You are on page 1of 32

Hyperfine Interactions 117 (1998) 3970 39

Clays and clay minerals: What can Mossbauer


spectroscopy do to help understand them?
Enver Murad
Bayerisches Geologisches Landesamt, Concordiastrasse 28, D-96049 Bamberg, Germany

Mossbauer spectroscopy is a powerful technique for the characterization of materials


formed in the weathering environment. Mossbauer studies of clay-sized phyllosilicates,
however, are burdened with several problems: the samples are rarely monomineralic, they
may be poor in iron, and only few iron-rich species order magnetically above 4.2 K. Site
occupancies are difficult to determine, and cis and trans octahedral-OH site assignments are
normally not possible. Unequivocal information that can be gained from such work thus is
often restricted to the determination of the oxidation state of iron and average structural site
distortions.
Mossbauer data on iron oxides are generally more straightforward to interpret because
these can be studied in the magnetically ordered state. A further asset of Mossbauer spec-
troscopy when studying iron oxides lies in its high sensitivity for magnetically ordered
phases. Adverse effects ensuing from small particle size, interparticle interactions, non-
stoichiometry and foreign-element substitution that often affect the Mossbauer parameters
of iron oxides occurring in clays and soils can be at least partly offset by taking spectra at
low temperatures.

1. Introduction

Clay minerals are the most complex of all minerals they mostly are colloidal in
size, often in metastable equilibrium with their environment, and they have a very
complex internal structure that reflects where theyve been and where they might
go. (W.F. Moll Jr., CMS News 8, 1996).
Phyllosilicates are the dominant, though not exclusive, mineral constituents of
clays, but the term clay is also in use as a particle-size designation [1]. A recent
controversial discussion attests to the lack of comprehensive agreement on use of the
terms clay and clay mineral [2,3]. The present paper deals with two groups of
minerals in the clay-size fraction (<2 m as used here) that are independent of se-
mantic discussions common components of soils and sediments and the predominant
constituents of industrial clays: phyllosilicates and iron(III) oxides.
The first definitive Mossbauer data on clay-sized phyllosilicates were published
some 30 years ago. Weaver et al. [4] showed Mossbauer spectra and gave parameters
for 16 minerals including each one goethite and siderite. In a different approach,
Malden and Meads [5] presented Mossbauer spectra of two commercial kaolins and

J.C. Baltzer AG, Science Publishers


40 E. Murad / Clays and clay minerals study by Mossbauer spectroscopy

critically analyzed the data with respect to the influence of ancillary mica on the
parameters of the spectra.
Numerous studies addressing the applications of Mossbauer spectroscopy to dif-
ferent aspects of clay mineralogy (e.g., properties of specific minerals, influence of
weathering and oxidation, industrial applications) have appeared in the sequel, and
results of this work on phyllosilicates have been summarized in reviews by Coey [6]
and Heller-Kallai and Rozenson [7]. More recent work involving well-characterized
materials and high-quality data, however, has cast doubt upon some earlier assign-
ments of spectral features. Mossbauer spectra of clay-sized phyllosilicates often have
significantly broadened resonant lines, inciting the use of multiple doublets to attain
statistically good fits. This carries the risk of over-interpretation, a problem that will
be addressed in more detail below.
The study of phyllosilicates by Mossbauer spectroscopy is associated with a
variety of potential pitfalls that are often not adequately appreciated, and is therefore
no simple matter. Typical potential problems that may have to be overcome are the
following:
Clays are hardly ever monomineralic. This precludes an unequivocal assignment of
spectral features to phyllosilicate structures more often than is generally realized (or
sometimes admitted). Usually selective pretreatments followed by thorough control
(e.g., by slow step-scanned X-ray diffraction) are necessary to ensure the absence
of deleterious ancillary phases.
Phyllosilicates may be poor in iron. This leads to low count rates and increases
the risk of interference by ancillary phases. Low iron contents can also bring on
adverse physical effects such as paramagnetic relaxation.
Only few very iron-rich phyllosilicates order magnetically above 4.2 K. This re-
stricts the parameters provided by Mossbauer spectra to the isomer shift and quadru-
pole splitting. The immediate information that can be derived from these parameters
are the oxidation state of iron, in favourable cases its coordination, and the iron
site distortion.
Because of their platy morphology, phyllosilicates tend to preferred orientation
(texture). This will usually result in unequal doublet line intensities, which needs
to be taken into consideration when fitting spectra, and unless spectra are taken
under the magic angle (i.e., with the sample oriented at an angle of 54.7 to the
ray) or at different temperatures, may also be misinterpreted as resulting from a
GoldanskiiKaryagin effect.
Although iron oxides have attracted the attention of the Mossbauer community
at an initiatory stage, publication of systematic work on samples of small particle
size [8,9] started at about the same time as that on clay-sized phyllosilicates. Initial
work on natural poorly-crystalline iron oxides by Janot et al. [10] has not received
adequate attention, probably because it was published in French. Early Mossbauer
studies on iron oxides have been comprehensively reviewed by Bowen [11] and more
recent work by Murad and Johnston [12].
Table 1
Mossbauer parameters of naturally-occurring Fe3+ oxides and oxyhydroxides and common Fe3+ -rich acid mine drainage minerals.
Mineral Occurrence Composition TN , TC MAGa Bhf /Fe Bhf
Room temperature 4.2 K
Magnetite common Fe3 O4 850 fim 49.2 0.26 6 |0.02| 50.6 0.00

E. Murad / Clays and clay minerals study by Mossbauer spectroscopy


46.1 0.67 6 |0.02| 3652b 1.18(0.79)

Hematite very common -Fe2 O3 955 wfm 51.8 0.37 0.20 53.5 0.20
afm 54.2 0.41

Maghemite common -Fe2 O3 950 fim 50.0 0.23 6 |0.02| 52.0 6 |0.02|
50.0 0.35 6 |0.02| 53.0 6 |0.02|

Goethite very common -FeOOH 400 afm 38.0 0.37 0.26 50.6 0.25

Akaganeite very rare -FeOOH 299 afm 0.38 0.55 47.3 0.81
0.37 0.95 47.8 0.24
48.9 0.02

Lepidocrocite common -FeOOH 77 afm 0.37 0.53 45.8 0.02

Feroxyhite very rare -FeOOH 450 fim 41 0.37 0.06 53 0.0


52 0.0

Ferrihydrite common Fe5 HO8 4H2 O 115 spm 0.35 0.62d 50d 0.07
25c 0.35 0.78d 47d 0.02

Bernalite very rare Fe(OH)3 427 wfm 41.5 0.38 6 |0.01| 56.2 6 |0.01|

Jarosite common RFe3 (OH)6 (SO4 )2 54 afm 0.38 1.13 47.9 0.19

Schwertmannite common Fe8 O8 (OH)6 SO4 75 afm 0.39 0.65d 45.0d 0.35
0.33 0.65d
45.3d 0.01
a
Magnetic character (fim = ferrimagnetic, afm = antiferromagnetic, wfm = weakly ferromagnetic, spm = speromagnetic).
b
Several magnetic B-site subspectra below the Verwey transition at 120 K.
c
Range of superparamagnetic blocking temperatures, which vary as a function of crystallinity.
d
Maximum probabilities of quadrupole-splitting and hyperfine-field distributions.

41
42 E. Murad / Clays and clay minerals study by Mossbauer spectroscopy

In the context of this paper the term iron oxides refers to three oxides, five
oxyhydroxides, and one hydroxide of Fe3+ , which have been identified in nature to
date (table 1). The interpretation of Mossbauer spectra of iron oxides is generally
more straightforward than for phyllosilicates for several reasons:
Because of high iron contents about 7258 wt.%, iron oxides bring on a relatively
high resonant absorption. This increases the sensitivity of Mossbauer spectroscopy
to the iron oxides relative to other, possibly interfering iron-bearing minerals.
All bulk and most superparamagnetic iron oxides order magnetically at temperatures
substantially above 4.2 K. Thus, parameters relating to their magnetic properties,
e.g., magnetic hyperfine fields, magnetic ordering temperatures, and angles between
the directions of the magnetic hyperfine field and the principal axis of the electric
field gradient, can usually be obtained at a passable expenditure of time and cost.
The magnetic properties of iron oxides vary systematically as a function of different
factors such as particle size and substitution of iron by other elements, particularly
aluminium. Although a direct assessment of such factors will normally not be
possible (because the influences of these factors add up), reduced hyperfine fields
can serve as a general indication for deviations from crystalline and/or chemical
perfection.
In natural samples, iron oxides are commonly associated with other minerals. As a
consequence particle morphology, which is less pronounced than for the phyllosil-
icates at any rate, will usually not play a major role.
Mossbauer spectroscopy thus can serve to unequivocally identify iron oxides
occurring in natural samples and distinguish these from possibly associated silicates
in samples of complex mineralogy. However, as for the clay-sized phyllosilicates,
microcrystalline iron oxides often also exhibit broadened resonant lines, complicating
the attribution of unequivocal parameters to the spectra. Problems may also arise
if a sample contains different iron oxides that have similar magnetic hyperfine fields,
requiring spectra to be taken over a range of temperatures or the application of physical
or chemical separation procedures.
The present paper complements recent reviews [1320] on the Mossbauer spectra
of clays, soils and related materials, and presents an assessment of the present state of
knowledge on the Mossbauer spectra of clays and clay minerals sensu lato.

2. Clay-sized phyllosilicates

2.1. General properties

The basic structural elements of all phyllosilicates are similar: sheets of corner-
sharing oxygen tetrahedra with Si4+ and generally some Al3+ as central cations, and
sheets of edge-sharing octahedra in which oxygens and hydroxyls surround a variety
of possible di- or trivalent cations (mainly Mg2+ , Fe2+ , Al3+ and Fe3+ ). These sheets
E. Murad / Clays and clay minerals study by Mossbauer spectroscopy 43

