and Nowik, I. (1967) Nuclear Phys., A93, 689.
[24] Seidel, E. R., Kaindl, G., Clauser, M. J. and Massbauer, R. L. (1967)
Phys. Letters, 25A, 328.
[25] Wignall,J. W. G.(1966)J. Chem. Phys., 44,2462.
[26] Cox, M., Fitzsimmons, B. W., Smith, A. W., Larkworthy, L. F. and
Rogers, K. A. (1969) Chem. Comm., 183
[27] Blume, M. (1967) Phys. Rev. Letters, 18,305.
[28] Buckley, A. N., Herbert, I. R., Rumbold, B. D., Wilson, G. V. H. and
Murray, K. S. (1970) J. Phys. Chem. Solids, 31,1423.
[29] Fitzsimmons, B. W. and Johnson, C. E. (1970) Chem. Phys. Letters, 6,
267.
[30] Johnson, C. E., Cranshaw, T. E. and Ridout, M. S. (1964) Proc. Int.
Cant on Magnetism, Nottingham, (Inst. Phys. Physical Soc., London,
1965), p. 459.
[31] Oosterhuis, W. T. and Lang, G. (1969)Phys. Rev., 178,439.
[32] Kundig, W., Bammel, H., Constabaris, G. and Lindquist, R. H. (1966)
Phys. Rev., 142,327.
[33] van der Kraan, A. M. and van Loef, J. J. (1966) Phys. Letters, 20,614.
CHAPTER SEVEN
159
160 Principles of M8ssbauer Spectroscopy
plicated derivatives. Some topics involving dynamic effects such
as electron hopping, paramagnetic relaxation and superparamagne-
tism have been described separately in Chapter 6.
I
I
I
I I I
I /- I
;'
I'-"---_ r " <
~ ----~-/~~---- -~
--+~ -+-jq-;;--~;' :
I B I
I I
I I
I I
Fig. 7.1 The normal spinel structure, A [B21 04, contains eight octants of
alternating A04 and B404 units as shown on the left; the oxygens build up
into a face-centred cubic lattice of 32 ions which co-ordinate A tetrahedrally
and B octahedrally. The unit cell, Ag[B161 032, is completed by an encom-
passing face-centred cube of A ions, as shown on the right in relation to two
B404 cubes.
3·0
2·5
C
2·0
, D 'E,
~ 1·5
E
II Sr'g SEg
6 ~
Free Ion Octahedral Trigonal 19
1·0 field field
0·5
°0~--~----~----~--~-----10~00--~
100"--~"'-- 294K
98
..
96
94 170K
92
90
.......~...~'.;..~J.:~'
.l 423 K
.'
..
..
'
I I
,.'
" , , I
-1 0 1 2
Velocity/(mm s-')
3
"
Fig. 7.4 The Mossbauer spectrum of FeO.9400 recorded during the precipi-
tation of Fe304 by disproportionation. The sample was held at successive
temperatures for about 2 -3 hours. Note the gradual appearance of an intense
line due to Fe304 on the right, and the narrowing of the Fel_xO line as the
phase tends towards stoichiometry. The line components due to Fe304 are
indicated. ([ 101, Fig. 2) .
168 Principles ofM6ssbauer Spectroscopy
upon the growing Fe304 resonance. Variations in the intensity of
the various spectra with time enable the kinetics of the reactions
to be followed. The spectra of Fel_xO at 1074 K where the com-
pound is thermodynamically stable comprise an unsplit broad line
because of rapid diffusion of the vacancies (see Chapter 6).
6'
Pen) = ---' -- (1 - c)6-n c n
n!(6-n)!
I ,
( l
o ~
I
c:
o
';;;
'"
'e .'.
.,c:'" ....
~
Co FtZ04 quenched
-9 -8 -7 -6
Velocity I (mm 5- 1 )
Fig. 7.5 One of the lines in the Mossbauer spectrum of CoFe204 (quenched
sample) in an applied flux density of 1.7 T. The A site is relatively unaffected
by the cation distribution, but contributions from iron atoms at B sites with
0, I, 2 and 3 Co2+ cations on the six nearest neighbour A sites are identified.
([9], Fig. 4).
Oxides and Related Systems 171
the magnetic cations are diluted by a non-magnetic substituent.
At low temperatures the spectrum of such an oxide containing Fe 3+
ions is usually a six-line spectrum, if somewhat broadened, but as
the temperature increases there tends to be a progressive inward
collapse, often with the appearance of an apparently paramagnetic
central component before the ordering temperature is reached.
This is illustrated in Fig. 7.6 for some data for (Zno.sFeo.s)
[Co o.s Fel.5] 0 4 where the most probable cation distribution is of
three Fe 3+ and three Zn 2+ cations about each B site [12]. How-
ever, if the cation distribution is truly random, no less than 11 %
:. ,
:,:.,
! " ;:\:,
.~ ""!<:< : .,; .;!~t{'
l~'f~q~ ;.!tt.~.:~:~
/'1:/387 K
:.",~i'f,,'!')·,,//\r\,:~ 222 K
,..,.
'il? 367 K
:~~4i:~~-) :. a.:.;JJ.:,,-::.;
J '
Velocity
85
':.:
(Q)
80
51
-',: ;,'.... .....~. . :::.:...........0.-·
.:~ : ....
.
-:' ::.:.. .:::01'::" .::,.•...: ..:.::. ~.:.........:':. .... . .,.:.......~:- '::.:
'"
I 49 "
S2
....
" :
~ 47
C
:::I
o
(.)
(b)
.......... ...... :..._.
192 - ••.:..,........... ,
... ',' .....
...........'_.
0" ,"
..
.....
..··0, ....... .:..: ...... .,....... ..... . '.,~': :"'.
168 :.:
t i
184 A B BA
180
(e)
-10 -8 -6 -4 -2 0 2 4 6 8 10
Velocityf(mm s-')
Fig. 7.7 57Fe Mossbauer spectra for CoFe204 at 4.2 K in (a) zero magnetic
field (b) a flux density of 8 T applied along the direction of observation and
(c) C00.4ZnO.6Fe204 at 4.2 K in an externally applied flux density of 8 T.
The variation in the intensities of the outer lines allows a determination of
the site occupancy, and the weak intensity of the Ilm = 0 lines in (b) and
particularly in (c) indicates the presence of spin canting. ([ 14] , Figs. 1, 3).
174 Principles of Mossbauer Spectroscopy
(ZnO.6Fe0.4) [CoO.4Fe 1.6] 04· As can be seen the cobalt is quickly
displaced from the A sites by zinc. The f:..m = 0 lines are weakest
in CoFe204, but increase rapidly in intensity above x = 0.4, and
are very intense for x = 0.8 (not shown). Although CoFe204
approximates to a Neel ordered ferrimagnet, spin canting is pre-
sent at both sites and increases with zinc substitution. Detailed
study of the lineshapes for x = 0.8 provides evidence that individual
B-site spins become increasingly canted as the number of Zn 2+ ions
in the nearest neighbour A sites increases. Therefore in any given
sample, the spins are canted by different angles with a sta tistical
distribution.
;. : ..........
'.'
',:. ::'
.:": x =0,15
.. ':'
'.'
.. ..
....: :
(.'r.:...: . . . :.:..":. ~':""':. .. . . . .
....r
~: . ':...
. :~:. . " ....'.... x =0-75
• ':.0.:
",'::.:'
. ........... ...... -:.~:':. ',::',. .
'0.:-, ............
.. . '::.' .
.:.:::: ',:::
c:::
"
'~ " .--. ....
x = 1·2
·e "", '',..
".
','
'"c:::
"'-
I-
. ....
. : / . ........ ,,',',,:. x=1·5
.. :.' ' . :'... • . ... 0'
" ",
Velocity
the europium and iron sites. All the c-sites are not in fact equiva-
lent because the iron sub1attice magnetization is in the [111] direc-
tion, and there are two 151 Eu hyperfine flux densities of 56.1 and
62.3 Tat 4.2 K [19]. Substitution of gallium for iron takes place
preferentially at the d sites to give {EU3} [Fe2] (GaxFe3-x)012. The
exact site occupancy can again be determined from the s7Fe spectra,
and in fact 85% of the gallium is in d sites for x = 0.66 and 73% for
x = 3.03 [20]. It is interesting to note that the a- and d-site mag-
netizations are opposed to each other, and the net magnetization
vanishes at X""" 1.4 because of 'magnetic compensation'.
The lSlEu resonance at 4.2 K is altered dramatically with increase
in x, and as can be seen in Fig. 7.1 0, the magnetic splitting collaps-
178 Principles of Mossbauer Spectroscopy
es completely. The europium field is generated by an exchange
interaction with the two nearest-neighbour Fe atoms at d sites,
which will therefore be weakened by the gallium substitution.
The observed spectra can be simulated successfully using a simple
model. Each europium ion can have one of three environments:
(a) two Fe 3+ neighbours at d sites (statistical weight y2 where y is
the concentration of iron on d sites determined from the 51Fe
spectra); (b) one Fe 3+ and one Ga 3+ (weight 2y[ 1 - y]); (c) two
Ga 3+ neighbours (weight [1 - y]2). The exchange mechanisms
produce a flux density at the Eu 3+ of Be, O.52Beand 0 respectively.
The solid curves in the figure were calculated using essentially
1'14
1'12
0,42
0'41
0·62
X=1'26
0'60
§0'59
~
.~ X=I'60
c:
E
... 0'57
0'55
X=2,37
0·53
1,40
X=3·03
1·37
'"
~
!
c:
"
o
u
- 15 o 15
Velocity/(mm s-' )
order between x = 0.9 and 1.2 results in a rapid collapse of the flux
density at both the 57 Fe and 119Sn nuclei.
The cation Sb 5+ is isoelectronic with Sn 4+, and transferred hyper-
fine interactions are also known for antimony in the spinels
Nil+2xFe2-3xSbx04 [25]. The antimony substitutes at the octa-
hedral B sites, but here the average number of Ni 2+ neighbours at
A sites alters with increasing substitution. The result is a more
complex spectrum in which each possible combination of A-site
neighbours produces a different hyperfine flux density at the
antimony.
Transferred hyperfine interactions can also occur at the anion,
and an example of this can be found in the ferromagnetic spinel
CuCr2Te4 where the 125Te resonance reveals a flux density of 14.8
T at 80 K [26] .
Oxides and Related Systems 181
References
[1] Evans, B. J., Hafner, S. S. and Weber, H. P. (1971)J. Chern. Phys., 55,
5282.
[2] Eibschutz, M., Ganiel, U. and Shtrikman, S. (1966) Phys. Rev., 151,
245.
