Download as pdf or txt
Download as pdf or txt
You are on page 1of 6

Powder Technology 148 (2004) 32 – 37

www.elsevier.com/locate/powtec

Nonionic surfactants adsorption onto activated carbon.


Influence of the polar chain length
C.M. González-Garcı́aa, M.L. González-Martı́nb,*, J.F. Gonzáleza,
E. Sabioa, A. Ramiroa, J. Gañána
a
Department of Chemical and Energy Engineering, University of Extremadura, 06071-Badajoz, Spain
b
Department of Physics, University of Extremadura, 06071-Badajoz, Spain

Available online 12 October 2004

Abstract

The adsorption of a nonionic surfactants series characterized by a different length of the hydrophilic tail on an activated carbon has been
studied over a wide concentration range. Adsorption isotherms show two steps limited by the critical micelle concentration (c.m.c.) of the
surfactants. The adsorption extension depends on the oxyethylenic chain length, in such way that the amount adsorbed decreases with
increasing the chain length, although the effect is much lower for the longest polar chains. In the concentration range below the c.m.c. of the
surfactants, the molecules are adsorbed by direct interactions with the activated carbon surface, but with different configuration of the
polyoxyethylene chain (POE) directed to the aqueous phase. For concentrations above the c.m.c., the adsorption takes place by the
interactions between the adsorbed surfactants molecules and differences related to the length of the hydrophilic chain are also found.
D 2004 Elsevier B.V. All rights reserved.

Keywords: Nonionic surfactants; Activated carbon; Adsorption; Langmuir equation

1. Introduction layer attains a high degree of structure and interactions


among adsorbed molecules are similar to those in the
It is generally accepted that surfactant adsorption at solid/ micellar state. Equilibrium concentrations higher enough
liquid interfaces may involve hydrophobic interactions than c.m.c. correspond in some cases to the appearance of a
between the adsorbed molecules similar to those in bulk plateau.
micelle formation, if their concentration exceeds the limit Several authors have already described this model of
given by the critical micellar concentration (c.m.c.) [1–4]. adsorption for nonionic as well as for ionic surfactants [4,8].
In general, surfactant adsorption isotherms from aqueous However, an equation to express the adsorption equilibrium
solutions onto solids show several steps when equilibrium relationship is often needed. This is usually done by fitting
concentration increases [3,5–7]. At the first retention stage, experimental data to an isotherm model. Nevertheless, the
well below the surfactant c.m.c., adsorbed molecules tend to different kind of interactions that progressively dominates
lie flat on the surface and interactions between them are the surfactants adsorption process suggests very unlikely
negligible. When adsorption increases, a progressive reor- that a single adsorption equation—which comes at the end
ientation of molecules to a more tilted position on the from an assumed adsorption model dictated by a single kind
adsorbent surface is then possible. The next stage of of interactions—could fit the whole range covered in the
adsorption, around the c.m.c. region, is dominated by isotherm.
cooperative interactions (adsorbate–adsorbate); the adsorbed The aim of our work is to describe the adsorption of a
nonionic surfactants series characterized by a different
length of the hydrophilic tail on an activated carbon by a
* Corresponding author. Tel.: +34 924289532; fax: +34 924289651. suitable equation that represents the adsorption systems.
E-mail address: mlglez@unex.es (M.L. González-Martı́n). This study has been extended to the equilibrium concen-
0032-5910/$ - see front matter D 2004 Elsevier B.V. All rights reserved.
doi:10.1016/j.powtec.2004.09.017
C.M. González-Garcı́a et al. / Powder Technology 148 (2004) 32–37 33

