Sei sulla pagina 1di 22

Clay Minerals, (2011) 46, 539559

Characterization of modified kaolin from


the Ranong deposit Thailand by XRD,
XRF, SEM, FTIR and EPR techniques
N. WORASITH1,{, B. A. GOODMAN2,{, J. NEAMPAN3, N. JEYACHOKE1
P. THIRAVETYAN1,*

AND

1
School of Bioresources and Technology, King Mongkuts University of Technology, Thonburi, Bangkhuntien,
Bangkok, Thailand, 2 Health and Environment Department, Environmental Resources &Technologies, Austrian
Institute of Technology, A-2444 Seibersdorf, Austria, and 3 Department of Geology, Chulalongkorn University,
Bangkok, Thailand

(Received 13 July 2010; revised 2 May 2011; Editor: Eric Ferrage)

AB ST R ACT : Various physical and analytical techniques (XRD, XRF, SEM, FTIR and EPR) have
been used to investigate the effects of chemical and/or physical modification of Ranong kaolin,
which has been proposed as a potential bleaching clay for vegetable oils. Acid treatment after
grinding resulted in major changes compared with acid treatment of the original mineral sample or
mechanical treatment alone. Previous work has shown that the combined treatments produce
increases in surface area and new porous structures, and the present measurements show reductions
in Al:Si ratios. These are accompanied by a major reduction in O H stretching vibrations as a result
of grinding, although acid treatment produced little subsequent effect on the O H bands in the FTIR
spectra. However, acid treatment resulted in a reduction in the Al OH Al bending vibrations and
the appearance of Si O bands associated with newly synthesized material; these effects were much
greater with samples that had been ground prior to the acid treatment. There were appreciable
qualitative differences in the way in which the EPR spectra of Fe and Mn were affected; the Fe signal
was sensitive to mechanical treatment, but little subsequent change was induced by acid extraction,
whereas the Mn peaks were sensitive to the both the pH and the chemical nature of the acid used.
These results therefore indicate that the Fe and Mn are in different types of site in the kaolin
structure. Little change was observed in the main oxygen-based free radical centre associated with Si
atoms, but that associated with Al was lost as a result of the treatments. Such mineral
characterization is of fundamental importance to understanding the modification of kaolins and
their uses as adsorbents in the food and environmental sciences.

KEYWORDS: kaolin, XRD, XRF, SEM, FTIR, EPR, grinding, acid activation, pH, bleaching clay, natural oil
decolourization, Ranong deposit, Thailand.
The kaolin group minerals, kaolinite, dickite,
nacrite and halloysite, are 1:1 layer silicates, in

which each layer consists of a combination of one


sheet of tetrahedral and one sheet of octahedral

* E-mail: paitip.thi@kmutt.ac.th
{
On leave from Department of Chemistry, Faculty of Science and Technology, Rajamangala University of
Technology Krungthep, 2 Nang Lin Chi Road, Soi Suan Plu, Sathorn, Bangkok, Thailand 10120.
{
Current address: State Key Laboratory for Conservation and Utilization of Subtropical Agro-Bioresources,
Guangxi University, Nanning, 530004 Guangxi, Peoples Republic of China.
DOI: 10.1180/claymin.2011.046.4.539

# 2011 The Mineralogical Society

540

N. Worasith et al.

cations linked by bridging oxygen atoms (Murray,


2007). The surface of the tetrahedral sheet consists
of oxygen atoms, whereas that of the octahedral
sheet contains hydroxyl groups. In kaolinite, dickite
and nacrite the neighbouring layers are held
together by hydrogen bonds and these minerals
differ only in the stacking arrangements of the
layers. Halloysite can contain a variable number of
water molecules in this interlayer region; the d001
peak of fully hydrated halloysite is at lower angle
(8.752y) than the other kaolin minerals (12.372y),
but it collapses to about 12.282y when the
humidity of halloysite is below 30%. These
minerals are all dioctahedral, and thus nominally
have only two thirds of their octahedral cation sites
occupied.
With an ideal composition, the tetrahedral cations
are Si and the octahedral cations Al. Such a
structure is charge neutral, i.e. there is no net
layer charge but, in common with other aluminosilicate minerals, various isomorphous substitutions
can occur. Substitution of Al3+ or Fe3+ for Si in the
tetrahedral sheet generates an electron surplus,
which may be compensated by either the generation
of a layer charge (which yields a cation exchange
capacity), or the protonation of oxygen atoms to
generate hydroxyl groups on the edge surfaces of
the tetrahedral sheet (e.g. Huertas et al., 1998).
Note that the possible substitution of tetravalent
ions, such as V4+ or Mn4+ for Si, would not
generate a surface charge but, like Fe, these ions
are capable of existing in more than one oxidation
state, so any process which alters their redox status
would be expected to influence the surface properties of minerals containing them. Various substitutions in the octahedral sheet, such as V4+, Mn4+,
Fe3+, Fe2+, Mg2+, Mn2+, etc. may or may not lead
to the generation of a surface charge, or a change in
the degree of surface protonation; furthermore,
there is the possibility of maintaining charge
neutrality by substituting three divalent ions for
two trivalent ions, since one third of the octahedral
sites are vacant in an ideal structure.
Accurate knowledge of the detailed composition
of natural kaolin samples is difficult to obtain. Xray diffraction (XRD) is the fundamental method
for identifying the crystalline mineral phases, but it
requires careful work to discriminate between the
various kaolin group minerals and also some
impurity phases, such as illites or smectites.
Furthermore, XRD may not detect the presence of
poorly crystalline minerals such as oxides, even

when present in substantial quantities, since their


diffraction patterns are broad and weak. Such
phases, however, may have a major role in
determining surface properties, such as sorption
processes, because of their large surface areas.
The sensitivity of XRD is such that it often
cannot detect small variations in lattice parameters
that arise from isomorphous substitutions, especially
when they occur at relatively low levels, and their
presence is inferred from a combination of chemical
analyses and measurement of ion exchange capacities. Thus, the presence of even small quantities
of impurity phases can result in misleading
conclusions concerning the compositions of
minerals, and this approach is of little value in
determining the chemical compositions of individual components in mixed mineral specimens.
Consequently, a variety of chemical and spectroscopic methods has been developed for characterizing aluminosilicate minerals (e.g. Hawthorne,
1988; Coyne et al., 1998). These include electron
microscopic techniques, transmission electron
microscopy (TEM) and scanning electron microscopy (SEM), and various spectroscopic methods,
such as nuclear magnetic resonance (NMR),
infrared spectroscopy (IR), etc., which have
general applications to mineral characterization. In
addition, with samples such as kaolins that have
relatively low levels of paramagnetic ions, electron
paramagnetic resonance (EPR) spectroscopy (e.g.
Mabbs & Collison, 1992) can provide valuable
information on the nature and distribution of
paramagnetic transition metal ions and on free
radical defects (e.g. Calas, 1988), which might be
expected to play important roles in determining
sorption properties.
Because of its great abundance in the Earths
crust, kaolin has a long history of use by man, the
most important of these uses being in the ceramics,
paint and paper industries. However, there is
currently considerable interest in using kaolins as
the source material for the production of highervalue products for a wide range of industrial uses,
and acid activation of kaolinites has been reported
to yield products for the removal of anions (Gogoi
& Baruah, 2008), cations (Bhattacharyya & Sen
Gupta, 2007) and dyes (Karaoglu et al., 2010) from
waters. Furthermore, there are also several recent
reports of kaolinites being used as the starting
materials in the production of materials with novel
physical and chemical properties (e.g. Frost et al.,
2001a,b, 2004; Temuujin et al., 2001; Belver et al.,

Characterization of modified kaolinite

2002; Meenakshi et al., 2008; Vagvolgyi et al.,


2008; Steudel et al., 2009; Panda et al., 2010).
In the edible oil industry, it is often desirable to
decrease the concentrations of pigments, such as
chlorophyll and b-carotene, and acid-activated
bentonites are commonly used for this purpose
(e.g. Christidis et al., 1997; Wu et al., 2006). In a
search for lower cost bleaching clays, we have
investigated the performance of acid-activated
kaolin from the Ranong deposit in southern
Thailand, but it was found to be ineffective for
the removal of pigments from rice bran oil
(unpublished results). However, after modification
by a combination of physical and chemical
treatments, the sorption properties of this kaolin
improved to such an extent (~80% decolourization
of rice bran oil compared to ~82% with a
commercial bleaching clay under the same conditions (Worasith et al., 2011)) that it has potential
use for bleaching vegetable oils. It was to develop a
detailed understanding of kaolin alteration that the
present investigations were undertaken. This was
deemed necessary because different samples of
bleaching clay activated by similar treatments
have been reported to have considerable differences
in performance (Woumfo et al., 2007), and it seems
likely that (perhaps subtle) variations in physical
properties and/or chemical composition can have a
significant impact on the bleaching performance of
clay minerals. Detailed physical and chemical
characterization is especially important for investigations of reactions that occur on the surfaces of
natural mineral samples, since these are influenced
by factors that may vary appreciably from one
deposit to another. Also, natural mineral samples
often contain associated poorly crystalline phases
whose effects on adsorption reactions may be
relatively greater than their concentrations in the
specimens.
The current paper describes the characterization
of material produced by modifications of kaolin
from the Ranong deposit in Thailand, and which
was used in the rice bran oil bleaching studies of
Worasith et al. (2011). A mineral sample from this
deposit, which is thought to have been produced by
the hydrothermal alteration of granite (Kuentag &
Wasuwanich, 1978), has recently been the subject
of basic physical and chemical characterization
(Nuntiya & Prasanphan, 2006), using TEM, X-ray
fluorescence spectroscopy (XRF), and XRD. The
present work builds on these initial measurements,
but with an emphasis on characterizing products

541

obtained from this kaolin after physical and


chemical modification, and including additional
information that can be obtained with spectroscopic
techniques, such as FTIR and EPR.