[3] Gibb, T. C., Greatrex, R., Greenwood, N. N., Puxley, D. C. and
Snowdon, K. (1973) Chern. Phys. Letters, 20, 130.
[4] Tanaka, M., Tokoro, T. and Aiyama, Y. (1966)J. Phys. Soc. Japan,
21,262.
[5] Ono, K., Chandler, L. and Ito, A. (1968) J. Phys. Soc. Japan, 25, 175.
[6] Eibschutz, M., Hermon, E. and Shtrikman, S. (1967)J. Phys. Chern.
Solids, 28, 1633.
[7] Armstrong, R. J., Morrish, A. H. and Sawatzky, G. A. (1966) Phys.
Letters, 23,414.
[8] Gallagher, P. K., MacChesney, J. B. and Buchanan, D. N. E. (1964)
J. Chern. Phys., 41,2429.
[9] Sawatzky, G. A., van der Woude, F. and Morrish, A. H. (1969)Phys.
Rev., 187, 747.
[10] Greenwood, N. N. and Howe, A. T. (1972) 1. Chern. Soc. (A), pp. 110,
116 and 122.
[11] van der Woude, F. and Sawatzky, G. A. (1971)Phys. Rev., 4, 3159.
[12] Iyengar, P. K. and Bhargava, S. C. (1971)Phys. Stat. Sol. B, 46, 117.
[13] Coey, J. M. D. and Sawatzky, G. A. (1971) Phys. Stat. Sol. B, 44,673.
[14] Pettit, G. A. and FOf€!ster, o. W. (1971) Phys. Rev. B, 4,3912.
[I5] Gilleo, M. A. and Geller, S. (1958) Phys. Rev., 110, 73.
[16] van Loef, J. J. (1968) 1. Appl. Phys., 39, 1258.
[17] Lyubutin, I. S. and Lyubutina, L. G. (1971) Soviet Physics-Crystallo-
graphy, 15,708.
[I8] Lyubutin, I. S., Belyaev, L. M., Vishnyakov, Y. S., Dmitrieva, T. V.,
Dodokin, A. P., Dubossarskaya, V. P. and Shylakhina, L. P. (1970)
Soviet Physics-JETP, 31,647.
[19] Stachel, M., Hufner, S., Crecelius, G. and Quitmann, D. (1969) Phys.
Rev., 186,355.
[20] Nowik, I. and Ofer, S. (1967) Phys. Rev., 153,409.
[21] Baurninger, E. R., Nowik, I. and Ofer, S. (1967) Phys. Letters, 29A,
328.
[22] Atzmony, U., Bauminger, E. R., Mustachi, A., Nowik, I., Ofer, S. and
Tassa, M. (1969) Phys. Rev., 179,514.
[23] Goldanskii, V. I., Trukhtanov, V. A., Devisheva, M. N. and Belov, V. F.
(1965) ZETF Letters, 1, 19.
[24] Belov, K. P. and Lyubutin, I. S. (1966) Soviet Physics-JETP, 22,518.
[25] Ruby, S. L., Evans, B. J. and Hafner, S. S. (1968) Solid State Cornrn.,
6,277.
[26] Ullrich, J. F. and Vincent, D. H. (1967) Phys. Letters, 25A, 731.
CHAPTER EIGHT
182
Alloys and Intermetallic Compounds 183
be a contribution from unpaired conduction-electron spin-density
which usually has considerable s-character and therefore also con-
tributes significantly to the observed magnetic flux density. This
is referred to as the 'conduction-electron polarization'. The rela-
tive importance of these two effects and the way in which they
are influenced by change in composition of the alloy can often
be determined by Mossbauer spectroscopy. The temperature
dependence of the magnetic flux density below the ordering tem-
perature is essentially similar to the behaviour found in ionic com-
pounds (see p. 112) and will not be discussed further.
It is convenient to consider any metallic phase as belonging to
one of two classifications. A disordered alloy is one in which two
or more elements occupy the same crystallographic sites with a
random probability. Such a phase frequently has a wide range of
composition. An intermetallic compound differs in that each ele-
ment shows strong preference for a particular lattice site. The
result is a regular structure with a composition close to a simple
stoichiometry and a very restricted range of composition. How-
ever, there are intermediate examples where an alloy shows a vari-
able amount of atomic order/disorder according to its previous
thermal and mechanical history.
1 n nr,!0 (, ,',.-
:i V ~. V \' \. ,
~
'{
i V J/
"'" ", r \ ,-... r~: r'\
'. r;it....-.. r
\/ \/
~: ".
V V \:
Pt ; j
\!
:;
~ +. "-; ~
i
-6 -4 -2 0
V.locity I (.,," ,-I)
Fig. 8.1 The 57Fe spectra at room temperature of alloys derived from a-iron
by substitution of (top to bottom) 2.2 at.% Cr, 1.0 % Mo, 5.0 % V, 3.0 % Pt.
([2], Fig. I) .
-
I-..
z
HS
c:
HO
o
'Vi
'"
'£1
VI
c:
d
~
3-48
3-44 ErFez
I' I
I
-4 -2 0 2 4
Velocity I (mm S-I)
Fig. 8.3 The 57Fe Mossbauer spectra at 77 K of HoFe2 and ErFe2 showing
that only one iron site exists in the former, but two in the latter because of a
different direction of the magnetic axis. ([6], Figs. 1,3).
A~Mn ~H~
c/o ·2·55 c/o - 2·56
Fig.8.4 The structures of AU2Mn, AuMn2, and the t1 and t2 phases of AuMn.
The manganese spins are aligned parallel within any shaded plane, and anti-
parallel in adjacent planes (ignoring any spiral effects) to give an antiferro-
magnetic structure. AuMn2 is non-magnetic.
ture (40-53 at.% gold). Typical 197Au spectra at 4.2 K are shown
in Fig. 8.6. Upon cooling, the cubic high-temperature phase trans-
forms initially to a tetragonally distorted t1 phase (cIa < 1), but
for less than 50 at.% Au there is a further phase transition to a
second tetragonal phase, t2 (cIa> 1), without any change in cell
volume. Both tetragonal phases are antiferromagnetically ordered
with alternating ferromagnetic sheets of Mn atoms perpendicular
to a short axis (see Fig. 8.4). In the ordered 50 at.% alloy the
1·01, ~
......
~
..,
-
I::l
;::
I::l..
~
~
~
~, :::::
;;; ("')
-.
c:
.~
'" 0·98 g
'"
-}; ~
'1;:j
0
0; a
:::::
'" ;::
0·97 I , ~
AU2Mn
0·96
I
0·95' , , " '"
- 12 -8 -4 0 ~. - .-
Velocity/(mm 5· ')
Fig.8.5 The 77.34-keV 197 Au resonance at 4.2 Kin AU2Mn. The eight component lines are
from an Ig = -} ~ Ie = transition with E2/M 1 mixing. The magnetic hyperfine splitting appears
1.
asymmetnc because ot a small quadrupole interaction. ([9), Fig.2).
.-
\0
W
194 Principles of Mossbauer Spectroscopy
.. J......:..
. .
,..
:
\ ... .. ' .-.;, . ..... . . -',- , . '.-
.r .'
" _:,' '... _',- 53'/. Au
, :
'r '.
"
- 10 -5 0 10 15
V.loc;ty/(mIlU-')
100
.
99 (a) ' \ r'F~:1
98 annealed
97
96
9S I
(;1
100
98
(b)
96 cr~5hed
94
co 92
'iii 90 , .. .
.~ 100
c
.-. .,... .
"
~ 98 (c)
Fe,., Al 0.Q
96 annealed
9-4
92
90
88
86
84 -8 -6 -4 -2 0 2 4 6 8
Velocity f(mm 5-')
Fig. 8.7 The 57Fe spectra at 4.2 K of (a) ordered FeA1 (b) crushed FeA1
(c) Fel.1A10.9. ([10], Fig. 2).
196 Principles of Mossbauer Spectroscopy
stic of an inhomogeneous magnetic hyperfine interaction. Plastic
deformation of the body-centred cubic lattice will generate slip
planes, giving antiphase boundaries or stacking faults containing
an excess of like-atom pairs. This increase in the number of con-
tiguous iron atoms is responsible for the magnetic behaviour.
An indirect verification of this interpretation can be obtained
by examining a non-stoichiometric ordered alloy of the type
Fel+xAll-x. Each iron atom on an Al site has eight Fe nearest
neighbours, while each Fe site has dominantly between 6 and 8 Al
nearest neighbours. The spectrum for an alloy with x = 0.1 is
shown in Fig. 8.7. The magnetic area in the wings of the spectrum
(about 9% of the total area) corresponds closely to the fraction
of iron on AI sites which are thus centred in an Fe9 cluster. The
eight neighbouring iron atoms on Fe sites are less strongly coupled,
but are nevertheless affected as seen by the broadening of the cen-
tral component. The ferromagnetic behaviour of nearly stoichio-
metric FeAl is therefore determined by the presence of magnetic
clusters, the number of which may be varied by appropriate treat-
ment of the alloy.
References
[1] Cordey Hayes, M. and Harris, 1. R. (1967) Phys. Letters, 24A, 80.
[2] Vincze, I. and Campbell, I. A. (1973) J. Phys. F, 3, 647.
[3] Cranshaw, T. E. (1972)J. Phys. F, 2, 615.
[4] Stearns,M. B.(1963) Phys. Rev., 129,1136.
[5] Tansil, J. E., Obenshain, F. E. and Czjzek, G. (1972)Phys. Rev. B, 6,
2796.
[6] Bowden, G. J., Bunbury, D. st. P., Guimaraes, A. P. and Snyder, R. E.
(1968)J. Phys. C, 1,1376.
[7] Bowden, G. J. (1973) J. Phys. F, 3,2206.
[8] Longworth, G. and Window, B. (1971) J. Phys. F, 1,217.
[9] Thompson,1. 0., Huray, P. G., Patterson, D. O. and Roberts, L. D.
(1968) In Hyperfine Structure and Nuclear Radiations, ed. E. Matthias
and D. A. Shirley, p. 557, North Holland, Amsterdam.
[10] Wertheim, G. K. and Wernick, J. H. (1967) Acta Met., 15,297.
CHAPTER NINE
Analytical Applications
197
198 Principles of M6ssbauer Spectroscopy
and the need to know the recoilless fraction accurately, tend to
make the experimental result subject to errors larger than those
obtained by more conventional analysis. However, there are
many examples where the elemental analysis cannot fully charac-
terize a compound because two or more elements have alternative
oxidation states.
