Vous êtes sur la page 1sur 6

Journal of Colloid and Interface Science 299 (2006) 910–915

www.elsevier.com/locate/jcis

A phenol-induced structural transition in aqueous cetyltrimethylammonium


bromide solution
J.P. Mata a,∗ , V.K. Aswal b , P.A. Hassan c , P. Bahadur d
a Department of Chemistry, VNSG University, Surat 395007, India
b Solid State Physics Division, BARC, Mumbai 400 085, India
c Novel Materials and Structural Chemistry Division, BARC, Mumbai 400 085, India
d Chemical Engineering Department, Lamar University, Beaumont, TX, USA

Received 23 November 2005; accepted 15 February 2006


Available online 10 March 2006

Abstract
The effect of phenol on the structure of micellar solution of a cationic surfactant, cetyltrimethylammonium bromide (CTAB) was investigated
using viscosity, dynamic light scattering (DLS), small angle neutron scattering (SANS) and nuclear magnetic resonance (NMR) techniques. The
relative viscosity and apparent hydrodynamic diameters of the micelles in CTAB solution increase initially and then decrease with addition of
phenol. SANS studies indicate a prolate ellipsoidal structure of the micelles. The axial ratio of the prolate ellipsoidal micelles increases and then
decreases with addition of phenol, consistently with DLS and viscosity measurements. NMR studies confirm the solubilization of phenol to the
palisade layer and growth of the micelles at high concentration of phenol as revealed from the broadening of peaks.
© 2006 Elsevier Inc. All rights reserved.

Keywords: Viscosity; DLS; SANS; Micellar transition

1. Introduction ution or condensation on the micellar surface [14]. Weican et


al. [15], by means of laser light scattering, 1 H NMR and flu-
Surfactants are widely used in many industrial and commer- orescence measurements probe reported that CTAB (10 mM)
cial applications. Besides their surface active nature, cationic forms rod-like micelles at 0.1 mol L−1 KBr and the worm-like
surfactants offer additional advantages owing to their antibac- micelles are formed at above 0.2 mol L−1 KBr. Several tech-
terial properties [1–4]. The sphere-to-rod transition in micelles niques, viz., light scattering, NMR spectroscopy, small-angle
of quaternary ammonium type cationic surfactants in water has X-ray/neutron scattering, electrical conductivity, viscosity, sol-
been studied extensively [5–8]. While the micelles are gener- ubilization, fluorescence quenching, osmometry, ultrasonic ve-
ally spherical at concentration near the critical micelle con- locity, etc., have been employed to examine sphere-to-rod tran-
centration (CMC) of surfactants, rod-like micelles do form at sitions in ionic micelles. The presence of neutral salts [6,9–11],
higher surfactant concentration. Nature and size of the hydro- slightly polar cosolutes like medium chain alcohols [16,17,33]
carbon tail, polar head group and the counterion of the sur- and amines [18] have shown a marked shift in phase transi-
factant strongly influence sphere-to-rod transitions. It is well tion to lower concentration range. 1 H NMR is widely used to
known that salts such as KBr and sodium salicylate (NaSal) in- investigate the solubilization of additives in micellar solutions
[34,35]. Various models for the loci and orientation of organic
duce pronounced growth of CTAB micelles due to screening of
additives in micelles by using 1 H NMR spectroscopy are well
the electrostatic repulsion of head groups [9–13]. The size and
reported [34]. Many reports show that when solubilizate mole-
shape of the micelles are influenced by the counterion distrib-
cules possess a phenyl group and up-field shift of surfactant
protons in the 1 H NMR spectra is observed with an increase of
* Corresponding author. Fax: +91 261 2256012. solute concentration, the chemical shifts linearly decreased as
E-mail address: jitendramata@yahoo.com (J.P. Mata). the solute concentration increased [36,37]. Cationic surfactant
0021-9797/$ – see front matter © 2006 Elsevier Inc. All rights reserved.
doi:10.1016/j.jcis.2006.02.032
J.P. Mata et al. / Journal of Colloid and Interface Science 299 (2006) 910–915 911