MATERIALS AND METHODS


Sulphuric acid (H2SO4) and oxalic acid (H2C2O4)
used for chemical treatments were AR grade and
purchased from Merck, Germany, and Ajax
Finechem, Australia, respectively. The kaolin
sample from Ranong Province in southern
Thailand was supplied by Had Som Pan Co., Ltd.
This natural kaolin, pale yellow in colour, was
initially washed with distilled water and dried
overnight in an oven at 80C. Although this
temperature was chosen to minimize the possibility
, it is possible that this
of damaging halloysite-10 A
, because of
phase was converted to halloysite-7 A
the ease with which water can be removed from this
mineral. This sample was designated K.

Chemical treatments
Acid treatments were performed by adding dried
natural kaolin that had been gently crushed to pass
through a 0.074 mm sieve to 18% or 30% w/w
sulphuric acid; the ratio of clay:acid was 1 g:50 ml.
Samples were refluxed at 90C under mechanical
stirring (using an IKA hotplate stirrer model C-MAG
HS7) at about 250 rpm for 4 h. Then the supernatants were removed and the residues washed with
two litres of distilled water; the washing was
repeating until the pH of the clay suspension was
53. The sample was then dried at 100C for 24 h,
this temperature being chosen because of the planned
use of the product as a vegetable oil bleaching agent.
These samples were designated KS18 and KS30. An
additional sample was prepared using 6% w/v oxalic
acid, but it was not fully investigated because it
became apparent early in the work that physical
treatment was an essential aspect of the kaolin
modification process to produce a bleaching clay
(which was the ultimate objective of this work).

Physical treatment
Dried natural kaolin, designated K, initially
crushed to 40.074 mm as for the chemical
treatments described above, was ground using a
planetary ball mill (Retsch model S 100). For this
physical treatment, 20 g of clay were added to a

542

N. Worasith et al.

500 ml grinding jar containing twenty 20 mm


grinding balls (the weight ratio of balls to kaolin
was 30:1); both the pot and milling media were
stainless steel. The clay samples were ground for
1 h at 300 rpm, and the product designated as GK.

Combined physical and chemical treatments


Samples of ground kaolin GK were treated with
18% or 30% w/w sulphuric acid as described for the
preparation of the corresponding unground kaolin
samples, and designated GKS18 and GKS30. A
similar sample was prepared using oxalic acid
instead of sulphuric acid with acid concentrations
6% w/v and designated as GKO6. These acid
concentrations were used for the current investigations on the basis of their performance as rice bran
oil bleaching agents (Worasith et al., 2011).

Surface modification of the products from


combined physical and chemical treatments
It has been reported that kaolinite surface
properties are influenced by the synthesis pH
(Fialips et al., 2000; Worasith et al., 2011),
indicating the occurrence of protonation/deprotonation reactions at surface oxygen atoms. Therefore,
the influence of pH on the physical/spectroscopic
properties of samples GKS18 and GKO6 was also
investigated by re-suspending the modified kaolin
samples in solutions of sulphuric or oxalic acids at
pH 2.0, 3.0, 3.5, 4.0 and 5.0. After separation by
centrifugation, these kaolin samples were dried at
100C for 24 h. These samples were designated
GKS18 P2.0, GKS18 P3.0, GKS18 P3.5, GKS18
P4.0, GKS18 P5.0, GKO6 P2.0, GKO6 P3.0, GKO6
P3.5, GKO6 P4.0 and GKO6 P5.0.

X-ray diffraction studies


A Bruker AXS model D8 Advance X-ray
diffractometer employing Ni-filtered Cu-Ka radiation was used to investigate the mineralogy of the
natural kaolin sample and its modified products.
Kaolin powder was placed in a flat holder, with
20 mm diameter and 15620 mm irradiated area,
and diffraction patterns were then collected in the
range of 2y = 5 to 60 at a scanning speed 1 per
2y min 1, and working at 40 kV and 30 mA.
Mineral components in the X-ray diffractograms
were identified by comparison with standards in the
JCPDS Powder Diffraction File.

Additional measurements were made with


oriented samples of the natural kaolin in an
,
attempt to discriminate between halloysite-7 A
kaolinite and illite components. These oriented
samples were subjected to four treatments: air
drying, glycolation, intercalation with formamide,
and heating to 550C for 1 h. The structural order
of natural kaolinite was estimated using the
Hinckley index (HI), and the crystallinities of the
kaolin minerals in the various treated products were
also estimated from the widths and heights of peaks
in the XRD patterns, since peak widths are
inversely related to the crystallinity of the sample.

X-ray fluorescence
A Bruker AXS model S4 Pioneer XRF spectrometer with a scintillator detector was used to
determine the chemical compositions of the kaolin
samples. Wax was added to the clay samples as a
binder before fusing into pellets (weight ratio of
clay:wax about 7:3), and compositions are
expressed as relative concentrations in the form of
oxides.

Scanning electron microscopy


The surface morphology of the kaolin samples
that had been dried at 80C and coated with gold to
enhance conductivity was investigated using a
scanning electron microscope (JEOL model JSM5410 LV) with an accelerating voltage of 15 kV
and a vacuum of 10 5 Pa. The samples were
inserted into the SEM chamber, transferred to the
path of the electron beam, and then scanned
automatically. Various magnifications were used
to compare the textures and shapes of modified
kaolins before and after treatments but only those of
620,000 are presented. In addition, chemical
analyses were performed by using a link to an
energy dispersive X-ray analysis system (EDX)
which incorporated ISIS series 300 software.

FTIR
Attenuated Total Reflectance Fourier Transform
Infrared Spectra (ATR-FTIR) were recorded on a
Bruker Tensor 27 FTIR spectrometer incorporating
a 1.8 mm Ge crystal. A sample of the clay powder
was placed on the Ge crystal and 1024 scans were
acquired for each spectrum at a resolution of
1 cm 1 in the mid-IR range (4,000 600 cm 1).

Characterization of modified kaolinite

EPR spectroscopy
EPR spectra were recorded using various Bruker
and JEOL spectrometers operating at X-band
frequencies and using Gunn diodes as microwave
sources. All samples were studied at room
temperature (~22C) using 100 kHz modulation
frequency with other acquisition parameters determined by the line widths and saturation properties
of the component signals. Spectra were first
recorded as first derivatives of the microwave
absorption over the scan range 0 500 mT using
10 mW microwave power and 1 mT modulation
amplitude to obtain a general overview of the
signals. Separate spectra were then recorded using
parameters that were optimized for the individual
signals. Free radical components were recorded as
both first and second derivatives using microwave
powers in the range 2 10 mW and modulation
amplitudes in the range 0.2 1.0 mT over field
scans ranging between 5 and 40 mT. Receiver gain,
number of scans, conversion time and time constant
were adjusted individually for each spectrum
depending on the intensity of the signal being
characterized. Diphenylpicrylhydrazyl (DPPH) (g =
2.0036) was used as an external standard for the
determination of g-values.

