Anda di halaman 1dari 6

Geosciences Journal Vol. 7, No. 3, p.

203 208, September 2003

Study on the exchange reaction of HDTMA with the inorganic cations in reference montmorillonites
Seung Yeop Lee* Soo Jin Kim

} School of Earth and Environmental Sciences, Seoul National University, Seoul 151-747, Korea
spectroscopy, and atomic force microscopy (AFM) have been used to investigate the adsorption of cationic surfactant to the mineral-water interface as a function of surfactant concentrations, with and without added electrolytes, and as a function of pH (Manne and Gaub, 1995; Ducker and Wanless, 1999; Schulz and Warr, 2000; Atkin et al., 2001; Yui et al., 2002). From the effort, it appeared that the adsorption of cationic surfactants onto non-swelling silicates was well rationalized and could be satisfactorily described by theoretical models. In contrast, adsorption onto swelling clay such as montmorillonite was considered to be complicated by the interlayer intercalation of surfactants and the evolution of the aggregation state of the clay sheets upon adsorption of cationic surfactants (Brahimi et al., 1992). Accordingly, the observation for the surfactantmodified swelling clays has received little attention. Although many scientists have performed experiments for the surfactant treated clays, there is not sufficient research on the surface change via intercalated surfactants through scanning electron microscopy (SEM). In this paper, our intention is therefore to lead to an improved characterization for organoclays treated with long alkyl chain surfactants. 2. EXPERIMENTS 2.1. Clay Samples Wyoming Na-montmorillonite (SWy) and Arizona Camontmorillonite (SAz) were obtained from the Source Clay Repository (The Clay Minerals Society, Columbia, MO). The montmorillonite samples were used without further purification. The characteristics of the samples are shown in Table 1. 2.2. HDTMA Adsorption Isotherms Batch equilibration isotherms were determined on 20 mg clay samples weighed into 50 mL centrifuge tubes. To these tubes, 40 mL of solutions containing the quantities of HDTMA bromide (Sigma, St. Louis, MO) equivalent to 0.02 to 1 times the cation exchange capacity (CEC) of clay samples were added. These suspensions were then shaken 7 days on a reciprocating shaker under ambient conditions. The mixture was centrifuged at 8000 rpm for 20 min to yield a clear

ABSTRACT: The adsorption of hexadecyltrimethylammonium (HDTMA) cations on swelling layer silicates (montmorillonites) was studied by adsorption isotherms, Xray diffraction, and electron microscopy. At low HDTMA concentrations, HDTMA ions started to be adsorbed on interlayer sites of SWy montmorillonite, causing a preferential release of interlayer Na+ compared with Ca2+, while the lateral attraction between adsorbed HDTMA cations at edges or external surfaces highly prevailed in SAz montmorillonite. Montmorillonite clay surfaces appeared as foliated and irregular aggregate structures at high HDTMA loadings. Besides the clay surface charge, the interlayer inorganic cations appeared to have a substantial influence on the expansion behavior of silicate layers as well as the evolution of the surface aggregates as a function of HDTMA surface coverage. Key words: adsorption, hexadecyltrimethylammonium, SWy montmorillonite, SAz montmorillonite

1. INTRODUCTION Organically modified layered silicates have long been used in numerous industrial applications. Understanding cationic surfactant adsorption on layer silicates is of great importance due to the widespread use of those compounds in both household and commercial activities (Theng, 1975; Hackett et al., 1998; Lee and Kim, 2002). In aqueous solutions, surfactants self-assemble into a variety of aggregate shapes including spherical and rod like micelles and planar bilayers (Wanless et al., 1997). In particular, electrostatics and surfactant association at the solid-liquid interface have been established as important interactions in ionic surfactant-charged solid systems (Chandar et al., 1987). Mica has been regarded as a typical hydrophilic substrate and has been implicitly used as a model for clay minerals and even for oxide minerals. Mica has been also employed as a model for clays in swelling studies. Recently, cylindrical aggregates typically observed on mica were not seen on either kaolinite or montmorillonite, where sphere-forming cationic surfactants in solution also formed spherical aggregates at the interface (Warr, 2000; Schulz and Warr, 2000). They found that adsorbed layers of cationic surfactants on mica had decidedly different morphologies from those on other typical clays, kaolinte and montmorillonite. For these kinds of experiments, optical reflectometry,
*Corresponding author: lsyblue@hanmir.com

