Vous êtes sur la page 1sur 9

Chemical Papers 68 (7) 950958 (2014)

DOI: 10.2478/s11696-013-0525-3

ORIGINAL PAPER

Synthesis and characterization of a silylated Brazilian


clay mineral surface
a
a

Mrcia M. F. Silva, a Saloana S. S. Gomes, a Maria G. Fonseca*,


Kaline S. Sousa, a Jos G. P. Espnola, b Edson C. da Silva Filho
a Chemistry
b LIMAV

Department of Paraba Federal University, Jo


ao Pessoa, 58051-970, Paraba, Brazil

Chemistry Department of Piau Federal University, Teresina, 64049-550, Piau, Brazil


Received 12 February 2013; Revised 25 October 2013; Accepted 29 October 2013

Clay mineral containing kaolinite, illite and montmorillonite was organofunctionalized with silylating agents: (3-aminopropyl)triethoxysilane, 3-[2-(2-aminoethylamino)ethylamino]propyl-trimethoxysilane and (3-mercaptopropyl)trimethoxy-silane, to yield three hybrids labelled Clay1, Clay2
and Clay3, respectively. These solids were characterized using elemental analysis, thermogravimetry, X-ray diractometry, infrared spectroscopy, scanning electron micrograph, and 29 Si and 27 Al
solid state NMR. Immobilized quantities of the organic groups were 0.66 mmol g1 , 0.48 mmol g1
and 0.88 mmol g1 for Clayx (x = 13), respectively. X-ray diraction patterns conrmed the immobilization of silanes onto the surface without changes in the textural properties of the clay mineral
as noted from the SEM images. Spectroscopic measurements were in agreement with the covalent
bonding between the silanes and the hydroxyl groups deposited on the surface. The new hybrids
were utilized as adsorbents of cobalt in aqueous solution, with retention values of 0.78 mmol g1 ,
1.1 mmol g1 and 0.70 mmol g1 for Clayx (x = 13), respectively.
c 2013 Institute of Chemistry, Slovak Academy of Sciences

Keywords: clay mineral, silylation, adsorption, inorganicorganic hybrids

Introduction
Clay minerals constitute a large family of important solids with a wide range of technological and academic applications. Their unusual structure and the
presence of silanols on the broken edges, which consists of an arrangement of layers, provides convenient
chemical modication with organic moieties that can
generate multifunctional compounds (Bergaya et al.,
2006). Thus, the new solids obtained are useful for a
range of applications in specic areas such as chromatography, electrochemistry, adsorption and catalysis (Ruiz-Hitzky, 2004).
The success involving chemical procedures is closely related to the silylation processes, which have been
only poorly explored for clay minerals. The silylation
of layered silica was initiated by Ruiz-Hiztky and Rojo
(1980). From the experimental point of view, the use

of silylating agents favours a covalent bond formation


by the displacement of hydrogen in the free hydroxyl
groups on the surface as occurring on kaolinite, and
by the reaction of organosilyl derivatives of precursor
silane. This type of reaction can occur in aqueous and
non-aqueous conditions, the latter being the most effective route for obtaining inorganicorganic hybrids
(da Silva et al., 2007). Thus, the silylation route is
similar to that with silica gel and mesoporous silica,
where several of the organic anchored groups yield a
representative value above 1.0 mmol g1 (Ian et al.,
2007). Further, silylations reactions on synthetic and
natural clay minerals have been proposed (Itagaki &
Kuroda, 2003; Murakami et al., 2004, Alkan et al.,
2005; He et al., 2005; Ishii et al., 2005; Wheeler et
al., 2005; Frost & Mendelovici, 2006; Shanmugharaj
et al., 2006; Wang et al., 2007; Zhu et al., 2007).
Other examples reported silylation of talc in inert at-