tration range above the c.m.c. to elucidate interactions at the (V ma=0.28 cm3 g1) volume. Just prior to use as adsorbent,
solid–liquid interface between adsorbed surfactants mole- the carbon was dried in an oven at 110 8C for 24 h.
cules. In order to analyze the extension at which each of the The adsorption isotherms were established by the change
mechanisms of nonionic surfactant adsorption can be the of the surfactant concentration before and after the contact
main driven force of the process, and its relative weight with the solid in an end-over-end stirrer, at 11 rpm, in an air
depending on the length of the surfactant hydrophilic tail, bath at 20.0F0.1 8C for 72 h. Approximately, 0.02 g of
we have modeled the adsorption process as a combination of activated carbon is added to 10 ml of the surfactant solution
two simple contributions each of them characterized by a of a given initial concentration. The supernatant was
simple Langmuir equation, by testing the whole set of separated from the solid by centrifugation at 4000 rpm for
experimental results against a bi-Langmuir equation, which 10 min (Centrifuge Digicen-R). Determination of the
allows to get a single monolayer coverage and a single concentration was carried out by UV absorption at 278 nm.
adsorption free energy for each of the assumed processes.
The proposed method was successfully applied to the
adsorption characterization of a given ionic or nonionic 3. Results and discussion
surfactant on carbon blacks and activated carbons [9–13],
showing how topographical differences on samples affected The adsorption isotherms of TX-100, TX-305 and TX-
the adsorption process. It is expected that the application of 405 surfactants from aqueous solutions onto the activated
the bi-Langmuir equation to our systems can lead to a better carbon Darco are shown in Fig. 1, expressed as the amount
understanding of the role of the hydrophilic chain length in of surfactant adsorbed per gram of adsorbent (X) vs. the
adsorption for the analyzed surfactants. logarithm of the equilibrium concentration (C). Isotherm
shapes display two clear sections limited around the c.m.c.
of the surfactants (logC c.m.c. equal to 3.5 for TX-100, 3
2. Materials and experimental methods for TX-305 and 2.9 for TX-405), which indicate the
presence of at least two different adsorption mechanisms
Nonionic surfactants of the Triton X series (TX), involved in the whole adsorption process. As we have
polyoxyethylenic octylphenol surfactants type, with an previously indicated, we will analyze the adsorption
average number of ethylene oxide units of 9.5 for Triton behavior with the help of a bi-Langmuir equation, but, as
X-100 (TX-100), 30 for Triton X-305 (TX-305) and 40 for a first step, we have studied the most diluted concentration
Triton X-405 (TX-405) have been used [14]. The surfactants range (below the c.m.c. of the surfactants), where it could be
were used as supplied by Fluka (TX-100, purityN99%) and expected that the adsorption mechanism should be domi-
Union Carbide (TX-305 and TX-405). Aqueous solutions of nated near exclusively by one kind of interaction, adsor-
surfactants were prepared with ultrapure water [Milli-Q Plus bate–adsorbent.
(Millipore)]. Its critical micelle concentration in water at 20 The first section of the adsorption isotherms can be
8C [15] is given in Table 1. The molecular dimensions of the analyzed with the help of a Langmuir equation, assuming
surfactant molecules have been evaluated from the surface that in the low concentration range, there is not interaction
area of each group of the surfactants [16,17]; that is, among adsorbed molecules and there is a constant average
hydrocarbon chain (HC), phenyl group (P) and polyoxy- adsorption free energy value. Langmuir equation [19]:
ethylene chain (POE) (Table 1).
Xm Cb
The activated carbon Darco (D, Aldrich) was used as X ¼ ð1Þ
1 þ Cb
adsorbent for this study. Some of the surface characteristics
of the carbon had been evaluated in a previous work [18]: provides an estimate of the maximum adsorption capacity,
the nitrogen adsorption isotherms at 77 K were used to assumed as a complete monolayer coverage, X m, and
calculate the BET-specific surface area accessible to N2 through the parameter b a measure of the adsorbate–
(S N2=611 m2 g1), the micropore (V mi=0.25 cm3 g1) and adsorbent interaction free energy, DG ads:
mesopore (V me=0.41 cm3 g1) volumes and mercury
DGads ¼  RT lnðbxÞ ð2Þ
porosimetry data were used to determine the macropore
where x refers to the water concentration in solution (55.55
mol l1 at 20 8C).
Table 1 Langmuir equation matches the experimental results very
Characteristics of the surfactants accurately up to a concentration value around 1.8104 mol
Surfactant l1 (TX-100), 8.2104 mol l1 (TX-305) and 8.4104
TX-100 TX-305 TX-405 mol l1 (TX-405), below but near the c.m.c. of the
c.m.c.a (mol l1) 3.0104 1.0103 1.2103 surfactants, after which it gets worse. Nevertheless, the
Contactable surface 1.4829 3.4817 4.4567 curves obtained from this adjustment have been extrapolated
area (nm2 molecule1) for the entire range of C studied (Fig. 1), showing clearly
a
Taken from Ref. [15]. that contributions due to adsorption mechanisms different
34 C.M. González-Garcı́a et al. / Powder Technology 148 (2004) 32–37