RESULTS
X-ray diffraction
Composition of the original kaolin, K. The XRD
pattern from the original Ranong kaolin K is shown

543

in Fig. 1. As reported by Kuentag (2001) and


Nuntiya & Prasanphan (2006), the majority of the
peaks can be accounted for by the kaolin minerals;
correspond
the strong peaks at 7.14, 3.57 and 1.49A
to the kaolinite d001, d002 and d060 reflections and
the peaks at 4.45, 2.56, 2.49, 2.34, 1.99, 1.66, and
1.49 are all consistent with kaolinite. Positive
identification of halloysite from these data is
difficult, because the main peaks from halloysite at 10A
and halloysite-7 A
at 4.42 A
both
10 A
overlap with peaks from illite, and the peak at
can correspond to either halloysite or
4.34 A
kaolinite (Brindley & Brown, 1980). Illite, which
is a common impurity in kaolin-group minerals, is
present in appreciable quantities in the Ranong
kaolin, as seen by the presence of major peaks at
in addition to those mentioned
9.92 and 2.56 A
above. Other peaks consistent with illite are
. The presence of
observed at 4.99 and 2.99 A
,
quartz is seen by its principal reflection at 3.34 A
and associated minor peaks at 4.25, 2.29, 1.82, and
. In addition, weak peaks at 4.18 and 3.25 A

1.54 A
could correspond to the main reflections from
goethite and microcline, respectively, so these
minerals may be present as minor impurities, but
the absence of other diffraction peaks prevents a
positive identification.
The clay mineral groups were distinguished by
the behaviour of their characteristic basal reflections in response to various chemical and heat
treatments (Fig. 2). The insensitivity of the peaks at
to ethylene glycol, and heating to 550C
~10.0 A
mean that they correspond at least mainly to the

d001 reflections from illite, and that halloysite-10 A

FIG. 1. Powder X-ray diffraction pattern of untreated kaolin, sample K, and the assignment to its component
minerals. K = kaolinite, H = halloysite, Q = quartz, I = illite, G = goethite and M = microcline

544

N. Worasith et al.

FIG. 2. Powder X-ray diffraction patterns of oriented kaolin, sample K, as a result of (a) air-drying,
(b) glycolation, (c) intercalation with formamide, and (d) heating to 550C.

makes only a minor contribution (if any) to their


and kaolinite are unafintensity. Halloysite-7 A
fected by glycolation, but become amorphous to
X-rays after heating to 550C, and Fig. 2 shows the
and ~3.5 A
to an
conversion of peaks at d ~7 A
amorphous phase at ~12.382y at this temperature.
Finally, since intercalation with formamide results
in an increase in the d001 reflection of halloysite-7
to ~10.0 A
, the peak at 7.17 A
in the original
A
kaolin corresponds primarily to halloysite. These
results are thus qualitatively similar to those
reported by Nuntiya & Prasanphan (2006) for
their kaolin sample from Ranong, but the sample
in the present work contained smaller amounts of
impurity phases. The Hinkley index (HI) (Hinckley,
1963) of about 0.32 suggests that the kaolinite is of
only moderate crystallinity, whereas both the FTIR
and EPR results (see below) indicate that the
mineral is well ordered. However, Brindley et al.
(1986) report that the HI is directly correlated with
kaolinite Fe content, so its relationship to crystallinity may be complicated by isomorphous substitutions. Furthermore, the presence of overlapping
peaks from illite and quartz with those of kaolinite
used for measuring the HI complicates its accuracy
for kaolinite crystallinity determination in the
present sample.
Influence of individual chemical and physical
treatments on the kaolin crystallinity. XRD patterns

for Ranong kaolin before and after grinding or


sulphuric acid treatment are shown in Fig. 3. They
show that the mineral is resistant to the acid
treatment (Fig. 3b,c) since the d001 reflection at
, the d002 reflection at 3.57 A
, and the d060
7.14 A
remained sharp, even under
reflection at 1.48 A
strongly acidic conditions (Table 1). However,
grinding the clay resulted in a major decrease in
the relative intensities of the peaks from the kaolin
mineral and illite phases (Fig. 3d), consistent with
appreciable structural damage to these minerals.
Furthermore, there was also an increase in the
background in the 20 302y range which is typical
of the presence of amorphous phases.
Using the Scherrer formula (Langford & Wilson,
1978) the FWHM data for d001 indicates that the
mean number of layers in a kaolinite stack is ~90
for samples K, KS18 and KS30. This value
decreases to ~70 (i.e. a reduction of ~25%) after
grinding, but then increases again to ~90 for
GKS18 and GKO6. Grinding had very little effect
on the d060 peak and samples K and GK both have
~380 planes in a stack. However, it is of great
importance after acid attack, with the number of
layers in the b dimension decreasing to ~280 for
KS18 and KS30, and ~190 for GKS18 and GKO6.
Effects of combined physical and chemical
treatments. Further changes in the XRD patterns
were observed after treating the ground sample with

545

Characterization of modified kaolinite

FIG. 3. Powder X-ray diffraction patterns for Ranong kaolin, sample K, and modified samples, KS18, KS30, GK,
GKS18 and GKO6. Peaks corresponding to the identified mineral phases are shown on the diffractogram of the
untreated sample.

either sulphuric or oxalic acids (Fig. 3e,f); apart


from the peak at ~26.742y, which probably
corresponds to quartz, there is an appreciable
reduction in intensity in all regions. Al compounds
resulting from grinding-induced kaolin breakdown
were removed by the acid treatments as indicated
by the XRF results, presumably as a result of the
formation of soluble Al salts. Although the peaks
corresponding to the kaolinite/halloysite minerals
were very weak in these XRD spectra, they were
still quite sharp (Table 1), suggesting that the

residual fraction of these minerals is relatively


unaltered.

X-ray fluorescence spectra


Table 2 shows the major elements determined by
XRF analyses of the original kaolin, the original
and ground kaolins treated with 18% w/w sulphuric
acid and the ground kaolin treated with 6% w/v
oxalic acid. For comparison, this table also shows
the results from analysis of a second set of samples

TABLE 1. Variation of positions and line-widths (2y) of the main kaolinite XRD peaks in the various Ranong
kaolin samples.
Sample

d001

FWHM

d002

FWHM

d060

FWHM

K
KS18
KS30
GK
GKS18
GKO6

7.09
7.11
7.12
7.15
7.15
7.14

0.27
0.29
0.27
0.37
0.27
0.28

3.56
3.56
3.57
3.57
3.57
3.57

0.29
0.32
0.28
0.41
0.40
0.31

1.49
1.49
1.49
1.49
1.49
1.49

0.32
0.44
0.42
0.31
0.60
0.69

546

N. Worasith et al.

obtained from this deposit. Si and Al are the


dominant elements in the original kaolin, along
with minor amounts of K and Fe; the concentrations
of Mn, Ti, Mg and Ca are all small. Decreases in
the Al:Si and Fe:Si ratios were observed as a result
of acid treatment of the unground clay, but much
larger decreases were observed after acid leaching
the ground kaolin. Since the main minerals in the
sample are halloysite, kaolinite, illite and quartz
with minor amounts of mica and orthoclase and
iron oxides, it is possible to use the data in Table 2
to get a rough idea of the composition of the
samples. By using the theoretical chemical compositions for these phases, and attributing all the Al to
kaolinite/halloysite, we find that this component
accounts for ~85% of the original kaolin sample
and this decreases to ~40% in GKS. In ground and
acid-treated products, the main phase is probably
amorphous silica (450%) which is consistent with
the strong development of Si O bands in the FTIR
spectra (see below).

Scanning electron microscopy


Results from SEM/EDX studies of the Ranong
kaolin and its modified products are shown in
Fig. 4. The original clay sample K shows the
presence of both platy and tubular shapes that are
characteristic of kaolinite and halloysite respectively. However, these structures were largely

destroyed by grinding (sample GK). Hot acid


treatments of the ground sample then resulted in
the formation of products with globular morphology
(samples GKS18 and GKO6). Chemical analysis of
sample surfaces performed using EDX indicate that
Si, Al, O are the major elements, and K and Fe are
minor components, as was found with the XRF
analyses of the bulk samples. Also as observed with
XRF, acid treatment of the ground samples resulted
in an appreciable decrease in the Al content.

FTIR spectroscopy
The main absorptions in the FTIR spectra of the
kaolin samples are shown in Fig. 5. In the original
kaolin sample, K, the O H stretching vibrations
consist of two well defined peaks at 3694 and
3621 cm 1 flanking two weaker peaks at 3669 and
3652 cm 1. This spectrum is typical of kaolinite (e.g.
van der Marel & Krohmer, 1969, report bands at
3693, 3668, 3652 and 3620 cm 1). In contrast to the
HI from the XRD results, the good resolution of
these peaks suggests that the mineral is quite well
ordered, and provides support for the analysis of the
d001 and d060 peaks in the XRD results. As proposed
by Farmer (1974, 1998), and subsequently confirmed
by theoretical calculations (Balan et al., 2005), the
band at the lowest wave-number corresponds to
vibrations of the inner OH group, that at the highest
wave-number to the in-phase motion of the three

TABLE 2. Chemical composition (wt.%) of Ranong kaolin samples determined by X-ray fluorescence.
Oxide
(wt.%)
SiO2
Al2O3
K2O
Fe2O3
MnO
TiO2
CaO
MgO
Na2O
LOI
Al:Sia
Fe:Sia

Sample
K
KS18
GKS18
GKO6
43.30 (46.70)
37.20 (37.90)
1.36 (1.23)
0.66 (0.54)
0.05 (0.04)
0.03 (0.03)
0.01 (0.01)
0.04 (0.03)
0.04 (0.03)
17.31 (13.49)
1.01 (0.95)
1.14% (0.85%)