*Corresponding author, e-mail: mgardennia@quimica.ufpb.br

Brought to you by | Indian Institute of Technology - Mumbai


Authenticated
Download Date | 9/17/15 4:43 PM

M. M. F. Silva et al./Chemical Papers 68 (7) 950958 (2014)

mosphere with aminated alkoxisilanes for the immobilization of porphyrine (de Faria et al., 2004). Xue
et al. (2006) described the functionalization of palygorskite with dierent silanes and clay minerals with
aminopropyltrimethoxisilane (Ha et al., 2008), whose
new compounds were initially synthesized as reinforcing agents for nanocomposites with polymers. These
materials are important as potential sorbents in the
removal/speciation of various pollutants in waste waters (Ogawa et al., 1998; Fujita et al., 2005; Ide &
Ogawa, 2007; Xue et al., 2010) and they emerge as a
potential new area of research in the waste water treatment technologies (Lee & Tivari, 2012). Silylation of
sodium montmorillonite was performed in glycerol using three aminosilanes of dierent chain length. It was
found that the degree of silylation of sodium montmorillonite increased with the increasing length of the
aminosilane organic moieties, the overall aminosilane
concentration, and temperature. The same benecial
eects were observed on the silicate d-spacing, as its
value increases with the increasing silane concentration and reaction temperature (Piscitelli et al., 2010).
Silylation of montmorillonite surfaces with 3aminopropyltriethoxysilane was carried out in polar
protic and nonpolar solvents. Silylated montmorillonites prepared in nonpolar solvents showed a higher
amount of loaded silane and a higher extent of condensation among dierent silane molecules compared
with those prepared in polar-protic solvents with high
dielectric constant (Su et al., 2013).
Successful silylation of clay mineral surfaces
strongly depends on the reactivity of the clay mineral surfaces, including internal surfaces, external surfaces and broken edges (Bergaya et al., 2006). Phyllosilicate surfaces consist in two basic types, siloxane
and hydroxyl surface. Thus, the 2 : 1 clay minerals
(e.g., smectite group minerals) only contain siloxane
surfaces while the 1 : 1 clay minerals (e.g., kaolinite
group minerals) contain both kinds of surfaces (He et
al., 2013). The hydroxyl surfaces as such Al-octahedral
surface in kaolinite or SiOH in leached montmorillonites are excellent sites for grafting. Therefore, different studies have shown that the broken edges are
the most reactive sites for silane grafting rather than
external and internal surfaces for the 2 : 1 type synthetic swelling clay minerals. For expandable clay minerals, all internal-surface, external surface and broken
edge provide possible sites for silane grafting (He et
al., 2013). However, many of these silylations were developed using a pure clay mineral, so the inuence of
dierent phases on the grafting reaction has not been
explored.
Other relevant aspects are the conditions of the
silylation reaction: nature of the solvent and the silane,
silane concentration, reaction time and temperature of
the system. All these parameters aect the structure
and properties of the grafted materials (Herrera et al.,
2005).

951

In the attempt to contribute to the understanding of the reaction between silanes and clay mineral surfaces containing SiOH and AlOH groups,
the present investigation reports exploring the reactions between three silanes and a Brazilian clay mineral. Physical measurement methods were employed to
characterize the compounds obtained and to evaluate
the new adsorption capacities for cobalt in aqueous
solutions.

Experimental
The Brazilian company Uni
ao Brasileira de Mineracao from the city of Santa Rita, State of Paraiba,
Brazil, supplied the clay mineral sample. Chemical
analyses of the sample were performed by Atomic
Absorption Spectroscopy using a PerkinElmer 5100
Model (PerkinElmer, USA) instrument with an
air-acetylene ame, where the samples were prepared in an acidic mixture (HFHCl). The silylating agents, (3-aminopropyl)triethoxysilane, 3-[2-(2aminoethylamino) ethylamino]propyl-trimethoxysilane and (3-mercaptopropyl)trimethoxysilane (Aldrich,
Germany) and other chemicals used, such as cobalt
nitrate hexahydrate (Fluka, Germany), ethanol, and
xylene (Merck, Brazil), were all of reagent grade. They
were not further puried prior to use.
The clay mineral was initially puried to eliminate
organic impurities. Thus, raw solid (100.0 g) was mechanically stirred for 72 h with 120.0 cm3 of H2 O2 .
The solid was separated by centrifugation, washed
three times with deionized water, and dried at 373 K.
Clay mineral saturated with sodium was obtained after mechanically stirring a sample with
1.0 mol dm3 NaCl for 12 h at 323 K. The solid was
separated by centrifugation and the process was repeated. Finally, the solid was washed with water, dried
in vacuum at 373 K and labelled Clay0.
The clay mineral (Clay0) was initially activated by
heating under vacuum at 393 K for 48 h. Silylation occurred when 5.0 g of activated solid were mixed with
100.0 cm3 of dry xylene and stirred under nitrogen atmosphere at 373 K. Using a syringe, 5.0 cm3 of silane
were added to this suspension and the reaction mixture was maintained under these conditions for 1 h.
Finally, the products were separated by centrifugation at 400 min1 for 20 min, washed with ethanol,
and dried under vacuum at 323 K for further 24 h.
The new hybrids were labelled Clayx (x = 13) for
the reaction products associated with the three silylating agents: (3-aminopropyl)triethoxysilane, 3-[2-(2aminoethylamino)ethylamino]propyl-trimethoxysilane and (3-mercaptopropyl)trimethoxysilane, respectively. Reproducibility of these reactions was checked
using a larger quantity of silylating agents to successfully yield the same products.
X-ray diraction (XRD) patterns were obtained on
a Shimadzu XD3A model diractometer (Shimadzu,

