ArticlePDF Available

Abstract and Figures

This study aims to synthesize and characterize organoclays developed from an Argentinian montmorillonite (Bent) using hexadecyltrimethylammonium bromide (HDTMA-Br) as the intercalation agent. Subsequently, an adsorption mechanism is proposed. The obtained organoclays were more hydrophobic than the starting clay. Surfactant molecules were adsorbed initially through cation exchange in sites placed in the interlayer space of the clay. Adsorption in such sites continued until the interlayer space was saturated. Depending on the surfactant loading introduced during the intercalation process, different organizations of surfactant in the interlayer were obtained. Further adsorption of surfactant occurred in the mesopores generated by tactoids in the “house of cards” organization. This process kept surfactant molecules relatively free and out of the interlayer space.
Content may be subject to copyright.
Quim. Nova, Vol. XY, No. 00, 1-6, 200_
Artigo
doi number
*e-mail: pnaranjo@unsa.edu.ar
SYNTHESIS AND CHARACTERIZATION OF HDTMA-ORGANOCLAYS: INSIGHTS INTO THEIR
STRUCTURAL PROPERTIES
Pablo M. Naranjoa,*, José Molinab, Edgardo Ling Sham a,c and Elsa M. Farfán Torresa,b
aInstituto de Investigaciones para la Industria Química, Consejo Nacional de Investigaciones Científicas y Técnicas, Av. Bolivia
5150 (A4408FVY) Salta, Argentina
bFacultad de Ciencias Exactas, Universidad Nacional de Salta, A4408FVY Salta, Argentina
cFacultad de Ingeniería, Universidad Nacional de Salta, A4408FVY Salta, Argentina
Recebido em 13/03/2014; aceito em 16/09/2014; publicado na web em 13/11/2014
This study aims to synthesize and characterize organoclays developed from an Argentinian montmorillonite (Bent) using
hexadecyltrimethylammonium bromide (HDTMA-Br) as the intercalation agent. Subsequently, an adsorption mechanism is proposed.
The obtained organoclays were more hydrophobic than the starting clay. Surfactant molecules were adsorbed initially through cation
exchange in sites placed in the interlayer space of the clay. Adsorption in such sites continued until the interlayer space was saturated.
Depending on the surfactant loading introduced during the intercalation process, different organizations of surfactant in the interlayer
were obtained. Further adsorption of surfactant occurred in the mesopores generated by tactoids in the “house of cards” organization.
This process kept surfactant molecules relatively free and out of the interlayer space.
Keywords: nanoclay; surfactant; adsorption; cation exchange.
INTRODUCTION
Organic Pillared clays, also known as organoclays or nanoclays,
are obtained by cation exchange of interlayer cations, component
of 2:1 type clay minerals, with quaternary ammonium cations. This
process leads to the formation of materials with hydrophobic charac-
teristics that can be applied in a variety of processes, such as catalytic
processes, rheological control agents in paintings and lubricants;
polymer and plastics matrix reinforcing; adsorbents for effluent
treatments, oil spilling, releasing active matrix, etc.1-6
The application of these materials depends strongly on the ob-
tained structure, which, at the same time, conditions its properties.
Therefore, understanding the relation between both the method and
synthesis parameters used, and the developed structure, is of the
utmost importance.
Among variables that affect the synthesis process, the properties
of the employed clay must be considered: chemical composition,
cation exchange capacity (CEC), charge density, specific surface
area, etc.; and also those related to the synthesis process itself: kind
and concentration of surfactant, time of contact, temperature, etc.
Some authors have demonstrated that the structure and properties
developed by these materials are correlated directly with the organiza-
tion of organic cations in the interlayer space,3 showing the importance
of determining the type of arrangement developed in organoclays.
The purpose of this work was to synthesize and characterize
organoclays from an Argentinian montmorillonite (Bent) with he-
xadecyltrimethylammonium bromide (HDTMA-Br) as intercalation
agent. The influence of the amount of intercalated hexadecyltri-
methylammonium cation (HDTMA+) characterizing the obtained
products was evaluated. Their structures were correlated with models
of distribution and ordering of HDTMA+ molecules intercalated in
the intra and extra layer space.
EXPERIMENTAL
Materials
The used clay was an Argentinian bentonite sample (Bent) from
Rio Negro Province, with mineralogy: 84% Na-montmorillonite,
4% quartz and 12% feldspars, determined by Rietveld method;7 the
CEC was 0.837 mol kg-1, determined by the Cu(EDA)22+ method.8
HDTMA-Br was supplied by Merck, (99%), used without further
purification. All solutions and dispersions were prepared using
deionized water obtained with SQC 3 Reverse Osmosis, from Water
Factory System® equipment.
Samples preparation
Organoclays were obtained by putting in contact a 5% w/v sus-
pension of starting clay with a determined volume of HDTMA-Br
5 × 10-3 mol L-1 solution for 2 h with orbital shaking at room
temperature.“Surfactant loading” (SL) is defined as the x fraction of
the CEC that is replaced by HDTMA+:
SL = (mol HDTMA+ / mol exchange cations) CEC = x CEC
To obtain different SL, different volumes of HDTMA-Br
5 × 10-3 mol L-1 solution were used. Solid products were isolated by
centrifugation and repeatedly washed with distilled water until they
were free of bromide ions (determined by AgNO3 test). Furthermore,
the solids were dried at 60 ºC over night, milled in agate mortar, kept
under conditions of controlled humidity, and denoted as Bent followed
by the SL (Bent-0.4; Bent-0.8, etc.).
Desorption of the surfactant was analyzed through leaching
tests. A suspension of each organoclay was prepared in distilled
water and then shaken for 0.5 h on an orbital shaker at room tem-
perature, after which it was separated by centrifugation. The leach-
ing process was repeated “n” times. Solids obtained in this way are
indicated with the name of the initial organoclay followed by “Ln”
(Bent-0.4-L3, Bent-0.8-L7, etc).
Naranjo et al.
2Quim. Nova
Materials characterization
Differential thermal analysis (DTA) was performed with a
Rigaku TAS 1100 from room temperature to 1200 °C, using about
20 mg samples, with a heat temperature rate of 20 °C min-1 in static
air atmosphere.
X-ray diffraction (XRD) studies were obtained on oriented sam-
ples by spreading the sample suspension on glass slides and further
drying (48 h) at room temperature with a relative humidity of 0.47.