22]. Then, the minimum area occupied by each molecule


will be dependent on the stereochemical requirements of the
oxyethylene chains protruding into the aqueous phase, since
the possible configurations of the flexible POE chain in
solution and at the solid–liquid interface are numerous
[8,23]. Conformations of the oxyethylene chain of adsorbed
surfactants depend on the degree of hydration of the chains
[8], its length and its interaction with the solid surface [23].
This conformation, especially for long POE chains, can
range from a flat to a coil configuration.
The X m sequence shown in Table 2, indicates that despite
the amount of TX adsorbed decreases with increasing the
length of the oxyethylenic chain, it does especially when the
change goes from 9.5 to 30 oxyethylene groups, but
remaining nearly constant when the number of these group
passes from 30 to 40. According to the results, the
adsorption extension does not depend on the length of the
hydrophilic moiety when the molecule is long enough,
which could indicate that the longest POE chains show a
random coil configuration which projects a similar excluded
area for the incoming molecules. Also, the very different
values of X m found between TX-100 surfactant (5.83104
mol g 1 ) and the other two ones (1.08 10 4 and
1.05104 mol g1, for TX-305 and TX-405, respectively)
can indicate that the random coil conformation of the long
polar chains blocks the more internal positions of the
adsorbent and the retention would be prevented, because of
porosity and pore size distribution of the carbon. This fact
can also be confirmed from the area per adsorbed TX
molecule, A m, evaluated according to:
S
Am ¼ ð3Þ
Xm NA
where S is the accessible surface area of the carbon
determined by nitrogen adsorption at 77 K, S N2, and N A
is Avogadro’s number. The A m values obtained (Table 2)
Fig. 1. Adsorption isotherms of TX surfactants onto the activated carbon can be compared with the contactable surface area of Triton
(., TX-100; E, TX-305; x, TX-405). molecules evaluated from their molecular dimensions (Table
1). These values are lower than that of A m for all cases, in
than direct adsobate–adsorbent interaction are gaining accord with the non-accessibility to the whole S N2 for the
importance as the concentration increases above c.m.c. TX molecules. However, it is worth noting that whereas for
The X m values obtained (Table 2), show that the amount TX-100 surfactant the A m value is only slightly higher than
of TX adsorbed decreases as the number of ethylene oxide its molecular dimensions, for samples TX-305 and TX-405,
units in the TX molecules increases; according to the its values are much higher, with a difference of more than 6
general conclusion that the excluded area created by the nm2 molecule1. This fact indicates the lesser contribution
polyoxyethylene chain directed to the solution phase is the from non-accessible surface in narrowest pores for TX-100
principal parameter which limits the amount adsorbed [20– than for the other two surfactants.