Brought to you by | Indian Institute of Technology - Mumbai


Authenticated
Download Date | 9/17/15 4:43 PM

952

M. M. F. Silva et al./Chemical Papers 68 (7) 950958 (2014)

Table 1. Contents of carbon, hydrogen, sulphur, nitrogen and the amount of organic groups (q) and calculated (and theoretical)
carbon to nitrogen and carbon to sulphur molar ratios for hybrids (Clayx)
C
Clay

Clay1
Clay2
Clay3

% (mmol g1 )
2.30 (1.92)
3.36 (2.80)
5.21 (4.34)

1.31
1.73
1.47

0.92 (0.66)
1.33 (0.95)

q
mmol g1

2.80 (0.88)

Japan) equipped with monochromatic CuKa operating at 40 kV and 30 mA. The diraction patterns
were recorded from 1.4 to 70 with the scan rate of
0.67 s1 .
FTIR spectra of samples dispersed in KBr disks
were recorded at room temperature on a Bomem spectrometer (MB series) (ABB Bomem, Canada) over the
range of 4000400 cm1 at the resolution of 4 cm1
and 30 scans for each run.
Absorption atomic measurements for cobalt were
obtained on a PerkinElmer instrument 5100 model
(PerkinElmer, USA) with an air-acetylene ame.
Thermal analyses were carried out using a DuPont
model 1090 B (DuPont, USA) thermogravimetric apparatus coupled to a thermobalance 951 heated to
1273 K at the heating rate of 10 C min1 in the atmosphere of dry nitrogen. The samples varied in weight
from 15.0 mg to 30.0 mg.
The surface area measurements were obtained on a
Flowsorb II 2300 Micromeritcs instrument (Micromeritcs, USA). Surface areas were determined using BET
equations.
Carbon, nitrogen, sulphur, and hydrogen quantities were determined using a PerkinElmer PE-2400
microelemental analyzer (PerkinElmer, USA).
29
Si NMR measurements were obtained on an AC
300/P Bruker spectrometer (Bruker, Germany) at
room temperature. The measurements were recorded
at the frequency of 59.6 MHz using the HPDEC
technique with pulse repetition duration of 60 s and
the pulse width of 45 . The PeakFit V4.12 software
was employed for deconvolution using tted GaussianLorentzian line shapes. 27 Al spectra were obtained at
the frequency of 78.39 MHz and the magic angle spinning (MAS) of 4 kHz.
Scanning electron microscopy images were used
to study the morphology. The clays, being nonconducting materials, were examined by sputter coating samples with conducting layers of gold and carbon (Plasma Science, USA). Secondary electrons were
detected using a microscope Jeol JSTM-300 (Jeol,
Japan).
A series of samples of 50 mg of each solid with
20.0 cm3 of 0.01 mol dm3 cobalt solution were stirred
for 2 h at (298 1) K. Then, the samples were centrifuged and the cation quantities in the initial and
nal solutions were analyzed by atomic absorption

0.66
0.48
0.88

C:N

C:S

2.89 (3.0)
2.95 (3.0)

4.93 (5.0)

spectroscopy. Adsorption capacity (Nads ) was calculated as mmol of cation per gram of modied solid
using the expression Nads = (Ni Ns )/m, where Ni
and Ns are the initial and nal amounts of the Co2+
in the solution, respectively, and m is the mass of
the solid. In a separate control experiment, the same
procedure was applied for the precursor clay mineral
(Clay0).

Results and discussion


Total chemical composition of the pristine mineral clay (mass %) was: SiO2 (48.1), Al2 O3 (25.5),
Fe2 O3 (2.17), MgO (2.24) and K2 O (2.48) and heat
loss (19.51). From the present data, clay mineral composition can be considered as rich in aluminum. Identication of all constituents was done by XRD, where
a series of three peaks at 2 = 8.85, 12.5, and 6.96
were indexed to (001) plan of illite, kaonilite, and
montmorillonite (Brindely & Brown, 1980; Moore &
Reynolds, 1997). The specic surface area calculated
by the BET method was 25 m2 g1 .
The amount of organic moieties was determined
by the CHNS elemental analysis (Table 1). Based on
their nitrogen and sulphur percentages, the number
of pendant moles of organic groups in hybrids (q)
was calculated as 0.66 mmol g1 , 0.48 mmol g1 and
0.88 mmol g1 for Clayx (x = 13), respectively. Likewise, the values of the carbon to nitrogen and carbon to sulphur ratios (calculated and theoretical) were
very close to the predicted values indicating that organic chains were maintained during the immobilization.
The values of q and the surface area of 25 m2 g1 ,
were used to estimate the degree of modication
by silylating agents considering the calculated value
of SiOH density of 5 nm2 (SiOH or AlOH)
(Hermosn & Cornejo, 1986). Thus, the quantities of
molecules per nm2 in the clay mineral surface were
16, 11.6 and 21.2 for Clay1, Clay2 and Clay3, respectively. These data illustrate that the grafting was more
eective when using the mercapto compared to the
aminosilanes.
X-ray diraction patterns for the precursor (Fig. 1)
showed reections at 2 = 6.96 , 8.85 and 12.5 indexed to (001) plan, which corresponds to the interlayer distance of 1.26 nm, 1.0 nm and 0.710 nm. Eval-