55.35 (51.40)
29.63 (33.63)
1.46 (1.46)
0.49 (0.41)
0.04 (0.03)
0.02 (0.02)
n.d. (n.d.)
0.03 (0.03)
0.03 (0.03)
12.94 (12.90)
0.63 (0.77)
0.66% (0.60%)

74.80 (72.77)
16.80 (13.50)
1.81 (1.48)
0.44 (0.29)
0.04 (0.03)
0.03 (0.03)
n.d. (n.d.)
n.d. (0.03)
0.04 (0.04)
6.08 (11.76)
0.26 (0.22)
0.44% (0.27%)

70.70 (67.84)
20.40 (16.13)
2.01 (1.51)
0.50 (0.27)
0.04 (0.03)
0.04 (0.03)
0.03 (n.d.)
n.d. (0.02)
n.d. (0.04)
6.08 (14.09)
0.34 (0.28)
0.53% (0.30%)

The numbers in brackets correspond to analyses of a second set of samples obtained from the same deposit
atomic ratios
n.d. = not detected

Characterization of modified kaolinite

547

FIG. 4. Scanning electron microscope micrographs and representative energy dispersive X-ray analysis spectra
from small areas on the surface of Ranong kaolin samples K, GK, GKS18, and GKO6. Scale bars = 1 mm.

surface OH groups, and those at intermediate


positions to out-of-phase motion of these groups.
The bands at ~935 and 913 cm 1 correspond to
the inner surface and inner Al OH Al bending

vibrations (Farmer & Russell, 1964) and those at


1116, 1030 and 1007 cm 1 to Si O stretching
vibrations in kaolinite/halloysite, but there was only
a hint of a shoulder at ~1100 cm 1 where the Si O

548

N. Worasith et al.

FIG. 5. Fourier transform infrared spectra of the main absorption regions in the FTIR spectra of modified Ranong
kaolin samples.

(apical) band is expected (e.g. Frost et al., 2002).


The band at 1030 cm 1 could also contain a
contribution from muscovite, which was detected
by Nuntiya & Prasanphan (2006) as an impurity in
their Ranong kaolin, but was not seen in the XRD
patterns from our present sample. The peak at
793 cm 1 is probably associated with quartz, which
was identified as an impurity in the XRD traces of
both Nuntiya & Prasanphan (2006) and the present
work. There was no evidence for a discrete
absorption at ~3425 cm 1, the frequency at which
the O H stretching vibration occurs in smectites
which are also common impurities in kaolin
minerals. The bands at 751 and 687 cm 1

correspond to those assigned to Si O stretching


vibrations (Ekosse, 2005).
Little change was observed in the spectra as a
result of sulphuric acid treatment, indicating that
the kaolin minerals are resistant to acid treatment,
although weak new peaks were seen at ~1220 and
828 cm 1 and a shoulder at ~1075 cm 1; since
peaks were observed in these positions, but with
higher intensity in the ground samples after acid
treatments, they indicate that the acid treatment did
induce some modification of the unground kaolin.
There was also some decrease in the relative
intensities of the bands at ~935 and 913 cm 1
from Al OH Al bending vibrations. In contrast,

549

Characterization of modified kaolinite

grinding reduced the intensities of the kaolinite


O H stretching vibration bands by ~65%, indicating that physical modification of the mineral was
accompanied by extensive dehydroxylation as
observed by Miller & Oulton (1970). However,
the Al OH Al bending vibration bands were
reduced by a smaller amount, and the O H and
Al OH Al peaks that remained after grinding
were unshifted. Thus, they indicate that the
kaolinite that remained was essentially unaltered
from that in the original sample.
Frost et al. (2001a,b) have also reported IR
results for a mechanochemically-treated kaolinite,
in which there were both similarities and differences from the measurements reported above. Both
our measurements and those of Frost et al. showed
decreases in the intensity of the OH stretching
vibrations (3695 and 3619 cm 1) and deformation
modes (935 and 913 cm 1), but Frost et al. also
observed an increase in the bending mode at
1650 cm 1 from water bound to the surface of the
modified kaolinite; this was extremely weak in the
present measurements (probably because of the
drying of the sample). Also, whereas we observed
little change in the Si O stretching vibrations,
Frost et al. reported a significant reduction in

intensity that was accompanied by an increase in a


new band at 1113 cm 1, which was assigned to the
surface of the newly synthesized product. It should
also be noted that the Si O stretching vibrations
reported by Frost et al. (2001b) were at 1034 and
1056 cm 1, which are considerably higher than the
values of 1011 and 1030 cm 1 that were observed
in the present work.
Acid treatment of the ground sample did not
produce any further changes in the OH stretching
frequencies, and the differences in the relative
intensities of the peaks are not considered
significant because of the low overall intensity in
this region of the spectra. There were, however,
appreciable changes in other regions of the spectra.
The intensities of the Al OH Al bending vibrations were reduced, and are roughly correlated with
the reduction in the Al content (Table 2). There
were also significant differences in the relative
intensities of new bands in the sulphuric and oxalic
acid-treated samples, although their positions were
similar (Table 3). Intense broad bands generated at
~1205 cm 1 and 1067 or 1077 cm 1 (depending on
the acid used) probably correspond to Si O
stretching bands associated with newly synthesized
material, since weak bands in similar positions were

TABLE 3. FTIR frequencies (cm 1) and assignments in Ranong kaolin samples.


NK

KS18

KS30

GK

GKS18

GKO6

3694
3669
3652
3621

3692
3674
3649
3621

3692
3672
3650
3621

3695
3670
3646
3623

3692
3669
3648
3621
3397
(v broad)
1646
1633
1205
1116

3692
3669
3648
3621
3346
(v broad)
1650
1634
1205
1116
1077
1035
1013

~1650
~1560
1116
1030
1007
935
913
793
751
687
648

1654

1650

~1220
1116
~1075
1030
1007

~1220
1116
~1075
1030
1007

935
913
828
796
751
687
647

935
913
828
794
751
687
647

1650
1632
1116
1033
1011
935
913
794
751
687
647

1067
1036
1011
940
916
828
796
754
696
649

937
914
828
796
754
696
642

Assignment
O H

O H
Surface water
Si O
Si O

Al OH Al
Quartz
Si O

550

N. Worasith et al.

observed in the spectra of the unground clay after


acid treatment. These are significantly different
from the results reported by Frost et al. (2001), who
described the formation of a new band at
1113 cm 1; the narrow peak at 1117 cm 1 in the
natural kaolin (and at 1121 cm 1 in the ground
sample) in the present work, was lost completely in
the acid treatment experiments of Frost et al.
(2001a,b, 2002, 2003, 2004). However these authors
found that the presence of quartz increased
appreciably the rate of grinding-induced breakdown
of the kaolinite, and this may also be a factor in
determining the physical characteristics of the
breakdown products.
There was a reduction in the relative intensities
of the Si O deformation bands, which occurred at
1007 and 1030 cm 1 in the original kaolin sample,
and these were also shifted to slightly higher wavenumbers after grinding and acid treatment, but
overall the changes in this region of the spectrum
were relatively minor. Larger changes were
observed with the Al OH Al bending vibrations
at 935 and 913 cm 1; although their intensities
were reduced after grinding, much greater additional reductions were apparent after the acid
treatment, consistent with dissolution of the

octahedral sheet which contains most of the O H


groups.

EPR spectroscopy
The wide scan EPR spectra of the Ranong kaolin
before and after grinding are shown in Fig. 6. As is
common for clay mineral samples with a relatively
low overall content of paramagnetic ions, these
spectra contain signals from four different sources;
(i) a group of peaks in the field range 50 250 mT
which originate from isolated Fe3+ ions, (ii) a broad
feature centred on 350 mT from magnetically
interacting ions, probably also Fe3+, (iii) a sextet
component with peak separation ~9.5 mT centred
on 350 mT from Mn (55Mn, I = 5/2), and (iv) a
narrow feature centred on 350 mT from free radical
defect centres in the mineral structure. Similar
features have been reported in previous studies of
kaolin samples (e.g. Goodman & Hall, 1994), and
are discussed individually below.
Only the low field signals were affected by
physical treatment of the original kaolin. These
signals consist of peaks at ~80, ~140, ~195 mT and
a cross-over at 163 mT; the last of these signals
increased and the others decreased as a result of

FIG. 6. Wide scan electron paramagnetic resonance spectra of unground and ground Ranong kaolin samples
before and after treatment with 18% w/w sulphuric or 6% w/v oxalic acid.