Analyses were performed using a Philips PW 1710 diffractometer
using CuKα radiation (0.154 nm). Measuring conditions were: a
power supply of 40 kV and 30 mA; 1° (2θ) divergence and detector
slits, 0.02° (2θ) step size, counting time of 10 s step-1 and patterns
collected from 3° to 70° (2θ). In cases where overlapping peaks
were detected, mathematical deconvolutions were performed with
OriginPro 8 software. The positions and peak areas were calculated
using the Gaussian functions, applying a fitting algorithm for non-
linear least squares.
KBr pressed discs of dried montmorillonite and organoclays
with a sample to KBr relationship of 1:100 approximately, ground
in agate mortar and pressed at a pressure of 3 ton, were analyzed by
FTIR in a Spectrum GX Perkin Elmer Infrared spectrometer, between
4000 and 400 cm-1.
N2 adsorption isotherms were obtained on a Micromeritics sorp-
tometer ASAP 2020 V3.03 E, at -196 °C, previously outgassing solids
at 100 ºC for 10 h (or until a pressure lower than 8 × 10-6 mmHg was
reached). BET surface was calculated with at least 5 points with a
minimum linear correlation coefficient of 0.999. Micropore Volume
(VμP) was calculated with the t-plot method using Harkins-Jura-Boer
equation; Total Pore Volume (VTP) by Gurvitch method, interpolat-
ing at p/p0 = 0.98; and the Mesopore Volume (VmP) by the difference
between VTP and VμP. Pore size distribution was conducted using
Halsey equation over desorption branch.
Using Mopac 2009 software, a computational simulation of the
surfactant molecule in all-trans conformation was performed. A semi-
empirical method with a PM6 base was used for its optimization. In
this way, molecular dimensions and surface area were calculated.
Thickness values for the multilayer structures were also calculated
through techniques of computational calculation.
RESULTS AND DISCUSSION
TGA
Thermal analysis of Bent samples showed the two typical re-
gions associated with water-loss in smectites (Figure. 1).9 The water
adsorbed by external surfaces and hydration water of the interlayer
cations were removed between 80 and 150 °C, while structural water
was released at temperatures higher than 500 °C. Thermal treatment
of organoclays samples produced further mass loss at temperatures
of 180 to 780 °C,10,11 corresponding to organic matter oxidation and
charcoal formation.12 Temperature ranges were directly related with
the SL of each sample.
To determine the mass loss scheme in each sample, mi,α was
defined as the mass loss produced in the α temperature range for the
organoclays with an i surfactant load.
Range 1: Room Temperature – 180 ºC. Within this range of
temperatures, physisorbed and interlayer water and/or of water
of hydration of exchangeable cations (non structural water) were
eliminated,13 both in the starting clay and in the organoclays. This
mass loss was associated with the presence of endothermic peaks
in the DTA curve.
Range 2: 180 °C – 800 °C. In this range, a mass loss of 3% was
observed in Bent clay that corresponds to dehydroxylation of clay
layers.10,14,15
For organoclays, three mass losses in this range of temperatures
were observed, associated with several (from 3 to 5) exothermic
events (Figure 1). Initial and final temperatures of each one of these
mass losses varied with the surfactant loading, but fell among the
following subdivisions of range 2 (Table 1).
Range 2-a: Mass loss was associated with a very important
exothermic event (Figure 1-B); its position depends on the SL. The
amount of mass eliminated in this stage increased with the SL up to
2.0 CEC (m0.4,2-a = 2.5%, m0.8,2-a = 4.0% and m2.0,2-a = 22.5%)
and then kept constant (m3.0,2-a = 22.0%).
Range 2-b: The second stage of mass loss was associated with
various exothermic events, whose positions depended on the SL
(Figure 1). The amount of mass eliminated in this stage increased
with surfactant loading up to 0.8 CEC (m0.4,2-b = 4.5%, m0.8,2-b =
8.2%,) and kept constant up to 2.0 CEC, whereas for higher SL mass,
loss decreased slightly (m2.0,2-b = 8.5% and m3.0,2-b = 7.5%). This
slight decrease could be attributed to the fact that as SL increased, a
higher amount of heat was released by combustion of organic groups
in the 2-a range, which provided enough energy for the combustion
of molecules that should be eliminated in range 2-b.
Range 2-c: A mass loss associated with a weak exothermic peak
was observed (Figure 1). In this stage, mass loss due to dehydro-
xylation of layers and the third stage of mass loss associated with
surfactant combustion were overlapped.
To calculate the mass of organic compounds eliminated in this
stage, the fact that mass corresponding to dehydroxylation of layers
represents a 3% of Bent clay mass was taken into account.
The amount of organic compounds eliminated in this stage follows
the same trend that the one observed in Range 2-b. The proportion of
mass loss at this range of temperatures corresponding only to com-
bustion of organic compounds (after subtracting the contribution of
the dehydroxylation of layers) was 3.6%, 6.6%, 6.9% and 6.8% for
Bent-0.4, Bent-0.8, Bent-2.0 and Bent-3.0 respectively.
Range 3: 800 ºC – 1200 ºC. For Bent sample, an endothermic
maximum at 867 °C and two exothermic maximums at 906 and
1145 °C were found, all of them without any associated mass loss
(Figure 1). These peaks were attributed to structural rearrangements
and/or formation of new phases such as spinel, cristobalite or
mullite.10
To perform a deeper analysis of the processes of mass
loss observed at the different temperature ranges, the relation
Ri(α/β)=(mi,α)/( mi,β) was defined, where i and mi,α have the
meaning indicated in section 3-1, and β is a range of temperatures
different from α.
Figure 1. TGA (A) and DTA (B) of indicated samples. Temperature range limits
are indicated with vertical lines. Note different scales in y axis of figure 1-B
Synthesis and characterization of HDTMA-organoclays 3
Vol. XY, No. 00
Ri(2-a/2-b) and Ri(2-a/2-c), increased with the surfactant loading,
whereas Ri(2-b/2-c) kept practically constant (Figure 2). The slight
decrease of Ri(2-b/2-c) with SL was due to the strong release of heat
that took place in stage 2-a, which was discussed before.
Thermograms of leached organoclays were also obtained. In
general, they showed a similar behavior to starting organoclays. For
surfactant loadings up to 0.8 CEC, values of m2-a, m2-b and m2-c
were not modified with leaching treatments. For superior SL, the
m2-a decreased with the lecheates, whereas m2-b and m2-c kept
constant (Figure 3).
Taking into account Figures 2 and 3, m2-b and m2-c showed the
same behavior. To facilitate the discussion of the results, stages 2-b
and 2-c will be analyzed together, and be named as stage 2-(b+c).