204

Seung Yeop Lee and Soo Jin Kim

Table 1. Mineralogical properties of the montmorillonite samples used in this study. Sample designation SWy SAz
1 2

Specimen Low charge SWy-2 montmorillonite High charge SAz-1 montmorillonite

Cation exchange capacity (meq/100 g)1 85 123

Layer charge (cmolc/kg)2 0.87 0.97

Na2O/CaO ratio3 1.25 0.02

Obtained from Borden and Giese (2001). Obtained from Mermut and Lagaly (2001). 3 Obtained from Mermut and Cano (2001).

supernatant solution and the HDTMA concentrations were measured using a total organic carbon analyzer (TOC-5000A, Shimadzu, Tokyo, Japan). Adsorbed HDTMA was determined as the difference between the added HDTMA and that remaining in the solution. Additionally, the concentration of inorganic cations (Na+ and Ca2+) released from the montmorillonites via the exchange with HDTMA was measured using an inductively coupled plasma atomic emission spectrophotometer (ICPQ 1000, Shimadzu, Tokyo, Japan). 2.3. HDTMA Clay Preparation Organoclays were prepared by adding the amounts of HDTMA bromide equivalent to 0.02 to 1 times the CEC on SWy and SAz. The HDTMA bromide was dissolved in warm deionized water and added to clay suspensions. After mixing for 7 days, the HDTMA clay suspensions were centrifuged and washed gently with deionized water. 2.4. Xray Diffraction Analysis Ten grams of organoclay samples were dried at 25oC as oriented aggregates on glass slides. The d(001) values were recorded using CuK radiation from an automated Xray diffractometer (XRD) (Rigaku Geigerflex RAD3-C, Tokyo, Japan) with a graphite monochromator. The operating condition of XRD was at 40 kV/30 mA in a step scan mode. The scanning speed was 0.02o 2 s1 and the slit set was 1o, 0.15 mm, 0.5o. 2.5. Electron Microscopy The clay samples were dried at room temperature and coated with Au under vacuum in an argon atmosphere for scanning electron microscopy (SEM) studies (LEO 1455VP SEM, Cambridge, UK). 3. RESULTS 3.1. Adsorption Isotherms of HDTMA Adsorption isotherms of HDTMA by SWy and SAz montmorillonites are shown in Figure 1. Their adsorption isotherm patterns were significantly different at initial surfactant loadings. SWy showed that adsorption amount of HDTMA progressively diminished until about 0.4 CEC, and then increased as the loading amount increased (Fig. 1). Meanwhile, the adsorption isotherm by SAz showed a remarkable HDTMA adsorption increase at slight loading levels ( 0.2 CEC), and then a progressive decrease of sorption with the increase of HDTMA loading. When the loading amount of HDTMA was more than 0.6 CEC, the equilibrium surfactant concentration of SAz was almost the same as that of SWy. For exchangeable cations, as HDTMA sorption by montmorillonite proceeded, inorganic cations (Na+ and Ca2+) existing in the interlayers of the clay were released into water, resulting from the cation exchange reaction (Xu and Boyd, 1995) (Fig. 2). In that case, exchangeable inorganic cations on SWy were greatly released at low concentration of loaded HDTMA (Fig. 2A), accompanying Na+ ions released more than Ca2+ ions, while there was not a great release of inorganic cations from SAz (Fig. 2B). Additionally, the cation release curve for SAz was characterized as being linear with HDTMA loading, but the cations on SWy were considerably released more again from 0.4 CEC, which was

Fig. 1. Adsorption of HDTMA by SWy and SAz montmorillonites.