Table 2
Parameters found from the fitting of the experimental results of adsorption of TX surfactants onto the activated carbon by Langmuir equation in the
concentration range below the c.m.c. of the surfactants
Surfactant Xm b Am DG ads Dg ads
(104 mol g1) (103 l mol1) (nm2 molecule1) (kJ mol1) (1020 J molecule1)
TX-100 5.83 18.5 1.74 33.7 5.6
TX-305 1.08 5.7 9.37 30.8 5.1
TX-405 1.05 4.8 9.64 30.4 5.1
C.M. González-Garcı́a et al. / Powder Technology 148 (2004) 32–37 35

When the equilibrium concentration increases, adsorp-


tion progresses on more open areas of the surface, and
around the c.m.c., lateral interactions at the adsorbed layer
could occur [1–4], which is usually displayed by a sharp
increase in adsorption isotherms [1]. Despite of differences
in the adsorption process at the lowest concentrations, the
presence in all the adsorption isotherms of an inflexion point
located at values of C around the c.m.c of each surfactant
(Fig. 1) suggests that the adsorption progress is similar in all
systems. Therefore, it would be reasonable to assume that
the rising part in the adsorption isotherms indicates that
adsorbate–adsorbate interactions at the solid–liquid interface
take place. Then, two adsorption mechanisms could be
adopted to explain the whole adsorption process of Triton
surfactants on the carbon surface.
As we have previously discussed, we will try to analyze
the adsorption behavior with the help of a double scheme of
Langmuir. For this purpose, the experimental results of the
Fig. 2. Fitting of the experimental results of TX-100 adsorption onto the
activated carbon by a double Langmuir scheme (curve 1) and each of the
surfactants adsorption onto the activated carbon for the
summed curves (curves 2 and 3). entire range of C studied were fitted to equation:
Xm2 Cb2 Xm3 Cb3
The accessibility of TX-100 to smaller pores than TX- X ¼ þ ð4Þ
1 þ Cb2 1 þ Cb3
305 and TX-405 is also confirmed by the values of the
adsorption free energy obtained. The values of the free where the parameters have the habitual meaning of the
energy of adsorption calculated from Eq. (2), DG ads, as well Langmuir equation. In Figs. 2–4, the different adsorption
as the values per molecule, Dg ads (Table 2), show the isotherms, for the entire range of concentration studied,
stronger interaction for TX-100 molecules and also the same Cc1.0103 mol l1, have been expressed as X vs. C. The
sequence of variation as X m values is found, i.e., a higher curves resulting from that fit, curve labeled as 1, have been
absolute value for TX-100 than for TX-305 and TX-405, plotted in Figs. 2–4 together with each of the summed
which have both the same value, and reflect that TX-100 curves (curves 2 and 3); hence, curve 1 is the sum of curve 2
adsorption occurs in pores small enough to allow the and 3. It can be seen that curve 1 matches the experimental
adsorbed molecule experiences some enhanced potential, dates very accurately and that each of the curves 2 and 3 has
due to simultaneous interactions of the adsorbed molecule a similar shape for all the adsorption systems. Curve 3 has a
with both pore walls. nearly vertical branch at the lowest C but quickly reaches a

Fig. 3. Fitting of the experimental results of TX-305 adsorption onto the Fig. 4. Fitting of the experimental results of TX-405 adsorption onto the
activated carbon by a double Langmuir scheme (curve 1) and each of the activated carbon by a double Langmuir scheme (curve 1) and each of the
summed curves (curves 2 and 3). summed curves (curves 2 and 3).
36 C.M. González-Garcı́a et al. / Powder Technology 148 (2004) 32–37

Table 3
Parameters found from Eq. (4)
Surfactant X m2 X m3 X m3/X m3+X m2 b2 b3
(104 mol g1) (104 mol g1) (%) (102 l mol1) (103 l mol1)
TX-100 8.03 4.60 36 2.8 27.5
TX-305 2.88 0.51 15 2.4 7.8
TX-405 2.16 0.34 14 5.4 7.1