An interesting example is provided by a blue iron-titanium
anhydrous double sulphate which approximates in composition
to the formula FeTi(S04h with iron in the Fe 2+ oxidation state
[1]. Repeated attempts to prepare a stoichiometric compound
have failed; in every case the product appears to be a single phase
material, but the chemical analyses always show an excess of iron
and a significant but variable proportion of Fe 3+. The Mossbauer
spectra at 295 K are comparatively simple, with a quadrupole
doublet from Fe 2+ (A = 0.98, D = 1.26 mm s-l with respect to Fe
metal) and a broad singlet from Fe3+ (D = 0.64 mm s-l). However,
in every preparation the proportion of iron in the Fe 3+ oxidation
state, as estimated approximately from the relative absorption
areas, is found to be appreciably higher than indicated by chemical
analysis (the figures for the sample closest to stoichiometry being
28% by Mossbauer spectroscopy as opposed to 5.5% by analysis).
The formula of the hypothetical ideal compound is
(Fe 2+)(Ti4+)(SOl-)3' and indeed the ratio of Fe + Ti to sulphate is
always found to be very close to 2 :3. In order to incorporate the
excess of iron,a formulation of (FetxFe~1(Fe~+Titx)(SOl-)3 must
be adopted, which is then consistent with the analyses, but not with
the Mossbauer data. The solution to the problem is to propose the
existence of a proportion of Ti3+ in the solid phase. The new formu-
lation is (Fetx_yFe~!y)(Fe~irTitx_y)(SOl-)3' which in the limit
of apparent stoichiometry becomes (FetyFer)(Ti~~ity)(SOl-h.
The unstable Ti 3+ ion will reduce Fe 3+ to Fe 2+ on dissolution, so
that the solution analysis gives a low value for the Fe 3+ content of
the solid. Unfortunately, the crystal structure of this interesting
compound has yet to be determined.
A second example is provided by the solid-solution SrFexRul_x~_y
derived from SrFe03 and SrRu03, both of which have the perov-
skite lattice with M4+ cations but are antiferromagnetic and ferro-
magnetic respectively [2]. The perovskite lattice is often oxygen
deficient, and in characterizing the solid-solution phase it is im-
portant to know the oxygen content. Unfortunately, for this
Analytical Applications 199
combination of elements it is difficult to obtain analyses of suf-
ficient accuracy to determine the value of y. However, the oxygen
deficiency can be determined indirectly from the 99Ru and 57Fe
Mossbauer spectra which give the oxidation states of the cations.
Thus substitution of iron in SrRu03 takes place entirely as Fe 3+
and oxygen deficiency increases, so that compositions such as
3+ RuO.9
Sr F eO.1 4+ O2.95d
anSr F eO.2
3+ RuO.802.90
4+ .
are obtamed. From
about x = 0.3 it appears that the increase in oxygen deficiency is
halted by the introduction of some Ru 5+, and at x = 0.4 the com-
position is approximately SrFet4Rut39Ru5~2102.91. At x = 0.5
the oxygen deficiency appears to be minimal with the unexpected
composition of SrFet5Rug~503. Such detailed information could
not be obtained by conventional analysis.
One rarely reported but very useful application of the Mossbauer
spectrum is as a 'fingerprint' to check the efficacy of a chemical
preparation. This is particularly true for co-ordination and organo-
metallic compounds of iron and tin, where unwan ted products are
often easily detected by the presence of additional components
in the Mossbauer spectrum. A preliminary examination in this way
can be carried out quickly, and is completely non-destructive of
the material.
This non-destructive aspect of the technique can be particularly
useful. Furthermore, almost any size or shape of object can be
examined if the experiment is carried out in a scattering mode.
One unusual application is in the classification of works of art.
Many of the yellow and brown pigments used in oil paintings such
as ochres, siennas and umbers contain a substantial proportion of
iron, mainly in the form of the magnetic oxides a-Fe203 and a-
FeOOH. However, the naturally occuring deposits from which
they were originally derived are variable in quality and average
particle size. A systematic survey [3] of iron pigments has shown
that the room-temperature Mossbauer spectrum is often characteri-
stic, with a degree of superparamagnetic collapse of the hyperfine
pattern which relates directly to the particle size. Pigments have
been successfully identified in 18th-century oil paintings, and a
systematic study of the work of a particular artist should prove
valuable in the corroboration of doubtfully attributed paintings.
The scattering method is also applicable to analysis of terracotta
statuary whose colour derives mainly from the presence of iron
[3]. Typical spectra from two terracotta works are shown in
200 Principles of M6ssbauer Spectroscopy
Fig. 9.1. The appearance of magnetic components in the sample
TC-19 has been established to be the result of firing the clay to a
higher temperature than was the case for TC-3. Individual crafts-
men undoubtedly varied in their methods of furnace handling,
and this distinction is also useful in the attribution of works to a
particular artist. Fake statues made from different material can
be readily distinguished from the genuine articles without damage
to either.
The application of scattering techniques in the study of surface
effects is considered in a separate section later in the chapter.
0·25
0·20
0·1 5
TC-3
0·10
0·05
~' "t '",,~. ~l~ ~,'" .'.e~ ~ . &
.~
'"
>
~
o
"iii
a:
-0·05'--~-----l---...L----L _ _ _l....-...J
-10 -5 0 5 10
Velocity / (mm 5- 1)
Fig. 9.1 The 57Fe Mossbauer spectrum in a scattering geometry from two
samples of terracotta statuary. The magnetic components in the spectrum
from TC-19 establish that it was fired to a higher temperature than TC-3 .
([3], Figs. 20,21).
Analytical Applications 201
9.2 Silica te minerals
The many naturally occurring silicate minerals are all structurally
derived from the tetrahedral bonding of silicon to oxygen. The
discrete orthosilicate anion, SiO:-, is found in only a small pro-
portion of them. In the vast majority the Si04 tetrahedra are
joined by oxygen sharing into infinite chains or sheets. Examples
are the pyroxenes with single-strand chains of composition (SiO~-)n'
the amphiboles with double-strand cross-linked chains of composi-
tion (Si 40Yl)n and the micas with infinite sheets of composition
(Si20g-)no The silicon-oxygen anionic groupings are held together
by metal cations between the chains or sheets in sites with usually
four or six co-ordination to oxygen. The strength and rigidity of
the silicon-oxygen lattice allow an unusually large variation in the
nature of the cations which serve primarily to produce e1ectro-
neutrality. Partial replacement of silicon by aluminium results
in the closely related feldspars, zeolites and ultramarines. The
silicates therefore show a greater degree of compositional variation
than is found for example in simple oxides.
Many of the silicates contain a substantial proportion of iron as
Fe 2+ and Fe 3+ cations, and therefore give a good 57Fe Mbssbauer
absorption. The number of distinct cation sites is usually limited
to between one and four. The distribution of the various cations
within these sites is not completely random. A substantial degree
of ordering is often found, determined largely by the differing sizes
of the lattice sites and of the cations which occupy them. To
determine quantitative site populations is not an easy matter,
and Mbssbauer spectroscopy represents one of the most useful
methods for this purpose.
The effects on the Mbssbauer spectrum of the immediate site
co-ordination in an iron oxide are described in Chapter 7, and
these principles still hold in the silicates. The main difference lies
in that the somewhat random occupation of the near-neighbour
cation sites by widely different cations tends to produce a notice-
able broadening of the resonance lines. Magnetic effects are almost
unknown, and any deductions must be made from the chemical
isomer shift, quadrupole splitting, and intensity of absorption.
The large difference in chemical isomer shift between Fe 2+ and
Fe 3+ enables the approximate proportions of the two oxidation
states to be determined quite easily from the relative areas of the
202 Principles of Mossbauer Spectroscopy
absorption lines. The recoilless fractions at different sites are
usually similar, and the error introduced by assuming them to be
equal is small. The quadrupole splitting of an Fe 3+ component is
generally small, and it is difficult to distinguish multiple site sym-
metries for this cation. However, the comparatively large splitting
found for an Fe 2+ site and its sensitivity to environment usually
results in a resolved fine structure, leading directly to an estimate
of the relative site occupancy by the Fe 2+ ions. Both the chemical
isomer shift and quadrupole splitting of an Fe 2+ ion decrease as
the co-ordination decreases from eight to six to four. As a result
it is possible to make deductions concerning site symmetry and
occupation in silicates of unknown structure. Many mineral
samples are strongly textured (particularly the amphiboles and
micas) and cannot be made into an absorber without some residual
orientation. As a result the two components of a quadrupole
doublet may have an unequal intensity.
These general remarks can be illustrated by the spectra in Fig.
9.2. The mineral howieite is believed to have a chain structure and
(0] (b]
2 2
~3 ~3
c:: c:
0 0
A A'
'e. 4
0
~4
L..
0
VI VI
J:J J:J
'"" 5 '""5
6
6
7
7
8
-2 -1 0 2 3 -1 0 2 3
Velocity / (mm 5- 1]
Fig. 9.2 The room temperature 57Fe Mossbauer spectra of (a) howieite and
(b) deerite. ([4], Fig. 1).
Analytical Applications 203
has the approximate composition NaFe~+Mn3Fe~+Si12031(OH)13.
The room-temperature spectrum [4] shows two quadrupole doub-
lets, that with the larger splitting (AA') having the high chemical
isomer shift and large quadrupole splitting (see Table 9.1) typical
of high-spin Fe 2+ ions in octahedral co-ordination. The inner
doublet (BB') is typical of high-spin Fe 3+ ions, also in octahedral
co-ordination. The mineral deerite has the approximate composi-
tion FerjFefSi13044(OH)l1. The spectrum shows three quadrupole
doublets. AA' arises from Fe 2+ ions in distorted six co-ordination,
BB' from Fe 3+ in distorted six co-ordination, while the CC' lines
with a significantly lower shift than AA' but still within the range
for high-spin Fe 2+ are clearly from a four co-ordinated tetrahedral
site. Note that the two components of each doublet are not neces-
sarily equal in intensity because of texture in the absorber.
A better known class of silicates is the pyroxenes which contain
single-stranded chains of Si04 groups of overall composition
(SiO~-)n" The general formula can be written as ABSi206 where
A and B represent distinct cation sites known as M1 and M2
respectively. The M1 sites have a near-regular six co-ordination to
oxygen and are usually occupied by small cations such as Mg2+,
Al3+, Mn2+, Fe 3+ or Na+. The M2 site is grossly distorted, and being
larger tends to be occupied by Na+, Mg2+, Ca 2+, Mn 2+ and Fe 2+. All
pyroxenes do not have a common space group; for example
pigeonite has the space group P2dc, CaMgSi206 and CaFeSi206
belong to C2/c, and Mg2Si206 to Pbca, but the small differences
involved have only a minor influence on the Mossbauer spectra.