Characterization of modified kaolinite

grinding. There was, however, little further change


in these signals when the ground samples were
extracted with sulphuric or oxalic acids. In contrast
to the low field Fe3+ signals, the Mn signal showed
little change as a result of grinding alone, but its
shape was modified as a result of acid treatment.
This resulted in a sharpening, but little change in
the separation of the Mn lines, and the effect was
greater with the ground than with the unground
kaolin; it was also much more pronounced with
oxalic than with sulphuric acid treatment.
The free radical signal in the original kaolin is
shown in Fig. 7. The main feature is highly
anisotropic with g// ~2.050 and g\ ~2.008, and is

551

similar to the A-centre described by Angel et al.


(1974). It corresponds to electron holes trapped on
oxygen atoms (Cuttler, 1981), and has been
associated with radiation-induced defects in
natural kaolinites (Clozel et al., 1994). The g\
region in Fig. 7, however, clearly contains more
than one feature, and is consistent with the Q-band
EPR results of Clozel et al. (1994), who described
the presence of three distinct free radical centres in
kaolinite that had been subjected to g-irradiation.
Furthermore several (at least six) peaks are evident
between the g// and g\ features of the second
derivative spectrum and correspond to hyperfine
structure from a radical centre involving interaction

FIG. 7. Expansion of the free radical region of the 1st and 2nd derivative electron paramagnetic resonance spectra
from the Ranong kaolin samples K, KS18, GK, GKS18 and GKO6.

552

N. Worasith et al.

of unpaired electrons with 27Al (I = 5/2) nuclei.


This is consistent with the B-centre described by
Clozel et al. (1995) as corresponding to O centre
linking two Al atoms in octahedral sites, though
these authors were not able to determine whether
this was at a surface position, or associated with an
O atom linking the octahedral and tetrahedral
sheets. It should also be noted that the present
results cannot distinguish between a B-centre and
one with a single 27Al nucleus interacting with the
unpaired electron.
There was little influence of sulphuric acid
extraction on the A-centre free radical signal, but
there was an improvement in the resolution of the
27
Al hyperfine structure associated with the
B-centre. Since the most likely effect of the acid
treatments is removal of surface-adsorbed species,
this result suggests that the B-centre is associated
(at least partially) with a surface O atom on the

octahedral sheet. Furthermore, the absence of any


H hyperfine structure indicates that this oxygen is
not protonated.
Grinding resulted in a loss of the 27Al hyperfine
structure from the B-centre, although the signal from
the A-centre was little changed, either by grinding
alone, or subsequent acid treatment. However, as
seen by the variations in features labelled A and B
in the g\ region of the second derivative recording,
the combined treatments resulted in a change in the
relative amounts of the components that contributed
to this absorption. However, at least one of these
components probably corresponds to a defect centre
in quartz for which radiation-induced defects are
well known (e.g. Weeks, 1956).
There was little change in the low field Fe3+
signals after these kaolin samples had been treated
by re-suspension in either sulphuric or oxalic acid
at different pH values, but the widths of the Mn
1

FIG. 8. Effect of adjusting the surface pH on the electron paramagnetic resonance spectra from Ranong kaolin that
had been ground for 1 h in a ball mill and treated with either 18% w/w sulphuric acid or 6% w/v oxalic acid at
90C for 4 h, before being suspended in solutions with different pH values, then separating by centrifugation and
drying at 100C.

Characterization of modified kaolinite

peaks decreased with increasing pH; also this effect


was far more pronounced with oxalic acid than with
sulphuric acid (Fig. 8). This result suggests that this
treatment is not simply a protonation/deprotonation
reaction, and that the chelating ability of oxalic acid
may also be an important factor. In the sample
treated with sulphuric acid at pH 3.5, a new feature
with g ~2.5 was also generated. Since it was
reproduced in replicate experiments, it corresponds
to a genuine product, but its interpretation is
unclear. However, under certain conditions, Fe3+
can produce an isotropic spectrum in this region
(Golding et al., 1977; Mabbs & Collison, 1992).

DISCUSSION
Although the principal objective of this paper was
the characterization of the products formed by
combined physical and chemical treatments of
Ranong kaolin, it was deemed necessary first of
all to investigate the unaltered material and the
products of individual chemical and physical
treatments. This is because the reaction pathways
of the mineral are likely to be influenced by both its
mineral components and its chemical composition.
The XRD pattern of the untreated kaolin sample
(Fig. 1) shows that it is composed primarily of
kaolinite and halloysite with some quartz and illite
or mica, along with possible minor amounts of
microcline and goethite, the latter probably being
responsible for the yellow colouration of the sample.
The methods that can be used to distinguish between
kaolinite and halloysite have been reviewed by
Joussein et al. (2005) and, by preparing oriented
samples intercalated with formamide, we were able
to show that halloysite was the dominant mineral in
the unaltered kaolin sample. Thus the present result
is qualitatively similar to that of Nuntiya &
Prasanphan (2006), although these latter authors
observed the additional presence of muscovite and
appreciably greater amounts of quartz. However,
local variations in the composition of kaolin deposits
are common as a result of variations in the facies of
the parent rock, the degree of weathering, and
variations in micro-environmental condition during
kaolin formation (Duzgoren-Aydin et al., 2002).
Systematic changes in composition with particle size
have been reported by Lombardi et al. (1987) for
kaolin deposits in Europe and North America, for
example. Thus it is not surprising that there should
be some differences between the samples of the
kaolin used in the present work and that investigated

553

by Nuntiya & Prasanphan (2006), even though they


were obtained from the same general geographical
region and are described under the same name.
The XRF results for the kaolin samples used in
the present work also indicate some differences in
chemical composition compared to the sample
investigated by Nuntiya & Prasanphan (2006);
although there were only minor differences in the
SiO2 and Al2O3 concentrations, and the MnO
concentrations are virtually identical, the Fe2O3
content is much smaller in the present sample.
Indeed, our analysis of a second set of samples
from this deposit (Table 2) showed greater variations for the Fe2O3 contents than for any other
elements. Thus there are probably considerable
variations in the concentrations of associated iron
oxide phases in different samples from the deposit.
This is a factor that must be considered when
developing novel uses for the mineral, because such
phases may make disproportionately large contributions to the sorption properties (as a result of their
often poorly crystalline nature); it should be noted
that deferration treatments may have an adverse
effect on the sorption properties. Furthermore, the
appreciable differences between the FTIR spectroscopic results from the present investigations and
those of the mechanochemically-treated kaolin of
Frost et al. (2001a,b) are also an indication of
different chemical behaviour of samples from
different origins; these could be the consequence
of different chemical compositions or structural
order, or simply the amount of quartz in the mineral
sample (Mako et al., 2001). Thus careful attention
must be paid to physical characterization when
developing samples for practical uses, such as
adsorbents in the food industry, and caution
should be exercised in extrapolating properties
from one specimen to another without detailed
physical characterization.
The tubular morphology observed in the SEM
result from the unaltered sample (Fig. 4) is
indicative of the presence of halloysite. However,
made only small contributions to
halloysite-10 A
the XRD pattern, but measurements with intercalated formamide (Fig. 2) indicated that the
corresponded
reflection at around 7.1 7.2 A
. The absence of any
primarily to halloysite-7 A
tubular structures in the SEM images and the large
weakening of the peaks associated with kaolinite/
halloysite in the XRD patterns from the ground
kaolin samples (Fig. 3) indicate that the habits of
these minerals are largely destroyed by the physical

554

N. Worasith et al.

treatment. The SEM images show that grinding


produced a decrease in particle size (Fig. 4).
Increases in specific surface area and total pore
volume have been reported (Worasith et al., 2011),
and these observations are all consistent with the
formation of poorly crystalline or amorphous
phases. The FTIR spectroscopic results indicate
that OH groups were lost as a result of grinding,
presumably as a result of dehydroxylation caused
by heat generated during the grinding process.
Since these groups are associated with the surface
of the octahedral sheet, there would inevitably be a
major change in the coordination environment of
the Al and a decrease in the number of hydrogen
bonds that hold together the individual aluminosilicate layers. This result is also consistent with the
NMR study of Temuujin, et al. (2001), who
reported destruction of the octahedral layers of
kaolinite by prolonged grinding. However, the
present FTIR results indicated that the small
amount of residual kaolinite was largely unaffected
by the combined physical and chemical treatments.
The XRD, FTIR and EPR results all show only a
small effect of hot sulphuric acid treatment on the
unground kaolin (Figs 3, 5 and 6) and indicate that
the kaolinite and halloysite components have a high
degree of resistance to attack by mineral acids,
presumably because of the strong hydrogen bonds
in their structures. This result should be contrasted
with that of Steudel et al. (2009) who found that
activation of kaolinites by sulphuric acid at 80C
produced layered materials with a simple chemical
composition (silica) and large surface area as a
result of successive dissolution of octahedral sheets
by edge attack. However, in comparing the
behaviour of di- and trioctahedral minerals,
Steudel et al. (2009) also reported that this reaction
is aided by the presence of structural Mg or Fe,
which were removed faster than Al; thus the very
small Mg content of the Ranong kaolin (Table 2)
may be a reason for its resistance to acid.
Nevertheless, the kaolin was not completely inert
to acid attack. The XRF results (Table 2) show
respective reductions in the Al:Si and Fe:Si atomic
ratios from 1.01 to 0.63 and 1.14% to 0.66% for
treatment of the unground kaolin with 18%
sulphuric acid, and there was a reduction in the
relative intensity of the broad g = 2.0 component in
the EPR spectra, which is consistent with the
removal of associated poorly crystalline iron oxide
phases. Treatment of the original kaolin sample
with hot sulphuric acid had little or no influence on