such as CTAB is well known for antibacterial properties while output power of 2 W. The average decay rate (Γ ) of the elec-
phenol solutions are disinfectants/antiseptics. Typical formula- tric field autocorrelation function, g 1 (τ ), was estimated us-
tion containing CTAB and phenol may serve specialized cleans- ing the method of cumulants. The apparent diffusion coeffi-
ing action, e.g., in surgical instruments, lavatories, etc. Since cients (D) of the micelles were obtained from the relation
phenol is sparingly soluble in water (∼7.7%, w/w) and the Γ = Dq 2 (q is the magnitude of the scattering vector given
derivatives of phenol like cresol, chlorocresols, chloroxylenols, by q = [4πn sin(θ/2)]/λ, n being the refractive index of the
and thymol are even less soluble, their use as disinfectants has solvent, λ, the wavelength of laser light and θ is the scatter-
led to the need for formulation in surfactant solutions [19]. ing angle) and the corresponding hydrodynamic diameters (d)
The effect of phenols on the alkali metal soap solution has were calculated using the Stokes–Einstein relationship. For all
been studied by Angelesen et al. [20,21]. Particularly cationic the solutions, Γ varies linearly with q 2 , indicating translational
surfactants that are positively charged form insoluble complex diffusion of the scatterers.
with negatively charged phenol in aqueous solution at high con-
centrations while at lower concentration they form viscous so- 2.2.3. Small angle neutron scattering
lutions. Cationic surfactants show spectacular changes in their The solutions were held in a 0.5-cm thick quartz cell with
solution properties in presence of phenol derivatives, particu- Teflon stoppers. The SANS experiments were performed using
larly iodophenol [8], naphthols [22,23] and salicylic acid [8,22, a SANS diffractometer at the Dhruva reactor, BARC, Trombay,
23] and its sodium salts, [24] even at very low concentration. India [25]. The diffractometer uses a polycrystalline BeO fil-
The effect of phenol on the micellization and structural tran- ter as a monochromator. The mean wavelength of the incident
sition of cetyltrimethyammonium bromide (CTAB) micelles neutron beam is 5.2 Å with a wavelength resolution (λ/λ) of
was examined using viscosity, dynamic light scattering (DLS), approximately 15%. The angular distribution of the scattered
small angle neutron scattering (SANS) and nuclear magnetic neutron was recorded by a linear 1-m-long He3 position sensi-
resonance (NMR) measurements. The results are interpreted in tive detector (PSD). The data were recorded in the Q range of
terms of solubilization of phenol and consequent changes in 0.02–0.24 (Å)−1 . All the measured SANS distributions were
the structure of micelles. Effect of salt NaBr on CTAB–phenol corrected for the background and solvent contributions. The
complex has been also checked. Systematic studies help to un- data were normalized to the cross-sectional unit using standard
derstand the CTAB–phenol complex behavior. procedures [25].

2. Materials and methods 2.2.4. Nuclear magnetic resonance


1 H NMR spectra were recorded with a Varian 300 MHz in-
2.1. Materials strument. All spectra for solutions of various compositions were
recorded in one series at around 25 ◦ C.
The cationic surfactant cetyltrimethylammonium bromide
(CTAB) was purchased from Lancaster, Leeds, England. The 3. Results and discussion
purity of CTAB was ascertained by surface tension measure-
ments. No minimum was observed in the surface tension–log Fig. 1 shows the variation of relative viscosities of cetyltri-
concentration plot. Phenol was from E. Merck (Anala R grade) methylammonium bromide solutions with concentration at var-
and salt NaBr was from Fluka, Switzerland, and were used as ious temperatures. A sharp increase in the relative viscosity as
received. Triply distilled water from an all-Pyrex glass appara-
tus was used for the preparation of solutions for viscosity mea-
surements. For DLS measurement Milli-Q water having spe-
cific resistance 18.2 M cm was used. The samples for SANS
and NMR experiments were prepared in D2 O.