In all the pyroxenes the Fe 2+ cations at the more distorted M2
sites show a smaller quadrupole splitting than at the M1 sites. This
Table 9.l Mossbauer parameters for howieite and deerite at room tem-
perature
* relative to Fe metal.
204 Principles of M6ssbauer Spectroscopy
is probably because of a large 'lattice' contribution to the electric
field gradient tensor which is opposite in sign to the 'valence' term
from the 3d6 configuration (see p. 103). The 57Fe spectra comprise
two superimposed doublets which are partially resolved (ignoring
for the present a possible third contribution if Fe 3+ ions are also
present; some examples of spectra from orthopyroxenes are shown
in Fig. 9.4).
It is convenient to limit further discussion to the orthopyroxenes,
which have a composition close to the system MgSi03-FeSi03.
The Fe 2+ cations prefer to occupy the M2 sites, but considerable
disorder can be induced by heating to 1270 K or above. Anneal-
ing at lower temperatures causes a preferential occupation of the
M2 sites by Fe 2+ which increases with decreasing temperature
until at about 750 K the ionic diffusion is apparently frozen [5,6].
If the relative occupations of the MI and M2 sites by Fe 2+ are Xl
and X2 respectively, then a parameter k can be defined by
AGE = -RTln k
o·g
0·8 • •
.".
•
+
lit
~
•
+
0.7·. '9.
0·6
004
0·3 •
0'2
0·1
o 0·1 0·2 0·3 0'4 0'5 0·6 0·7 0·8 O·g 1·0
x1
Fig. 9.3 A plot of the measured site occupancies x I and x2 of the M 1 and
M2 sites in a selection of natural orthopyroxenes. The open circles represent
natural samples reheated to 1270 K and quenched. The filled circles repre-
sent natural slowly cooled metamorphic and plutonic rocks. The crosses re-
present quickly cooled volcanic rocks. ([ 6] , Fig. 1).
0·06
0·08
0·10
0·12
Fez+(M z)
0·14 Fe2+(M 111
....- - - - - - - '
c:
o
'';:;
..
a.
L.
Sl
.0
« 0·02 (bl
0·04
0·06
0·08
0·10
0·12
0·14
Velocity / (mm 5- 1I
sion [9]. The spectrum for scattering from a steel plate exposed
to fumes of HCl in air is shown in Fig. 9.Sa. The dominant doub-
let in the spectrum can be identified from the chemical isomer
shift and quadrupole splitting as being that of j3-FeOOH and is
distinct from the spectra of the other oxides and oxyhydroxides
of iron. The intensity of the scattering can be related to the depth
of the surface layer by evaluation of the appropriate scattering
integral, and in this case gives an approximate thickness of a bout
2 x 10- 3 cm. The additional weak components on the baseline are
the magnetic hyperfine pattern of the underlying steel. Fig. 9.Sb
208 Principles of Mossbauer Spectroscopy
(a)
(e)
+12 +8 +4 0 -4 -8 -12
Veloeity/(mm S-I)
:. . ~ : ~:~ ;\ j ~ ,.
- ~V\;~\/ '....J, .
(dl
.. .,
(el
,.,
:: :;;; . \ ;:
"
.! :~
; v\JV\. h./ .
.'
.. ..
" (bl
(QI
.'
:JJvJJ ~
-10'0 -5-0 0-0 5·0 10·0
VfIOCily/( mm ,-'I
References
[1] Gibb, T. C., Greenwood, N. N., Tetlow, A. and Twist, W. (1968) J.
Chern. Soc. (A), 2955.
[2] Gibb, T. C., Greatrex, R., Greenwood, N. N. and Snowdon, K. G.
(1974) J. Solid State Chern., 14,00.
[3] Keisch, B. (1973) Archaeornetry, 15, 79.
[4] Bancroft, G. M., Burns, R. G. and Stone, A. J. (1968) Geochirn.
Cosrnochirn. Acta, 32,547.
[5] Virgo, D. and Hafner, S. S. (1970) Arner. Mineral, 55,201.
[6] Hafner, S. S. and Virgo, D. (1969) Science, 165,285.
[7] Schiinnann, K. and Hafner, S. S. (1972) Proc. Third Lunar Sci. Conf.,
Geochirn. Cosrnochirn. Acta, Suppl. 3, vol. 1,493.
[8] Dundon, R. W. and Hafner, S. S. (1971) Science, 174,581.
[9] Terrell, J. H. and Spijkerman, J. J. (1968) Appl. Phys. Lett., 13, 11.
[10] Simmons, G. W., Kellennan, E. and Leidheiser, H. (1973) Corrosion,
29,227.
[11] Davies, D. E., Evans, U. R. and Agar, J. N. (1954) Proc. Roy. Soc.,
225A,443.
[12] Bonchev, Zw., Jordanov, A. and Minkova, A. (1969) Nuclear Instr.
Methods, 70,36.
[13] Hobson, M. C. and Gager, H. M. (1970) J. Catalysis, 16,254.
[14] Hobson, M. C. and Gager, H. M. (1970) J. Colloid & Interface Sci.,
34,357.
[15] Gager, H. M., Lefelhocz, J. F. and Hobson, M. C. (1973) Chern Phys.
Letters, 23,387.
[16] Delgass, W. N., Garten, R. L. and Boudart, M. (1969)J. Phys. Chern.,
73,2970.
CHAPTER TEN
213
214 Principles of Mossbauer Spectroscopy
standard absorber also represents a higher energy. For this reason
a hyperfine emission spectrum is the mirror image of a correspond-
ing absorption spectrum. The sign of the chemical isomer shift
must therefore be reversed for comparison with conventional
absorption data.]
However, the radioactive decay of the parent nucleus may cause
drastic chemical modification of the environment of the daughter
nucleus before the M6ssbauer -y-emission takes place. If the host
matrix is metallic, then the high mobility of the electrons ensures
that any after-effects of the decay are obliterated within a time
much less than the M6ssbauer excited-state lifetime. Therefore
the doping of metals with for example 57Co or 119m Sn can be
regarded simply as a means of producing very low concentrations
of iron or tin impurity atoms. If on the other hand the host matrix
is an insulator, then after-effects may be observed. These will
be discussed separately. The after-effects of nuclear decay may be
studied advantageously by M6ssbauer spectroscopy because the
technique is specific to the site of the disturbance and gives infor-
mation concerning the atomic environment within 10- 7 s of the
primary event.
. ..
0 · 91~ -
0·86 r- -
e
''E:
0
..
.e
0
• (a)
....
"-
0·81 t I I I
0 2
1·01 r---:------..,.-----..,.---,----- , . . . - - - - - - - - - ,
"
l
\ , 8
0·95 '.;. 0:
". (b)
0 .92L---_~1--~0----Ll-----!2~--~--4L---..L5-....J
Veloc i ty I (mm S-I)
Fig. 10.1 The Mossbauer spectra of CdO.99FeO.01F2 at (a) 296 K and (b) 5 K
The weak doublet ~B) seen at low temperatures corresponds to a near-
neighbour Fe 2+-Fe + cation pair, while the central resonance (A) corresponds
to isolated Fe 2+ ions. ([ 1] , Fig. 1).
V Cr Mn
-0·5
2
"j'
'"
E
E
~
:.c
'" 0
L..
E
0
'"
.!!!
a 8
.~
E
~
(J
'"
Fe Co Au
~
RU Rh N, Cu
4 w Os
Ir Pd
Ag
PI
o Au
3 5 7 9 11
Number of outer electrons
Fig. 10.2 The chemical isomer shifts of 99Ru, 193Ir and 197 Au impurity
atoms in the 3d-, 4d- and Sd-metals. The values were obtained from the emis-
sion spectra of 99Rh, 1930s and 197Pt, and are quoted with respect to Ru, Ir
and Au metal respectively but in the sense of a conventional absorption ex-
periment. 8R/R is positive in all cases, so that a more positive shift cor-
responds to a larger s-electron density at the nucleus. ([3 J, Fig. I).
c
o
·iii
·e
VI
VI~_ _ _
c
.='"
..
\i I 5%
absorption
1 -1
Velocity/(mm S-l)
Fig. 10.3 The 57Co emission spectra at room temperature of doped alloys
near the composition Ni3Al. ([4], Fig. 1).
......
.... . .. ...
. ,.... ".,
. .,"'. ... ....... .
. -.
, ..
1001-
.
.,-,'~...'..
,. :. I • ••
.' . l'
'" . # Kr/Fe -136
'"
.:s;;; 99,5-
..
'E
." .....
..•
<
.,
c:
...........
"
I
,=
... .-
~
lOOt-- •
..... /', .<1.' _~
Ie· .--.-,
t~ • •."., '. :
.....
•
.~.
.
..
'10 ,
\
. ,,.,.'
).... I.
:
. , Kr/Fe=44
99·5 -
..
,I
-,.
-2 -1 0 2
Velocity / (mm 5- 1)
.~ '485
e
...
~
".
1"70~L-__-L____~__~"____~__~____~__~__~
3 z o -I -2 -3 -4
Velocityf(mm 5- 1)
Fig. 10.6 The Mossbauer emission spectrum at 295 K for 57Co doped into
MgF2' 38% of the decays produce an Fe 3+ ion, while the rest give Fe 2+. The
spectrum has been reversed to resemble the equivalent absorption spectrum.
([ 11 ], Fig. O.
....
. .'
•••••• (0)
......
~
.'
....
°0 ••
(b)
-4 -2 0 2 4
Velocity/(mm 5- 1)
. . ..:.::..:.,,-'/
......'"\
...........
... ...
-.:
..' '.
......
:- .,- .
.,:I ••
:.
220 f-
..
:
.."'.,-..
~
.
200
0:'
'.:.
I I I I I J I I
-6 -4 -2 0 2 4 6 8
Velocity I (mm 5- 1 )
Fig. 10.8 The 119mSn emission spectrum (upper) and 119Sn absorption
spectrum (lower) from a matrix of SnS04 doped with 119mSn. Note the
production of some Sn4+ species by the isomeric transition. The velocity
scale of the source experiment has been reversed to facilitate comparison.
([ 14], Fig. 1).
Impurity and Decay After-effect Studies 229
The concentration of these defects is too low to register in the
absorption spectrum. This oxidation is attributed to the conse-
quences of an autoradiolysis process which accompanies the Auger
cascade. However, decay after-effects are not very common in
tin because the covalent bonds are usually stronger than in iron
compounds.
If the primary decay is by t3-emission, then the effects on the
atomic environment are less drastic. The increase in the nuclear
charge by +e effectively increases the oxidation state of the atom.