well-defined plateau. Curve 2 has a small slope at the lowest results, analysis of Langmuir equation at concentrations
C, but it changes smoothly with C in the whole range. Thus, below the c.m.c. and curve 3 from double Langmuir
the behavior of curve 1 is controlled by curve 3 and 2 at the scheme, which suggests that the treatment of the exper-
lowest and the highest concentrations, respectively. This imental results with this scheme gives a good representation
mathematical behavior allows the analysis of the whole of the adsorption systems.
Triton surfactants adsorption onto the activated carbon into The second adsorption mechanism that can be assumed
two steps: the surfactant adsorption by direct interaction in our systems is that represented by curve 2 (Figs. 1–4) and
with the carbon surface, which can be assumed as through A m2 and Dg ads2 values (Table 4). The A m2 values
completed when concentration reaches values close to the are much lower that those corresponding to the process
surfactant c.m.c. (curve 3); and a second process, mainly represented by curve 3 (A m3). The lowest A m2 value
above the c.m.c., for which the newly adsorbed molecules corresponds, as expected, to TX-100 surfactant. If these
are retained by interaction with those previously adsorbed last values are compared with the molecular dimensions of
(curve 2), both of them described by a Langmuir equation. surfactants (Table 1), it is obtained that both are similar for
The X m and b values found from both curves are reported TX-305 and TX-405, and lower for TX-100. Bearing in
in Table 3, with the sub-index representing the curve from mind that an important part of the BET-specific surface area
which they were obtained. The value of area occupied per of the adsorbent is not accessible to the TX molecules, these
molecule in each of the two proposed processes, A m3 and results reflect some arrangement of adsorbed molecules on
A m2, and the adsorption free energy, DG ads3 and DG ads2, as the carbon surface, as adsorption progresses probably by the
well as their values per molecule, Dg ads3 and Dg ads2, are formation of surfactant aggregates in the adsorbed layer by
also given in Table 4. interactions of the newly adsorbed molecules with those
As it was mentioned, at the lowest equilibrium concen- already fixed onto the solid surface.
trations, the adsorption process takes place by direct This possibility of aggregate formation can be confirmed
adsorbate–adsorbent interactions, but with different config- by the adsorption free energy. Since micellization takes
uration of the polar tail depending on its length. If a coil place by the interactions of surfactants molecules through
conformation for the longest POE chains limits its adsorp- water, a measure of the interaction free energy of surfactant
tion in the low concentration range, this behavior must be molecules in the aqueous phase can be obtained from their
manifested in our treatment by the curve representative of micellization free energy, DG mi. In aqueous solutions where
such process (curve 3). the c.m.c. is lower than 101 mol l1, DG mi can be
For each sample X m3 is lower than X m2, although it can calculated according to [24]:
be distinguished important differences related to its relative
DGmi ¼ RT lnðc:m:c:=xÞ ð5Þ
contribution on curve 1 (Table 3). As can be seen, the
influence of curve 3 on curve 1 is much lower for TX-305 Taking into account the c.m.c. of surfactants at 20 8C
and TX-405 than that for TX-100, since for the last one the (Table 1), the values per molecule, Dg mi, obtained are
contribution of both curves at the highest concentrations is 4.91020, 4.41020 and 4.31020 J molecule1
quite similar. Also, the values of Dg ads3 are similar to those for TX-100, TX-305 and TX-405, respectively. From
calculated from a simple Langmuir equation, Dg ads (Table comparison between the values of Dg ads2 (Table 4) and
2), since both are related to the same type of interaction, and Dg mi, it can be seen that the values of Dg ads2 are lower in
they reflect the stronger adsorption of TX-100 surfactant. As absolute value than Dg mi. If we assume that the adsorption
it can be seen from Tables 2 and 4, there is a very good process related to curve 2 takes place by lateral interactions
correlation in all systems between these two groups of between TX molecules, the lower value of Dg ads2 indicates