can be linked in different manners, e.g., one tetrahedral sheet to one octahedral sheet
(forming a layer termed 1 : 1) or one octahedral sheet sandwiched between two tetra-
hedral sheets (2 : 1). The phyllosilicates are commonly classified by a sequence of
characteristics, the primary attribute being the layer type (which is directly reflected
in the basal spacing). This is followed by the type of interlayer (individual cations,
hydrated cations, octahedral hydroxyl sheets), the layer charge (resulting from isomor-
phous substitutions in the tetrahedral or octahedral sheets), and the occupancy of the
octahedra (ideally either two out of three octahedra should be occupied by trivalent or
all octahedral positions by divalent cations, arrangements that are termed dioctahedral
and trioctahedral, respectively).
Because of isomorphous substitutions within either or both the octahedral and
tetrahedral sheets, many phyllosilicates have a usually negative charge, which must
be offset to achieve electric neutrality. Depending on the magnitude of the layer charge,
this can be compensated for in the different manners outlined above: among the more
common phyllosilicates, micas typically have a layer charge close to 1 per formula
unit and K+ in the interlayer, vermiculites and smectites have layer charges from
0.90.6 and 0.60.2, respectively, and can accommodate a variety of materials in their
interlayers. Chlorites have variable layer charge and octahedral hydroxyl interlayers
(brucite sheets), whereas kaolinites and serpentines possess neither layer charge nor
interlayers. It must, however, be pointed out that this ideal classification scheme is
complicated in practice by the existence of structural polytypes and interstratification
of different layer structures.
The most effective technique for the identification of phyllosilicates is X-ray dif-
fraction, commonly combined with various treatments to determine the layer charge
and character of the interlayer [21,22]. Figure 1a shows the differences in X-ray dif-
fraction between a kaolinite and an illite, two of the most common 1 : 1 and 2 : 1
clay minerals. Among other instrumental techniques that lack the predicative power
of XRD but are nevertheless commonly used in phyllosilicate studies are thermal
methods (figure 1b), infrared spectroscopy (figure 1c) and Mossbauer spectroscopy
(figure 1d). Magic angle spinning (MAS)-NMR can produce high-quality data on Al
and Si coordination and site occupancies in phyllosilicates, but the fact that paramag-
netic components cause extensive line broadening limits its applicability to samples
containing iron at concentrations even below the percent range [23].
Fe3+ has an ionic radius of 64.5 pm in octahedral coordination and 49 nm in
tetrahedral coordination [24], and can therefore not only enter the octahedra, but also
substitute to some extent for Si4+ (which has an ionic radius of 26 pm) in the tetrahedra.
Fe2+ has a significantly larger ionic radius of 78 pm in octahedral and 63 pm in
tetrahedral coordination, and therefore fits only into the octahedra. Depending on the
locations of OH in the octahedra, two sites can be distinguished: cis (M2) in which
the hydroxyls are adjacent to another and trans (M1) in which they occupy opposite
corners of the octahedra.
Because of the paramagnetic nature of essentially all phyllosilicates, Mossbauer
spectra of these typically consist of one or more doublets ensuing from Fe in the
44 E. Murad / Clays and clay minerals study by Mossbauer spectroscopy

Figure 1. (a) X-ray diffraction traces of a kaolinite (Amazon) from the Jari River deposit, Brazil, and
an illite (OECD #5) from Le Puy, France (modified from [28,36]). The poor resolution of the 020, 11l
peaks of the kaolinite (2225 2) and high line width of the first basal peak of the illite indicate both
minerals to have a low degree of structural order [163,164]. (b) Differential thermal analysis (DTA) and
thermogravimetry (TG) of the Amazon kaolinite and the OECD #5 illite. Note the sharp dehydroxylation
peak of kaolinite at 590 C and the associated well-defined loss in mass, which is in distinct contrast to
the dehydration peak of the illite around 135 C and the more gradual loss in mass at higher temperatures.
(c) Transmission infrared spectra of the Amazon kaolinite and the OECD #5 illite. Note the differences
in the OH-stretching (36003700 cm1 ) and lattice vibrational regions (<1105 cm1 ). (d) Room-
temperature Mossbauer spectra of the Amazon kaolinite and the OECD #5 illite (modified from [28,36]).
The only manifest difference is the absence of Fe2+ in the kaolinite and its presence in the illite.

di- or trivalent state and, where applicable, on different structural sites. Thus the
Mossbauer spectrum of an illite could theoretically comprise five subspectra ensuing
from Fe3+ and Fe2+ in cis and trans octahedral sites and Fe3+ in tetrahedral sites.
In practice this complex situation will be rendered even more complicated because
isomorphous substitutions in both the octahedra and tetrahedra lead to a variety of
non-identical environments for each site, thus causing the individual nuclear energy
levels to be smeared out. This is the principal cause for the broadened resonant lines
often exhibited by clay-sized phyllosilicates, and it is not surprising that a variety of
different models, ranging from only one doublet each for Fe2+ and Fe3+ to as much as
five different doublets, has been used to fit Mossbauer spectra of such minerals, often
making a comparison of data even on one and the same sample all but impossible.
In the majority of earlier studies, Mossbauer spectra of phyllosilicates were fit-
ted with as many doublets as possible, the doublets being subsequently assigned to
E. Murad / Clays and clay minerals study by Mossbauer spectroscopy 45

Figure 1. (Continued.)

possible structural sites (e.g., cis and trans octahedral, tetrahedral, and in some cases
interlayer sites). Assignment of Fe3+ doublets with which Mossbauer spectra of iron-
rich dioctahedral phyllosilicates have been fitted to cis and trans octahedral sites in the
structure, once considered self-evident, however, are in conflict with X-ray and elec-
tron diffraction data, which indicate only the cis sites to be occupied [25,26]. A critical
examination of the parameters of such multiple doublets furthermore often shows al-
leged differences between the individual parameters to lack statistical significance [27],
and exacting Mossbauer work carried out on micas and clay-sized phyllosilicates has
recently shown such assignments to be untenable in the majority of cases [28,29].
Attributions of spectral components that are based merely on 2 improvements in the
course of spectral fitting, i.e., in the absence of visible spectral features, to tetrahedral
Fe3+ have also been lately disputed [30].
The following possibilities of fitting Mossbauer spectra of clay-sized phyllosili-
cates have been made use of in current studies of these minerals:
A minimal model-free solution, fitting only as many Lorentzians as are absolutely
necessary. In general, one pair of Lorentzians each for Fe2+ and Fe3+ (which are
readily distinguished on the basis of their different isomer shifts) will suffice, and
only in selected cases will a third pair of Lorentzians be necessary to account for
tetrahedral Fe3+ . This method may not always produce perfect fits, but the results
are consistent and serviceable.
The use of distributions of quadrupole-split doublets to account for broadened
lines [32,39]. In contrast to the assignment of multiple doublets to structural sites,
46 E. Murad / Clays and clay minerals study by Mossbauer spectroscopy

which is questionable in most cases, the fitting of distributions produces physically


unbiased data of high quality.
The use of models derived from the actual clay structures, under consideration
of all the local structural environments of the iron atoms derived from computer
simulation [33,34]. This approach, however, requires supporting data from other
techniques (e.g., infrared spectroscopy) and has rigorous requirements regarding the
monomineralic composition of samples.
A phenomenon that is often ignored in Mossbauer studies of iron-poor phyl-
losilicates is paramagnetic relaxation. This results from the fact that relaxation of the
electron spins in magnetically dilute systems is no longer necessarily faster than the
lifetime of the excited nucleus, leading to the development of magnetically split spec-
tra. In the case of Fe3+ , paramagnetic relaxation produces complex spectra which are
usually characterized by a high magnetic field of about 55 T at low temperatures (e.g.,
4.2 K). At room temperature paramagnetic relaxation leads to magnetic fields of a few
Tesla at room temperature that show up as a more or less broad background which
is easily overlooked. Paramagnetic relaxation appears to be a common phenomenon
in kaolinites [35,36], and must be taken into consideration when fitting Mossbauer
spectra of this mineral.
In the phyllosilicates Fe3+ is in the high-spin state, and thus has no inherent elec-
tric field gradient. The quadrupole splitting is therefore determined solely by external
charges; it will be zero for cubic symmetry (e.g., in perfect octahedral coordination),
and increase proportionally with the iron site distortion. As a consequence, systematic
variations have been reported for structurally related minerals such as the dioctahe-
dral 2 : 1 phyllosilicates, in which the octahedral site distortion and consequently the
quadrupole splitting decreases with increasing iron content [7]. In favourable cases
Mossbauer spectroscopy can also reveal specific characteristics of individual minerals;
thus, the quadrupole splitting of a series of kaolinites was observed to increase with the
Fe2+ proportion, probably because the introduction of this large ion into the structure
leads to an increasing average site distortion [36].
Because of the difficulty of inducing magnetic order, only few Mossbauer studies
have been carried out on magnetically ordered phyllosilicates. In a study on six non-
tronites, Cardile et al. [37] observed magnetic order in three samples at 4.2 and in all
six at 1.3 K, with complex spectra the components of which were assigned to two cis
octahedral and a tetrahedral site. Townsend et al. [38] also observed a nontronite to
order magnetically between 4.2 and 1.3 K, and attributed the low ordering temperature
to either spin-glass behaviour or a partially-disordered, triangular-spin configuration
(frustration). Murad et al. [39] were able to generate magnetic order in a nontronite by
applying external magnetic fields up to 9 T at temperatures 537 K, and concluded that
frustration due to tetrahedral Fe3+ lowered the magnetic ordering temperature from a
bulk Neel temperature of about 20 K to below 7 K. By pillaring a nontronite with an
iron oxide, Gangas et al. [40] were able to raise the magnetic ordering temperature of
this sample from below 10 K to above 44 K.
E. Murad / Clays and clay minerals study by Mossbauer spectroscopy 47

There have also been several attempts to introduce extrinsic cations containing
Mossbauer isotopes into the interlayers of smectitic clays [4143]. Such work is of
interest not only with reference to properties of the intercalated ions, but also because
it can disclose characteristic properties of the adsorbent minerals.
The element-selectivity of Mossbauer spectroscopy thus renders it effective for
the characterization of any iron that phyllosilicates may contain (and, where relevant,
other nuclides with which such minerals may have been treated). The differences in
Mossbauer parameters between the individual clay mineral species, however, are not
sufficiently pronounced to allow an identification of these to be made in samples of
complex mineralogy, and attempts to apply Mossbauer spectroscopy to the identifica-
tion of phyllosilicates (e.g., [44]) have not received widespread acceptance. It must,
therefore, be stressed that Mossbauer spectra of clay-sized phyllosilicates should al-
ways be viewed in combination with data from other techniques if reliable structural
information is required.
The difficulty of obtaining monomineralic clays has been pointed out above.
The majority of studies on natural clay-sized materials (including the vast majority
of commercial clays) is concerned with samples of composite mineralogy. When
such samples are studied by Mossbauer spectroscopy, crystallographic assignments of
spectral features will no longer make sense. Fe2+ /Fe3+ ratios, however, may reflect
environmental conditions [4547]; phyllosilicates can be distinguished from iron ox-
ides (usually requiring spectra taken at low temperatures); and selective concentration
procedures, e.g., particle size fractionation and/or gravity separation [48], can raise
the proportions of specific minerals or mineral groups to a level where they produce
meaningful Mossbauer spectra (figure 2).
Mossbauer spectroscopy has been successfully applied to the elucidation of
changes induced in clay minerals and clays in the course of oxidation (e.g., [49,50])
and firing. The latter topic will be treated in detail in a subsequent paper.