If the 'daughter' environment is isoelectronic and isostructural
with a stable compound, then the co-ordinate geometry is un-
likely to be affected. The primary influence appears to be the
chemical stability of the daughter product in the environment of
the parent lattice. In some instances it is possible to characterize
previously unknown ionic species trapped in the host matrix.
Thus the {3-decay of 129Te in (NJ4hTeCI6 populates the 27.8-keV
Mossbauer transition of 129 1 [15]. A single-line 129 1 emission is
observed at -6.08 mm s-l relative to a Zn 129 Te source (Le. at
+6.08 mm s-l for an absorption). The regular octahedral TeCll-
anion appears to decay to an isostructural ICI6" anion. There is
also an alternative source precursor. 129mTe decays by an isomeric
transition to 129Te prior to the ensuing t3-decay. A source of
(NH4hTeCI 6 doped with 129m Te gives the same single-line spectrum
as from 129Te. In this instance neither the isomeric transition nor
the t3-decay cause molecular fragmentation, showing that the 106"
anion is comparatively stable.
Similar experiments have been used to prepare molecules of
unknown xenon compounds. The 39.58-keV Ig= 4- ~ Ie= f reso-
nance of 129Xe can be easily observed using a source of 1291 in a
matrix of NaI, K104 or Na3H2I06, all of which show a single-line
absorption with solid xenon or xenon hydroquinone clathrate as
the absorber. In contrast, absorbers of XeF 2, XeF 4, and Xe03
show well resolved quadrupole doublets with Il = 39,41, and 11
mm s-l respectively because withdrawal of electron density from
the xenon by the ligands causes an asymmetry in the valence
orbital occupation.
Unexpectedly, the three sources mentioned above show small
but significant chemical isomer shifts when compared with the
same clathrate absorber, and it can be argued that the t3-decay of
the source generates a xenon species with some degree of chemical
230 Principles of Mossbauer Spectroscopy
binding to the lattice. More dramatically, the decay of 1291°3 in a
matrix of Nal03 gives a spectrum with a clathrate absorber show-
ing a quadrupole splitting identical to that from an Nal source and
an Xe03 absorber, so that it is an obvious inference that the i3-decay
has created an {Xe03} molecule trapped in the iodate lattice.
Similar experiments [16, 17] with KICI4.H20, KICI2.H20 and
KIBr2 sources show quadrupole split spectra attributable to trap-
ped {XeC14}, {XeCI2} and {XeBr2} molecules, none of which are
known as stable compounds but which may be presumed to be
isoelectronic with IC14", ICI 2 and IBr2 respectively. The observed
splittings are f1 = 25.6, 28.2 and 22.2 mm s-1 respectively. Com-
parison of these values with those of XeF 2 and XeF4 (see Fig. 10.9)
show consistency with the predicted bonding in these molecules.
{XeCI4} is square-planar, and {XeCI2} and {XeBr2} are linear mole-
cules. In the tetrahalides, fluorine is more electron withdrawing
than chlorine, and in the dihalides the order is fluorine> chlorine
> bromine as predicted.
However, not all 1291 i3-decay sources give a chemically bonded
xenon atom. Both 12912 and Cs129IBr2 for example appear to break
down to atomic xenon within the time scale of the Mossbauer
event (10- 9 s). It should be emphasized that the {XeBr2} mole-
cule in KIBr2 may in fact have a lifetime of less than a microsecond.
An example of a change in oxidation state following i3-decay is
given by the decay of 1930s in K2[OsCI 6] and (NH4h[OsCI6] in
which the Os4+ has a low-spin 5d 4 configuration. The decay popu-
lates the 73-keV level of 193Ir, and the oxidation state of the Ir
daughter nucleus proves to be +4 (Sd 5) [18]. This configuration
is known to be stable in K2[IrCI 6] and (NH4h[IrCI6] , and pre-
sumably the [IrCI 6] - ion which is initially formed is rapidly reduc-
ed to the more stable [IrCI6]2- configuration.
The effects on the Mossbauer spectrum of an energetic nuclear
reaction such as a-decay, Coulombic excitation by 2S-Me V oxygen
ions, or an (n, 'Y) reaction are often surprisingly small despite the
fact that the nucleus is displaced from its original site. In all cases
the nucleus seems to reach its final lattice site in a time much less
than the Mossbauer lifetime (10- 8 s), and in metallic matrices
the spectrum recorded usually appears identical to that from a con-
ventional absorption experiment. Few data are available for non-
metallic compounds, and it is in these that after-effects are likely
to be severe. A reduction in the apparent recoilless fraction is
Impurity and Decay After-effect Studies 231
4·0
-4,0
(a)
-8,0
-12'0
4'0
-4·0 (b)
{Xeel z }
_ -8'0
~
4·0
"
0
.~
0
...
~
.Q
-4·0
-8,0
-12'0
-16·0
4'0
0
...
-4·0
XeF z (d)
-8'0
Fig. 10.9 The 1291 Mossbauer emission spectra of (a) {XeCI4} trapped in
K1C4, (b) {XeCI2} trapped in K1C12, and the 129Xe absorption spectra of
(c) XeF4and(d) XeF2. ([161, Fig.I).
References
[1] Steger, J. and Kostiner, E. (1973) 1. Chern. Phys., 58,3389.
[2] Abeledo, C. R., Frankel, R. B., Misetich, A. and Blum, N. A. (1971)
J. Appl. Phys., 42,1723.
[3] Wagner, F. E., Wortmann, G. and Kalvius, G. M. (1973) Phys. Letters,
42A,483.
[4] Liddell, P. R. and Street, R. (1973) J. Phys. F, 3, 1648.
[5] McNab, T. K., Micklitz, H. and Barrett, P. H. (1971) Phys. Rev. (B),
4,3787.
[6] Micklitz, H. and Barrett, P. H. (1972) Phys. Rev. (B), 5, 1704.
[7] Bos, A., Howe, A. T., Dale, B. W. and Becker, L. W. (1972) Chern.
Cornrn .• 730.
Impurity and Decay After-effect Studies 233
[8] Micklitz, H. and Barrett, P. H. (1972) Phys. Rev. Letters, 28, 1547.
[9] Data presented by Micklitz, H. and Litterst, F. J. at the International
Conference on the Applications of the Mbssbauer Effect at Bendor,
September 1974.
[10] Regnard, J. R. (1973) Solid State Cornrn., 12,207.
(11] Cruset, A. and Friedt, J. M. (1971) Phys. Stat. Sol (B), 47,655.
[12] Baggio-Saitovitch, E., Friedt, J. M. and Danon, J. (1972) J. Chern.
Phys., 56, 1269.
[13] Friedt, J. M., Shenoy, G. K., Abstreiter, G. and Poinsot, R.(1973)
J. Chern. Phys., 59,3831.
[14] Llabador, Y. and Friedt, J. M. (1971) Chern. Phys. Letters, 8, 592.
(15] Jones, C. H. W. and Warren, J. L. (1970)J. Chern. Phys., 53, 1740.
[16] Perlow, G. J. and Perlow, M. R. (1968) J. Chern. Phys., 48,955.
(17] Perlow, G. J. and Yoshida, H. (1968) 1. Chern. Phys., 49,1474.
(18] Zahn, U., Potzel, W. and Wagner, F. E. (1973) Perspectives in Mbssbauer
Spectroscopy, ed. Cohen, S. G. and Pasternak, M. (1973) Plenum Press,
New York, p. 55.
[19] Vogl, G., Schaeffer, A., Mansel, W., Prechtel, J. and VogI, W. (1973)
Phys. Stat. Sol. (B), 59, 107.
CHAPTER ELEVEN
Biological Systems
234
Biological Systems 235
but nevertheless with na tural concentrations of iron the Moss-
bauer resonance is always comparatively weak. However, in many
instances it is possible to cultivate organisms with a diet enriched
in 57Fe, and the 40-fold improvement in the resonant cross-section
which can be achieved results in an adequate absorption intensity.
Alternatively, for some proteins it is possible to remove the iron
and then to re-constitute the protein with enriched 57Fe without
affecting the bio logical activity.
The preparation of separated protein material is often a major
task in itself, and particular care must be taken to ensure that
accidental denaturing of the protein does not take place. Pure
compounds in crystalline form are rarely obtainable, but experi-
ence has shown that satisfactory results may be obtained on frozen
solutions of protein concentrates. Freeze-drying or lyophilization
is also often used. Nevertheless, there remains the possibility of a
conforma tional change in the solid so that the 'active' solution
form is not observed.
The Mossbauer spectrum of the iron site reflects the ligand-field
of the surrounding organic groups, and therefore the principles
described in Chapters 4 and 5 are directly applicable. The major
innovation is that individual iron atoms are now separated by
distances of the order of 25 A (2500 pm), and as shown in Chapter
6 this has an important influence on the relaxation properties of
paramagnetic configurations. In such cases the Mossbauer spectra
at low temperatures usually show a paramagnetic hyperfine split-
ting which slowly collapses with rising temperature. The spin-
spin mechanism of electronic relaxation between neighbouring iron
nuclei is comparatively ineffective, and in some instances is domi-
nated by the very weak interactions between the electrons on the
resonant atom and the nuclear spins of the ligand atoms. When
this occurs, the zero-field Mossbauer spectra are extremely diffi-
cult to interpret. The application of an external magnetic field
then results in a considerable simplification of the observed spec-
trum, from which many of the interaction parameters appropriate
to the zero-field spectra may be obtained.
Some biologically active compounds such as vitamin B12 (cyano-
cobalamin) contain cobalt, and although it is possible to dope with
57Co, this technique has not been widely exploited because of two
major disadvantages. It is unlikely that the electronic configuration
of the 57Fe daughter nucleus will be related to that of the cobalt
236 Principles of Mossbauer Spectroscopy
parent, and furthermore there is also a danger of local damage by
autoradiolysis (see Chapter 10).
More detailed examples to illustrate these general principles are
taken from the haemoproteins and ferredoxins.
11.1 Haemoproteins
The haemoproteins are well known as the oxygen carriers in blood.
The oxygen is transported by chemical binding to an iron atom,
which is itself in the centre of a planar porphyrin structure known
as protoporphyrin IX (illustrated in Fig. 11.1). In the protein
haemoglobin, the porphyrin is attached by co-ordination of the
iron to a fifth nitrogen in a histidine unit of the protein chain
below the plane of the four nitrogen ligands. The sixth co-ordina-
tion position of the iron on the opposite side of the porphyrin is
thus left vacant, and forms a site for labile bond formation with
small molecules.