Brought to you by | Indian Institute of Technology - Mumbai


Authenticated
Download Date | 9/17/15 4:43 PM

M. M. F. Silva et al./Chemical Papers 68 (7) 950958 (2014)

Fig. 1. XRD patterns of Clay0 and hybrids: Clay1, Clay2, and


Clay3. M, K, I and Q are montmorillonite, kaolite, illite
and quartz, respectively.

Table 2. Estimation of the ratios of the intensity of basal peak


of the desired (montmorillonite, illite and kaolinite)
and the quartz phase (Iquartz /Iclay )
Iquartz /Iclay
Clay

Clay0
Clay1
Clay2
Clay3

Montmorillonite

Illite

Kaolinite

4.2
7.1
6.8
5.1

3.4
3.4
3.4
3.5

2.4
2.4
2.5
2.5

uation of the crystanillity was accomplished by the


ratios of the intensity of the basal peak of the desired
phase (montmorillonite, illite and kaolinite) and the
quartz one, (Iquartz /Iclay ) as shown by the data in Table 2. A signicant change in the ratios was observed
for montmorillonite, suggesting an alteration in the
original crystallinity of the inorganic matrix during
the silylation process. On the other hand, changes of

953

reections localized at 2 = 8.85 and 12.5 related to


the illite and kaolinite phases were observed, implying
possible immobilization of silanes onto their surfaces.
The relative reduction of the reection at 6.96 for
montmorillonite suggests alteration in the organization of the solid without any alteration in the basal
spacing, as shown for Clay3. For Clay1 and Clay2, an
increase in the basal spacing occurred for 1.88 nm and
2.03 nm, (Figs. 1b and 1c), indicating possible entry of
silanes into the interlayer region of montmorillonite.
A comparison of thermogravimetric curves between raw clay mineral and the grafted ones shows
a lot of changes. Thermogravimetric curves for the
extent of mass loss vary considerably (Fig. 2). The
temperature intervals of thermal decomposition were
based on DTG curves and attributions as obtained
from literature (Balek & Murat, 1996; Hedley et al.,
2007). For the precursor, two weight losses were detected: the rst one in the 298347 K region was attributed to the loss of 1.7 % of physisorbed water, the
second corresponding to 6.0 % in the 6661000 K interval was assigned to the displacement of water in the
interlayer space which is also related to the dehydroxilation of structural OH groups. For Clay1 and Clay2,
the weight loss of physisorbed water occurred in the
298400 K region, whereas the oxidation of the organic matter and dehydroxylation occurred between
4001072 K. The weight losses of about 10.5 % and
13.5 % were attributed to the coordination of water,
dehydroxilation of the remaining structural OH, and
degradation of the organic moieties anchored onto the
surface. Clay3 required a more complex thermal decomposition process, involving four steps corresponding to the loss of: i) 1.7 % of adsorbed water between
300370 K, ii) 4.3 % of coordination water and partial
organic moieties between 389660 K, iii) 4.2 % due to
the loss of the reaming organic groups in the 727
783 K region and iv) 5.3 % related to the dehydroxilation of structural OH groups in the 7971101 K re-

Fig. 2. Thermogravimetric curves for original and hybrid solids (a) and TG and DTG curves for Clay3 (b).

Brought to you by | Indian Institute of Technology - Mumbai


Authenticated
Download Date | 9/17/15 4:43 PM

954

M. M. F. Silva et al./Chemical Papers 68 (7) 950958 (2014)

Fig. 3. Infrared spectra of Clay0 and hybrids: Clay1, Clay2, and Clay3, in the 40002070 cm1 (a) and 1300400 cm1 (b) regions.