2.2. Specific minerals

The following section gives a brief outline of Mossbauer data on three phyllosil-
icates that belong to different groups and vary widely in a number of properties such
as iron content and particle size. The data on these minerals can thus serve to illustrate
the possibilities and limitations of Mossbauer spectroscopy for the study of clay-sized
phyllosilicates.
Kaolinite, Al4 Si4 O10 (OH)8 , usually with a minor substitution of Fe for Al, has
a variety of applications in industry, of which employment as paper filler and coating,
and for the production of ceramics are the most important (the term kaolin refers
to commercial clays that consist predominantly of kaolinite or its polytypes, dickite
and nacrite). Particularly the latter use requires a material with a low iron content,
which automatically limits possibilities for study by Mossbauer spectroscopy. The
already mentioned pioneering work of Malden and Meads [5] showed up some typical
complications and possibilities to overcome these.
48 E. Murad / Clays and clay minerals study by Mossbauer spectroscopy

Figure 2. Room-temperature Mossbauer spectra of different particle-size fractions of a Vertisol from


Cameroun: (a) 20.6 m, (b) 0.60.2 m and (c) <0.2 m. Fe3+ subspectra are plotted with solid
lines, Fe2+ with broken lines, and non-phyllosilicate components (ilmenite and epidote) with dotted lines
(modified from [48]).
E. Murad / Clays and clay minerals study by Mossbauer spectroscopy 49

Kaolinites usually have relatively low structural iron contents, ranging from es-
sentially nil in some commercial samples to a maximum of about 2 wt.% in soil
kaolinites [51]. Associated (super)paramagnetic iron-rich phases may therefore ex-
ert a significant influence on the Mossbauer spectra of kaolinites, and care must be
taken to ensure the absence of these if the spectra are to be interpreted in struc-
tural terms. Heller-Kallai and Rozenson [7] noted a disturbingly wide spread of
published Mossbauer data on kaolinites, for which such ancillary minerals may, at
least in part, be responsible. The specificity of Mossbauer spectroscopy to magnet-
ically ordered iron oxides, on the other hand, makes it a particularly suitable tech-
nique for the identification of such ancillary minerals if they are magnetically or-
dered [52], and the sensitivity of Mossbauer spectroscopy to iron oxides is such that
as little as 0.1% goethite associated with a kaolinite could be unequivocally identi-
fied [36].
In a comprehensive study, Murad and Wagner [36] reviewed the state of knowl-
edge on the Mossbauer spectra of kaolinite and presented novel data on 10 kaolinites
selected on the basis of mineralogical purity (i.e., the absence of potentially interfer-
ing phases) and a wide range of degrees of structural order and iron contents. The
Mossbauer spectra showed the majority of samples to contain divalent and trivalent
iron in octahedral coordination, i.e., isomorphously substituting for aluminium. The
room-temperature parameters for Fe3+ (isomer shift 0.35 mm/s relative to metallic
iron and quadrupole splitting 0.51 mm/s for Fe2+ -free kaolinite, increasing to about
0.65 mm/s for samples with high Fe2+ /Fe3+ ratios) and Fe2+ (isomer shift 1.11 mm/s
and quadrupole splitting 2.54 mm/s) showed only moderate variation and closely re-
sembled those published by St. Pierre et al. [51]. It is thus conceivable that the spread
of data in the literature, rather than reflecting a genuine variation of parameters, may
result from one or several of the following causes: interference by ancillary con-
stituents, inadequate counting statistics, inappropriate fitting procedures (in particular,
disregard of paramagnetic relaxation) and, where relevant, disregard of the possible
influence of structural Fe2+ .
Illite, theoretically Kx Al2 (Si4x Alx )O10 (OH)10 (a typical composition was given
by Bailey et al. [53], as K0.75 (Al1.75 R2+0.25 )(Si3.5 Al0.5 )O10 (OH)2 ), is a common clay-
sized mica that has more silicon and less potassium than muscovite, and consequently
a somewhat lower layer charge. Illite is a common constituent of soils, rocks and
shales; it is also the main source of alkali in clays. Illites invariably contain iron in
concentrations ranging from <1 to over 8 wt.% [54,55]. Glauconite and celadonite are
iron-rich variants of illite, the definitions of which, however, are somewhat nebulous.
A distinction between these can be effected on the basis of layer charge thus that the
layer charge of glauconite results from isomorphous substitution in both octahedral
and tetrahedral sites, whereas about half the octahedral sites in celadonite are occupied
by divalent cations [56].
Mossbauer spectra of illite have relatively broad, non-Lorentzian lines. As a
result such spectra can be fitted using various models (one, two and three discrete
Fe3+ doublets and distributions of quadrupole-split Fe3+ doublets, and one or two
50 E. Murad / Clays and clay minerals study by Mossbauer spectroscopy

Table 2
Room-temperature Mossbauer parameters of the illite OECD #5 (cf. figure 4) fitted with various models.
Spectrum taken in the velocity range 4.00 mm/s. Data from Murad and Wagner [28] and Murad
(unpublished). Errors on the last digit are given in parentheses.
Fe3+ Fe2+
1 /Fe2 W3 A4 1 /Fe2 W3 A4 2

0.606 (2) 0.356 (1) 0.51 (1) 0.950 (6) 2.63 (2) 1.110 (8) 0.28 (2) 0.050 (4) 1.48

0.43 (3) 0.349 (3) 0.41 (2) 0.46 (8) 2.67 (2) 1.08 (1) 0.33 (3) 0.058 (4) 1.08
0.80 (3) 0.358 (2) 0.43 (2) 0.48 (8)

0.40 (4) 0.350 (3) 0.40 (3) 0.46 (13) 2.64 (7) 1.10 (4) 0.44 (4) 0.070 (6) 0.97
0.76 (3) 0.356 (2) 0.36 (6) 0.40 (16)
1.22 (5) 0.36 (2) 0.29 (7) 0.07 (4)

0.42 (3) 0.348 (3) 0.41 (2) 0.44 (8) 2.57 (3) 1.06 (1) 0.24 (5) 0.035 (7) 1.02
0.79 (2) 0.360 (2) 0.43 (2) 0.50 (8) 2.89 (3) 1.14 (2) 0.20 (5) 0.021 (6)

0.605 0.354 (1) 0.93 (7) 2.56 (2) 1.11 (1) 0.07 (3) 1.08
1
Quadrupole splitting (in mm/s).
2
Isomer shift w.r.t. metallic Fe (in mm/s).
3
Full width at half maximum (in mm/s).
4
Relative area.
5
Distribution of quadrupole doublets. Quadrupole splitting and isomer shift of Fe2+ component from
Lorentzian doublet fit.

Fe2+ doublets), which renders a direct comparison of many data virtually impossible.
Table 2 lists the parameters ensuing from such different fits to the Mossbauer spectrum
of the OECD #5 illite (figure 1d).
The current state of knowledge on the Mossbauer spectra of illite was outlined
and Mossbauer data on eight pure illites with iron contents 0.88.4% were presented
by Murad and Wagner [28]. All samples contained divalent and trivalent iron, the
proportion of the former varying between 4 and 44%. Tetrahedral Fe3+ was detected
only in the most iron-rich sample (that had 8.41 wt.% Fe), and a distinction of Fe sites
with cis and trans-OH coordination was not possible. The Fe3+ quadrupole splitting
was higher (0.73 mm/s) in the Fe-poor illites (<1 to 3 wt.% Fe) than in the Fe-
rich (>5 wt.% Fe) samples (0.59 mm/s). The iron-poor illites, as in the case of the
kaolinites, exhibited paramagnetic relaxation. The presence of iron oxides that were
superparamagnetic at room temperature was established in two samples at 4.2 K. The
contributions of the iron oxides to the Mossbauer spectra which would have gone
unnoticed at room temperature amounted to 75% of the spectral area in one case,
corresponding to 1% ferrihydrite in the sample.
Nontronite, ideally M+ x Fe2 (Si4x Alx )O10 (OH)2 (the average composition was
given by Koster [57] as M+ 3+
0.4 (Fe1.6 Al0.2 Mg0.3 )(Si3.6 Al0.4 )O10 (OH)2 ), is the iron-rich
end-member dioctahedral smectite. The particle sizes of nontronites, as of smectites
E. Murad / Clays and clay minerals study by Mossbauer spectroscopy 51

Figure 3. Mossbauer spectra of nontronites: (a) CMS SWa-1, (b) A.P.I. H33a, and (c) from Hundsangen,
Germany. Spectra (a) and (b) were taken at 77 K, (c) was taken at 120 K. All spectra show the presence
of ancillary goethite with relative areas amounting to 3%, 5% and 13% in the above order (modified
from [39]).
52 E. Murad / Clays and clay minerals study by Mossbauer spectroscopy

in general, are significantly smaller than those of the two preceding minerals, the
thicknesses of some samples amounting to about only 6 unit cells [58].
Mossbauer studies on a variety of nontronites have shown iron in these to be
predominantly to exclusively trivalent. The formulae show that the layer charge of
nontronite arises essentially from tetrahedral substitution. A feature that is not imme-
diately apparent from the formulae, however, is the possible presence and extent of
tetrahedral Fe3+ . Similarly to cis and trans-OH assignments, this has been a subject
of some controversy [59]. As an example, the proportion of tetrahedral Fe3+ in the
Garfield nontronite has been given as very low (<2%) [6062]; as 5% by [39,63];
and as 9% [64,65]. In this context it is worth noting that a spectrum of the Garfield
nontronite taken by Bonnin et al. [60] at 77 K indicated in distinct contrast to the
room-temperature spectrum the presence of a component of relative area 16% that
had a low isomer shift of 0.299 mm/s and could thus easily pass as resulting from
tetrahedral Fe3+ . Mossbauer spectra of a series of synthetic nontronites consisted of
broad doublets that were assigned to octahedrally-coordinated Fe3+ [66]. Neither Fe2+
nor tetrahedral Fe3+ were observed, although chemical analyses and infrared spectra
did indicate the existence of the latter.
Mossbauer spectra taken at low temperatures (6120 K) have furthermore revealed
a common association of nontronites with iron oxides that are superparamagnetic at
room temperature [39,67], cf. figure 3, making crystallographic assignments that are
based on room-temperature spectra only questionable. The variation of tetrahedral iron
ratios reported for the Garfield nontronite in the mentioned studies thus may, at least
in part, ensue from the presence of unlike amounts of associated goethite in different
aliquots of this sample, the presence of which leads to broadened lines (which are less
well defined and more difficult to fit) at room temperature.
Ancillary iron oxides associated with nontronites cannot be simply removed by
dithionite reduction, because this treatment will not only reduce structural Fe3+ in the
nontronites [68], but may also effect a modification of the reoxidized product [69,70].
Mossbauer spectra taken at liquid nitrogen temperature allow provision to be made
for associated iron oxides without having to remove these, and therefore constitute a
practicable alternative to deleterious chemical treatments [39].
As in the case of illite, Mossbauer spectra of nontronite have broad lines, and a
variety of fitting models exists in the literature, again complicating the comparison of
published data. The most distinctive feature of nontronite Mossbauer spectra is a low
average Fe3+ quadrupole splitting of about 0.4 mm/s, in line with the comparatively
low site distortion characteristic for iron-rich dioctahedral phyllosilicates [7,71].