It is possible to isolate compounds of iron with protoporphyrin
IX, and although it might be expected that these would be model
compounds for study of the haemoproteins themselves, this pro-
mise is not entirely fulfilled. The iron(lII) chloride derivative of
CH z
II
CH
Me
Me
5 4 0.83
Haemin 2" 0.36
bispyridyl haeme 0 4 1.14 0.36
HbCO 0 4 0.36 0.26
Hb02 0 1.2 2.24 0.24
77 2.19 0.26
195 1.89 0.20
Hb (reduced) 2 4 2.40 0.91
5
HiF 2" 1.2-195 broadened
HiCN 1 195 1.39 0.17
2"
* The abbreviation Hb is used for an Fe(I1) haemoglobin compound and Hi for an Fe(lII)
haemoglobin compound.
238 Principles of Mossbauer Spectroscopy
However, this diamagnetic configuration cannot show relaxation
and therefore gives a sharp symmetrical spectrum at all
temperatures.
The majority of the data on haemoglobin derivatives have been
obtained using the blood from rats injected with ferric citrate
solution containing iron enriched to 80% in 57Fe [2]. It is not
always easy to isolate pure compounds, and in this case spectra
have usually been recorded in frozen solutions of red-cell concen-
trates or haemoglobin extracts. The oxidation state of the iron
is very dependent on the nature of the sixth ligand. The carbon
monoxide and oxygen derivatives of haemoglobin (HbeO and
Hb02) are both S = 0 low-spin iron(II) compounds. The spectra
comprise a simple quadrupole doublet, and typical parameters
are given in Table 11.1. The diamagnetic S = 0 state may be
distinguished from paramagnetic configurations by the absence
of any induced hyperfine splitting when a small external magnetic
field is applied.
The applied-field method of determining the sign of e 2 qQ has
been used to show that e2qQ is negative in Hb02, but unfortunately
the magnitude of f/ is not known. The large quadrupole splitting
and low chemical isomer shift of Hb02 (compared for example
to the bispyridyl haeme) show that the oxygen-iron bond is strong-
ly covalent. A possible geometrical arrangement and bonding
scheme [2] are shown in Fig. 11.2. The oxygen molecule is
assumed to lie parallel to the porphyrin plane to give a strong
overlap between the d yz metal orbital and the 7T*-antibonding or-
bital on the oxygen. The resultant separation of the molecular
orbitals causes spin-pairing of the otherwise paramagnetic oxygen
electrons to give the diamagnetic configuration observed. The
initially non-bonding dyz electrons thus become intimately in-
volved in the bonding. The molecular orbital (dyz + 7TiJ/..j2 has
only half the character of a dyz orbital, and this results in an effec-
tive decrease in the occupation of this orbital. As a result, the
principal axis of the electric field gradient tensor should lie along
the x axis of Fig. 11.2, and e2qQ should be large and negative (as
observed experimentally). The quadrupole splitting is unusually
temperature dependent for a diamagnetic configuration (see
Table 11.1). This has been suggested to indicate a possible libra-
tional motion of the oxygen molecule at higher temperatures.
The paramagnetic derivatives show various configurations; e.g.
Biological Systems 239
0, Complex Fe
Fig. 11.2 A possible bonding scheme in Hb02 showing how the interaction
of the1r:molecular orbital on the oxygen with the dyz orbital on the iron can
remove the degeneracy and cause spin pairing in the bound state.
-10 o 5 10
Velocity I (mm S-I)
·::/~~::i;::·:1~.\. /.;-:~~~~:~.:t~~:
(a) .'
/:.
.:"~'.
".
-3 -2 -1 0 1 2 3 4
Velocity / (mm 5- 1)
Fig. 11.4 Mossbauer spectra at 195 K of (a) the oxidized and (b) the reduc-
ed forms of Scenedesmus ferredoxin enriched with 57Fe. ([41, Fig. 6).
Scenedesmus ferredoxin
reduced form 195
{~ 5
0.59
2.75
+0.22
+0.56
oxidized form 195 2" 0.60 +0.20
Chromatium high-potential
iron protein
reduced form 77 ? l.l2 +0.42
oxidized form 77 ? 0.79 +0.32
(Et4Nh[Fe4S4(SCH2Ph)41 ? 1.26 +0.32
Biological Systems 243
r
z
CYS CYS'l
"JJ
W 0
f-
-<
~ ~
z
YCys s Cysr-J
Fig. 11.5 Representation of the active site in the reduced form of a plant
t
ferredoxin. The S = and S = 2 spin-states couple antiferromagnetically
to give a resultant of;:} = -}. In the oxidized form, both spin-states are S = ~
and couple antiferromagnetically to give S = o.
/J! ,odor"
(cYS)S~
S---Fe
/ / -----.S(CYS)
Fe 5
(CYS)S/
doxins from bacteria and from plants has proved useful in the bio-
logical classication of the intermediate algae. The blue-green algae
and photosynthetic bacteria both lack a nucleus within the cell
envelope and carry out photosynthesis throughout the cell. In
contrast, the green algae and higher plants localize this function
within a chloroplast inside the cell. However in other ways the
biological activity of the ferredoxin in blue-green algae appears to
be more closely related to the higher plants. The M6ssbauer spec-
trum of a sample extracted from the planktonic blue-green alga
Microcystis flos-aquae and chemically reconstituted with 57Fe
enriched iron is identical with those described earlier from the
higher plant forms Scenedesmus and spinach ferredoxins [7].
The establishment of this rela tionship of the blue-green algae
ferredoxins to those of the higher plants provides useful evidence
in support of the hypothesis that the chloroplasts in the cells of
higher plants are the descendents of blue-green algae that lived
symbiotically within the early plant cells.
References
[1] Moss, T. H., Bearden, A. J. and Caughey, W. S. (1969) 1. Chem Phys.,
51,2624.
[2] Lang, G. and Marshall, W.(1966) Proc. Phys. Soc., 87,3.
[3] Lang, G. (1968) Phys. Letters, 26A,223.
[4] Rao, K. K., Cammack, R., Hill, D. O. and Johnson, C. E. (1971)
Biochem. J., 122,257.
[5] Evans, M. C. W., Hall, D. O. and Johnson, C. E. (1970) Biochem. J.,
119,289.
[6] Herskovitz, T., Averill, B. A., Holm, R. H., Ibers, J. A., Phillips, W. D.
and Weiher, J. F. (1972) Proc. Nat. Acad. Sci. U.S.A., 69, 2437.
[7] Rao, K. K., Smith, R. V., Cammack, R., Evans, M. C. W., Hall, D. O.
and Johnson, C. E. (1972) Biochem. 1., 129, 1159.
Bibliography
Books
Mossbauer spectroscopy - an introduction for inorganic chemists and geo-
chemists, Bancroft, G. M. (1973) McGraw-Hill, London, 252 pp.
Mossbauer effect and its applications, Bhide, V. G. (1973) New DeIhl,
Tata McGraw-Hill.
Mossbauerspectroscopy, Greenwood, N. N. and Gibb, T. C. (1971) Chapman
and Hall, London, 659 pp.
Chemical applications of Mossbauer spectroscopy, Goldanskii, V. I. and
Herber, R. H. (eds.) (1968) Academic Press, New York, 701 pp.
An introduction to Mossbauer spectroscopy, May, L. (ed.) (1971) Plenum
Press, New York, 202 pp.
Der Mossbauer-Effekt und seine Anwendung in Physik und Chemie,
Wegener, H. (1965) Bibliographisches Institut, Mannheim, 214 pp.
Mdssbauer effect: principles and applications, Wertheim, G. K. (1964)
Academic Press, New York, 116 pp.
Mossbauereffectdata index 1958-1965, Muir,A. H., Ando, K. J. and
Coogan, H. M. (1966) Interscience, New York, (a catalogue of references
up to early 1966).
Mossbauer effect data index, Stevens, J. G. and Stevens, V. E. (eds.) IFI/
Plenum Data Corp., New York, (an annual publication listing references
from 1969 onwards).
Mdssbauer effect methodology, Vols. 1-8, Gruverman, I. J. (ed.) Plenum
Press, New York, (an annual publication from 1965 containing a collection
of contributed papers and reviews biased towards instrumentation)
Perspectives in Mossbauer spectroscopy, Cohen, S. G. and Pasternak, M. (eds.)
(1973) Plenum Press, New York, 259 pp, (collected conference papers).
Proceedings of the conference on the application of the Mossbauer effect -
Tihany (Hungary), 1969, Dezsi, I. (ed.) (1971) Akademiai Kiado,
Budapest, 803 pp.
Mossbauer spectroscopy and its applications, a Panel Proceedings of the
International Atomic Energy Agency, Vienna (1972) 421 pp, (a collection
of 16 review articles).
Spectroscopic properties of inorganic and organometallic compounds, vols.
1-7, Greenwood, N. N. (ed.) The Chemical Society, London, (containing
an annual review of the literature from 1967 onwards).
245
246 Principles of Mossbauer Spectroscopy
Reviews
Bancroft, G. M. and Platt, R. H. (1972) Mossbauer spectra of inorganic com-
pounds: bonding and structure,Adv. Inorg. Chem. and Radioch em. , 15,
59-258.
Bearden, A. J. and Dunham, W. R. (1972) Iron electronic configurations in
proteins: studies by Mossbauer spectroscopy, Structure and Bonding, 8,
1-52.
Cohen, R. L. (l972) Mossbauer spectroscopy: recent developments, Science,
178,828-835.
Devoe, J. R. and Spijkerman, J. J. (1970) Mossbauer spectrometry, Analyt.
Chem., 42, 366R-388R.
Friedt, J. M. and Danon, J. (1972) Applications of the Mossbauer effect in
radiochemistry, Radiochimica Acta, 17, 173-190.
Gibb, T. C. Applications of Mossbauer spectroscopy to organometallic
chemistry. In Spectroscopic methods in organometallic chemistry, ed.
George, W.O., p. 33-60, Butterworths, London.
Go1danskii, V. I., Kbrapov, V. V. and Stukan, R. A. (1969) Application of
the Mossbauereffect in the study of organometallic compounds,
Organometallic Chem. Rev. A, 4, 225-261.
Herber, R. H. (1967) Chemical applications of Mossbauer spectroscopy,
Progr.lnorg. Chern., 8, 1--41.
Hobson, Jr., M. C. (l972) The Mossbauer effect in surface science, Progr.
Surface Membrane Sci., 5, 1-61.
Johnson, C. E. (1971) Applications of the Mossbauer effect in Biophysics,
J. Appl. Phys., 42, 1325-1331.
Kurkjian, C. J. (1970) Mossbauer spectroscopy in inorganic glasses, 1. Non-
cryst. Solids, 3, 157-194.
Lang, G. (1970) Mossbauer spectroscopy of haem proteins, Quart. Rev.
Biophys., 3, 1-60.