The surfactant mass that was eliminated in stage 2-a decreased
with lecheates and led to the conclusion that the m2-a corresponds
to molecules of surfactant weakly attached to the clay. On the other
hand, in stage 2-(b+c), the loss of surfactant molecules more strongly
attached to the surface was produced, as m2-(b+c) did not decrease
with lecheates.
XRD studies
X-ray diffraction patterns (Figure 4) of different organoclays
showed values of the X-Ray diffraction peak corresponding to d(001)
distance that increased from 12.1 Å for the starting clay up to 19.2 Å
for the organoclay with higher SL (3.0 CEC). This increase (Table 2)
indicated that at least a fraction of the cationic surfactant has replaced
the hydrated interlayer cations.7
In Bent-0.8 the X-Ray diffraction peak d(001) was actually com-
posed by the overlapping of more than one component. In Bent-2.0
and in Bent-3.0 the presence of other X-Ray diffraction peaks of
lower intensity was also observed. This behavior is an indicator of
interstratification. The area of those peaks is related to the amount
of layers with a determined value of d(001).
In table 2, the positions of d(001) X-Ray diffraction peaks obtained
by deconvolution of XRD pattern are shown. From these distances,
the basal spacing was calculated, taking into account that the thick-
ness of the layer was 9.5 Å.16 In the cases in which there are more
than one d(001) line, the percentages of the area of each one of them
in respect to the total area were calculated.
Table 1. Temperatures of the different steps observed in TG, and position of the centre of the ATD peaks related to the steps
Sample Limits (°C) Maximum T (°C)
Range 2-a Range 2-b Range 2-c Range 2-a Range 2-b Range 2-c
Bent - - 448 – 690 - - 662*
Bent 0.4 220 - 343 343 - 573 573 – 784 291 322 655
Bent 0.8 212 - 321 321 - 560 560 - 779,5 311 329, 401, 534 650
Bent 2.0 184 - 334 334 - 554 554 – 758 300 346, 417, 516 668
Bent 3.0 185 - 313 313 - 544 544 – 782 282 342, 548 648
*Endothermic event that corresponds to dehydroxylation of layers.
Figure 2. Ri(α/β) (equation 4) in function of surfactant loading. (): Ri(2-
-a/2-b), (): Ri(2-a/2-c), (): Ri(2-b/2-c)
Figure 3. Washing effect from a Bent-2.0 sample. Squares indicate mass loss
in range 2-a, circles in range 2-b and triangles in range 2-c
Figure 4. X-ray diffraction pattern of indicated samples: Bent-0.4 (a), Bent-
0.8 (b), Bent-2.0 (c) and Bent-3.0 (d)
Naranjo et al.
4Quim. Nova
To determine the type of structure developed by the surfactant
intercalated within clay layers, the molecule of HDTMA+ was mod-
elled in an all-trans conformation.
Considering both the computational models and the interlaminar
distances experimentally obtained, different structures that the surfac-
tant could present in the interlamelar space were proposed (Table 2).
N2 adsorption
For expandable clays, N2 adsorption studies give a sub-evaluation
of the specific surface area value because they only provide informa-
tion about the external surface.7,17 They also provide information about
the meso and macropores that could be generated by the ordering of
tactoids in the structures of the “house of cards” type18 or, in the case
of mixture of clays with other materials of bigger grain size, by the
spatial ordering of particles.19
The N2 adsorption isotherm of Bent sample presented a hysteresis
loop type B, indicating mesopores presence (Figure 5).20 This hys-
teresis loop decreased with the SL until it disappeared for Bent-2.0
and Bent-3.0 samples. Pore size distribution (Figure 6-A) displayed
a maximum of pore size of 20.6 Å for Bent sample, increasing up to
25.1 Å for Bent-0.4 and Bent-0.8 samples. A maximum in the curve
of pore size distribution for Bent-2.0 and Bent-3.0 samples was not
observed. The value of BET surface area decreased from Bent (56 m2
g-1) to almost 10 m2 g-1 for clays with higher SL (Table 3).
Total pore volume (VTP) increased with the SL from Bent to
Bent-0.8, and then decreased for organoclays with higher SL (Table
3). Studies of scanning electron microscopy (not shown) have shown
that the starting clay presents a rather compact structure with a great
fraction of the layers presenting a face to face ordering. On the other
hand, organoclays present a more opened structure called “corn
flakes”. This change in the structure generated the biggest porosity
detected in organoclays. When the SL exceeded a certain value, the
excess of surfactant occupied mesopores or blocked their entries,
and, in this way, porosity decreased.
XRD results indicated that the interlayer space was never higher
than 10 Å, so that observed mesopores (within 20.6 and 25.1 Å)
corresponded to the pores generated by the “house of cards” order-
ing type of tactoids. This indicate that, apart from the amount of
Table 2. Peak position, basal spacing and structure proposed for the surfactant
in the interlayer space. Values obtained by XRD analyses
Sample (2θ)d(001)
(Å)
l
(Å)
a
(%) Stucture
Bent 7.3 12.1 2.6 100 -
Bent-0.4 6.22 14.2 4.7 100 Monolayer
Bent-0.8 6.08 14.5 5.0 33 Monolayer
5.27 16.8 7.3 67 Ps. Bilayer/Bilayer
Bent-2.0
9.28 9.5 0.0 4 Collapsed
6.54 13.5 4.0 4 Monolayer
4.78 18.5 9.0 92 Bilayer/Ps. Trilayer
Bent-3.0
9.06 9.8 0.3 6 Collapsed
6.30 14.0 4.5 2 Monolayer
4.60 19.2 9.7 92 Bilayer/Ps. Trilayer
l = Basal spacing; a = % of X-Ray diffraction peak area.
Figure 5. N2 adsorption-desorption isotherms. Symbols indicate: () Bent;
() Bent-0.4; () Bent-0.8; () Bent-2.0 and () Bent-3.0 samples
Table 3. BET surface area, C constant, micropore (VμP), mesopore (VmP)
and total pore (VTP) volume and fractal dimension of indicated samples.