3. Iron(III) oxides

3.1. General properties

This section is concerned with the oxides of trivalent iron that occur in the weath-
ering environment (table 1). Data are also given on iron oxyhydroxysulfates that are
E. Murad / Clays and clay minerals study by Mossbauer spectroscopy 53

commonly associated with iron(III) oxides in specific weathering environments such


as acid sulfate soils and mine drainage precipitates. Minerals that contain cations
other than iron (e.g., all spinels with the exception of magnetite and maghemite) are
excluded, as are synthetic species like the high-pressure phase -FeOOH, and indeter-
minate varieties that are not recognized mineral species, e.g., limonite, defined by
Fleischer and Mandarino [72] as a general field term for hydrous iron oxides, mostly
Goethite or hydrohematite, a term occasionally used to specify hydroxyl-containing
hematite (which Stanjek and Schwertmann [73] showed to be nothing unusual any-
way). Materials with designations such as amorphous ferric hydroxides or iron
oxide gels have been often used when referring to minerals that are not genuinely
amorphous, but simply poorly crystalline. These materials produce broad and weak
X-ray diffraction maxima that are easily overlooked unless measurements are carried
out under suitable experimental conditions (e.g., using slow step-scanning and peak
profile analysis). Typical examples for minerals that have been frequently misnamed
in such a manner are ferrihydrite [74] and schwertmannite (cf. section 4.3).
Iron oxides form in the course of release of iron during the weathering of
iron-containing minerals, followed by reprecipitation under oxidizing conditions in
the supergene zone. All of the iron oxides listed in table 1 have been described
in nature. Goethite and hematite are by far the most common iron oxides, but, de-
pending on environmental conditions, others may also be locally abundant. Thus
the formation of akaganeite is known to require the presence of either chloride or
fluoride ions (this mineral is correspondingly rare in nature), the formation of fer-
rihydrite is promoted by the presence of silica and organics, whereas lepidocrocite
will form only if the carbonate activity is adequately low. Thorough reviews of the
physico-chemical conditions for iron oxide formation and their occurrences in natural
assemblages have been presented by Schwertmann and Taylor [75] and Cornell and
Schwertmann [76].
The basic structural elements of all iron oxides are similar: Fe3+ ions are located
in the interstices of hexagonal or cubic close packed arrangements of oxygens and/or
hydroxyls. Akaganeite and the spinel-type oxides magnetite and maghemite deviate in
part from this configuration: akaganeite has a body-centered anion arrangement and
the spinel oxides ideally have 1/3 of their iron in tetrahedral coordination.
When free from foreign-element substitution and of bulk crystallinity, the different
iron oxides order magnetically over a wide range of temperatures extending from 955 K
(hematite) to 77 K (lepidocrocite), cf. table 1. The saturation magnetic hyperfine fields
and quadrupole shifts of the individual iron oxides also differ noticeably, so that a
distinction of the individual species is generally a straightforward affair. Because
of its element-selectivity, Mossbauer spectroscopy has proven very effective for the
identification of iron oxides even at concentrations well below the percent range in
mineralogically complex systems (cf. the above mentioned iron oxides associated with
phyllosilicates). It is therefore not surprising that this technique has been extensively
used for the study of these minerals in materials formed in terrestrial weathering
54 E. Murad / Clays and clay minerals study by Mossbauer spectroscopy

environments such as clays and soils, and its application in the course of interplanetary
missions has also been suggested [7779].
Iron oxides formed under ambient environmental conditions have characteristic
features, in particular small particle size and commonly substitution of other elements
for iron, that set them apart from their counterparts formed under other conditions, e.g.,
in igneous and metamorphic rocks. These features can have a profound influence on
magnetic properties [19,80,81], and must be taken into regard during the interpretation
of Mossbauer spectra of such minerals.
The substitution of other elements, in particular aluminium, for iron in its oxides
has been recurrently documented [76]. The frequency of Al-for-Fe substitution results
from the abundance of aluminium in the earths crust, where its atomic abundance
is more than threefold that of iron, and, as in the converse substitution of iron for
aluminium in the phyllosilicates, the similar ionic radii of Al3+ and Fe3+ .
The effects of foreign-element substitution in the iron(III) oxides are dominated
by the fact that Fe3+ in these is in the high-spin state, having five unpaired 3d electrons
and the highest possible magnetic moment that a fourth-period transition element can
have. Al3+ , in contrast, is a diamagnetic ion with the electronic configuration of neon
(a closed 2p shell), so that the substitution of Al3+ for Fe3+ will lead to a maximum
reduction of magnetic interactions. The only other transition-element ions with a 3d5
configuration are Cr+ and Mn2+ , both of which are not expected to substitute for
Fe3+ in its oxides. The majority of potential transition-element substituents will thus
also reduce magnetic interactions, though (with the exception of Ti4+ , which is also
diamagnetic) to a lesser extent than Al3+ . Further complications can arise from the
fact that the substitution of other elements for iron may also have an influence on the
particle sizes of iron oxides.
The particle sizes of iron oxides formed under surficial conditions can extend
down to the nanometer range, giving rise to an assortment of relaxational and surface-
related effects, and in extreme cases reductions of the recoil-free fraction down to
zero. Many (though not all) of these effects can be counteracted by taking Mossbauer
spectra at low temperatures, commonly at 77 or 4.2 K (measurements at the latter
temperature, however, bear the risk that other iron-rich constituents may also order
magnetically, leading to interferences with the desired data).
The classic particle-size related relaxational phenomenon is superparamag-
netism [82], which results from spontaneous reversals of the magnetization direction
in small particles. Superparamagnetic relaxation can be described by the equation
= 0 e(KV /kT ) ,
where is the relaxation time, 0 a constant in the order of 1010 s, K is the anisotropy
constant (a measure for the resistance of the particle to spin direction reversals), V is
the particle volume, k is the Boltzmann constant and T is the temperature. If the spin
reversals occur more rapidly than the nucleus can follow (L 108 s for 57 Fe), the
result will be a net reduction of the hyperfine field, down to zero for sufficiently small
particles. At room temperature this is the case for hematites <8 nm and goethites
E. Murad / Clays and clay minerals study by Mossbauer spectroscopy 55

<20 nm in size [83], indicating hematite to have a significantly larger anisotropy


constant than goethite [84]. Because of particle size distributions, most superparamag-
netic materials show a breakdown of magnetic order over a range of temperatures;
the temperature at which the magnetically ordered and superparamagnetic components
observed in Mossbauer spectra are equal is called the magnetic blocking temperature.
It is immediately obvious from the above formula that superparamagnetic fluctua-
tions can be slowed down by cooling samples. However, because temperature reduction
will also have an effect on other dynamic phenomena, its influence is not limited to su-
perparamagnetism. The application of an external magnetic field, in contrast, has been
shown to restore magnetic order in superparamagnetic particles, and can be considered
diagnostic for this phenomenon [14,85].
In small particles, the magnetic vector may perform rapid fluctuations around its
characteristic direction significantly above the magnetic blocking temperature. Such
fluctuations, termed collective magnetic excitations [86], can lead to reductions with
a maximum extent about 515% of the time-averaged magnetic hyperfine field. The
extent of hyperfine-field reduction depends on the particle size and anisotropy constant,
and approaches zero at very low temperatures.
The role of interparticle interactions is not well understood, although the im-
portance of such effects has been long recognized. Thus hyperfine-field reductions of
up to 50% and an unexpectedly rapid transition from the magnetically ordered to the
superparamagnetic state were observed in closely intergrown goethite crystallites [87].
This was attributed to magnetic exchange coupling between the crystallites, a phenom-
enon that Mrup et al. [87] termed superferromagnetism. The strength of magnetic
interactions between superparamagnetic particles can have a notable influence on the
magnetic behaviour, increasing strength of interactions, for example, leading to a de-
crease of the superparamagnetic blocking temperature [88]. Substantial differences in
the magnetic properties of nanocrystalline maghemites have been observed not only
as a function of particle size, but also in dependence of the interparticle distances
of isolated particles and the degree of aggregation [89]. Deviations between the data
and model calculations can be attributed to possible influences of surface effects and
particle morphology.
Other dynamic phenomena that may exert an influence on the magnetic proper-
ties of fine particles are possibly different magnetization direction probabilities (called
superferromagnetism by Rancourt and Daniels [90]), and anomalous recoil-free frac-
tions that, in extreme cases, can lead to the complete disappearance of Mossbauer
resonances [91].
Effects that can influence the magnetic properties of microcrystalline materials but
are not temperature-dependent are spin canting, speromagnetism, and reduced exchange
coupling due to vacancies, spurious constituents and high surface-to-volume ratios.
Spin canting, first described by Coey [92], is not unusual in small ferrimagnetic
particles. Mossbauer spectra of such materials taken under longitudinally applied
external magnetic fields show an incomplete suppression of the mI = 0, i.e., the
second and fifth, peaks of sextet spectra. In a study of a natural ferric gel (probably
56 E. Murad / Clays and clay minerals study by Mossbauer spectroscopy

ferrihydrite), Coey and Readman [93] described a magnetic behaviour that indicated
random freezing of the spins. This behaviour, initially termed sperimagnetic, may
be considered an extreme case of spin canting.
In a study of a fine-grained goethite by three methods that operate over vastly
different time scales, a reduced Neel temperature of 358 K (ideally 400 K), observed
during all measurements, was attributed to a high concentration of vacancies [94].
Similar reductions of the Neel temperature in akaganeite [95] and lepidocrocite [96]
have been attributed to the presence of excess water in the structures.
The already mentioned reduction of magnetic hyperfine fields as a result of the
substitution of aluminium for iron has incited various attempts to determine the extent
of aluminium substitution in iron oxides by Mossbauer spectroscopy. Golden et al. [97],
however, observed the magnetic hyperfine field of goethite to vary not only with alu-
minium substitution, but also with the surface area. Fysh and Clark [98] claimed that
such deviations from a linear relation between the magnetic hyperfine field reduction
and the degree of aluminium substitution in goethite result from an influence of su-
perparamagnetism that persists down to 77 K. They recommended taking Mossbauer
spectra at 4.2 K, at which superparamagnetic relaxation should no longer play a role
in most materials. Murad and Schwertmann [99], however, showed that such a simple
correlation constitutes an inadmissible simplification, because particle size also has a
temperature-independent influence on the magnetic properties of microcrystalline ma-
terials that increases as the particles become smaller, thus constituting an effect that
cannot be neglected at any temperature. An apparently less pronounced influence of
particle size on the hyperfine field of natural goethites at 4.2 K could result from a
cumulative influence of uncontrollable variables (e.g., substitution by a multitude of
elements), which could obscure specific correlations in natural samples [100].
Finally, studies concerned with the magnetic properties of mixed crystals, e.g.,
in the Fe2 O3 Al2 O3 system, generally imply a random distribution of cations. The
concentration of Fe3+ clusters in the Al-rich side of this system, however, has been
shown to exceed that predicted on the basis of statistical calculations by a factor of
30 [101]. It is not unreasonable to deduce that the converse can also take place, i.e., that
aluminium can form clusters in Al-substituted iron oxides, thus affecting the magnetic
properties in a completely different manner than if it were dispersed at random.