Maddock, A. G. (1972) Mossbauer spectroscopy in the study of the chemical
effects of nuclear reactions in solids, MTP International Review of
Science, Vol. 8, p. 213-250, Butterworths, London.
Parish, R. V. (1972) The interpretation of 119Sn-Mossbauer spectra,Progr.
Inorg. Chem., 15, 101-200.
Reiff, W. M. (1973) Magnetically perturbed Mossbauer spectra of iron and
tin co-ordination compounds, Coord. Chem. Rev., 10,37-77.
Stevens, J. G., Travis, J. C. and DeVoe, J. R. (1972) Mossbauer spectrometry,
Analyt. Chem., 44, 384R--406R.
Zuckerman, J. J. (1970) Applications of 119mSn Mossbauer spectroscopy to
the study of organotin compounds, Adv. Organometallic Chem., 9,
21-134.
Observed Mossbauer Resonances
247
Index
Absolute calibration of velocity, 15 Chemical isomer shift-cant.
Absorber thickness, 11-2 effect of pressure, 122
Absorption cross-section, 10 in europium compounds, 77,130
of s7Fe, 17 in ferredoxins, 241
Absorption intensity, II in ferrocene, 98-100
Adsorption, study of, 211 in gold compounds, 127-9
Angular dependence of absorption in haemoproteins, 237
intensity, 40-1 in iodine compounds, 26, 77-84
Anisotropy of recoilless fraction, in iridium compounds, 126-7
43,141-4 in iron carbonyls and derivatives,
Antiferromagnetic coupling (intra- 46-53,61-2
molecular), 154 in iron cyanide complexes,
Antiferromagnetic ordering (see also 58-9,93-6
magnetic hyperfine splitting) in iron halides, 99-100
in Au 2Mn, 191 in matrix-isolated iron atoms
in FeF 2, lIS 220-4
in FeF 3, 112 in neptunium compounds, 131-3
in MnF2' 216-7 in organotin compounds, 63-4,
in spinels, 168-74 89-93
Neel type, 168 in palladium-tin alloys, 183-4
spin canting, 168, 172-3 in ruthenium compounds, 124- 5
Yaffet-Kittel type, 168 in silicates, 201-6
Antimony (lUSb), 18,84-6 in spin cross-over compounds,
halides, 85-6 119-21
organoantimony compounds, in tellurium compounds, 84
85-7 in tin(II) compounds, 87-8
transferred hyperfine field in, in tin(IV) halides, 89
180 of 99Rh, 1930s, 197Pt impurities
Antiphase boundaries, 196 in the transition metals, 217-8
Asymmetry parameter, 31-6 partial chemical isomer shifts,
in (diphos)Fe(FOh, 50 96-8
in [(tr-C sH s )Fe(COhhSnCI 2, 91 second order Doppler shift,
in FeF 2, II7 26-8
sign of, 75
Bam (unit) defined, 31 Chi-squared distribution, 20
Beta decay, after-effects of, 229-32 Chromatium high-potential protein,
Binomial distribution, 170 243
Biological systems, 234-43 cis-/trans-geometrical isomerism,
Brillouin, function, 112 54-6
C1ebsch-Gordan coefficient, 39-41
Cation vacancies in 'Y-Fe203, 165 Cobalt (S7Co) impurity doping, 142
Centre shift (see chemical isomer diffusion of in copper, 147
shift) in biological systems, 235
Chemical isomer shift, 22-8 in CaO, 224
calibration of, 26,74-5 in coordination compounds,
effect of change in oxidation 226-7
state, 74-7 in inorganic halides, 224- 5
effect of electron correlation, 76 in Ni3Al, 217-9
248
Index 249
Cobalt (S7Co) impurity doping-cont. Electric field gradient tensor-cont.
in zinc metal, 142 in iodine compounds, 77-84
matrix isolation in xenon, 222-3 lattice term, 33-5
Combined magnetic and quadrupole point charge calculations,
interactions, 36-8 54,69-71
Compton scattering, 11 principal values, 31
Computation of data, 19 valence term, 33-5
Conduction-electron polarization, Electric monopole interaction, 22
183 Electric quadrupole interaction, 22
Conformations, determination of, Electric quadrupole (E2) transition,
60-3 40
Coordination number, determination Electromechanical drive, 13
of, 48-9, 64-9 Electron capture decay, effects of,
Core polarization, 182 222-7
Corrosion, study of, 206-11 Electron hopping, 145-7
Cosine-effect, 15 Electron spin resonance, 23
Coulombic interaction, 22 Electronic relaxation, effect on
Cross-section for resonant absorp- hyperfine field, III
tion, 10 Electrostatic potential, 31
Cryogenics, 15 Energies of nuclear and chemical
Cryostats, 15 interactions, 4
Curie temperature, 112, 169 Energy of Mossbauer 'Y-ray, 16
Curve-fitting, 20 Europium (lSIEu), 18, 130
electron hopping in EU3S4, 145
Debye model for lattice vibrations, impurity in ice, 148
8, 139-41 in europium iron garnets, 177-8
Debye temperature, 8, 17, 139 in inorganic compounds, 130
Decay schemes (for S7Co, 119Sn, range of chemical isomer shifts,
1291, 193Ir), 17-8 77,130
Detection system, 14 Exchange interactions, 168-79
Diffusion in solids, 145-8 and nonstoichiometry, 169-74
Dipolar contribution to internal in rare-earth iron garnets, 174-9
field, 110 in spinels, 168-74
Disorder in alloys, 183- 8, 195-6
Disorder caused by shock, 206 Fermi contact interaction, 110
Disproportionation in FeO, 167 Ferredoxins, 241-4
Distortion, determination of, from Focussing ~-spectrometer, 209
quadrupole splittings, 104-7
Doppler broadening, 5 - 7 Gaussian distribution,S
Doppler energy, 4 Geometrical isomerism, 54-65
Doppler scanning, 9 Gold (197 Au), 18
Doppler velocity, 9-12 impurity in the transition metals,
Dysprosium (lsIDy), 18 217-8
paramagnetic relaxation in, in gold-manganese alloys, 189-94
150-2 in inorganic compounds, 127-9
Goldanskii-Karyagin effect, 43,
Effective thickness, 11
143-4
Einstein solid, 6, 138-40
Einstein temperature, 138
Electric dipole (E I) transition, 40 Haemoproteins, 236-41
Electric field gradient tensor, 31-6 Half-life of excited state, 3, 16
asymmetry parameter, 31 Harmonic oscillator approximation, 8
250 Principles of Moss bauer Spectroscopy
Heisenberg uncertainty principle, 2 Iron (57Fe)-cont.
Hyperfine interactions, 22-43 high-spin iron(III) compounds,
109,112-3
impurity diffusion in ice, 147
Instrumentation, 13-6 impurity in CdF 2, 214-6
Intensity of absorption lines, impurity in MnF 2, 216-7
38-43
impurity in MnS2, 122
Interferometer for calibration, 15
impurity studies (see also under
Intermetallic compounds, 188-96
cobalt)
Internal conversion, 9-11
in FeF 2, 115-7
Internal conversion coefficient,
in FeF 3, 112-3
11, 16-18
Internal field, 30 in FeO, 166-7
in FeTi(S04h, 198
Internal standard, 15
Iodine (1271, 1291), 18,67-9,77-84 intermetallic compounds, 188-9,
. after-effects of {3-decay, 229 195-6
intramolecular antiferromagnetism,
calibration of chemical isomer
shift, 77-80 154
effect of covalency, 77-84 ion-exchange resins, 211
iron alloys, 184- 7
effect of i.-character, 83
iron(II) halides, 99-100
in frozen solutions of iodine, 80
iron(II) low-spin compounds, 97
in halide derivatives, 67-8
iron phthalocyanine, 109, 119
in inorganic iodides, 82-3
isocyanide complexes, 54
in iodomethanes, 82
lunar materials, 205-6
in Na3H2I06, 26 matrix isolation, 220-3
in solid iodine, 80
N, N-dialkyldithiocarbamate
Ion-exchange resins, 211
derivative, 118
Iridium (1 93Ir), 18
paramagnetic relaxation in,
after-effects of {3-decay, 230 148-9
after-effects of (n, 'Y) reaction, perovskites, 165,198-9
232 pigments, 199-200
impurity in the transition metals, ranges of chemical isomer shifts,
217-8
76
in inorganic compounds, 126- 7 rare-earth iron garnets, 174-8
Iron (57Fe), 18
relaxation in Fe(acac hCl, 152-3
adsorption of gases, 211 relaxation in FeCl3 . 6H20, 153
carbony1s and derivatives, relaxation in haemin derivatives,
46-53,60-3 153
cyanide complexes, 56-60, relaxation of impurity in Al20 3,
93-6,117,140-3 155
cyc100ctatetraene compound, silicates, 201-6
60-1 single crystal of Cr alloy, 186
effect of pressure on, 122-4 spin cross-over compounds,
electron hopping in Fe304, 118-21
145-6 spinel oxides, 160-74
ferredoxins, 241-4 superparamagnetism in fine
ferrocene, 98-100 particles, 156-7
frozen solutions, 147-8 surface layers, 206-11
haemin derivatives, 237 tetrahedral iron(II) compounds,
haemoglobin derivatives, 238-40 107-8
high-spin iron(II) compounds, Isomer shift (see chemical isomer
102-9 shift)
Index 251
Isomeric transition, effects of, 227-9 Magnetic hyperfine splitting- cant.