Study on HDTMA with reference montmorillonites

205

Fig. 3. d-spacings of HDTMASWy and SAz montmorillonites as HDTMA loading increased.

3.3. SEM Observation Figure 4 shows the morphologies of montmorillonites (SWy and SAz) modified with HDTMA and provides a basis for comparison between the two clays. Characteristically, the raw SWy (Na-montmorillonite) was more massive and adhesive (Fig. 4A) than the raw SAz (Ca-montmorillonite) (Fig. 4D) when dried after being dispersed in water, presumably because the Na+ interlayer ion was more easily hydrated than was Ca2+. According to Grim (1968), variations in the exchangeable cation carried by montmorillonites are reflected in their appearance and morphology. He showed that the differences between raw montmorillonites are due to variations in the attraction between the montmorillonite particles and in the relative strength of the attractive forces at the edges of the particles and on their basal planes, which are affected by the internal or external surface cations. Moll (2001) also reported that Na-montmorillonite is thin and more flexible than other layer silicates. This means that the exchangeable inorganic cations on clay minerals may play a role in the overall appearance or morphology of silicate layers. The morphology of raw SWy changed from entrailslike (i.e., broad undulating sheets) to cornflake-like as the HDTMA loading increased (Fig. 4AC). As shown in Figure 4C, micrographs of HDTMASWy revealed irregular flake-shaped aggregates which appeared to be stackings of units without regular outlines. Ca montmorillonite (SAz) appeared as irregular aggregates which increased in size as the HDTMA added increased (Fig. 4DF). The layers were relatively flat, except for a curling of the edges, and especially more aggregative at high loadings of HDTMA (Fig. 4F).

Fig. 2. The release curves of exchangeable inorganic cations (Na and Ca) from (A) SWy and (B) SAz montmorillonites during HDTMA loading.

ascribed to an enhanced Ca2+ release. 3.2. XRD Measurement The d-spacings of air-dried clays were not nearly changed at low HDTMA loadings ( 0.4 CEC) (Fig. 3). As the HDTMA loading increased from 0.4 CEC to 1 CEC, the dspacings of SWy increased from 14.2 to 18.5 , while those of SAz increased from 14.7 to 23.7 , indicating a transition from parallel to vertical orientation of alkyl ammonium (Lagaly, 1994) (Fig. 3).

206

Seung Yeop Lee and Soo Jin Kim

Fig. 4. SEM micrographs of montmorillonite particles before and after HDTMA addition. (A) Raw NaSWy, (B) 0.15 CEC HDTMA SWy, (C) 0.9 CEC HDTMASWy, (D) raw CaSAz, (E) 0.15 CEC HDTMASAz, and (F) 0.9 CEC HDTMASAz.

Many montmorillonite particles can therefore be regarded as aggregates, and indeed, a foliated, lamellar aggregate structure was often obvious at high HDTMA loadings. 4. DISCUSSION 4.1. Characteristic Desorption of Inorganic Cations by HDTMA Adsorption For ionic surfactants at very low concentrations, adsorption occurs via electrostatic interactions between the clay surface and micelle, leading to neutralization of the surface charge (Atkin et al., 2001). In this experiment, HDTMA was also adsorbed mostly by cation exchange as indicated by the release of considerable inorganic cations (Fig. 2), similar to that of HDTMA isotherm on montmorillonite performed

by Xu and Boyd (1995). Inasmuch as the HDTMA cations were strongly preferred by the exchange sites over the inorganic cations, the basal plane surface of clay sheets became hydrophobic and progressively aggregated depending on the degree of surfactant loading (Fig. 4). In SWy montmorillonite, the Na+ in the interlayer was easily exchanged with HDTMA (Fig. 2A), and this active HDTMA access to the interlayer Na+ sites might make its random distribution in the interlayers, which minimized the lateral interactions among the alkyl groups of neighboring adsorbed HDTMA cations. In addition, this preferential release of interlayer Na+ by HDTMA made the distinctive HDTMA adsorption isotherm shape different from that of SAz, showing the HDTMA equilibrium concentration increase with HDTMA loading up to about 0.4 CEC (Fig. 1). Meanwhile, the decrease in the equilibrium concentration of