the low field Fe3+ EPR signal (Fig. 6), indicating


that any removal of (surface-bound) material by
acid treatment had no significant effect on the local
environment of structural Fe. Also, although there
was a reduction in the relative intensity of the broad
g = 2.0 component, it retained appreciable intensity
after acid treatment, indicating the presence of
magnetically interacting ions in the aluminosilicate
structure; this could correspond either to illite, or an
Fe oxide phase. Finally, the FTIR spectra show the
formation of new features at ~1220, 1075 and 828
cm 1, which are probably associated with an acidmodified component, and there was some improvement in the resolution of the B-centre radical, which
suggests that this is associated with oxygen atoms
on the surface of the octahedral sheet.
Hot acid treatments of the ground kaolin sample
resulted in large increases in the specific surface
area and the total pore volume (Worasith et al.,
2011), and this was accompanied by a major
decrease in the Al:Si ratios to 0.22 and 0.29 for
the 18% w/w sulphuric acid and 6% w/v oxalic acid
treatments, respectively (see Table 2). This was also
accompanied by reductions in the Fe:Si ratios to
0.6% and 0.7% for the corresponding treatments
with 18% w/w sulphuric acid and 6% w/v oxalic
acid. Thus, after breakdown of the aluminosilicate
structures by physical treatment, the components of
the original octahedral sheet were selectively
extracted by acid, leaving a Si-rich phase derived
from the tetrahedral sheet. This latter phase was
characterized by a new, but as yet unidentified,
peak in the XRD diffractogram, and major new
Si O stretching peaks in the FTIR spectra.
The EPR spectra provide an insight into the
structural changes that affect the Fe, Mn and free
radical components, all of which make relatively
minor contributions to the overall composition of
the kaolin, but which nevertheless represent
potential sites for chemical reactions. In the EPR
spectra of kaolinites, low field features from Fe3+
are common and correspond to two distinct
environments for the Fe3+ (Mestdagh et al., 1980;
Gaite et al., 1997). Balan et al. (1999) have shown
that both signals correspond to substitution of Fe3+
for Al in octahedral sites within the kaolinite
structure. The signal designated Fe(I) by Balan et
al. (1999), which corresponds to the central feature
at 163 mT in Fig. 6, has been assigned to a
disordered mineral structure. The remaining lowfield peaks in Fig. 6 are all derived from the signal
designated Fe(II) by Balan et al. (1999), and

Characterization of modified kaolinite

correspond to Fe3+ in a well-ordered crystalline


phase with a small number of stacking defects in its
structure.
It is clear that grinding has a major influence on
the relative proportions of iron in the Fe(I) and Fe(II)
sites. However, on the basis of the results in the
present work, it seems likely that the increased
fraction of the Fe3+ in the Fe(I) sites is the
consequence of dehydroxylation of the octahedral
sheet and not a change in the long-range order of
the kaolin-group minerals, which is the usual
assignment for such a component. The fraction of
the Fe3+ that remains in the Fe(II) sites is therefore
likely to correspond to the unaltered kaolinite
component that was seen in the XRD and FTIR
spectra. In contrast to the signals from the structural
Fe3+, the broad absorptions from magneticallyinteracting ions, along with those from the Mn
and free radical components, are not affected
significantly by grinding. The absence of a major
increase in the free radical signal as a result of the
grinding is somewhat surprising, since such
treatment is well-known to generate free radical
defects in mineral structures (e.g. Vallyathan et al.,
1988).
Although acid treatment of the original kaolin
had little effect on the distribution of Fe between
the Fe(I) and Fe(II) sites (see above), treatment of
the ground kaolin with either hot sulphuric or oxalic
acid (Fig. 6) increased the Fe(II) contribution to the
total Fe3+ signal, although it still remained much
lower than in the unground specimens. This
provides further evidence to support the conclusion
that these acid treatments result in selective
dissolution of the octahedral sheet from those
mineral components that were altered (dehydroxylated) during the grinding process and as a result
the concentrations of the unaltered fractions were
increased by subsequent acid treatments.
All of the spectra contain features attributable to
interactions of unpaired electrons with 55Mn nuclei,
identifiable by the characteristic sextet structure
from the I = 5/2 nucleus. Although not a common
substituent of kaolinites, there have been previous
reports of EPR spectra of Mn in natural kaolin
specimens, but it is usually assigned either to the
presence of impurity phases containing Mn2+ (e.g.
Sengupta et al., 2006), or Mn2+ adsorbed on
kaolinite surfaces (e.g. McBride et al., 1975). The
55
Mn hyperfine splitting of ~9.3 mT in the original
kaolin is typical of that for Mn2+ octahedrally
coordinated to oxygen atoms, as for example in

555

Mn(H2O)2+
6 on the exchange sites in clay minerals
(McBride et al., 1975) or in oxide crystals, such as
CaO or CdO (Title, 1963). Unlike the low-field
Fe3+ components, little grinding-related change was
seen in the Mn signal, which would appear to
exclude their presence in the octahedral sheet,
although the 55Mn hyperfine splitting increased
slightly to ~9.5 mT. However, the shape of the Mn
signal was very sensitive to acid treatment and the
chemical nature of the acid used, and both the
unground and ground samples indicate that the Mn
is in sites that are accessible to acids. However, the
lack of any significant decrease in overall Mn
signal intensity and the similar concentrations in the
XRF measurements before and after acid treatment
would appear to exclude surface adsorption of Mn2+
as the explanation for this signal; thus the Mn is in
structural sites. Association of the Mn with an
impurity phase cannot be specifically excluded,
although XRD provided no evidence for the
presence of significant amounts of likely candidate
phases, such as muscovite which was seen in the
Ranong kaolin sample investigated by Nuntiya &
Prasanphan (2006). Furthermore, the relatively
sharp peaks indicate that any impurity phase
would need to have the Mn in a magnetically
dilute form (i.e. low concentration), and located
close to the surface (but not exchangeable) in order
to explain the effect of acid treatment on its spectral
shape.
Further changes in the Mn signal were observed
after the modified kaolin had been re-suspended in
solutions of different pH values. This adds further
support to the conclusion that the Mn is located in
positions in the lattice that are different from those
that are occupied by the Fe3+ ions, which were
unaffected by such treatment. Thus, by a process of
elimination we propose that the Mn is located in the
tetrahedral sheet and therefore is probably in the
Mn4+ oxidation state. The interpretation of a Mn
signal in the EPR spectra of aluminosilicate
minerals as corresponding to Mn4+ in tetrahedral
sites is unusual, and such signals are commonly
assigned to Mn2+ in octahedral sites (e.g. Martin et
al., 1999). However, the ionic radii for Mn4+ and
for tetrahedral
Al3+ are essentially identical (0.39 A
coordination), somewhat larger than that of Si4+
) but smaller than that of tetrahedral Fe3+
(0.26 A

(0.49 A) (Shannon, 1976). Thus, on size considerations, Mn4+ can be accommodated in any site that
can accommodate Al3+, and on charge considerations Mn4+ can replace Si4+ without creating any