Values obtained by N2 adsorption/desorption isotherms
Surfactant Loading
Bent Bent-0.4 Bent-0.8 Bent-2.0 Bent-3.0
SBET (m2/g) 56 25 27 11 12
C (BET) 348 72 100 29 29
VμP (cm3/g) 0.008 0.000 0.000 0.000 0.000
VmP (cm3/g) 0.077 0.098 0.116 0.052 0.059
VTP (cm3/g) 0.085 0.098 0.116 0.052 0.059
D 2.71 2.51 2.50 2.46 2.45
Pore size (Å) 20.6 25.1 25.0 - -
Figure 6. Pore size distribution (A) and Fractal dimension (B). Symbols
indicate: () Bent; () Bent-0.4; () Bent-0.8; () Bent-2.0 and ()
Bent-3.0 samples
Synthesis and characterization of HDTMA-organoclays 5
Vol. XY, No. 00
HDTMA+ adsorbed in the interlayer space, there was another frac-
tion of surfactant adsorbed in the external surface, occupying and/or
blocking the mesopores.
To examine if an effect of surface coverage was present, frac-
tal dimensions (D) were calculated using equations developed by
Frenkel, Halsey and Hill (FHH Theory).21 D can be considered as an
operational measure of surface roughness. Generally, D is placed be-
tween 2 (soft and regular surface) and 3 (surface extremely irregular).
In this work, as the starting solid (Bent) presents hysteresis, the
equation used for calculating the fractal dimension was D = 3 – 1/m,
where (1/m) was the absolute value of the slope obtained in the
Ln(Ln(P0/P)) vs Ln(Vads/VM) graph (Figure 6-B).
Starting clay presented a high value for the fractal dimension
(DBent = 2.71), similar to the value obtained by Wang for Saz-1 clay
(DWang = 2.74),21 indicating a very irregular surface. The D value of
organoclays decreased as surfactant loading increased. This decrease
was produced in two stages: the first stage for Bent-0.4 and Bent-0.8
(DBent-0.4 = 2.51, DBent-0.8 = 2.50), that depends on the elimination of
micropores as well as on the “softness” of the surface by the surfactant
and the second stage for Bent-2.0 and Bent-3.0 (DBent-2.0 = 2.46, DBent-3.0
= 2.45) due to the increase in the amount of molecules of surfactant
in the organoclay and the blocking of mesopores.
As the fractal dimension decreased and Total Pore Volume in-
creased together with the SL for Bent-4.0 and Bent-0.8 solids, the
un-intercalated surfactant covered the external surface as well as the
internal surfaces of the walls of pores formed by tactoids grouping
in “house of cards” type structures.
Consequently, from nitrogen adsorption isotherms it could be
concluded that cationic surfactant, besides intercalating within clay
layers, was adsorbed over the external surface and over the internal
walls of mesopores, coating and decreasing surface roughness. It
also fills or blocks mesopores, decreasing hysteresis and making it
disappear from organoclays with higher SL.
FTIR
IR spectra of samples (Figure 7) showed that the absorption bands
corresponding to the starting clay were also seen in organoclays, ap-
proximately in the same positions.
Symmetric bending bands, rocking bands, and symmetric and
asymmetric stretching of methylene (νs(CH2) and νas(CH2) respec-
tively) were observed in all organoclays. ν(N-H) peak starts to be seen
from a SL of 2.0 CEC, as corresponds to free HDMTA+ molecules.22
This result agrees with the results obtained in N2 adsorption studies,
indicating that for Bent-2.0 and Bent-3.0 organoclays, part of the
surfactant is adsorbed out of the interlayer space, filling or blocking
mesopores.
The more intense bands in the spectrum of surfactant correspond
to symmetric (νs (CH2)) and asymmetric (νas (CH2)) stretching of
methylene groups of the carbon chain of surfactant. As SL increases,
the intensity of the bands increases and the frequency decreases, ac-
cording to literature.21,23-27
Modelling the developed structures
Molecular modelling calculation showed that the wider surface
that a HDTMA+ molecule can cover is of approximately 100 Å2.
Thermogravimetry indicates the amount of molecules that were
absorbed for each SL, and from DRX results the structure of the
interlayer space of each organoclay was obtained. Using these data,
the surface that can cover those surfactant molecules (SHDTMA+) could
be calculated in the following way:
SHDTMA+ (m2/g) = f*nHDTMA+ (mol)*NA*sHDTMA+2)*10-20 (m2Å-2) /
Mclay (g) (1)
where SHDTMA+ is the specified surface (in m2 g-1) that n moles of
adsorbed surfactant (nHDTMA+) can cover, taking into account the
surfactant structure in the interlayer space (monolayer, bilayer, etc);
NA is the Avogadro number; sHDTMA+ is the surfactant molecule area
2); 10-20 is the conversion factor between m2 and A2; Mclay is the
clay mass (g), and f is a factor that takes into account the amount
of surfactant layers present in the interlayer space of the clay. The f
factor is obtained knowing that, for example, if a bilayer is formed,
the surface covered by the surfactant will be equivalent to the clay
surface, as the surface of the inferior layer and the superior layer will
both be considered, and therefore fbilayer = 1. In the same way, if a
monolayer is formed the factor must be fmonolayer = 2, and for a trilayer
ftrilayer = 2/3. For the intermediate structures, factors are calculated as
the average of the structures that make them up. So, fps-bilayer = (2+1)/2
= 1.5, and fps-trilayer = (1+2/3)/2 = 5/6.
In the case of Bent-0.4 organoclays, the structure that was formed
is a monolayer and the factor that was applied is fmonolayer = 2. For the
other cases, the structures were interstratified, so that the factor that
was applied is a weighted average, taking into account the percentages
of the different structures present in each organoclay. For Bent-2.0
and Bent-3.0 cases the percentage of collapsed layers were also taken
into account (Table 2).
Equation (1) can be used with the total mass of absorbed organic
compounds or with the mass of any of the fractions determined by
thermogravimetry. In each case, the surface that the surfactant mol-
ecules of the employed range can cover will be obtained. The results
obtained employing the total amount of the adsorbed organic com-
pounds are shown as SHDTMA+, Total, and the amount absorbed in range
2-(b+c) as SHDTMA+, 2-(b+c) (Table 4). The percentage that this surface
represents in respect to the total surface of starting clay (621 m2 g-1)10
is also indicated.
For SL higher than 0.8, SHDTMA+, Total exceeded 100% of covered
surface, which corroborated the fact that a fraction of surfactant is
adsorbed out of the interlayer surface in the mesopores generated
by tactoids ordering, causing the decrease of the pore volume and
the elimination of the hysteresis loop that was observed in the N2
adsorption isotherm. On the other hand, SHDTMA+, 2-(b+c) never exceeded
100%, achieving 99% of surface covering for a SL of 0.8. The result
confirmed that the surfactant corresponding to stage 2-(b+c) was
Figure 7. IR spectra of indicated samples. Bent (a), Bent-0.4 (b), Bent-0.8
(c), Bent-2.0 (d), Bent-3.0 (e) and HDTMA-Br (f)
Naranjo et al.