Table 4
Free energy of adsorption, its value per molecule and area per molecule from curves 2 and 3
Surfactant DG ads2 DG ads3 Dg ads2 Dg ads3 A m2 A m3
(kJ mol1) (kJ mol1) (1020 J molecule1) (1020 J molecule1) (nm2 molecule1) (nm2 molecule1)
TX-100 23.5 34.7 3.9 5.8 1.23 2.20
TX-305 23.1 31.6 3.8 5.2 3.52 19.79
TX-405 25.1 31.4 4.2 5.2 4.68 29.29
C.M. González-Garcı́a et al. / Powder Technology 148 (2004) 32–37 37

that the arrangement of TX molecules on the adsorbed layer [7] Z. Bu-Yao, G. Tiren, Z. Xiaolin, General isotherm equation for
is worse than in micelles. However, the correlation between adsorption of surfactants at solid/liquid interfaces: Part 2. Applica-
tions, J. Chem. Soc., Faraday Trans. I 85 (11) (1989) 3819 – 3824.
both values is better as the length of the oxyethylenic chain [8] A. Gellan, C.H. Rochester, Thermodynamics of adsorption of o-n-
increases, reaching very similar values for TX-305 surfac- dodecylpentaethylene glycol and O-n-dodecyloctaethylene glycol
tant and then suggesting that as the length of the POE chains from aqueous solutions on to graphitised carbon, J. Chem. Soc.,
increases the two interaction mechanism are more similar. In Faraday Trans. I 81 (1985) 1503 – 1512.
[9] C.M. González-Garcı́a, M.L. González-Martı́n, V. Gómez-Serrano,
accordance with the proposal that above the c.m.c., the
J.M. Bruque, L. Labajos-Broncano, Determination of the free energy
adsorption mechanism of TX molecules takes place through of adsorption on carbon blacks of a nonionic surfactant from aqueous
interactions with those previously adsorbed, it is expected solutions, Langmuir 16 (8) (2000) 3950 – 3956.
that the differences found in the first adsorption layer are [10] C.M. González-Garcı́a, M.L. González-Martı́n, V. Gómez-Serrano,
also reflected on the second adsorbed layer. The longest TX J.M. Bruque, L. Labajos-Broncano, Analysis of the adsorption
molecules retained by direct interaction with the carbon isotherms of a non-ionic surfactant from aqueous solution onto
activated carbons, Carbon 39 (2001) 849 – 855.
surface have a more tilted orientation to the surface than that [11] C.M. González-Garcı́a, M.L. González-Martı́n, A.M. Gallardo-
for the TX-100 surfactant, because of its weaker interaction Moreno, V. Gómez-Serrano, L. Labajos-Broncano, J.M. Bruque,
predicted by Dg ads and Dg ads3 (Tables 2 and 4), their Free energy of interaction of sodium dodecyl sulfate in aqueous
configuration allowing a interaction between surfactant solution with carbon black surfaces, J. Colloid Interface Sci. 248
molecules more similar than that in micelles, in reference (2002) 13 – 18.
[12] C.M. González-Garcı́a, M.L. González-Martı́n, A.M. Gallardo-Mor-
to TX-100 surfactant, and then decreasing the differences eno, V. Gómez-Serrano, L. Labajos-Broncano, J.M. Bruque, Removal
between Dg ads2 and Dg mi values. of an ionic surfactant from wastewater by carbon blacks adsorption,
Sep. Sci. Technol. 37 (12) (2002) 2823 – 2837.
[13] C.M. González-Garcı́a, M.L. González-Martı́n, R. Denoyel, A.M.
4. Conclusions Gallardo-Moreno, L. Labajos-Broncano, J.M. Bruque, Ionic surfactant
adsorption onto activated carbons, J. Colloid Interface Sci. 278 (2)
(2004) 257 – 264.
Based on our results, it can be concluded that the use of a [14] E.A. Dennis, A.A. Ribeiro, M.F. Roberts, J. Robson, in: K.L. Mittal
double Langmuir scheme for the analysis of the studied (Ed.), Solution Chemistry of Surfactant, Plenum, New York, 1979, pp.
adsorption system provides a deeper information on the 175 – 194.