3.2. Specific minerals

The Mossbauer spectra of most iron oxides when chemically pure and of bulk
crystallinity are relatively straightforward. Among those exhibiting more complex spec-
tra, the ferrimagnetic oxides have two unequal sublattices and thus two in the case
of feroxyhite and maghemite poorly resolved subspectra, and Mossbauer spectra of
akaganeite in the magnetically ordered state require fitting with three sextets. Parame-
ters listed in table 1 refer to pure and, wherever applicable, well-crystallized minerals.
However, because iron oxides associated with clays are often poorly-crystalline, they
may have parameters that deviate noticeably from those given in table 1. Such devi-
E. Murad / Clays and clay minerals study by Mossbauer spectroscopy 57

ations include higher quadrupole splittings and line widths in the paramagnetic state
(reflecting higher site distortions and variabilities of iron site geometry), and lower
magnetic ordering temperatures and saturation hyperfine fields in the magnetically
ordered state.
Data on two iron oxides with contrasting magnetic properties are given below:
hematite, which has the highest Neel temperature (955 K) and the second-highest
saturation hyperfine field (54.2 T at 4.2 K) of all iron oxides, and ferrihydrite, which
always occurs in small particles that are prone to many of the deviations from ideal
magnetic behaviour outlined above.
Hematite, -Fe2 O3 , was the material on which Kistner and Sunyar [102], in a
seminal paper, first demonstrated the existence of an isomer shift and quadrupole shift
by Mossbauer spectroscopy. Shortly afterwards it was demonstrated that the spin flip
at the Morin transition [103] causes the quadrupole shift in hematite to double and
change sign [104].
Early work on the Mossbauer spectra of non-ideal hematites showed that the
substitution of aluminium for iron leads to a marked reduction of the magnetic hyperfine
field [105]. Kundig et al. [8] studied hematites of small particle size (down to <10 nm)
and showed that the reduction of particle size also causes both the magnetic hyperfine
field to decrease and lowers the temperature of the Morin transition. This particle-size
dependence of the hyperfine field, observed to persist down to 10 K [8], precludes a
simple linear correlation between the magnitude of the hyperfine field and the degree
of aluminium substitution except for hematites of bulk crystallinity [106].
The Morin transition has been the subject of numerous Mossbauer studies. Above
the Morin temperature, TM 264 K for pure hematite of bulk crystallinity [107],
the spins are located at right angles to the crystallographic c-axis. A slight canting
of the spins brings on a weakly ferromagnetic behaviour. When the temperature is
lowered below TM the spins change their direction, making an angle of about 7
with the c-axis [108]. The spins are now exactly antiparallel, and the behaviour is
antiferromagnetic. The electric field gradient is parallel to the c-axis throughout. The
quadrupole shift in the magnetically ordered state (m ) is related to the quadrupole
splitting in the paramagnetic state (p ) by the equation
1 
m = p 3 cos2 1 ,
2
where is the angle between the principal axis of the electric field gradient and the
magnetic hyperfine field. A consequence of this dependence is a change of m from
0.20 above to 0.41 mm/s below the Morin transition (table 1).
Small particle size and aluminium substitution cause the Morin transition to be
spread out, so that the spin flip takes place over a range of temperatures. Thus alu-
minous hematites can exhibit coexisting components that have and have not passed
through a Morin transition over a range of temperatures, and that the spins may be
canted away from the c-axis by an angle which varies with the degree of aluminium
substitution [109].
58 E. Murad / Clays and clay minerals study by Mossbauer spectroscopy

A further characteristic feature of aluminium substitution in hematites is that the


recoil-free fraction shows a maximum at about 4% substitution [110]. A similar mini-
mum of X-ray diffraction line widths has been attributed to the relief of strain at minor
degrees of aluminium substitution [111]. In analogy to other properties, the recoil-free
fraction has also a marked particle-size dependence, the mentioned maximum being
developed only in hematites of small particle size, whereas bulk hematites show a
more or less linear increase of the recoil-free fraction with the degree of aluminium
substitution [31].
Ferrihydrite, Fe5 HO8 4H2 O, occurs in particles only 27 nm in size. This
mineral is thus always poorly crystalline, and probably the most complex of all iron
oxides. The magnetic behaviour of ferrihydrite is governed by particle-size effects
such as superparamagnetism, random orientations of electron spins in the magnetically
ordered state, and distributions of hyperfine parameters both in the superparamagnetic
and magnetically ordered states.
One effect of the small particle size of ferrihydrite is the limited number of
X-ray diffraction lines. These vary in the d-value range above 0.14 nm from six (for
the better-crystalline, i.e., 7 nm large particles) broad lines to two (for the most
poorly-crystalline samples) extremely broad maxima. Samples with an intermediate
number of lines also exist in nature [112] and can be synthesized in the laboratory
(Schwertmann, personal communication, 1987), and the number of lines is often used
to characterize ferrihydrites with respect to their crystallinities.
Several Mossbauer studies of ferrihydrite predate the approval of this as an in-
dividual mineral species [113]. Thus van der Giessen [114,115] studied a precipitate
with essentially the same XRD pattern as 6-XRD-line ferrihydrite (incidentally listing
five additional diffraction peaks with d values smaller than 1.4 nm). Mossbauer spec-
tra showed this to be superparamagnetic from room temperature down to 77 K, but to
exhibit magnetic hyperfine splitting at 4.2 K. Subsequent Mossbauer work showed a
similar material to have a magnetic blocking temperature of 45 K [116]. Finally, Coey
and Readman [93] observed a natural ferric gel to have a superparamagnetic blocking
temperature of 13 K (the true Neel temperature was estimated to be about 100 K), and
a magnetic ordering behaviour characterized by a distribution of hyperfine fields with
a maximum at about 46 T.
The first Mossbauer data on material with a confirmed identity as ferrihydrite
was published by Murad and Schwertmann [117]. This work verified observations of
high quadrupole splittings (about 0.75 mm/s) at room temperature, and low magnetic
hyperfine fields and quadrupole shifts (about 49 T and 0.03 mm/s, respectively)
at 4.2 K. At room temperature the quadrupole splitting of maximum probability de-
creases from 0.78 to 0.62 mm/s and the quadrupole-splitting half width from 0.86
to 0.71 mm/s, respectively, from 2- to 6-XRD-line ferrihydrites, and the superpara-
magnetic blocking temperatures increase concurrently from about 28 to 115 K as the
crystallinity improves from that of a 2- to a 6-XRD-line material [118]. At 4.2 K
2- and 6-XRD-line ferrihydrites can be distinguished on the basis of their hyperfine
E. Murad / Clays and clay minerals study by Mossbauer spectroscopy 59

fields of maximum probability of 4750 T and hyperfine-field distribution half widths


of 84.5 T, respectively [119].
Room-temperature Mossbauer spectra cannot be considered diagnostic for ferri-
hydrite. Numerous other minerals, including other iron oxides if these are extremely
poorly crystalline, also produce broad doublets with similar quadrupole splittings.
Changes in Mossbauer spectra following the removal of ferrihydrite by dissolution
with acid ammonium oxalate [120] are also not specific, since this treatment may also
affect other very poorly crystalline iron oxides. Mossbauer spectra of magnetically
ordered ferrihydrite, in contrast, are distinctively different from those of the other iron
oxides and can serve to identify this mineral.
Mossbauer data on ferrihydrite can be integrated into a coherent model on the
basis of the information provided at the beginning of this section. As ferrihydrite
crystallinities decrease, the site distortions and variabilities increase, and magnetic in-
teractions become weaker on account of an increasing influence of surface effects.
Even in better crystalline samples, X-ray diffraction does not indicate the existence of
distinctly different core and surface sites. For this reason the use of two-doublet fits to
Mossbauer spectra of superparamagnetic ferrihydrites and two-sextet fits for their mag-
netically ordered counterparts must be discouraged. The only fits that do justice to the
fact that the iron site geometries are variable throughout are distributions of parame-
ters. As a simplified alternative, single-doublet or sextet fits give Lorentzian-averaged
parameters that, although subject to systematic deviations from the true maxima of the
distributions, can serve for a firsthand comparison between samples.
Confirmation of the above model is provided by studies on supported fer-
rihydrites, i.e., ferrihydrites that are closely intergrown with other minerals. Such
intergrowths would reduce iron site distortions, i.e., lead to an apparent improvement
of ferrihydrite crystallinity. However, if the supporting material contains cations with
a smaller magnetic moment than that of high-spin Fe3+ , magnetic interactions will
remain low, and the composite will have magnetic properties indicative of a ferrihy-
drite of poor crystallinity. Such a behaviour has been observed for ferrihydrite-silica
intergrowths developed on highly-weathered volcanic rocks [121] and for deep-sea
ferromanganese crusts consisting of intergrowths of ferrihydrite with manganese ox-
ides [122].
Several materials with Mossbauer spectra that resemble those of ferrihydrite have
been described in the literature. One of these is nanophase hematite, which can be
prepared in the laboratory by thermal decomposition of trinuclear acetato hydroxy-
iron(III) nitrate [123]. This differs from ferrihydrite mainly by a low water content
of about 3.5 wt.%, and even this may be physically held. Another ultrafine iron oxy-
hydroxide, termed nanocat, consists of 3 nm large particles with a bulk density of
only 0.08 g/cm3 , and has Mossbauer parameters resembling those of ferrihydrite at
10 K. However, unless pressed, this material shows a dramatic decrease of the recoil-
free fraction as temperatures increase, ending up at zero resonant absorption above
50 K [91,124]. Finally, ferritin is an iron storage protein with a core about 8 nm in
diameter that can contain a variable number of iron atoms (up to 4500) in the form of a
60 E. Murad / Clays and clay minerals study by Mossbauer spectroscopy

ferrihydrite-like phase. The Mossbauer spectra of ferritin closely resemble those of fer-
rihydrite, the magnetic properties such as the superparamagnetic blocking temperature
and hyperfine field being controlled by the degree of iron loading [125,126].

4. Composite materials

4.1. Clayiron oxide associations

Iron oxides can have a substantial effect on the properties of clays with which
they are associated. Their presence can limit the value of commercial clays, especially
when they are occluded in the phyllosilicates and cannot be removed by standard
extraction procedures. On the other hand, the presence of iron oxides usually betokens
the absence of sulfides, which have a considerably more deleterious influence on the
firing behaviour of clays. Because the conditions under which iron oxides form are
relatively well known [76], they can also serve as genetic indicators.
The strong pigmenting power of iron oxides makes them conspicuous in associa-
tions with other minerals, but poor crystallinity can hamper their identification by X-ray
diffraction, especially when they are present in low concentrations. Chemical extrac-
tion techniques, specifically dithionite reduction [127], which dissolves all iron oxides,
and complexation with acid ammonium oxalate in the dark [120], which removes only
magnetite and the most poorly crystalline minerals ferrihydrite and lepidocrocite, can
provide an indication of the quantity of these mineral groups, but do not enable specific
iron oxides to be identified. Two instrumental techniques which could serve for this
purpose are diffuse optical reflectance and Mossbauer spectroscopy.
Diffuse reflectance optical spectroscopy has been suggested as a technique suit-
able for the determination of iron speciation in kaolins [128]. The results presented in
this study, however, are not consistent, for example, indicating goethite to be associated
with the Georgia kaolin KGa-2, but akaganeite or ferrihydrite with the dithionite-treated
sample. The statement that iron oxides in minor amounts in kaolins cannot be detected
with other methods is untenable. An additional conclusion of this study, namely that
no evidence for iron substitution in the kaolinite structures could be established (a
statement that is in conflict with evidence to the contrary from other techniques), at-
tests to an intrinsic limitation of the usefulness of diffuse reflectance for a reliable
determination of iron speciation in kaolins.
As mentioned above, Mossbauer spectra cannot only reveal the presence of minute
amounts of iron oxides associated with phyllosilicates, but moreover enable an un-
equivocal identification of mineral species and their quantification at levels below 1%.
Examples of such associations containing only minor quantities of iron oxides have
been given by Murad and Wagner [28,36]. It is self-evident that the characterization of
iron oxides will be considerably less demanding if these are present in higher propor-
tions, e.g., [13,129], although spectra usually have to be taken at low temperatures in
extreme cases at 4.2 K or below to counteract relaxational effects ensuing from small
particle size. In an example for such a behaviour, the magnetic contribution to the
E. Murad / Clays and clay minerals study by Mossbauer spectroscopy
Figure 4. Mossbauer spectra of an illitic clay taken (a) at room temperature and (b) at 4.2 K. The latter spectrum is dominated by a sextet resulting
mainly from goethite. Spectra of the dithionite-treated sample (c) at room temperature and (d) at 4.2 K show a higher Fe2+ /Fe3+ ratio, the absence of
iron oxides, and the effects of paramagnetic relaxation at 4.2 K (modified from [130]).