6Quim. Nova
Table 4. Covered surface (%) obtained applying equation 1
Sample Factor Range 2 (Total Organic) Range 2-(b+c)
sHDTMA+, Total (m2/g) % covered surface sHDTMA+, 2-(b+c) (m2/g) % covered surface
Bent-0.4 2.00 475 77 383 65
Bent-0.8 1.50 701 113 579 99
Bent-2.0 0.96 1081 174 506 86
Bent-3.0 0.94 983 158 447 76
adsorbed between layers. SHDTMA+, 2-(b+c) decreased for the organoclays
with higher SL (2.0 and 3.0) which could be due to the excessive
release of heat that was generated by the combustion of surfactant
in range 2-a, as was explained before.
CONCLUSIONS
Obtained organoclays resulted to be more hydrophobic than the
starting clay.
Surfactant, both in molecular and cationic form, was adsorbed
in the Bent clay in at least two kinds of different sites, indicated by
the different thermal stabilities.
From characterization results and molecular modelling calcula-
tion, a mechanism of adsorption is proposed. The surfactant was
adsorbed initially, in his cationic form, in sites placed in the interlayer
space of the clay through cation exchange. Adsorption in these kind
of sites continued until the interlayer space was saturated. Depending
on the quantity introduced in the intercalation process, different
organizations of surfactant in the interlayer, both in molecular and
cationic form, were obtained, varying from monolayer in Bent-0.4
to Bilayer/Pseudo-trilayer in Bent-3.0.
Further adsorption of surfactant, principally in molecular form,
occurred in the mesopores generated by tactoids ordered in the “house
of cards” type structure. This process left surfactant molecules relati-
vely free, out of the interlayer space. This fact could be favourable or
unfavourable, and must be analyzed for each potential application in
particular (catalytic processes, rheological control agents in paintings
and lubricants; polymer and plastics matrix reinforcing; adsorbents
for effluent treatments, oil spilling, releasing active matrix, etc.).
ACKNOWLEDGEMENTS
The funding for this work was granted by “Consejo de
Investigación de la UNSa” (Project Nº 1632) and SECyT FONCyT-
ANCyP (Project 1360). Authors acknowledge Lic. L. Davies, Ing.
S. Locatelli, Dra. D. Acosta and Ing. J. Villarroel Rocha for their
technical assistance and clarifying discussions. Pablo Naranjo thanks
CONICET and Chubut Province for their fellowships.
REFERENCES
1. Azejjel, H.; del Hoyo, C.; Draoui, K.; Rodríguez-Cruz, M. S.; Sánchez-
Martín, M. J.; Desalination 2009, 249, 1151.
2. Carrado, K. A.; Appl. Clay Sci. 2000, 17, 1.
3. de Paiva, L. B.; Morales, A. R.; Valenzuela Díaz, F. R.; Appl. Clay Sci.
2008, 42, 8.
4. Delbem, M. A.; Valera, T. S.; Valenzuela-Díaz, F. R.; Demarquette, N.
R.; Quim. Nova 2010, 33, 309.
5. Cavalcanti, J. V. F. L.; de Abreu, C. A. M.; Sobrinho, M. A. M.; Baraúna,
O. S.; Portella, L. A. P.; Quim. Nova 2009, 32, 2051.
6. Teixeira-Neto, E.; Teixeira-Neto, A. A.; Quim. Nova 2009, 32, 809.
7. Reid-Soukup, U.; SSSA Book Series: 7 Soil Mineralogy with Environ-
mental Applications, Soil Science Society of America, Inc.: Madison,
2002.
8. Bergaya, F.; Vayer, M.; Appl Clay Sci. 1997, 12, 275.
9. Hedley, C. B.; Yuan, G.; and Theng, B. K. G.; Appl. Clay Sci. 2007, 35,
180.
10. Xi, Y.; Frost, R. L.; He, H.; J. Colloid Interface Sci. 2007, 305, 150.
11. Yariv, S.; Lapides, I.; J. Therm. Anal. Calorim. 2005, 80, 11.
12. Mackenzie, R. C.; Differential Thermal Analysis, Academic Press: Lon-
don, 1970.
13. Moronta, A.; Solano, R.; Ferrer, V.; Sánchez, J.; Choren, E.; Ciencia
2003, 11, 130.
14. Xie, W.; Gao, Z.; Pan, W. P.; Hunter, D.; Singh, A.; Vaia, R.; Chem.
Mater. 2001, 13, 2979.
15. Magnoli, A. P.; Tallone, L.; Rosa, C. A. R.; Dalcero, A. M.; Chiacchiera,
S. M.; Torres Sanchez, R. M.; Appl. Clay Sci. 2008, 40, 63.
16. He, H.; Frost, R. L.; Bostrom, T.; Yuan, P.; Duong, L.; Yang, D.; Xi, Y.;
Kloprogge, J. T.; Appl. Clay Sci. 2006, 31, 262.
17. Michot, L. J.; Villieras, F. In Handbook of Clay Science; Bergaya, F.;
Theng, B. K.; Lagaly, G., eds.; Elsevier: Amsterdam, 2006, chap. 12.9.
18. Tessier, D.; Doctoral Thesis, Institut National de la Recherche
Agronomique, France, 1984.
19. Přikryl, R.; Weishauptová, Z.; Appl. Clay Sci. 2010, 47, 163.
20. Gregg, S. J.; Sing, K. S. W.; Adsorption, Surface Area and Porosity,
Academic Press: London, 1982.
21. Wang, C. C.; Juang, L. C.; Hsu, T. C.; Lee, C. K.; Lee, J. F.; Huang, F.
C.; J. Colloid Interface Sci. 2004, 273, 80.
22. Wang, C. C.; Juang, L. C.; Lee, C. K.; Hsu, T. C.; Lee, J. F.; Chao, H.
P.; J. Colloid Interface Sci. 2004, 280, 27.
23. Mandalia, T.; Bergaya, F.; J. Phys. Chem. Solids 2006, 67, 836.
24. Patel, H. A.; Somani, R. S.; Bajaj, H. C.; Jasra, R. V.; Appl. Clay Sci.
2007, 35, 194.
25. Praus, P.; Turicová, M.; Študentová, S.; Ritz, M.; J. Colloid Interface
Sci. 2006, 304, 29.
26. Zhu, J.; He, H.; Zhu, L.; Wen, X.; Deng, F.; J. Colloid Interface Sci.
2005, 286, 239.