556

N. Worasith et al.

charge imbalance. Furthermore, Mn4+ is more


easily accommodated in tetrahedral sites than
Fe3+, yet tetrahedral Fe3+ is an accepted substituent
in several aluminosilicate minerals (e.g. Goodman
et al., 1976; Rancourt et al., 1994)). In contrast, the
ionic radius of Mn2+ in octahedral coordination is
, considerably larger than the values of
0.83 A
and 0.645 A
for octahedral Al3+ and Fe3+
0.535 A
respectively. It is also appreciably larger than the
for Mg2+ in octahedral coordinavalue of 0.72 A
tion; thus there are difficulties in accommodating
the larger Mn2+ ion in aluminosilicate mineral
structures. However, there is a further complicating
factor of the possible presence of Mn3+, which is
not easily detected by EPR (see for example Mabbs
& Collison, 1992), especially at trace concentrations, but whose ionic radius is very close to that of
Fe3+.
One may assume that the kaolinite structure
might contain trioctahedral regions in which Mn2+
could be located, but this interpretation is not
consistent with the presence of a well-resolved
55
Mn hyperfine structure. EPR spectra from
samples with clusters of Mn2+ ions would indeed
induce a broad peak as a result of exchange
interactions and would, therefore, contribute to the
broad g = 2.0 signal. It also seems unreasonable to
propose that Mn2+ is a minor component in
trioctahedral clusters in which the major divalent
ions are diamagnetic, because the analytical results
do not indicate the presence of sufficient quantities
of such ions that would be necessary for such a
structure.
This hypothesis of the location of Mn4+ in
tetrahedral sites is in contrast to the location of
V4+ in Georgia kaolin, for which Gehring et al.
(1993) found strong evidence for it being located
primarily in the octahedral sheet. However the
Jahn-Teller effect (Jahn & Teller, 1937) in V4+
makes its accommodation in the symmetrical
tetrahedral sites difficult, and its ionic radius in
) (Shannon, 1976) is
octahedral coordination (0.58 A
intermediate between those of Al3+ and Fe3+, thus
supporting its easy substitution in the octahedral
sheet.
The main free radical signal corresponds to the
A-centre reported previously for defects induced by
exposure to g-irradiation (Clozel et al., 1994). Its
relative insensitivity to prolonged grinding is
consistent with the assignment to an electron hole
in an O atom associated with Si, and its stability
adds support to the proposed use of this resonance

for measuring the radiation history of the sample


(Clozel et al., 1990). The 27Al hyperfine structure
that is associated with the B-centre was not resolved
in the ground samples, consistent with grindinginduced damage to the octahedral sheet.
In recent years there has been growing interest in
the modification of kaolinite or kaolin minerals to
produce new materials with specific value-added
properties. These have involved one or a combination of acid, alkali, heat or mechanical treatments,
and the relative stability of these non-swelling
minerals means that it is necessary to use more
severe activation treatments than with swelling
minerals, such as bentonites. Nevertheless the
relatively low cost of kaolins means that even the
need for more complex modification procedures can
still be cost effective. The material described in the
present paper is an example where there is now
good evidence that it can be a low-cost alternative
to bentonite bleaching clay for the decolourization
of rice bran oil (Worasith et al., 2011).

CONCLUSIONS
Because of their potential for use as bleaching clays
in the food industry, the products from physical and
chemical treatments of the Ranong kaolin have
been characterized using a range of physical and
chemical techniques. Acid treatments alone had
little effect on the aluminosilicate mineral structures, but (dry) grinding produced a major reduction
in OH content, as a result of dehydroxylation of
the surface of the octahedral sheet. Subsequent acid
treatments then resulted in the production of
material with greatly increased surface area, low
Al:Si ratios, and amorphous structure, which is
presumably responsible for the enhanced sorption
properties of the product.
We were also able to obtain a detailed picture of
the paramagnetic ions and free radical defects
within the Ranong kaolin, and the extent to which
their environments are influenced by physical and
chemical treatments. Grinding-induced changes
primarily affected the Fe3+ (i.e. octahedral cation
sites), but not the Mn EPR signal. On the other
hand, acid-induced changes (i.e. in the degree of
protonation of surface oxygen atoms) were seen in
the shape of the Mn EPR signal, a result which may
suggest that the Mn is located in the tetrahedral
sheet, and hence probably is in the Mn4+ state.
Further investigations including X-ray absorption
spectroscopy in order to probe coordination of Mn

Characterization of modified kaolinite

atoms would be needed to fully validate such an


assumption. Finally, the major free radical centre,
the so-called A-centre which corresponds to defects
caused by exposure to low levels of g-irradiation,
was little affected by either the physical or
chemical treatments. However, despite all of this
information on the kaolin composition and behaviour, we still do not know to what extent these
various paramagnetic components contribute to the
chemical behaviour of the products from the
physical and chemical treatments of the kaolin.
ACKNOWLEDGMENTS

Special thanks are due to Arag Vitittheeranon, of the


Office of Atoms for Peace (OAP), Thailand, Dr.
Sakorn Suwan, Chulalongkorn University, Thailand
for use of their EPR spectrometers, and Sirikan
Noonpui, Pilot Plant Development and Training
Institute, King Mongkuts University of Technology,
Thonburi, Thailand for assistance with the FTIR
spectroscopy. We are also grateful to Drs. Katja
Emmerich, Claude Fontaine and an anonymous
reviewer for their thorough reviews and detailed
comments that have helped to improve the presentation
of this work.
REFERENCES

Angel B.R., Jones J.P.E. & Hall P.L. (1974) Studies of


doped synthetic kaolinite I. Clay Minerals, 10,
247 256.
Balan E., Allard T., Boizot B., Morin G. & Muller J.P.
(1999) Structural Fe3+ in natural kaolinites: new
insights from electron paramagnetic resonance
fitting at X- and Q-band frequencies. Clays and
Clay Minerals, 47, 605 616.
Balan E., Lazzeri M., Saitta A.M., Allard T., Fuchs Y. &
Mauri F. (2005) First-principles study of OHstretching modes in kaolinite, dickite, and nacrite.
American Mineralogist, 90, 50 60.
Belver C., Munoz M.A.B. & Vicente M.A. (2002)
Chemical activation of kaolinite under acid and
alkaline conditions. Chemistry of Materials, 14,
2033 2043.
Bhattacharyya K.G. & Sen Gupta S. (2007) Influence of
activation of kaolinite and montmorillonite on
adsorptive removal of Cd(II) from water. Industrial
& Engineering Chemistry Research, 46, 3734 3742.
Brindley G.W. & Brown G. (1980) Crystal Structures of
Clay Minerals and their Identification.
Mineralogical Society Monograph no. 5.
Mineralogical Society, London.
Brindley G.W., Cao C.C., Harrison J.L., Lipsicas M. &
Raythatha R. (1986) Relation between structural

557

disorder and other characteristics of kaolinites and


dickites. Clays and Clay Minerals, 34, 239 249.
Calas G. (1988) Electron paramagnetic resonance. Pp.
513 571 in: Spectroscopic Methods in Mineralogy
and Geology, (F.C. Hawthorne, editor). Reviews in
Mineralogy, Vol. 18, Mineralogical Society of
America, Washington, DC.
Christidis G.E., Scott P.W. & Dunham A.C. (1997) Acid
activation and bleaching capacity of bentonites from
the islands of Milos and Chios, Aegean, Greece.
Applied Clay Science, 12, 329 347.
Clozel B., Calas G., Muller J.-P., Dran J.-C. & Herve A.
(1990) Kaolinites as dosimeters: a new possibility of
tracing radionuclides migration. Chemical Geology,
84, 259 260.
Clozel B., Allard T. & Muller J.-P. (1994) Nature and
stability of radiation-induced defects in natural
kaolinites: new results and a reappraisal of published
works. Clays and Clay Minerals, 42, 657 666.
Clozel B., Gaite J.-M. & Muller J.-P. (1995) Al-O-Al
paramagnetic defects in kaolinite. Physics and
Chemistry of Minerals, 22, 351 356.
Coyne L.M., Blake D.F., McKeever S.W. & McKeever
S.W.S., editors (1998) Spectroscopic
Characterization of Minerals and their Surfaces.
ACS Symposium Series no. 415, American
Chemical Society, Washington D.C.
Cuttler A.H. (1981) Further studies of a ferrous iron
doped synthetic kaolin: dosimetry of X-ray induced
defects. Clay Minerals, 16, 69 80.
Duzgoren-Aydin N.S., Aydin A. & Malpas J. (2002)
Distribution of clay minerals along a weathered
pyroclastic profile, Hong Kong. Catena, 50, 17 41.
Ekosse G.-I. (2005) Fourier transform infrared spectrophotometry and X-ray powder diffractometry as
complementary techniques in characterizing clay
size fraction of kaolin. Journal of Applied Science &
Environmental Management, 9, 43 48.
Farmer V.C. (1974) The Infrared Spectra of Minerals.
Mineralogical Society, London.
Farmer V.C. (1998) Differing effects of particle size and
shape in the infrared and Raman spectra of kaolinite.
Clay Minerals, 33, 601 604.
Farmer V.C. & Russell J.D. (1964) The infrared spectra
of layered silicates. Spectrochimica Acta, 20,
1149 1173.
Fialips C.-I., Petit S., Decarreau A. & Beaufort D.I
(2000) Influence of synthesis pH on kaolinite
crystallinity and surface properties. Clays and Clay
Minerals, 48, 173 184.
Frost R.L., Mako E., Kristof J., Horvath E. & Kloprogge
J.T. (2001a) Mechanochemical treatment of kaolinite. Journal of Colloid and Interface Science, 239,
458 466.
Frost R.L., Mako E., Kristof J., Horvath E. & J.T.
Kloprogge E. (2001b) Modification of kaolinite
surfaces by mechanochemical treatment. Langmuir,

558

N. Worasith et al.

17, 4731 4738.