2.2. Methods

2.2.1. Viscosity
The viscosity measurements were carried out using an Ubbe-
lohde suspended level capillary viscometer. The viscometer was
suspended vertically in a thermostat with a temperature stabil-
ity of ±0.1 ◦ C. The flow time for constant volume of solution
through the capillary was measured with a calibrated stopwatch.

2.2.2. Dynamic light scattering


DLS measurements were performed using a Malvern 4800
Autosizer employing 7132 digital correlator. The light source Fig. 1. Relative viscosity of C16 TAB as a function of its concentration at differ-
was argon ion laser operated at 514.5 nm with a maximum ent temperatures: (!) 30, (Q) 40, (a) 50, (E) 60 ◦ C.
912 J.P. Mata et al. / Journal of Colloid and Interface Science 299 (2006) 910–915

Fig. 2. Relative viscosity of 50 mM CTAB as a function of phenol concentration Fig. 3. Hydrodynamic diameters and diffusion coefficient for CTAB micelles in
at different temperatures having fixed concentration of 0.01 M NaBr: (2) 25, water containing 0.01 M NaBr at 25 ◦ C.
(!) 30, (Q) 40, (a) 50, (E) 60 ◦ C.

a function of CTAB concentration is observed at different tem-


peratures due to the sphere-to-rod transition of micelles [4–8].
The increase in viscosity is less pronounced at higher tempera-
ture, consistently with the expectation that smaller micelles are
formed at higher temperatures. The concentration at which a
structural change takes place is shifted towards higher CTAB
concentration when the temperature is increased.
The relative viscosities for 50 mM CTAB as a function of
added phenol concentration at different temperatures having
constant concentration of 0.01 M NaBr are shown in Fig. 2.
The curve shows that at 25 ◦ C, the relative viscosity increases
drastically after a certain amount of phenol added. However, the
relative viscosities showed a decrease with further addition of
phenol, reflecting a maximum in the viscosity at intermediate
phenol concentrations. Such a decrease in the viscosity has also
been observed by Tominaga et al. [26] for the CTAB–hexanol– Fig. 4. Hydrodynamic diameters and diffusion coefficient for CTAB (50 mM)
water system. The viscosity maximum decreases with increase micelles at different concentrations of phenol in water containing 0.01 M NaBr
in temperature. As shown in Fig. 2, an increase in temperature at 25 ◦ C.
favors spherical micelles as observed in CTAB–water system in
Fig. 1. Fig. 5 shows the effect of NaBr on the apparent diameter of
Fig. 3 shows the variation in apparent hydrodynamic diam- CTAB micelles in the presence and absence of phenol at 25 ◦ C.
eter and diffusion coefficient of CTAB micelles in water con- The hydrodynamic diameter of the micelles in the presence of
taining 0.01 M NaBr. The diffusion coefficient decreases and phenol is much higher than that in the absence of phenol at all
the apparent diameter increases with an increase in CTAB con- NaBr concentrations. The size of CTAB micelles in the absence
centration. and presence of phenol increases with NaBr concentration. The
This is consistent with the relative viscosity measurements observed increase in apparent diameter of the micelles with
suggesting an increase in micellar dimension with concentra- phenol possibly arises from the incorporation of phenol mole-
tion, apparent hydrodynamic diameters of 50 mM CTAB at dif-
cules into the micelles and consequent their growth. The effect
ferent concentrations of phenol in the presence of NaBr (fixed
is more pronounced in the presence of electrolytes such as NaBr
0.01 M) are shown in Fig. 4.
due to screening of the electrostatic interaction of surfactant
The initial increase and the subsequent decrease of the ap-
parent diameter of CTAB micelles with addition of phenol is head groups. A decreased surface charge of the micelles and
consistent with relative viscosity measurements. For 100 mM screening of interaction by electrolytes leads to a decrease in
CTAB in the presence of phenol, the observed variation in dif- the head group area of surfactants, which in turn increases the
fusion coefficient is less pronounced. This is possibly due to packing parameter of the micelles. In other words, decreased
the presence of strong intermicellar interactions at high volume head group repulsion favors closer packing of the surfactant
fractions of micelles that will interfere with the DLS measure- monomers and hence induces a sphere-to-rod transition in mi-
ments. celles [27–29].
J.P. Mata et al. / Journal of Colloid and Interface Science 299 (2006) 910–915 913