27. Zhu, R.; Zhu, L.; and Xu, L.; Colloids Surf., A 2007, 294, 221.
... Because of that the extent of intercalation of the MMT interlayer spaces the adsorption remains practically unchanged. This suggests that the interactions between the montmorillonite and the modifying agents take place exclusively inside the mesopores formed by the clay particles ordered in the ''house of cards" type structure [64]. ...
... The strongest and complex band at near 1050 cm À1 (Fig. 6 a-c) is related to the Si À O in-plane stretching, while the broad ones at 526-529 and 468 cm À1 are due to the Al-O-Si and Si-O-Si bending vibrations, respectively [62,63,65,70]. The shoulder between 1300 and 1150 cm À1 is originated from the Si À O out-of-plane stretching vibrations and the broad peak at 797 and 693 cm À1 indicates the presence of quartz [63,64]. ...
Article
Mixtures of the cationic starch (CS), carboxymethyl cellulose (CMC) and guar gum (GG) with the pseudoamphoteric surfactant - cocamidopropyl betaine (CAPB) were used as stabilizers of aqueous suspensions of clay mineral (CM) of montmorillonite type (MMT). The spectrophotometric method (UV-VIS) was used in to estimate the mechanism of the polysaccharide/CAPB/MMT systems stabilization. Additionally, surface tension measurements were performed to investigate the interactions between the polymers and the surfactant and the complex formation. In order to obtain complete information about the studied systems: FT-IR, XRD, SEM as well as porosity measurements were also performed for polymer/clay mineral system as well as of those in the presence of surfactant. The results show that the MMT suspensions can be successfully stabilize using chosen polysaccharides (CS, CMC and GG) able to be adsorbed on the MMT surface. The best stability of the studied suspensions was obtained with a small concentration of polysaccharides (2 ppm) and a small amount of CAPB (0.04%). It seems that this surface active agent can be the ideal option for industrial applications because its small concentration ensures very large stability of the MMT/polysaccharide suspensions.
... The hysteresis loop is of type H3. The latter is obtained for particles with parallel plate shaped pores (Naranjo et al., 2015). The specific surface calculated by BET transform is estimated to be 30.396 ...
Article
Full-text available
This work includes studying the removal of methylthionine chloride on bridged bentonite, and conducting a batch system adsorption test. The influence of various parameters like the dose of adsorbent, pH, contact time and temperature on the behavior of the MC was studied. The pseudo second order kinetic model seems adequate and correlates with the experimental results. The adsorption isotherm fitted well to the Langmuir model with a maximum adsorption capacity adequate to 45.68 mg g −1. The values of the thermodynamic parameters (∆H • , ∆G • , and ∆S •) are negative indicating that the method of MC removal by the bentonite is exothermic, spontaneous, and with increasing order at the solid-solution interface. The results of the FTIR, XRD, SEM and BET characterizations show that this bentonite may be a mixture of Montmorillonite, Kaolinite, Illite, Quartz and Calcite, with a specific surface estimated at 30,3961 m 2 g −1 .
... The X-Ray diffraction patterns of modified clay GO/AC/HDTMA exhibited an interlayer spacing of 19.4 Å. This increase in the basal spacing for the modified clay demonstrates that at least a fraction of the cationic surfactant has replaced the hydrated interlayer cations [19]. ...
Article
Full-text available
In this work, graphene oxide/activated clay/Gelatin (GO/AC/G) composite blends were prepared by a simple solution mixing method. X-ray diffraction (XRD) and Fourier transform infrared (FTIR) spectroscopy were used to study the novelty in the structural characterization of the samples. The thermal stability of these materials was carried out using thermogravimetric analysis (TGA). The obtained results showed that a homogeneous mixture of AC, GO, and G was formed. XRD results indicated on successful formation of an intercalated structure in the composites. The disappearance of peaks at 2 = 8.1° and 2 = 13.5° were observed for montmorillonite and GO, respectively, indicating the homogenous distribution of the GO sheets into the activated clay structure. The interlayer spacing increased from 19.4 to 23.5 Å due to the insertion of gelatin molecules into the sheets of the clay. The IR spectrum of (GO/AC/G) composite revealed the presence of CO -C bonds, C=C bending, C-OH vibration, and C=O bending. These results show that GO was composited with AC structure. Furthermore, an intense band of N-H of gelatin at 3419 cm-1 was ameliorated via the combination with absorption bonds of O-H, indicating the interaction of gelatin with the clay. A comparison of the thermograms of GO/AC and GO/AC/G showed that the thermal stability was improved in the new prepared composite. High adsorption potential and regeneration capability make the GO/AC and GO/AC/G composites the potential environmentally friendly materials for reducing dye pollution.
... Depending on the type and amount of the organic cation used in the intercalation process, various arrangements of the surfactant in the interlayer space and also on the surface may be obtained, ranging from monolayers to bilayers or even pseudotrilayers (Heinz et al., 2007;Naranjo et al., 2015;Schampera et al., 2016). The use of organoclays prepared by alkylammonium salts to prepare polymer-clay nanocomposites may be limited by the organoclay properties. ...
Article
The stability of organoclays prepared from smectites and organic cations depends on the type of used cation, among other factors. This study provides a prediction of the structure, stability and dynamic properties of organoclays based on montmorillonite (Mt) intercalated with two types of organic cations – tetrabutylammonium (TBA) and tetrabutylphosphonium (TBP) – using first-principle density functional theory. The results obtained from simulations were also used in the interpretation of the experimental infrared spectrum of the TBP-Mt organoclay. Analysis of interatomic distances showed that weak C–O···H hydrogen bonds were important in the stabilization of both TBA- and TBP-Mt models, with slightly stronger hydrogen bonds for the TBP cation. Calculated intercalation and adsorption reaction energies (Δ Eint //Δ Eads* /Δ Eads** ) confirmed that TBP-Mt structures (–72.4//–32.8/–53.8 kJ/mol) were considerably more stable than TBA-Mt structures (–56.7//–22.6/–37.4 kJ/mol). The stronger interactions of the alkyl chains of the TBP cation with Mt basal surfaces in comparison to those of the TBA cation were also correlated with the positions of the calculated bands of the C–H stretching vibrations.
... For instance, SiO 2 and Al 2 O 3 peaks reduced in intensity, and the peak at 50° (2θ), which is attributed to Fe 2 O 3 , disappeared in the SMM XRD pattern. A similar interaction between HDTMA-Br with clay was reported by Pablo et al. (2015), whereby little or no substantial changes could be observed in the XRD patterns before and after modification. ...