Study on HDTMA with reference montmorillonites

207

HDTMA with HDTMA loading ( 0.2 CEC) on SAz (Fig. 1) may result from both the increased lateral interaction of adjacent HDTMA cations on external or edge sites (Xu and Boyd, 1995) and the rare opening of the high charge interlayer space of SAz. In other words, the lower exchange of interlayer Ca2+ ions by HDTMA probably maximized the binding strength of HDTMA by increasing the degree of the lateral interactions on the external surfaces. This was indicated as low aqueous phase HDTMA concentrations of SAz at the initial stage ( 0.2 CEC) (Fig. 1). Consequently, at these low loading levels, HDTMA had a higher affinity for high charged CaSAz montmorillonite rather than for low charged NaSWy montmorillonite. 4.2. Basal Spacings Change From Xray diffraction measurements on dry clay it has been shown that stepwise changes in basal spacing between the elementary sheets were obtained when a long-chain cation such as HDTMA was added in increasing amounts to montmorillonite up to the CEC. The d-spacings of HDTMA SWy were different from those of HDTMASAz (Fig. 3). For example, the d-spacings of HDTMASWy at 0.71.0 CEC (1818.5 ) were several angstroms less than that of HDTMASAz at the corresponding loadings, due to the lower layer charge of the clay (Lagaly, 1994). As shown in Figure 3, the d-spacings of HDTMASWy began to increase from about 0.4 CEC, which may be related to the gradual increase of HDTMA adsorption shown at the HDTMA adsorption isotherm (Fig. 1) as well as the substantial release of interlayer Ca2+ ions from that point (Fig. 2A). This means that as the opening of the inter-gallery of SWy silicate layers was facilitated by the continuous HDTMA intercalation, the remaining Ca2+ ions could come out of the interlayer. 4.3. Alteration of Montmorillonite Surfaces As previously mentioned, upon addition of a cationic surfactant in a quantity less than the CEC, into an aqueous clay suspension, exchange with initial counterions (inorganic cations) occurs and adsorbed surfactants neutralize the anionic clay sites. In that case, HDTMA cations are either intercalated between the elementary sheets or adsorb on the external or edge sites of clay. Subsequently, the neutral and hydrophobic character of the clay particle-water interface makes a flocculation of clay particles as shown in Figure 4. It appeared that the different charge densities and inorganic interlayer cation types of the clays have an effect on the evolution of the surface aggregates as a function of HDTMA surface coverage. The clay particles were highly flocculated in high charge HDTMACaSAz montmorillonite rather than in low charge HDTMANaSWy montmorillonite (Fig. 4). As previously noted, in HDTMACaSAz, the active lateral interaction between adsorbed HDTMA

cations at external sites played an important role on such a flocculation. It seems that the flocculation process of clay particles therefore depends to a great extent on the adsorbed surfactant position or distribution according to the layer charge densities or interlayer cation types of clay surface. 5. CONCLUSION From the observation of adsorption isotherms, XRD and SEM, the internal and external surfaces of montmorillonites with different interlayer inorganic cations were characterized as clay sheets with heterogeneous expansion at initial HDTMA loadings. In SWy montmorillonite, the Na+ in the interlayer was easily exchanged with HDTMA, and this active HDTMA access to the interlayer Na+ sites made its random distribution in the interlayers. However, restricted access of HDTMA to the interlayer of SAz montmorillonite resulted in significant lateral interactions among alkyl groups at outer surface or edges even at low HDTMA loadings, and this progressively made clay particles greatly aggregated as HDTMA was added. Consequently, besides the clay surface charge, the interlayer inorganic cations appeared to have a substantial influence on the expansion behavior of silicate layers as well as the evolution of the surface aggregates as a function of HDTMA surface coverage.
ACKNOWLEDGMENTS: This work was supported by BK21 Project of the Korean Ministry of Education.