61
62 E. Murad / Clays and clay minerals study by Mossbauer spectroscopy

Mossbauer spectrum of an illitic clay increased from nil at room temperature to 78%
at 4.2 K (figures 4a, b) and even further to 90% at 1.8 K [130]. The Mossbauer para-
meters indicate this to result mainly from goethite. Following iron oxide removal with
dithionite, the Fe3+ /Fe2+ ratio had decreased from 91% to 75% (figure 4c). A spec-
trum taken at 4.2 K showed complex magnetic components that can be attributed to
paramagnetic relaxation (figure 4d); these had obviously been masked by the intense
magnetic resonance due to the iron oxides in the original sample.

4.2. Soils

The iron contents of soils can be grouped into iron in detrital minerals (i.e.,
such inherited from the parent rocks), in pedogenic minerals (such formed in soil
environments), and in organic matter. Soils thus may constitute significantly more
complex materials than the clay iron oxide associations described above.
Many detrital minerals contain no iron, precluding their study by 57 Fe Mossbauer
spectroscopy; they furthermore usually occur in larger particles than their pedogenic
counterparts, and can therefore be readily studied by other techniques. The study of
soil organic matter, on the other hand, is hampered by the fact that the relatively drastic
pretreatments which are often necessary to effect a separation of organic matter from
inorganic soil constituents may lead to the development of artefacts [131]. The ma-
jority of Mossbauer studies on soils has consequently been concerned with pedogenic
minerals, in particular the iron oxides.
The iron oxide mineralogy of soils reflects the conditions of soil genesis, it exerts
a dominant influence on soil colour, and it may have a profound effect on soil fertility.
It is therefore not surprising that an abundance of soil studies has been concerned
with iron oxides [75]. Many of these studies have made use of the inherent suitability
of Mossbauer spectroscopy for the characterization of magnetic phases in soils, and
several reviews concerned with this subject can be found in the literature [15,19,20,
131,133,134].
The identification of iron oxides in soils, as in their associations with clays,
is difficult when they are present in minor amounts, and Mossbauer spectroscopy
provided it is properly performed is the only method that can provide definitive data
on the iron oxides irrespective of crystallinity and at concentrations significantly below
the percent range. Thus Mossbauer spectra taken at 4.2 K showed that a mere 0.2 wt.%
of hematite at a hematite/goethite ratio of 0.03 are sufficient to impart a discernible red
coloration to soils formed under a temperate climate (mean annual temperature 7 C)
if these are adequately well drained [135].
Because pedogenic iron oxides in analogy to those associated with clays often
occur in small particles that are superparamagnetic at room temperature, Mossbauer
spectra of soils generally have to be taken at low temperatures. Table 1 shows that the
induction of magnetic order in all iron oxides (including the most poorly crystalline
ferrihydrite) requires spectra to be taken at temperatures below 25 K.
E. Murad / Clays and clay minerals study by Mossbauer spectroscopy 63

The existence of partly overlapping sextets can lead to problems when studying
soils that contain several iron oxides with similar Mossbauer parameters, e.g., goethite
and ferrihydrite, which both have magnetic hyperfine fields close to 50 T at 4.2 K. In
such cases it may be necessary to take Mossbauer spectra over a range of temperatures
(provided the magnetic ordering behaviour of the oxides differs) or to separate different
fractions using selective dissolution procedures.
A field in which Mossbauer spectroscopy has met with notable success and that
holds promise for future work is the study of assemblages that are unstable under
ambient conditions. Such work pertains mainly to materials formed under reducing
conditions and benefits from the fact that Mossbauer spectra taken under the exclusion
of oxygen or at temperatures low enough to significantly retard oxidation can enable
transient phases to be identified and quantified in their original states. Examples of
materials thus studied are anoxic river sediments from the Elbe in northern Germany,
which suffered significant oxidation of Fe2+ even when freeze-dried [136], and the
green rusts, metastable Fe2+ (Fe3+ , Al)-hydroxy salts. Early Mossbauer work on
synthetic green rusts by Murad and Taylor [137,138] showed that Fe2+ in these has a
very constant, oxidation-insensitive quadrupole splitting of 2.8 mm/s, and Mossbauer
spectroscopy has been instrumental in the recent identification of green rust in a natural
environment [139].
The anticipation of emplacing a Mossbauer spectrometer on the surface of Mars in
the near future has triggered a sizeable number of studies on possibly relevant weath-
ering processes and Mars analog materials. Examples of such work include the
weathering of basalt (e.g., [78,140,141]) and sulfides [142]. Papers involving iron ox-
ides have been referred to in section 3.1; some other materials considered relevant that
have been studied by Mossbauer spectroscopy specifically for the mentioned purpose
are magnetic minerals [143], heated clays [144] and clayiron oxide complexes [145].

4.3. Acid mine drainage precipitates

Acid mine drainage and the formation of related precipitates is a common and
particularly bothersome aftermath of mining sulfide-containing materials, often forming
extensive ochreous deposits in affected areas. Acid sulfate soils constitute a comparable
natural environment. The detrimental effects of these materials have been described
in numerous publications, including two detailed monographs [146,147].
Pyrite and marcasite, in which iron is divalent and in the low-spin state, oxidize
in several stages, the oxidation states of sulfur and iron allowing an assessment of the
degree of weathering. Early phases are characterized by the oxidation of sulfur and
the formation of Fe2+ sulfates. Thus melanterite (FeSO4 7H2 O), rozenite (FeSO4
4H2 O) and szomolnokite (FeSO4 H2 O) have been observed to occur in lignite mine
spoils [148].
As oxidation proceeds further, generally under the participation of acidophilic
bacteria such as thiobacillus ferrooxidans, Fe2+ oxidizes and Fe3+ minerals precipitate.
Depending on pH and the concentrations of sulfate and iron in solution, a variety of
64 E. Murad / Clays and clay minerals study by Mossbauer spectroscopy

minerals may form that have often been referred to by such inappropriate terms as the
already mentioned amorphous ferric hydroxide. Some of the most common products
of acid mine drainage, e.g., ferrihydrite and schwertmannite, are consistently very
poorly crystalline and difficult to characterize by standard mineralogical techniques
such as X-ray diffraction, but even the better crystalline minerals, e.g., goethite and
jarosite, occur in the clay fraction. With the exception of goethite, the forenamed
Fe3+ minerals are characterized by low magnetic ordering temperatures and saturation
hyperfine fields (table 1). They thus have characteristic Mossbauer parameters that
allow an unequivocal identification, and in favourable cases also their quantification
in mineralogically complex assemblages, by Mossbauer spectroscopy [149,150].

5. Conclusions and outlooks

The characterization of samples of complex mineralogy by physical methods


requires knowledge of the properties of the individual constituents. With regard to
the iron oxides, numerous studies on synthetic samples have provided a wealth of
information on their Mossbauer spectra both in the paramagnetic and in the magneti-
cally ordered states. In contrast, only few publications have addressed the Mossbauer
spectra of synthetic phyllosilicates, e.g., [151154] on iron-doped kaolinites and on
smectites [66]. This is a field which holds promise for future studies.
Recent developments that could prove useful for the characterization of clays
and clay minerals by Mossbauer spectroscopy include instrumentation that allows the
collection of spatially resolved spectra such as the Mossbauer milliprobe and imag-
ing Mossbauer spectroscopy, surface-sensitive methods such as integral low-energy
electron Mossbauer spectroscopy (ILEEMS), and applied field work.
The Mossbauer milliprobe [155,156] makes use of a narrowly collimated -ray,
and allows Mossbauer spectra of particles with diameters as small as 100 m to be
recorded, thus enabling the identification of inclusions in clays or the characterization
of spatial variations in heterogeneous samples. In imaging Mossbauer spectroscopy a
position-sensitive detector allows the simultaneous collection of 256 Mossbauer spectra
along a sample up to 50 mm in length, and thus a direct depiction of variations of
sample composition [157].
ILEEMS makes use of low-energy (1020 eV) electrons emitted in conjunction
with resonant -ray absorption and subsequent nuclear decay, and allows the character-
ization of a much thinner surface layer (<5 nm) than conventional conversion electron
spectroscopy with energies in the keV range. Thus comparing data from ILEEMS
and conventional transmission Mossbauer spectra, De Grave et al. [158] were able
to show that the surface layers of an unsubstituted and an aluminous hematite differ
significantly in their magnetic properties (e.g., Morin temperature and hyperfine field)
from the bulk (figure 5).
Externally applied magnetic fields have an assortment of uses for the character-
ization of materials with small particle sizes, in particular the restoration of magnetic
order in superparamagnetic materials and the identification of noncollinear spins in fer-
E. Murad / Clays and clay minerals study by Mossbauer spectroscopy 65

Figure 5. Integral low-energy electron Mossbauer spectrum (ILEEMS) and transmission Mossbauer
spectrum of an aluminous hematite (-Fe1.87 Al0.13 O3 ) taken at 130 K. Note the larger contribution from
weakly ferromagnetic hematite, i.e., that has not passed through a Morin transition, in the former spectrum
(from [158]).

rimagnetic materials (see section 3.1). Combined with sophisticated spectral analysis
programs that allow a noncorrelated distinction of distributions of parameters, applied-
field Mossbauer spectra can provide valuable information on poorly-crystalline iron
oxides associated with clays and soils [159161]. Applied-field work on clay-sized
66 E. Murad / Clays and clay minerals study by Mossbauer spectroscopy

phyllosilicates, in contrast to the iron oxides, is scarce to date, but could prove useful
for studies on iron-rich species [162].
Finally, a cautionary note. Measurement and fitting of Mossbauer spectra must
always be carried out bearing the objectives of the study in mind. Physically meaning-
less fits are to be avoided under all circumstances. While there is a general need for
good counting statistics (although there have been some disturbing instances of poor
counting statistics combined with fits that constitute a blatant over-interpretation of the
spectral data in the not too far past), even the best statistics do not necessarily endorse
the use of an unsubstantiated number of spectral components. The validation of results
using other techniques, while generally useful, is an indispensable requirement when
carrying out crystallographically-oriented studies.