gion. However, as the pendant organic chain increased


in size, the corresponding increase in the mass losses
detected was due to the quantity of such anchored
moieties bonded in the inorganic polymer; these data
are in agreement with the CHNS elemental analysis.
Infrared spectra of the precursor clay mineral and
hybrids (Fig. 3) presented the same set of absorption
bands related to the inorganic structure of the solids
(Tyagi et al., 2006). However, the hybrids showed
changes in the intensities of these bands and their
displacement at lower frequencies. In this case, the
bands were attributed to the structural AlOH at
3672 cm1 and 3651 cm1 for the precursor clay
changes to 3660 cm1 and 3620 cm1 for all hybrids.
The adsorbed water obscured the SiOH stretching
bands from illite and montmorillonite. The characteristic bands associated with aliphatic moieties appeared at 2930 cm1 and 2859 cm1 and were attributed to asymmetric and symmetric CH stretching vibration modes, respectively (Zhang et al., 2007).
Another absorption bands occurred at: 1080 cm1 and
1000 cm1 due to inplan and outplan of (SiOSi),
912 cm1 (AlOH and SiOH), 793 cm1 due to
(AlOSi) and 456 cm1 due to (SiO) (van der
Marel & Beutelspacher, 1976).
The 29 Si NMR spectrum detected only one peak
at 92 ppm related to Q3 , Si*(OH)(OSi)3 , species for
the precursor solid (Fig. 4a) and a very small peak
at 84 ppm of very low intensity assigned to Q1 ,
(SiO)Si*(OH)3 , species due to isolated SiOH on
the external surface of the inorganic framework (ElNahhal & El-Ashgar, 2007). The presence of two new
signals at 68 and 57 ppm was detected. The rst
one was attributed to the RSi (OSi)3 , T3 , species,
where R is the organic mercaptopropyl pendant chain
bonded to the inorganic framework (Jaber et al., 2002;
Gallgo et al., 2008). The second peak was attributed

Table 3. Amount of silicon species estimated by the 29 Si NMR


spectra of the pristine and modied clay minerals
Clay

Q3 /%

T2 /%

T3 /%

Clay0
Clay1
Clay2
Clay3

100
74
94
67

14

19

12
6
14

to silicon in the RSi (OSi)2 (OH), T2 , species.


For Clay2, only one peak at 68 ppm due to the
T3 species was detected (Sindorf & Maciel, 1983).
The relative amount of silicon species was estimated
by calculating the peak area of each species; the results are listed in Table 3. In agreement with the obtained percentages for the silicon species it was observed that Clay3 has the largest degree of immobilization (considering the quantity of the T2 and
T3 species) when compared with the other two hybrids. The reaction with aminosilane was strongly dependent on the chain size, as illustrated by the lowest degree of condensation for Clay2, where the T3
species were estimated to be 6 %. Therefore, steric
hindrance associated with the length of the organic
chain aects the accessibility of the silane to the surface.
27
Al NMR spectra (Fig. 4b) showed two types of
the Al species: the rst one at 2.0 ppm was attributed
to Al in octahedral sites present in the phases of kaolinite, montmorrilonite and illite and the second one at
58 ppm and 69 ppm related to tetrahedral Al of both
montmorillonite and illite/montmorillonite (Lipsicas
et al., 1984; Rocha, 1999). The last attribution is based
on previous studies, where it was established that
the 65 ppm distance typically separates tetrahedral

Brought to you by | Indian Institute of Technology - Mumbai


Authenticated
Download Date | 9/17/15 4:43 PM

M. M. F. Silva et al./Chemical Papers 68 (7) 950958 (2014)

Fig. 4.

955

29 Si

(a) and Al27 (b) solid state NMR spectra for the precursor and the silylated clays. Asterisks (*) denotes spinning side
bands.

Fig. 5. Proposed structures when aminopropyl (a), propylethylenediamino (b), and mercaptopropyl (c) groups are bonded onto
Clay1, Clay2, and Clay3, respectively.

A1 and octahedral sites in illite/smectite (Schroeder,


1993).
After the reactions with silanes, neither peak
showed alteration in the absorption suggesting unchanged chemical surroundings of aluminium after the
reactions for all solids.
The set of characterizations is in agreement with
the covalent immobilization of silanes onto the surface
edges of the clay minerals as shown in Fig. 5. The

covalently bound organic groups exhibit either amino


or thiol functions able to adsorb heavy metal cations
from solutions.
The scanning electron microscopy images (Fig. 6)
show the typical sheet morphology as visualized for
the precursor (Fig. 6a). For the hybrids, perceptible
changes in the textural properties were not observed
and the original arrangement was maintained even after the silane immobilization. The absence of poly-

Brought to you by | Indian Institute of Technology - Mumbai


Authenticated
Download Date | 9/17/15 4:43 PM

956

M. M. F. Silva et al./Chemical Papers 68 (7) 950958 (2014)

Fig. 6. SEM images for pristine clay (a), and hybrids: Clay1 (b), Clay2 (c) and Clay3 (d).