Frost R.L., Mako E., Kristof J. & Kloprogge J.T. (2002)
Modification of kaolinite surfaces through mechanichemical treatment
a mid-IR and near-IR
spectroscopic study. Spectrochimica Acta, 58A,
2849 2859.
.
Frost R.L., Horvath E., Mako E., Kristof J. & Redey A
(2003) Slow transformation of mechanically dehydroxylated kaolinite to kaolinite an aged mechanochemically activated formamide-intercalated
kaolinite study. Thermochimica Acta, 408, 103 113.
Frost R.L., Horvath E., Mako E. & Kristof J. (2004)
Modification of low and high defect kaolinite
surfaces: implications for kaolinite mineral processing. Journal of Colloid and Interface Science, 270,
337 346.
Gaite J.-M., Ermakoff P., Allard T. & Muller J.-P.
a sensitive probe for
(1997) Paramagnetic Fe3+
disorder in kaolinite. Clays and Clay Minerals, 45,
496 505.
Gehring A.U., Fry I.V., Luster J. & Sposito G. (1993)
The chemical form of vanadium (IV) in kaolinite.
Clays and Clay Minerals, 41, 662 667.
Gogoi P.K. & Baruah R. (2008) Fluoride removal from
water by adsorption on acid activated kaolinite clay.
Indian Journal of Chemical Technology, 15,
500 503.
Golding R.M., Singhasuwich T. & Tennant W.C. (1977)
An analysis of the conditions for an isotropic gtensor in high-spin d5 systems. Molecular Physics,
34, 1343 1350.
Goodman B.A. & Hall P.L. (1994) Electron paramagnetic resonance spectroscopy. Pp. 173 225 in: Clay
Mineralogy: Physical Determinative Methods (M.J.
Wilson, editor), Chapman & Hall, London.
Goodman B.A., Russell J.D., Fraser A.R. & Woodhams
F.W.D. (1976) A Mossbauer and infra-red spectroscopic study of the structure of nontronite. Clays and
Clay Minerals, 24, 53 59.
Hawthorne F.C., editor (1988) Spectroscopic Methods in
Mineralogy and Geology. Reviews in Mineralogy,
18, Mineralogical Society of America. Washington
D.C.
Hinckley D.N. (1963) Variability in crystallinity
values among the kaolin deposits of the coastal
plain of Georgia and South Carolina. Proceedings of
the 11th National Conference on Clays and Clay
Minerals, 229 235.
Huertas F.J., Chou L. & Wollast R. (1998) Mechanism
of kaolinite dissolution at room temperature and
pressure: Part 1. Surface speciation. Geochimica et
Cosmochimica Acta, 62, 417 431.
Jahn H. & Teller E. (1937). Stability of polyatomic
molecules in degenerate electronic states. I. Orbital
degeneracy. Proceedings of the Royal Society of
London. Series A, Mathematical and Physical
Sciences, 161, 220 235.

Joussein E., Petit S., Churchman J., Theng B., Righi D.


& Delvaux B. (2005) Halloysite clay minerals
a
review. Clay Minerals, 40, 383 426.
Karaoglu M.H., Dogan M. & Alkan M. (2010) Removal
of reactive blue 221 by kaolinite from aqueous
solutions. Industrial & Engineering Chemistry
Research, 49, 1534 1540.
Kuentag C. (2001) Clay deposits in Thailand. Pp. 28 43
in: The 2 nd Workshop on Clays and their
Applications, Department of Mineral Resources,
Thailand.
Kuentag C. & Wasuwanich P. (1978) Clay. Economic
Geology Bulletin No. 19, Economic Geology
Division, Department of Mineral Resources,
Thailand.
Langford J.I. & Wilson A.J.C. (1978). Scherrer after
sixty years. A survey and some new results in the
determination of crystallite size. Journal of Applied
Crystallography, 11, 102-113.
Lombardi G., Russell J.D. & Keller W.D. (1987)
Compositional and structural variations in the size
fractions of a sedimentary and a hydrothermal
kaolin. Clays and Clay Minerals, 35, 321 335.
Mabbs F.E. & Collison D. (1992) Electron paramagnetic resonance of d- transition metal compounds.
Elsevier, Amsterdam.
McBride M.B., Pinnavaia T.J. & Mortland M.M. (1975)
Electron spin relaxation and the mobility of
manganese(II) exchange sites in smectites.
American Mineralogist, 60, 66 72.
Mako E., Frost R.L., Kristof J. & Horvath E. (2001) The
effect of quartz content on the mechanochemical
activation of kaolinite. Journal of Colloid and
Interface Science, 244, 359 364
Martin F., Micoud P., Delmotte L., Marichal C., Le Dred
R., De Parseval P., Mari A., Fortune J.-P., Salvi S.,
Beziat D., Grauby O. & Ferret J. (1999) The
structural formula of talc from the Trimouns deposit,
Pyrenees, France. The Canadian Mineralogist, 37,
997 1006.
Meenakshi S., Sairam Sundaram C. & Sukumar R.
(2008) Enhanced fluoride sorption by mechanochemically-activated kaolinites. Journal of
Hazardous Materials, 153, 164 172.
Mestagh M.M., Vielvoye L. & Herbillon A.J. (1980)
Iron in kaolinite: II. The relationship between
kaolinite crystallinity and iron content. Clay
Minerals, 15, 1 13.
Miller J.G. & Oulton T.D. (1970) Prototrophy in
kaolinite during percussive grinding. Clays and
Clay Minerals, 18, 313 323.
Murray H.H. (2007) Applied Clay Mineralogy:
Occurrences, Processing, and Application of
Kaolins, Bentonites, Palygorskite-Sepiolite, and
Common Clays. Elsevier, Amsterdam.
Nuntiya A. & Prasanphan S. (2006) The rheological
behavior of kaolin suspensions. Chiang Mai Journal

Characterization of modified kaolinite


of Science, 33, 271 281.
Panda A.K., Mishra B.G., Mishra D.K. & Singh R.K.
(2010) Effect of sulphuric acid treatment on physicochemical characteristics of kaolin clay. Colloids and
Surfaces A, 363, 98 104.
Rancourt D.G., Christie I.A.D., Royer M., Kodama H.,
Robert J.-L., Lalonde A.E. & Murad E. (1994)
Determination of accurate [4]Fe3+, [6]Fe3+ and [6]Fe2+
site populations in synthetic annite by Mossbauer
spectroscopy. American Mineralogist, 79, 51 62.
Sengupta P., Saikia N.J., Bharali D.J., Saikia P.C. &
Borthakur P.C. (2006) ESR investigation of deferration treatment of iron-rich kaolinite clay from
Deopani, Assam, India. Current Science, 91, 86-90.
Shannon R.D. (1976) Revised effective ionic radii and
systematic studies of interatomic distances in halides
and chalcogenides, Acta Crystallogriphica A, 32,
751 767.
Steudel A., Batenburg L.F., Fischer H.R., Weidler P.G.
& Emmerich K. (2009) Alteration of non-swelling
clay minerals and magadiite by acid activation.
Applied Clay Science, 44, 95 104.
Temuujin J., Okada K., MacKenzie K.J.D. & Jadambaa
Ts. (2001) Characterization of porous silica prepared
from mechanically amorphized kaolinite by selective
leaching. Powder Technology, 121, 259 262.
Title R.S. (1963) Electron paramagnetic resonance
spectra of Cr+, Mn++, and Fe3+ in cubic ZnS.
Physical Review, 131, 623 627.

559

Vagvolgyi V., Kovacs J., Horvath E., Kristof J. & Mako


E. (2008) Investigation of mechanochemically modified kaolinite surfaces by thermoanalytical and
spectroscopic methods. Journal of Colloid and
Interfac Science, 317, 523 529.
Vallyathan V., Shi X., Dalal N.S., Irr W. & Castranova
V. (1988) Generation of free radicals from freshly
fractured silica dust. American Review of
Respiration Disease, 138, 1213 1219.
van der Marel H.W. & Krohmer P. (1969) O-H
stretching vibrations in kaolinite and related minerals. Contributions to Mineralogy and Petrology, 22,
73 82.
Weeks R.A. (1956) Paramagnetic resonance of lattice
defects in quartz. Journal of Applied Physics, 27,
1376 1381.
Worasith N., Goodman B.A., Jeyashoke N. &
Thiravetyan P. (2011) Decolorization of rice bran
oil using modified kaolin. Journal of the American
Oil Chemists Society. <doi:10.1007/s11746-0111872-2>
Woumfo D., Kamga R., Figueras F. & Njopwouo D.
(2007) Acid activation and bleaching capacity of
some Cameroonian smectite soil clays. Applied Clay
Science, 37, 149 156.
Wu Z., Li C., Sun X., Xu X., Dai B., Li J. & Zhao H.
(2006) Characterization, acid activation and bleaching performance of bentonite from Xinjiang. Chinese
Journal of Chemical Engineering, 14, 253 258.

Potrebbero piacerti anche