References

[1] S. Guggenheim and R.T. Martin, Clay Miner. 30 (1995) 257; Clays Clay Miner. 43 (1995) 255.
[2] D.M. Moore, Clays Clay Miner. 44 (1996) 710.
[3] S. Guggenheim and R.T. Martin, Clays Clay Miner. 44 (1996) 713.
[4] C.E. Weaver, J.M. Wampler and T.E. Pecuil, Science 156 (1967) 504.
[5] P.J. Malden and R.E. Meads, Nature 215 (1967) 844.
[6] J.M.D. Coey, Atom. Energy Rev. 18 (1980) 73.
[7] L. Heller-Kallai and I. Rozenson, Phys. Chem. Miner. 7 (1981) 223.
[8] W. Kundig, H. Bommel, G. Constabaris and R.H. Lindquist, Phys. Rev. 142 (1966) 327.
[9] T. Shinjo, J. Phys. Soc. Japan 21 (1966) 917.
[10] C. Janot, M. Chabanel and E. Herzog, Bull. Soc. Franc. Miner. Crist. 91 (1968) 166.
[11] L.H. Bowen, Mossbauer Effect Ref. Data J. 2 (1979) 76.
[12] E. Murad and J.H. Johnston (1987) in: Mossbauer Spectroscopy Applied to Inorganic Chemistry,
Vol. 2, ed. G.J. Long (Plenum, New York, 1987) p. 507.
[13] E. Murad and U. Wagner, Hyp. Interact. 45 (1989) 161.
[14] E. De Grave, R.E. Vandenberghe and L.H. Bowen, in: Condensed Matter Studies by Nuclear
Methods, eds. J. Stanek and A.T. Pedziwiatr (World Scientific, Singapore, 1990) p. 186.
[15] E. Murad, Adv. Soil Sci. 12 (1990) 125.
[16] R.E. Vandenberghe, E. De Grave, C. Landuyt and L.H. Bowen, Hyp. Interact. 53 (1990) 175.
[17] L.H. Bowen, E. De Grave and R.E. Vandenberghe, in: Mossbauer Spectroscopy Applied to Mag-
netism and Materials Science, eds. G.J. Long and F. Grandjean (Plenum, New York, 1993) p. 115.
[18] B.A. Goodman, in: Clay Mineralogy: Spectroscopic and Chemical Determinative Methods, ed.
M.J. Wilson (Chapman and Hall, London, 1994) p. 68.
[19] E. Murad, in: Transactions of the 15th World Congress of Soil Science 8a (Instituto Nacional de
Estadstica, Geografa e Informatica, Mexico, 1994) p. 85.
[20] E. Murad, Hyp. Interact. 111 (1998) 251.
[21] G.W. Brindley and G. Brown, Crystal Structures of Clay Minerals and Their X-Ray Identification
(Mineralogical Society, London, 1980).
[22] D.M. Moore and R.C. Reynolds, X-ray Diffraction and the Identification and Analysis of Clay
Minerals (Oxford Univ. Press, New York, 1997).
[23] S.P. Altaner, C.A. Weiss Jr. and R.J. Kirkpatrick, Nature 331 (1988) 699.
[24] R.D. Shannon, Acta Cryst. A 32 (1976) 751.
[25] S.W. Bailey, in: Micas, Reviews in Mineralogy, Vol. 13, ed. S.W. Bailey (Mineralogical Society
of America, Washington, DC, 1984) p. 13.
E. Murad / Clays and clay minerals study by Mossbauer spectroscopy 67

[26] V.A. Drits, in: 5th Meeting of the European Clay Groups, ed. J. Konta (Univerzita Karlova, Prague,
1985) p. 33.
[27] W.A. Dollase, Am. Miner. 60 (1975) 257.
[28] E. Murad and U. Wagner, Clay Miner. 29 (1994) 1.
[29] D.G. Rancourt, Phys. Chem. Miner. 21 (1994) 250.
[30] D.G. Rancourt, M.-Z. Dang and A.E. Lalonde, Am. Miner. 77 (1992) 34.
[31] E. Murad, Phys. Lett. A 111 (1985) 79.
[32] D.G. Rancourt, Phys. Chem. Miner. 21 (1994) 244.
[33] L.G. Dainyak, V.A. Drits and L.M. Heifits, Clays Clay Miner. 40 (1992) 470.
[34] V.A. Drits, L.G. Dainyak, F. Muller, G. Besson and A. Manceau, Clay Miner. 32 (1997) 153.
[35] S.A. Fysh, J.D. Cashion and P.E. Clark, Clays Clay Miner. 31 (1983) 285.
[36] E. Murad and U. Wagner, Neues Jahrb. Miner. Abh. 162 (1991) 281.
[37] C.M. Cardile, J.H. Johnston and D.P.E. Dickson, Clays Clay Miner. 34 (1986) 233.
[38] M.G. Townsend, G. Longworth, C.A.M. Ross and R. Provencher, Phys. Chem. Miner. 15 (1987)
64.
[39] E. Murad, Z. Pflanzenern. Bodenk. 150 (1987) 279.
[40] N.H.J. Gangas, J. van Wonterghem, S. Mrup and C.J.W. Koch, J. Phys. C: Solid State Phys. 18
(1985) L1011.
[41] A. Simopoulos, D. Petridis, A. Kostikas and N.H. Gangas, Hyp. Interact. 41 (1988) 843.
[42] C. Breen, K.C. Molloy and K. Quill, Clay Miner. 27 (1992) 445.
[43] D. Petridis and T. Bakas, Clays Clay Miner. 45 (1997) 73.
[44] H. Pollak and J.G. Stevens, Hyp. Interact. 29 (1986) 1153.
[45] D.S. Fanning, M.C. Rabenhorst, L. May and D.P. Wagner, Clays Clay Miner. 37 (1989) 59.
[46] J.L. LaCombe, P.J. Schurer and J.S. Mothersill, Hyp. Interact. 91 (1994) 747.
[47] V. Ernstsen, Clays Clay Miner. 44 (1996) 599.
[48] J. Breuer and E. Murad, Z. Pflanzenern. Bodenk. 155 (1992) 379.
[49] I. Rozenson and L. Heller-Kallai, Clays Clay Miner. 26 (1978) 88.
[50] H. Kodama, G. Longworth and M.G. Townsend, Can. Miner. 20 (1982) 585.
[51] T.G. St. Pierre, B. Singh, J. Webb and B. Gilkes, Clays Clay Miner. 40 (1992) 341.
[52] D.A. Jefferson, M.J. Tricker and A.P. Winterbottom, Clays Clay Miner. 23 (1975) 355.
[53] S.W. Bailey, G.W. Brindley, D.S. Fanning, H. Kodama and R.T. Martin, Clays Clay Miner. 32
(1984) 239.
[54] K. Norrish and J.G. Pickering, in: Soils: An Australian Viewpoint (Division of Soils CSIRO,
Melbourne, 1983) p. 281.
[55] J. Srodon and D.D. Eberl, in: Micas, Reviews in Mineralogy, Vol. 13, ed. S.W. Bailey (Miner-
alogical Society of America, Washington, DC, 1984) p. 495.
[56] H.M. Koster and U. Schwertmann, in: Tonminerale und Tone, eds. K. Jasmund and G. Lagaly
(Steinkopff, Darmstadt, 1993) p. 33.
[57] H.M. Koster, in: Internat. Clay Conf. 1981, eds. H. van Olphen and F. Veniale (Elsevier, Amster-
dam, 1982) p. 41.
[58] R.A. Eggleton, Clay Miner. 12 (1977) 181.
[59] J.W. Stucki, in: Iron in Soils and Clay Minerals, eds. J.W. Stucki, B.A. Goodman and
U. Schwertmann (Reidel, Dordrecht/Boston, 1988) p. 625.
[60] D. Bonnin, G. Calas, H. Suquet and H. Pezerat, Phys. Chem. Miner. 12 (1985) 55.
[61] D.M. Sherman and N. Vergo, Am. Miner. 73 (1988) 1345.
[62] V. Luca, Clays Clay Miner. 39 (1991) 467.
[63] G. Besson, A.S. Bookin, L.G. Dainyak, M. Rautureau, S.I. Tsipursky, C. Tchoubar and V.A. Drits,
J. Appl. Cryst. 16 (1983) 374.
[64] B.A. Goodman, J.D. Russell, A.R. Fraser and F.W.D. Woodhams, Clays Clay Miner. 24 (1976)
53.
68 E. Murad / Clays and clay minerals study by Mossbauer spectroscopy

[65] J.H. Johnston and C.M. Cardile, Clays Clay Miner. 33 (1985) 21.
[66] O. Grauby, S. Petit, A. Decarreau and A. Baronnet, Eur. J. Miner. 6 (1994) 99.
[67] P.R. Lear, P. Komadel and J.W. Stucki, Clays Clay Miner. 36 (1988) 376.
[68] I. Rozenson and L. Heller-Kallai, Clays Clay Miner. 24 (1976) 271.
[69] J.D. Russell, B.A. Goodman and A.R. Fraser, Clays Clay Miner. 27 (1979) 63.
[70] P. Komadel, J. Madejova and J.W. Stucki, Clays Clay Miner. 43 (1995) 105.
[71] M.V. Eyrish and A.A. Dvorechenskaya, Geochem. Int. 13(3) (1976) 79.
[72] M. Fleischer and J.A. Mandarino, Glossary of Mineral Species (The Mineralogical Record, Tucson,
AZ, 1995).
[73] H. Stanjek and U. Schwertmann, Clays Clay Miner. 40 (1992) 347.
[74] U. Schwertmann, D.G. Schulze and E. Murad, Soil Sci. Soc. Am. J. 46 (1982) 869.
[75] U. Schwertmann and R.M. Taylor, in: Minerals in Soil Environments, 2nd. ed., eds. J.B. Dixon
and S.B. Weed (Soil Science Society of America, Madison, WI, 1989) p. 379.
[76] R.M. Cornell and U. Schwertmann, The Iron Oxides (VCH, Weinheim, 1996).
[77] J.M. Knudsen, M.B. Madsen, M. Olsen, L. Vistisen, C.B. Koch, S. Mrup, E. Kankeleit, G. Klin-
gelhofer, E.N. Evlanov, V.N. Khromov, L.M. Mukhin, O.F. Prilutski, B. Zubkov, G.V. Smirnov
and J. Juchniewicz, Hyp. Interact. 68 (1991) 83.
[78] R.V. Morris, D.C. Golden, J.F. Bell III, H.V. Lauer Jr. and J.B. Adams, Geochim. Cosmochim.
Acta 57 (1993) 4597.
[79] J.L. Bishop and E. Murad, in: Mineral Spectroscopy: A Tribute to Roger G. Burns, eds. M.D. Dyar,
C. McCammon and M.W. Schaefer (Geochemical Society, Houston, TX, 1996) p. 337.
[80] S. Mrup, in: Magnetic Properties of Fine Particles, eds. J.L. Dormann and D. Fiorani (North-
Holland, Amsterdam, 1992) p. 125.
[81] E. Murad, Phys. Chem. Miner. 23 (1996) 248.
[82] L. Neel, Annales Geophys. 5 (1949) 99.
[83] C. Janot, H. Gibert and C. Tobias, Bull. Soc. Franc. Miner. Crist. 96 (1973) 281.
[84] L. Nalovic and C. Janot, Rev. Phys. Appliquee 14 (1979) 475.
[85] M.B. Madsen, S. Mrup and C.J.W. Koch, Hyp. Interact. 27 (1986) 329.
[86] S. Mrup, J. Mag. Mag. Mater. 37 (1983) 39.
[87] S. Mrup, M.B. Madsen, J. Franck, J. Villadsen and C.J.W. Koch, J. Mag. Mag. Mater. 40 (1983)
163.
[88] S. Mrup, Europhys. Lett. 28 (1994) 671.
[89] E. Tronc, P. Prene, J.P. Jolivet, F. dOrazio, F. Lucari, D. Fiorani, M. Godinho, R. Cherkaoui,
M. Nogues and J.L. Dormann, Hyp. Interact. 95 (1995) 129.
[90] D.G. Rancourt and J.M. Daniels, Phys. Rev. B 29 (1984) 2410.
[91] B. Ganguly, F.E. Huggins, Z. Feng and G.P. Huffman, Phys. Rev. B 49 (1994) 3036.
[92] J.M.D. Coey, Phys. Rev. Lett. 27 (1971) 1140.
[93] J.M.D. Coey and P.W. Readman, Earth Planet. Sci. Lett. 21 (1973) 45.
[94] S. Bocquet and S.J. Kennedy, J. Mag. Mag. Mater. 109 (1992) 260.
[95] D. Chambaere and E. De Grave, J. Mag. Mag. Mater. 42 (1984) 263.
[96] E. De Grave, R.M. Persoons, D.G. Chambaere, R.E. Vandenberghe and L.H. Bowen, Phys. Chem.
Miner. 13 (1986) 61.
[97] D.C. Golden, L.H. Bowen, S.B. Weed and J.M. Bigham, Soil Sci. Soc. Am. J. 43 (1979) 802.
[98] S.A. Fysh and P.E. Clark, Phys. Chem. Miner. 8 (1982) 180.
[99] E. Murad and U. Schwertmann, Clay Miner. 18 (1983) 301.
[100] J. Friedl and U. Schwertmann, Clay Miner. 31 (1996) 455.
[101] V.V. Viktorov, Phys. Status Solidi B 174 (1992) 529.
[102] O.C. Kistner and A.W. Sunyar, Phys. Rev. Lett. 4 (1960) 412.
[103] F.J. Morin, Phys. Rev. 78 (1950) 819.
[104] J. Gastebois and J. Quidort, Comptes Rendus Acad. Sci. 253 (1961) 1257.
E. Murad / Clays and clay minerals study by Mossbauer spectroscopy 69