Isomerization, 60 origins of internal field, 109-11
Isotopic enrichment, 17-9 sign of internal field, 30
selective, 57 temperature dependence, 112-7
Magnetic moment, 28
lahn-Teller distortion in spinels, 163 Magnetic perturbation method for
sign of e2qQ, 38
Kramers' doublets, 32 in ferrocene, 98
in [(1T-C sH s )Fe(COhhSnCI 2,
Larmor precession frequency, ISO 91-3
Lattice vibrations, 5-8 in iron carbonyl derivatives,
Line-intensities, 38-43 49-50
angular dependence, 40-1 in iron cyanide derivatives, 54
effect of polarization, 43 in iron(I) nitrosyl complex, 118
effect of texture, 43 in organotin compounds, 64-7
effect of thickness, 43 in paramagnetic iron (II) com-
Gold an skii-Karyagin effect, 43 pounds, 118
Lineshape, 10 in tin(II) compounds, 87
Linkage isomerism, 56 Magnetically induced distortion,
Lorentzian distribution, 10, 20 164
Matrix isolation, 220-1
Magnetic dipole interaction, 22 Mean-square vibrational amplitude,
Magnetic dipole (M I ) transition, 40 7,135-40
Magnetic flux density, 28 Mean-square vibrational displace-
Magnetic hyperfine splitting, 23, ment, 135-43
28-30,109-18 Molecular field theory, 172
combined with quadrupole Molecular orbital shceme for ferro-
interaction, 36-8 cene,99
dipolar contribution, 110-7 Mossbauer effect, 5-8
Fermi contact term, 110-7 Mossbauer spectrum, 9-12
in dysprosium compounds, 150-2 Multichannel analyser, 13
in erbium compounds, 152
in europium compounds, 130 Near-neighbour effects in alloys,
in FeF 2, 115-7 184-7,192-4
in FeF 3, 112-3 Neel temperature, 112, 169
in gold-manganese alloys, 189-93 Neptunium (237N p), 18, 131-3
in iridium compounds, 125-6 paramagnetic relaxation in,
in iron alloys, 184-7 131-3
in iron(II) halides, 117 range of chemical isomer shifts,
in iron oxides, 168-78 131
in K3Fe(CN)6, 117 Nickel (61 Ni), 18
in Laves phases, 189 in nickel-palladium alloys, 187-8
in neptunium compounds, 132-3 Non-resonant absorption, II
in nickel-palladium alloys, Non-stoichiometry, 159, 165-8
188-9 in FeO, 166-8
in samarium compounds, 152 in 'Y-Fe203, 165
in 119Sn , 29 in FeTi(S04h, 198
Influence of electronic relaxation in silicates, 201-6
time on, 111, 131-3, in SrRu03-SrFe03, 198-9
148-57 influence on magnetic exchange,
line intensities, 38-43 169-79
orbital contribution, 110-7 site occupancy in spinels, 169
252 Pr"mClp es 0
. 11M" ossbauer Spectroscopy
Nuclear magnet?n (unit) defined, 28 Quadrupole splitting-cant,
Nuclear magnetic resonance, 23 for a~-+2.transition 33-4
Nuclear parameters for selected 2 2 '
transitions, 18 in cis-/trans-isomers, 54-5
Nuclear quadrupole resonance in europium compounds, 130
23, 30 ' in Fe dimers in krypton, 220-1
Nuclear radius, change in 24-5 in ferredoxins, 241-2
Nuclear resonant absorption 2 in ferrocene, 98-100
Nuclear resonant fluorescen~e 2 in gold compounds, 127-9
Nuclear spin states, 1 7 - 8 ' in haemoproteins, 237-9
in iodine compounds, 67-9,
77-84
Orbital contribution to internal
field, 110 in iridium compounds, 126-7
Oxidation state, determination of in iron carbonyls and derivatives
46,56-60,74-7,85, 199 ' 46-53, 55,60-3 '
202-3 ' in iron(III) compounds 108-9
Oxygen deficiency in SrFe03, in iron cyanide comple~es,
165, 198-9 56-60,93-6
in iron(II) isocyanide complexes
54 '
Paramagnetic relaxation, 148-56
in matrix isolated iron atoms
in dysprosium compounds 150-2 220-4 '
in haemoproteins, 238-41'
in neptunium compounds, 131- 3
in iron compounds, 148-9
in organo-tin compounds,
in samarium compounds 152
Parity, 17-8 ' 54-5,63-7,89-93
in paramagnetic iron compounds
Partial quadrupole splitting, 96-8 102-9 '
Phonon energies, 6
in rare-earth iron garnets, 176
Photosynthesis, 240-4
in ruthenium compounds, 124-5
Pigment analysis, 199
in silicates, 201-6
Planktonic ferredoxins, 243-4
in spin cross-over compounds,
Poisson distribution, 20 119-21
Polarized absorber, 41 in tellurium compounds, 84
Pressure, effect of, 121-3 in tetrahedral iron(I1) com-
Pressure induced phase transitions
121-3 ' pounds, 107-8
in tin(II) compounds, 87-8
Principal values of electric field
in xenon compounds, 229-31
gradient tensor, 31
line intensities, 38-43
Protoporphyrin IX, 236
magnitude of q and 1'/, 35
Prussian blue, 58
partial quadrupole splitting,
Pyroxenes, 203-6
96-8
selection rules, 33
Quadrupole moment, 30
Quadrupole splitting, 30-6
at impurity sites in CdF 2 , 216 Rate ~f growth of surface layer, 209
combined with magnetic inter- Rayleigh scattering, 16
actions, 36-8 Recoil-energy, 3-4
determining sign of 35, 38-9 Recoil momentum 4
effect of distortions, 104-6 Recoil-free fractio~, 17, 135-143
effect of pressure, 122 Recoilless fraction, 17, 135-143
effect of spin-orbit coupling anisotropy in, 43, 141-4
104-6 ' in cyanide complexes of iron,
t
for a 1-+ transition, 33-4 140-1
Index 253
Recoilless fraction-cont. Spinel structure- cont.
in iron metal, 140 of 'Y-Fe203, 165
in sodium nitroprusside, 142 of Fe304, 145
low-temperature anharmonicity of oxides AB 20 4, 160-8
in, 141 of sulphides AB2S4, 164
zero-point motion and, 138 trigonal distortion in, 162-3
Recoiliess resonant absorption, 1 Statistics of gamma-ray counting,
Recoilless transitions, 7 13, 20
Relaxation, paramagnetic, 148-56 Sternheimer antishielding factor, 36
Ruthenium (99Ru), 18 Sternheimer shielding factor, 34
impurity in transition metals, Superconducting magnet, 15
217-8 Superparamagnetism, 156-7
in C02Ru04, 163
in inorganic compounds, 124-5 Tellurium (12STe), 18, 84
in SrRu03-SrFe03, 198-9 halides and oxides, 84
transferred hyperfine field in, 180
Samarium (l49Sm), 18 Thermodynamic equilibrium in
paramagnetic relaxation in, 152 pyroxenes, 204
Saturation with thickness, 11 Thickness broadening, 11
Scattering experiment, 16 Tin (119Sn), 18,54-5,63-7,77,
for corrosion studies, 206-11 86-93
for non-destructive analysis, decay after effects in, 227-9
199-200 Goldanskii-Karyagin effect in, 144
Scenedesmus ferredoxin, 241 in organic derivatives, 54-5,
Second-order Doppler shift, 63-7,89-93
26-8, 135-40 in tin(II) compounds, 87-8
high temperature limit of, 138 in tin(IV) halides, 89
in iron cyanide complexes, in palladium alloys, 183-4
140-1 in [(7T-C sH s )Fe(COhhSnCI2, 91-
zero-point motion in, 138-40 matrix isolation studies, 220-1
Selection rules, 28, 33 range of chemical isomer shifts, 77
Selective isotopic enrichment, 57 transferred hyperfine interactions
s-Electron density at the nucleus, in, 179-80
24-5,74-5 Transferred hyperfine interactions,
Shielding effects on chemical isomer 170, 179-80
shift,25 Transmission integral, 11
Shock, effect on disorder, 206 Transmitted intensity, 12
Silicates, 201-6 Turnbull's blue, 58
Source matrix, 17
Sources for Mossbauer experiments, Velocity calibration, 15
16-9 Velocity scanning, 9
Spectrometer, 13-6 Verwey transition in Fe304, 145-6
Spin canting, 168, 172-3
Spin cross-over, 119-21 Works of art, 199-200
Spin-lattice relaxation, 149
Spin-orbit coupling, 104-9 X-rays from internal conversion, 16-7
Spin-spin relaxation, 150 Xenon (129Xe), 18,229-30
Spinach ferredoxin, 241 after effects of {3-decay, 229-30
Spinel structure, 160-1
exchange interactions in, Yaffet-Kittelordering, 168
168-74
lahn-Teller distortion of, 163 Zero-phonon transitions, 7
254 Principles of Mossbauer Spectroscopy
Recoilless fraction-cant. Spinel structure- cant.
in iron metal, 140 of ,),-Fe203, 165
in sodium nitroprusside, 142 of Fe304, 145
low-temperature anharmonicity of oxides AB 20 4, 160-8
in, 141 of sulphides AB 2S4, 164
zero-point motion and, 138 trigonal distortion in, 162-3
Recoilless resonant absorption, 1 Statistics of gamma-ray counting,
Recoilless transitions, 7 13, 20
Relaxation, paramagnetic, 148-56 Sternheimer antishielding factor, 36
Ruthenium (99Ru), 18 Sternheimer shielding factor, 34
impurity in transition metals, Superconducting magnet, 15
217-8 Superparamagnetism, 156-7
in C02Ru04, 163
in inorganic compounds, 124-5 Tellurium (125Te), 18, 84
in SrRu03-SrFe03, 198-9 halides and oxides, 84
transferred hyperfine field in, 180
Samarium (149Sm), 18 Thermodynamic equilibrium in
paramagnetic relaxation in, 152 pyroxenes, 204
Saturation with thickness, 11 Thickness broadening, 11
Scattering experiment, 16 Tin (1l9Sn ), 18,54-5,63-7,77,
for corrosion studies, 206-11 86-93
for non-destructive analysis, decay after effects in, 227-9
199-200 Goldanskii-Karyagin effect in, 144
Scenedesmus ferredoxin, 241 in organic derivatives, 54-5,
Second-order Doppler shift, 63-7,89-93
26-8, 135-40 in tin(II) compounds, 87-8
high temperature limit of, 138 in tin(IV) halides, 89
in iron cyanide complexes, in palladium alloys, 183-4
140-1 in [(rr-C sH s )Fe(COhhSnCI2, 91-3
zero-point motion in, 138-40 matrix isolation studies, 220-1
Selection rules, 28, 33 range of chemical isomer shifts, 77
Selective isotopic enrichment, 57 transferred hyperfine interactions
s-Electron density at the nucleus, in, 179-80
24-5,74-5 Transferred hyperfine interactions,
Shielding effects on chemical isomer 170, 179-80
shift, 25 Transmission integral, 11
Shock, effect on disorder, 206 Transmitted intensity, 12
Silicates, 201-6 Turnbull's blue, 58
Source matrix, 17
Sources for Mossbauer experiments, Velocity calibration, 15
16-9 Velocity scanning, 9
Spectrometer, 13-6 Verwey transition in Fe304, 145-6
Spin canting, 168, 172-3
Spin cross-over, 119- 21 Works of art, 199-200
Spin-lattice relaxation, 149
Spin-orbit coupling, 104-9 X-rays from internal conversion, 16-7
Spin-spin relaxation, 150 Xenon (129Xe), 18,229-30
Spinach ferredoxin, 241 after effects of ~-decay, 229-30
Spinel structure, 160-1
exchange interactions in, Yaffet-Kittelordering, 168
168-74
Jahn-Teller distortion of, 163 Zero-phonon transitions, 7