Article
Full-text available
Montmorillonite modified with hexadecyltrimethylammonium bromide was used to remove vanadium (V) from synthetic and real mine water. Fourier transform infrared, X-ray diffraction, and scanning electron microscopy were used to characterise the adsorbent before and after adsorption, while the amount of V adsorbed was determined by ICP-OES. Batch adsorption was evaluated for dissolved V concentrations of 50–320 mg/L and V tailings seepage water from a South African mine. Adsorption capacity was affected by solution pH, temperature, sorbent mass, and the initial concentration. Electrical conductivity of the mine water before and after adsorption was measured to estimate the total dissolved solids. Equilibrium isotherm results revealed that V sorption follows the Freundlich isotherm, indicating that the sorbent surface was heterogeneous. A pseudo-second order kinetic model gave the best fit to the kinetic experimental data. The results of this study allow us to predict uptake efficiency of South African montmorillonite for V removal from mine water. However, the best adsorbent for the uptake of V or other contaminants will depend on the effluent to be treated.
Article
The mixtures of cationic cellulose (CC) or cationic guar gum (CGG) with the anionic sodium dodecyl sulfate surfactant (SDS) were used as stabilizers for the aqueous suspensions of montmorillonite (Mt). The stabilization processes and the stabilization mechanism were investigated using the UV-VIS. The obtained results show that both polysaccharides can be used as stabilizers of the water suspensions of montmorillonite due to the effective adsorption of CC and CGG with or without SDS on the Mt. surface. To obtain complete information on the studied systems, the additional measurements of the surface tension, zeta potential, FT-IR, XRD and SEM were made. The results prove that the intermolecular complexes formed between the polysaccharides and SDS can adsorb on the Mt. surface, change the structure of the electrical double layer and the stability properties of the studied suspensions.
Article
An organohalloysite was prepared using a new procedure. Halloysite (H) was pre-intercalated with dimethylsulfoxide and then mixed with a solution of hexadecyltrimethylammonium (HDTMA) having a concentration equivalent to six times the CEC of the starting Algerian halloysite. The novel nanohybrid obtained (HH6-d) was characterized and compared with a sample intercalated from the beginning by HDTMA (HH6). Intercalation of HDTMA cations in the interlayer space was evidenced by XRD with an expansion of the basal distance from 7.6 to 26.0 Å for an intercalation rate of 75% (HH6-d) versus 42% for HH6. All materials were used to remove diclofenac (DFC). pH influence, isotherm, thermodynamic data, reusability of the best adsorbent and the mechanism of interaction have been examined. The experimental isotherms were appropriately adjusted by the Redlich-Peterson equation. The best adsorbent adsorbs 154.3 mg g⁻¹, with an adsorption sequence as follows: HH6-d > HH6 > unmodified clay, i.e. according to the intercalated fraction, so that the HDTMA intercalating agent interacts with the diclofenac molecules. This interaction was entropically driven, endothermic and spontaneous. It would involve not only hydrophobic bond between the tail group of HDTMA represented by the hydrocarbon chain and the hydrophobic moiety of DFC represented by the aromatic groups, but also an electrostatic interaction between the negative charge of the carboxylate ion (COO⁻) and the positive charge of the aluminol groups coating the lumen.
Article
We describe the preparation of polymethylmethacrylate/(rice husk ash)/polypyrrole (PMMA/RHA/PPy) composite membranes and their use as active agents for the removal of hexavalent chromium (Cr(VI)) and the organic dyes tartrazine (E102) and indigo carmine (IC). We prepared the membranes following a two-step process: initially, using the electrospinning technique, we obtained PMMA:RHA mats with various relative concentrations, and afterward we used an in situ chemical polymerization to incorporate polypyrrole chains onto the surface of the membrane fibers. We characterized the membranes by SEM, water contact angle and fiber diameter measurements, tensile tests, UV–Vis and FTIR spectroscopy. We have found that the incorporation of RHA in the ratio 1:0.1 led to an enhanced mechanical strength of the polymeric fibers. At pH 2, the PMMA/RHA/PPy membranes exhibited a good removal capacity for all three contaminants, estimated as 360.5 mg/g (after 150 min), 165.7 mg/g (after 60 min) and 142.9 mg/g (after 70 min), for Cr(VI), E102 and IC, respectively. We suggest that the PMMA/RHA/PPy membrane is a promising active material for application in efficient water remediation protocols, for combining advantages as a simple preparation methodology and a high adsorption capacity toward different types of contaminants.
Article
Three organoclays were prepared by mixing an Algerian halloysite with a solution of hexadecyltrimethylammonium bromide (HDTMA-Br) equivalent to six times the cation-exchange capacity of our clay. Unlike a majority of studies which were focused on the initial concentration of the intercalating agent, this paper investigates the influence of the reaction time for a given initial concentration. Three intercalation times were examined: 2, 7, and 14 days. The resulting organoclays were analyzed by XRD, FTIR, TG–DTA, TEM, and N2 adsorption–desorption. The intercalation of HDTMA+ cations begins by a latency period up to 2 days, during which these cations interact with the external surface of halloysite. From 2 to 7 days, they migrate into the interlayer spaces, leading to an expansion of the basal distance from 7.3 to 26.0 Å. Between 7 and 14 days, the expansion remains unchanged for an intercalation rate around 42%. FTIR analysis proved that the surfactant interacts with the inner surface hydroxyl groups. From 200 °C, thermal analysis highlighted a succession of stages linked to the removal of HDTMA+. The TEM images showed a decrease in the outer diameter of the intercalated nanotubes with an enlargement of lumen diameter up to 20 nm. The arrangement of HDTMA+ species into interlayer spaces reflected a paraffin-type monolayer configuration. Knowing that the intercalation of organic compounds into the clay minerals changes their behavior from hydrophilic to hydrophobic, a nanotubular organohalloysite with a basal expansion of 26.0 Å could be a highly effective adsorbent for wastewater decontamination.
Article
Full-text available
The main goal of this research, was the preparation and use of a organophilic smectitic clay able to promoting the adsorption of phenol. In this work was used a natural, clay called Chocolate, from Campina Grande - PB (Brazil). The natural clay was treated with a solution of sodium, carbonate. After this the sodium clay was treated with quaternary ammonium salt. The adsorptive study was conducted by different values of pH and. temperature. The results showed a better performance in adsorptive at pH 7 and temperature 30 °C, with removal of more than 80% of phenol.