Fig. 7. Isotherms of divalent cobalt retention in aqueous solution at (298 1) K for the hybrids. Cs is the equilibrium
cation concentration in aqueous solutions and Nads is
the cobalt retention quantity.

merized silica is not evidenced as it is suggested by


the presence of the pure phase, which is supported by
the X-ray data.
The desired functionality assigned to the pendant

group of the organic moiety was designed taking into


account the assessment of the adsorption capacity of
the solid/liquid interface. This behaviour is to be expected because complexes can be formed on the surface by the basic nitrogen or sulphur of the organic
groups (Fonseca et. al., 2001). Thus, the three hybrids presented a signicant adsorption (a) in the order a(Clay2) > a(Clay3) > a(Clay1). The original clay
mineral gave an irrelevant adsorption under the same
experimental conditions. The maximum retention capacity (Nads ) values for cobalt were 0.78 mmol g1 ,
1.1 mmol g1 and 0.70 mmol g1 for Clay1, Clay2 and
Clay3, respectively. The highest value found for Clay2
relates to the available double nitrogen basic nuclei, as
also conrmed by the data provided by the elemental
analyses. The lower value for Clay3 is associated with
the softness of sulphur, which is more favourable for
the interaction with soft acids. The set of isotherms for
all solids showed the same shape (Fig. 7), suggesting
the same type of interactions forming complexes between cobalt and free Lewis basic centres. The nature
of the complexes formed was determined by relating

Brought to you by | Indian Institute of Technology - Mumbai


Authenticated
Download Date | 9/17/15 4:43 PM

M. M. F. Silva et al./Chemical Papers 68 (7) 950958 (2014)

the amount of maximum adsorbed cobalt, nitrogen,


and sulphur to the modied matrixes. It was observed
that the N : Co and S : Co atoms were 1 : 1, using the equilibrium concentration for the three hybrids
showed the formation of mono dentate complexes on
the solids in all cases.

Conclusions
A pristine clay mineral rich in aluminum phyllosilicates was used as a support for covalent immobilization of silanes indicating their eective incorporation
onto the surface. The obtained solids acted as good
adsorbents of cobalt from aqueous solutions, with a
better absorption performance than that of the original matrix. This new characteristic was associated
with the presence of organic moieties on their surface, which are free to interact with acidic nuclei at
the solid-liquid interface. All characterization results
conrmed the eectiveness of the reaction with a covalent bond formation between the inorganic polymers and each silylating agent, showing similar mechanisms, which is independent of the type of silane employed. However, the amount deposited is highly dependent on the size of the organic chain associated
with the depositing agent. The success of such a surface chemical modication is related to the similarity
of structural AlOH with SiOH, although the reaction involving silanol groups is quantitatively more
eective.
Acknowledgements. The authors thank CNPq and CAPES
for nancial support and fellowships.

References
Alkan, M., Tekin, G., & Namli, H. (2005). FTIR and zeta potential measurements of sepiolite treated with some organosilanes. Microporous and Mesoporous Materials, 84, 7583.
DOI: 10.1016/j.micromeso.2005.05.016.
Balek, V., & Murat, M. (1996). The emanation thermal analysis of kaolinite clay minerals. Thermochimica Acta, 282283,
385397. DOI: 10.1016/0040-6031(96)02886-9.
Bergaya, F., Theng, B. K. G., & Lagaly, G. (2006). Handbook
of clay science. Oxford, UK: Elsevier.
Brindely, G. W., & Brown, G. (1980). Crystal structures of clay
minerals and their X-ray identication. London, UK: Mineralogical Society.
da Silva, O. G., da Fonseca, M. G., & Arakaki, L. N. H. (2007).
Silylated calcium phosphates and their new behavior for copper retention from aqueous solution. Colloids and Surfaces
A: Physicochemical and Engineering Aspects, 301, 376381.
DOI: 10.1016/j.colsurfa.2006.12.072.
de Faria, A. L., Airoldi, C., Doro, F. G., Fonseca, M. G., & Assis, M. D. (2004). Anchored ironporphyrins the role of talcaminofunctionalyzed phyllosilicates in the catalysis of oxidation of alkanes and alkenes. Applied Catalysis A: General,
268, 217226. DOI: 10.1016/j.apcata.2004.03.035.
El-Nahhal, I. M., & El-Ashgar, N. M. (2007). A review on
polysiloxane-immobilized ligand systems: Synthesis, characterization and applications. Journal of Organometallic Chemistry, 692, 28612886. DOI: 10.1016/j.jorganchem.
2007.03.009.