[105] G. Shirane, D.E. Cox and S.L. Ruby, Phys. Rev. 125 (1962) 1158.
[106] E. Murad, J. Phys. E 17 (1984) 736.
[107] N. Amin and S. Arajs, Phys. Rev. B 35 (1987) 4810.
[108] A.H. Morrish, G.B. Johnston and N.A. Curry, Phys. Lett. 7 (1963) 177.
[109] E. De Grave, L.H. Bowen and G.G. Robbrecht, in: Solid State Chemistry (Studies in Inorganic
Chemistry), Vol. 3, eds. R. Metselaar, H.J.M. Heijligers and J. Schoonman (Elsevier, Amsterdam,
1982) p. 71.
[110] E. De Grave, A.E. Verbeeck and D.G. Chambaere, Phys. Lett. A 107 (1985) 181.
[111] U. Schwertmann, R.W. Fitzpatrick, R.M. Taylor and D.G. Lewis, Clays Clay Miner. 27 (1979)
105.
[112] L. Carlson and U. Schwertmann, Geochim. Cosmochim. Acta 45 (1981) 421.
[113] F.V. Chukhrov, B.B. Zvyagin, L.P. Ermilova and A.I. Gorshkov, in: Proc. of the 1972 Clay Conf.,
ed. J.M. Serratosa (C.S.I.C., Madrid, 1973) p. 333.
[114] A.A. van der Giessen, J. Inorg. Nucl. Chem. 28 (1966) 2155.
[115] A.A. van der Giessen, J. Phys. Chem. Solids 28 (1967) 343.
[116] Z. Mathalone, M. Ron and A. Biran, Solid State Comm. 8 (1970) 333.
[117] E. Murad and U. Schwertmann, Am. Miner. 65 (1980) 1044.
[118] E. Murad, L.H. Bowen, G.J. Long and T.G. Quin, Clay Miner. 23 (1988) 161.
[119] E. Murad, Neues Jahrb. Miner. Monatsh. (1982) 45.
[120] U. Schwertmann, Z. Pflanzenern. Dungung Bodenk. 105 (1964) 194.
[121] E. Murad and U. Schwertmann, Hyp. Interact. 41 (1988) 835.
[122] E. Murad and U. Schwertmann, Am. Miner. 73 (1988) 1395.
[123] R.V. Morris, H.V. Lauer Jr., D.G. Schulze and R.G. Burns, Lunar Planet. Sci. 22 (1991) 927.
[124] J. Zhao, F.E. Huggins, Z. Feng and G.P. Huffman, Clays Clay Miner. 42 (1994) 737.
[125] E.R. Bauminger and I. Nowik, Hyp. Interact. 50 (1989) 484.
[126] J. Webb and T.G. St. Pierre, in: Mossbauer Spectroscopy Applied to Inorganic Chemistry, Vol. 3,
eds. G.J. Long and F. Grandjean (Plenum, New York, 1989) p. 417.
[127] O.P. Mehra and M.L. Jackson, Clays Clay Miner. 7 (1960) 317.
[128] N. Malengreau, J.-P. Muller and G. Calas, Clays Clay Miner. 42 (1994) 137.
[129] B.A. Goodman and P.H. Nadeau, Clay Miner. 23 (1988) 301.
[130] U. Wagner, W. Knorr, A. Forster, E. Murad, R. Salazar and F.E. Wagner, Hyp. Interact. 41 (1988)
855.
[131] D.P.E. Dickson, L. Heller-Kallai and I. Rozenson, Geochim. Cosmochim. Acta 43 (1979) 1449.
[132] B.A. Goodman, in: Advanced Chemical Methods for Soil and Clay Minerals Research, eds.
J.W. Stucki and W.L. Banwart (Reidel, Dordrecht/Boston, 1980) p. 1.
[133] L.H. Bowen and S.B. Weed, in: Mossbauer Spectroscopy and Its Chemical Applications, eds.
J.G. Stevens and G.K. Shenoy (American Chemical Society, Washington, DC, 1981) p. 247.
[134] L.H. Bowen and S.B. Weed, in: Chemical Mossbauer Spectroscopy, ed. R.H. Herber (Plenum,
New York, 1984) p. 217.
[135] U. Schwertmann, E. Murad and D.G. Schulze, Geoderma 27 (1982) 209.
[136] I. Konig, H.-D. Knauth, C. Koopmann, F.E. Wagner and U. Wagner, Hyp. Interact. 41 (1988) 811.
[137] E. Murad and R.M. Taylor, Clay Miner. 19 (1984) 77.
[138] E. Murad and R.M. Taylor, in: Industrial Applications of the Mossbauer Effect, eds. G.J. Long
and J.G. Stevens (Plenum, New York, 1986) p. 585.
[139] F. Trolard, M. Abdelmoula, G. Bourrie, B. Humbert and J.-M.R. Genin, Comptes Rendus Acad.
Sci. 323 (1996) 1015.
[140] S. Mrup, O. Saksager, M.B. Madsen, M.D. Bentzon and C.J.W. Koch, Hyp. Interact. 57 (1990)
2269.
[141] D.W. Straub, R.G. Burns and S.F. Pratt, J. Geophys. Res. E 96 (1991) 18819.
[142] R.G. Burns and D.S. Fisher, J. Geophys. Res. B 95 (1990) 14415.
70 E. Murad / Clays and clay minerals study by Mossbauer spectroscopy

[143] M.B. Madsen, D.P. Agerkvist, H.P. Gunnlaugsson, S. Faurschou Hviid, J.M. Knudsen and L. Vis-
tisen, Hyp. Interact. 95 (1995) 291.
[144] S. Faurschou Hviid, D.P. Agerkvist, M. Olsen, C. Bender Koch and M.B. Madsen, Hyp. Interact.
91 (1994) 529.
[145] J.L. Bishop, C.M. Pieters, R.G. Burns, J.O. Edwards, R.L. Mancinelli and H. Froschl, Icarus 117
(1995) 101.
[146] J.A. Kittrick, D.S. Fanning and L.R. Hossner, eds., Acid Sulfate Weathering (Soil Science Society
of America, Madison, WI, 1982).
[147] C.A. Alpers and D.W. Blowes, eds., Environmental Geochemistry of Sulfide Oxidation (American
Chemical Society, Washington, DC, 1994).
[148] J.B. Dixon, L.R. Hossner, A.L. Senkayi and K. Egashira, in: Acid Sulfate Weathering, eds. J.A. Kit-
trick, D.S. Fanning and L.R. Hossner (Soil Science Society of America, Madison, WI, 1982) p. 169.
[149] E. Murad, U. Schwertmann, J.M. Bigham and L. Carlson, in: Environmental Geochemistry of
Sulfide Oxidation, eds. C.N. Alpers and D.W. Blowes (American Chemical Society, Washington,
DC, 1994) p. 190.
[150] J.M. Bigham and E. Murad, Adv. GeoEcol. 30 (1997) 193.
[151] B.R. Angel, A.H. Cuttler, K.S. Richards and W.E.J. Vincent, Clays Clay Miner. 25 (1977) 381.
[152] A.H. Cuttler, Clay Miner. 15 (1980) 429.
[153] S. Petit, A. Decarreau, J.-P. Eymery and J.-H. Thomassin, Comptes Rendus Acad. Sci. 307 (1988)
1961.
[154] S. Petit and A. Decarreau, Clay Miner. 25 (1990) 181.
[155] C.A. McCammon, V. Chaskar and G.G. Richards, Measur. Sci. Technol. 2 (1991) 657.
[156] C.A. McCammon, I.L. Chinn, J.J. Gurney and M.E. MCallum, in: Extended Abstracts of the 6th
Kimberlite Conf. (Novosibirsk, Russia, 1995) p. 359.
[157] P.R. Smith, J.D. Cashion and L.J. Brown, Hyp. Interact. 71 (1992) 1503.
[158] E. De Grave, C. Dauwe, L.H. Bowen and R.E. Vandenberghe, Hyp. Interact. C 1 (1996) 286.
[159] P.M.A. de Bakker, E. De Grave, R.M. Persoons, L.H. Bowen and R.E. Vandenberghe, Measur.
Sci. Technol. 1 (1990) 954.
[160] R.J. Pollard, C.M. Cardile, D.G. Lewis and L.J. Brown, Clay Miner. 27 (1992) 57.
[161] E. De Grave, P.M.A. de Bakker, L.H. Bowen and R.E. Vandenberghe, Z. Pflanzenern. Bodenk.
155 (1992) 467.
[162] E. Murad, J.D. Cashion and L.J. Brown, Clay Miner. 25 (1990) 261.
[163] D.N. Hinckley, Clays Clay Miner. 11 (1963) 229.
[164] B. Kubler, Rev. Inst. Franc. Petrole 19 (1964) 1093.

You might also like