Article
Full-text available
Chemical modification of clays is possible due to their ion-exchange and adsorption capacities, which allows the adjustment of the physicochemical properties of the surfaces of their layers. This modification makes possible the use of clays to produce a great number of new materials, which range from coarse applications such as oil based drilling fluids to refined applications such as pharmaceutical products. This article intends to expose where there is still space for research and investment aiming at the performance improvement of clay-based materials.
Article
Full-text available
In this work, a smectite clay from the State of Paraiba, Brazil, was treated with six different types of ammonium salts, which is an usual method to enhance the affinity between the clay and polymer for the preparation of nanocomposites. The clays, before and after modification, were characterized by X ray diffraction. The conformation of the salts within the platelets of the clay depended on the number of long alkyl chains of the salt. The thermal stability of the clays was also studied. The ammonium salts thermal decomposition was explained in light of their position within the organoclays.
Article
Full-text available
Chemical modification of clays is possible due to their ion-exchange and adsorption capacities, which allows the adjustment of the physicochemical properties of the surfaces of their layers. This modification makes possible the use of clays to produce a great number of new materials, which range from coarse applications such as oil based drilling fluids to refined applications such as pharmaceutical products. This article intends to expose where there is still space for research and investment aiming at the performance improvement of clay-based materials.
Article
Thermo-XRD-analysis is applied to identify whether or not the adsorbed organic species penetrates into the interlayer space of the smectites mineral. In this technique an oriented smectite sample is gradually heated to temperatures above the irreversible dehydration of the clay, and after each thermal treatment is diffracted by X-ray at ambient conditions. In the thermal treatment of organo-clays, under air atmosphere at temperatures above 250°C, the organic matter is in part oxidized and charcoal is formed from the organic carbon. In inert atmosphere e.g. under vacuum above 250°C the organic matter is pyrolyzed and besides small molecules, charcoal is formed. If the adsorbed organic compound is located in the interlayer space, the charcoal is formed in that space, preventing the collapse of the clay. A basal spacing of above 1.12 nm suggests that during the adsorption the organic compound penetrated into the interlayer space. Thermo-XRD-analyses of montmorillonite complexes with anilines, fatty acids, alizarinate, protonated Congo red and of complexes of other smectites with acridine orange are described. To obtain information about spacings of the different tactoids that comprise the clay mixture, curve-fitting calculations on the X-ray diffractograms were adapted.
Article
The thermal stability of seven organically modified montmorillonites (‘organoclays’) has been investigated using differential thermal analysis, differential scanning calorimetry, and thermogravimetry in conjunction with X-ray diffractometry. Six organoclays were synthesised by replacing the interlayer inorganic cations, initially present, with quaternary phosphonium and ammonium surfactant cations. The samples modified with tetrabutylphosphonium (TBP), and butyltriphenylphosphonium (BTPP) ions have an appreciably higher thermal stability than the octadecyltrimethylammonium (ODTMA)-modified clays. Thus, in the case of TBP- and BTPP-modified montmorillonites, the onset temperature of decomposition is close to 300 °C. Samples modified with hexadecyltributylphosphonium (HDTBP) ions have a lower onset temperature of decomposition of 225 °C. In comparison, the onset temperature for ODTMA-montmorillonites (obtained at different concentrations of ODTMA-bromide) ranges from 158 to 222 °C, being highest where the concentration of intercalated surfactant is lowest. The onset temperature for a commercial alkylsilane-treated quaternary ammonium-modified organoclay (S-BEN N-400FP) is 207 °C. The basal spacing of the TBP- and BTPP-modified clays is 1.7–1.8 nm, indicating a monolayer arrangement of quaternary phosphonium ions in the interlayer space, while the value of 2.5 nm for HDTBP-montmorillonite indicates a more open structure. The ODTMA-modified samples have basal spacings ranging from 1.9 to 2.1 nm, indicative of a bilayer to pseudo-trilayer arrangement. The exceptionally high basal spacing of 3.4 nm for the S-BEN N-400FP organoclay might be due to interlayer penetration of organosilane hydrolysis products during synthesis. The thermal properties of organoclays are apparently related to the nature of the surfactants and their arrangement in the interlayer space of montmorillonite.
Article
This work was to examine the relationship between the configuration and sorption characteristics of surfactant–clay complexes. Various amounts of cetyltrimethylammonium bromide (CTMAB) were intercalated into the bentonite matrixes with different layer charges. Packing densities of the adsorbed surfactants and sorption characteristics of the obtained CTMA–bentonite complexes towards phenol and naphthalene were examined. Experimental results indicated that packing density of the adsorbed surfactant was proportional to the surfactant loading amount and layer charge of the bentonite, and sorption capacities of these complexes had a close relationship with the surfactant packing density. That was, with the increase of surfactant packing density, the organic-carbon normalized sorption coefficient (Koc) first rose till the maximum, and then began to decrease as the packing density further increased. This could be interpreted that increase of surfactant packing density would render the surfactant phases more hydrophobic environment, and the hydrophobic affinity of the surfactant phases towards the solutes thus increased accordingly. However, in the high surfactant packing density region, the densely packed surfactants reduced the available free space for the solutes, resulting in decrease of sorption capacity for these complexes. Hence, with the increase of surfactant packing density, the adsorbed surfactants would form a series of partition phases showing different affinity to the solutes.
Article
Clay mineral–polymer nanocomposites are prepared by dispersing solid organo clay minerals in two different melted polyolefin matrices, namely polyethylene (PE) and ethylene vinyl acetate (EVA). The organo clay minerals are prepared by adding different amounts of surfactant corresponding to the CEC of the pristine clay mineral. The characteristics of the organo clay minerals are obtained by XRD, IR spectroscopy, TGA, and swelling volume measurements. The amount of added surfactant has a direct effect on the interlayer separation and organophilicity–hydrophilicity balance of the clay mineral, evidencing a particular behavior transition about OMt1.2 The intercalation of PE is found to be dependent on the interlayer distance of the organo clay minerals while EVA intercalates in the organo clay minerals whatever the amount of surfactant (> 0.5CEC), leading to the same interlayer spacing (4nm). The polymer intercalation is more homogeneous in clay minerals having high surfactant loading corresponding to 1.5 and 2 CEC. Cone calorimeter results of the studied nanocomposites show a PHRR reduction of 32% for PE–OMt1.5 and of 47% for EVA–OMt1. For both polymers, the best compromise between mechanical and thermal properties, is obtained for organoclay filler obtained with an amount of added surfactant in a range 1–1.5 CEC.