957

Fonseca, M. G., Oliveira, A. S., & Airoldi, C. (2001). Silylating


agents grafted onto silica derived from leached chrysotile.
Journal of Colloid and Interface Science, 240, 533538. DOI:
10.1006/jcis.2001.7663.
Frost, R. L., & Mendelovici, E. (2006). Modication of brous
silicates surfaces with organic derivatives: An infrared spectroscopic study. Journal of Colloid and Interface Science,
294, 4752. DOI: 10.1016/j.jcis.2005.07.014.
Fujita, I., Kuroda, K., & Ogawa, M. (2005). Adsorption of alcohols from aqueous solutions into a layered silicate modied
with octyltrichlorosilane. Chemistry of Materials, 17, 3717
3722. DOI: 10.1021/cm048023q.
Gallgo, J. C., Jaber, M., Mieh-Brendl, J., & Marichal,
C. (2008). Synthesis of new lamellar inorganicorganic talc
like hybrids. New Journal of Chemistry, 32, 407412. DOI:
10.1039/b713004j.
Ha, S. R., Rhee, K. Y., Kim, H. C., & Kim, J. T. (2008). Fracture performance of clay/epoxy nanocomposites with clay
surface-modied using 3-aminopropyltriethoxysilane. Colloids and Surfaces A: Physicochemical and Engineering
Aspects, 313314, 112115. DOI: 10.1016/j.colsurfa.2007.04.
082.
He, H. P., Duchet, J., Galy, J., & Gerard, J. F. (2005). Grafting
of swelling clay materials with 3-aminopropyltriethoxysilane.
Journal of Colloid and Interface Science, 288, 171176. DOI:
10.1016/j.jcis.2005.02.092.
He, H. P., Tao, Q., Zhu, J. X., Yuan, P., Shen, W., & Yang, S.
Q. (2013). Silylation of clay mineral surfaces. Applied Clay
Science, 71, 1520. DOI: 10.1016/j.clay.2012.09.028.
Hedley, C. B., Yuan, G., & Theng, B. K. G. (2007). Thermal
analysis of montmorillonites modied with quaternary phosphonium and ammonium surfactants. Applied Clay Science,
35, 180188. DOI: 10.1016/j.clay.2006.09.005.
Hermosn, M. C., & Cornejo, J. (1986). Methylation of sepiolite
and palygorskite with diazomethane. Clays and Clay Minerals, 34, 591596. DOI: 10.1346/ccmn.1986.0340514.
Herrera, N. N., Letoe, J. M., Reymond, J. P., & BourgeatLami, E. (2005). Silylation of laponite clay particles with
monofunctional and trifunctional vinyl alkoxysilanes. Journal of Materials Chemistry, 15, 863871. DOI: 10.1039/
b415618h.
Ian, P. B., Blitz, J. P., Gunko, V. M., & Sheeran, D. J. (2007).
Functionalized silicas: Structural characteristics and adsorption of Cu(II) and Pb(II). Colloids and Surface A, 307, 83
92. DOI: 10.1016/j.colsurfa.2007.05.016.
Ide, Y., & Ogawa, M. (2007). Interlayer modication of a layered titanate with two kinds of organic functional units for
molecule-specic adsorption. Angewandte Chemie International Edition, 46, 84498451. DOI: 10.1002/anie.200702360.
Ishii, R., Nakatsuji, M., & Ooi, K. (2005). Preparation of highly
porous silica nanocomposites from clay mineral: A new approach using pillaring method combined with selective leaching. Microporous and Mesoporous Materials, 79, 111119.
DOI: 10.1016/j.micromeso.2004.10.033.
Itagaki, T., & Kuroda, K. (2003). Organic modication of the
interlayer surface of kaolinite with propanediols by transesterication. Journal of Material Chemistry, 13, 10641068.
DOI: 10.1039/b211844k.
Jaber, M., Miehe-Brendle, J., & Le Dred, R. (2002). Mercaptopropyl AlMg phyllosilicate: Synthesis and characterization
by XRD, IR and NMR. Chemistry Letters, 31, 954955. DOI:
10.1246/cl.2002.954.
Lee, S. M., & Tiwari, D. (2012). Organo and inorganoorganomodied clays in the remediation of aqueous solutions:
An overview. Applied Clay Science, 5960, 84102. DOI:
10.1016/j.clay.2012.02.006.
Lipsicas, M., Raythatha, R. H., Pinnavaia, T. J., Johnson, I. D.,
Giese, R. F., Jr., Costanzo, P. M., & Robert, J. L. (1984). Sil-