REFERENCES
Atkin, R., Craig, V.S.J. and Biggs, S., 2001, Adsorption kinetics and structural arrangements of cetylpyridinium bromide at the silicaaqueous interface. Langmuir, 17, 61556163. Borden, D. and Giese, R.F., 2001, Baseline studies of The Clay Minerals Society Source Clays: cation exchange capacity measurements by the ammonia-electrode method. Clays and Clay Minerals, 49, 444445. Brahimi, B., Labbe, P. and Reverdy, G., 1992, Study of the adsorption of cationic surfactants on aqueous laponite clay suspensions and laponite clay modified electrodes. Langmuir, 8, 19081918. Chandar, P., Somasundaran, P. and Turro, N.J., 1987, Fluorescence probe studies on the structure of the adsorbed layer of dodecyl sulfate at the alumina-water interface. Journal of Colloid and Interface Science, 117, 3146. Ducker, W.A. and Wanless, E.J., 1999, Adsorption of hexadecyltrimethylammonium bromide to mica: nanometer-scale study of binding-site competition effects. Langmuir, 15, 160168. Grim, R.E., 1968, Clay Mineralogy. McGraw-Hill, New York, 596 p. Hackett, E., Manias, E. and Giannelis, E.P., 1998, Molecular dynamics simulations of organically modified layered silicates. Journal of Chemical Physics, 108, 74107415. Lagaly, G., 1994, Layer charge determination by alkylammonium inos. In: Mermut, A.R. (ed.), Layer Charge Characteristics of 2:1 Silicate Clay Minerals. The Clay Minerals Society, Aurora, Vol. 6, p. 146. Lee, S.Y. and Kim, S.J., 2002, Adsorption of naphthalene by HDTMA

208

Seung Yeop Lee and Soo Jin Kim

modified kaolinite and halloysite. Applied Clay Science, 22, 5563. Manne, S. and Gaub, H.E., 1995, Molecular organization of surfactants at solid-liquid interfaces. Science, 270, 14801482. Mermut, A.R. and Cano, A.F., 2001, Baseline studies of The Clay Minerals Society Source Clays: chemical analyses of major elements. Clays and Clay Minerals, 49, 381386. Mermut, A.R. and Lagaly, G., 2001, Baseline studies of The Clay Minerals Society Source Clays: layer-charge determination and characteristics of those minerals containing 2:1 layers. Clays and Clay Minerals, 49, 393397. Moll, W.F., 2001, Baseline studies of The Clay Minerals Society Source Clays: geological origin. Clays and Clay Minerals, 49, 374380. Schulz, J.C. and Warr, G.G., 2000, Adsorbed layer structure of cationic surfactants on clays (mica is not a typical substrate for adsorption studies). Langmuir, 16, 29952996.

Theng, B.K.G., 1975, The Chemistry of Clay-Organic Reactions. Wiley, New York, 343 p. Wanless, E.J., Davey, T.W. and Ducker, W.A., 1997, Surface aggregate phase transition. Langmuir, 13, 42234228. Warr, G.G., 2000, Surfactant adsorbed layer structure at solid/solution interfaces: impact and implications of AFM imaging studies. Current Opinion in Colloid and Interface Science, 5, 8894. Xu, S. and Boyd, S.A., 1995, Cationic surfactant adsorption by swelling and nonswelling layer silicates. Langmuir, 11, 25082514. Yui, T., Yoshida, H., Tachibana, H., Tryk, D.A. and Inoue, H., 2002, Intercalation of polyfluorinated surfactants into clay minerals and the characterization of the hybrid compounds. Langmuir, 18, 891896. Manuscript received June 4, 2003 Manuscript accepted August 5, 2003

Anda mungkin juga menyukai