evolution of the retention process and the differences in the [15] M. Lindheimer, E. Keh, S. Zaini, S. Partyka, Interfacial aggregation of
nonionic surfactants onto silica gel: calorimetric evidence, J. Colloid
behavior of the surfactants selected. At the lowest equili- Interface Sci. 138 (1) (1990) 83 – 91.
brium concentrations, surfactants are adsorbed directly to [16] B. Janczuck, M.L. González-Martı́n, J.M. Bruque, C. Dorado-
the carbon surface, allowing their polar tail in a more or less Calasanz, J. Moreno del Pozo, The relationship between the
coil configuration be directed to the liquid phase. This interfacial free energy and the free energy of micellization of triton
X-100 and sodium dodecyl sulfonate, J. Colloid Interface Sci. 176
arrangement in the first adsorption layer drives the
(1995) 352 – 357.
adsorption progress by surfactant–surfactant interactions [17] C.J. van Oss, R.J. Good, Relation between the apolar and polar
more similar to those taking place in micelle formation as components of the interaction energy between chains of nonionic
much the molecules adsorbed on the carbon surface have a surfactants and their CMC in water, J. Dispers. Sci. Technol. 12 (1)
longer part directed to the liquid phase. (1991) 95 – 105.
[18] V. Gómez-Serrano, C.M. González-Garcı́a, M.L. González-Martı́n,
Nitrogen adsorption isotherms on carbonaceous materials. Compar-
ison of BET and Langmuir surface areas, Powder Technol. 116 (2001)
References 103 – 108.
[19] A.W. Adamson, A.P. Gast, Physical Chemistry of Surfaces, Wiley-
[1] J.S. Clunie, B.T. Ingram, in: G.D. Parfitt, C.H. Rochester (Eds.), Interscience, New York, 1990, pp. 421 – 459.
Adsorption from Solution at the Solid/Liquid Interface, Academic [20] C. Ma, Y. Xia, Mixed adsorption of sodium dodecyl sulfate and
Press, London, 1983, pp. 105 – 152. ethoxylated nonylphenols on carbon black and the stability of carbon
[2] S. Partyka, S. Zaini, M. Lindheimer, B. Brun, The adsorption of non- blacks dispersions in mixed solutions of sodium dodecyl sulfate and
ionic surfactants on a silica gel, Colloids Surf. 12 (1984) 255 – 270. ethoxylated nonylphenols, Colloids Surf. 66 (1992) 215 – 221.
[3] I. Czinkota, R. Ffldényi, Z. Lengyel, A. Marton, Adsorption of [21] J.M. Douillard, S. Pougnet, B. Faucompre, S. Partyka, The adsorption
propisochlor on solids and soil components equation for multi-step of polyoxyethylenated octyl and nonylphenol surfactants on carbon
isotherms, Chemosphere 48 (2002) 725 – 731. black and sulfur from aqueous solutions, J. Colloid Interface Sci. 154
[4] P.E. Levitz, Adsorption of non ionic surfactants at the solid/water (1) (1992) 113 – 121.
interface, Colloids Surf., A Physicochem. Eng. Asp. 205 (2002) 31 – 38. [22] L. Bossolettti, R. Ricceri, G. Gabrielli, The adsorption of polystyrene
[5] L. Luciani, R. Denoyel, J. Rouquerol, Poly(ethoxy) anionic surfac- sulfonate and ethoxylated non-ionic surfactants at carbon black–water
tants: micellization and adsorption at the solid/liquid interface, interface, J. Dispers. Sci. Technol. 16 (1995) 205 – 220.
Colloids Surf., A Physicochem. Eng. Asp. 178 (2001) 297 – 312. [23] P.E. Levitz, Non-ionic surfactants adsorption: structure and thermo-
[6] Z. Bu-Yao, G. Tiren, General isotherm equation for adsorption of dynamics, C. R. Geosci. 334 (2002) 665 – 673.
surfactants at solid/liquid interfaces: Part 1. Theoretical, J. Chem. [24] M.J. Rosen, Surfactant and Interfacial Phenomena, Wiley Inter-
Soc., Faraday Trans. I 85 (11) (1989) 3813 – 3817. science, New York, 1989, pp. 108 – 169.

You might also like