Brought to you by | Indian Institute of Technology - Mumbai


Authenticated
Download Date | 9/17/15 4:43 PM

958

M. M. F. Silva et al./Chemical Papers 68 (7) 950958 (2014)

icon and aluminium site distributions in 2 : 1 layered silicate


clays. Nature, 309, 604607. DOI: 10.1038/309604a0.
Moore, D. M., & Reynolds, R. C., Jr. (1997). X-Ray diraction
and the identication and analysis of clay minerals. Oxford,
UK: Oxford University Press.
Murakami, J., Itagaki, T., & Kuroda, K. (2004). Synthesis of
kaolinite-organic nanohybrids with butanediols. Solid State
Ionics, 172, 279284. DOI: 10.1016/j.ssi.2004.02.048.
Ogawa, M., Okutomo, S., & Kuroda, K. (1998). Control of interlayer microstructures of a layered silicate by surface modication with organochlorosilanes. Journal of the American
Chemical Society, 120, 73617362. DOI: 10.1021/ja981055s.
Piscitelli, F., Posocco, P., Toth, R., Fermeglia, M., Pricl, S.,
Mensitieri, G., & Lavorgna, M. (2010). Sodium montmorillonite silylation: Unexpected eect of the aminosilane chain
length. Journal of Colloid and Interface Science, 351, 108
115. DOI: 10.1016/j.jcis.2010.07.059.
Rocha, J. (1999). Single and triple-quantum 27 Al MAS NMR
study of the thermal transformation of kaolinite. The
Journal of Physical Chemistry B, 103, 98019804. DOI:
10.1021/jp991516b.
Ruiz-Hitzky, E., & Rojo, J. M. (1980). Intracrystalline grafting
on layer silicic acids. Nature, 287, 2830. DOI: 10.1038/2870
28a0.
Ruiz-Hitzky, E. (2004). Organicinorganic materials: From intercalation chemistry to devices. In P. Gmez-Romero, &
C. Snchez (Eds.), Functional hybrid materials (pp. 1549).
Weinheim, Germany: Wiley. DOI: 10.1002/3527602372.ch2.
Schroeder, P. A. (1993). A chemical, XRD and 27 A1 MAS NMR
investigation of miocene gulf coast shales with application
to understanding illite/smectite crystal-chemistry. Clays and
Clay Minerals, 41, 668679. DOI: 10.1346/ccmn.1993.0410
605.
Shanmugharaj, A. M., Rhee, K. Y., & Ryu, S. H. (2006).
Inuence of dispersing medium on grafting of aminopropyltriethoxysilane in swelling clay materials. Journal
of Colloid and Interface Science, 298, 854859. DOI:
10.1016/j.jcis.2005.12.049.
Sindorf, D. W., & Maciel, G. E. (1983). Solid-state NMR
studies of the reactions of silica surfaces with polyfunctional chloromethylsilanes and ethoxymethylsilanes. Journal
of the American Chemical Society, 105, 37673776. DOI:
10.1021/ja00350a003.

Su, L., Tao, Q., He, H. P., Zhu, J. X., Yuan, P., & Zhu, R. L.
(2013). Silylation of montmorillonite surfaces: Dependence
on solvent nature. Journal of Colloid and Interface Science,
391, 1620. DOI: 10.1016/j.jcis.2012.08.077.
Tyagi, B., Chudasama, C. D., & Jasra, R. V. (2006). Determination of structural modication in acid activated montmorillonite clay by FT-IR spectroscopy. Spectrochimica Acta
Part A: Molecular and Biomolecular Spectroscopy, 64, 273
278. DOI: 10.1016/j.saa.2005.07.018.
van der Marel, H. W., & Beutelspacher, H. (1976). Atlas of
infrared spectroscopy of clay minerals and their admixtures.
New York, NY, USA: Elsevier.
Wang, S. F., Lin, M. L., Shieh, Y. N., Wang, Y. R., &
Wang, S. J. (2007). Organic modication of synthesized
clay-magadiite. Ceramics International, 33, 681685. DOI:
10.1016/j.ceramint.2005.12.005.
Wheeler, P. A., Wang, J. Z., Baker, J., & Mathias, L. J. (2005).
Synthesis and characterization of covalently functionalized
laponite clay. Chemistry of Materials, 17, 30123018. DOI:
10.1021/cm050306a.
Xue, S. Q., Reinholdt, M., & Pinnavaia, T. J. (2006). Palygorskite as an epoxy polymer reinforcement agent. Polymer,
47, 33443350. DOI: 10.1016/j.polymer.2006.03.036.
Xue, A. L., Zhou, S. Y., Zhao, Y. J., Lu, X. P., & Han, P.
F. (2010). Adsorption of reactive dyes from aqueous solution
by silylated palygorskite. Applied Clay Science, 48, 638640.
DOI: 10.1016/j.clay.2010.03.011.
Zhang, B., Li, Y. F., Pan, X. B., Jia, X., & Wang, X.
L. (2007). Intercalation of acrylic acid and sodium acrylate into kaolinite and their in situ polymerization. Journal of Physics and Chemistry of Solids, 68, 135142. DOI:
10.1016/j.jpcs.2006.09.020.
Zhu, L. H., Tian, S. L., Zhu, J. X., & Shi, Y. (2007). Silylated
pillared clay (SPILC): A novel bentonite-based inorgano
organo composite sorbent synthesized by integration of pillaring and silylation. Journal of Colloid and Interface Science, 315, 191199. DOI: 10.1016/j.jcis.2007.06.053.

Brought to you by | Indian Institute of Technology - Mumbai


Authenticated
Download Date | 9/17/15 4:43 PM

Vous aimerez peut-être aussi