The change in size and shape of micelles with addition of The differential scattering cross section per unit volume
phenol is further confirmed from SANS measurements. The (dΣ/dΩ) as a function of scattering vector Q, for a system of
SANS data for 50 mM CTAB solution containing varying monodisperse interacting micelles (dΣ/dΩ) is given by [27]
amount of phenol in 0.01 M NaBr at temperature 30 ◦ C are

reported in Fig. 6. The SANS distribution for 50 mM CTAB (Q) = n(ρm − ρs )2 V 2

in the presence of considerably lesser phenol concentration    2  
(0.025%) shows a well-defined peak at the wave vector transfer × F (Q)2 + F (Q) S(Q) − 1 + B, (1)
Q ∼ 0.07 Å−1 . This correlation peak is an indication of strong
repulsive interaction between the positively charged CTAB mi- where n denotes the number density of the micelles, ρm and ρs
celles. The peak occurs at Qm ∼ 2π/d, where d is the average are, respectively, the scattering length densities of the micelle
distance between the micelles and Qm is the value of Q at the and the solvent and V is the volume of the micelle. F (Q) is the
peak position. As shown in the figure, cross section increases single particle form factor and S(Q) is the interparticle struc-
and the peak position shifts to lower Q values with the in- ture factor. B is a constant term that represents the incoherent
crease in phenol concentration. This reflects an increase in the scattering background, which is mainly due to hydrogen in the
dimension of the micelles in the presence of phenol. The broad- sample.
ening of the correlation peak suggests a decrease of the effective The prolate ellipsoidal shape (a = b = c) of the micelles is
charge per micelle. Quantitative estimate of the micellar para- widely used in the analysis of SANS data because it also repre-
meters were obtained by model fitting the data. sents the other different possible shapes of the micelles such as
spherical (a = b) and rod-like (a  b). For such an ellipsoidal
micelle,
3(sin x − x cos x)
F (Q, μ) = , (2)
x3
 1/2
x = Q a 2 μ2 + b2 (1 − μ2 ) , (3)
where a and b are, respectively, the semimajor and semiminor
axes of the ellipsoidal micelle and μ is the cosine of the angle
between the direction of a and the wave vector transfer Q.
We have calculated S(Q) as derived by Hayter and Penfold
[30] from the Ornstein–Zernike equation and using the rescaled
mean spherical approximation. The micelle is assumed to be a
rigid equivalent sphere of diameter σ = 2(ab2 )1/3 interacting
through a screened Coulomb potential.
The dimensions of the micelle, aggregation number and
the fractional charge have been determined from the analysis.
Fig. 5. Hydrodynamic diameters of CTAB (0.050 M) micelles with and with- The semimajor axis (a), semiminor axis (b = c) and the frac-
out phenol (0.75%, v/v) at different concentrations of NaBr at 25 ◦ C: (") with tional charge (α) are the parameters in analyzing the SANS
phenol, (1) without phenol.
data. The aggregation number is calculated by the relation N =
4πab2 /3v, where v is the volume of the surfactant monomer.
Throughout the data analysis, corrections were made for in-
strumental smearing [31]. The parameters in the analysis were
optimized by means of nonlinear least-square fitting program
and the errors of the parameters were calculated by the standard
methods used [32]. The micellar parameters obtained from the
fit are summarized in Table 1, aggregation number for 50 mM

Table 1
Micellar parameters for 50 mM CTAB in presence of varying concentrations of
phenol at fixed (0.01 M) NaBr concentration at 30 ◦ C
[Phenol] Aggregation Fractional Semiminor Semimajor Axial ratio
(%, v/v) number, N charge, α axis b = c (Å) axis a (Å) a/b
0.00 145 (150) 0.26 24.0 33.7 1.40
0.25 313 0.10 21.3 105.5 4.95
0.50 773 0.10 22.1 271.9 12.30
0.75 647 0.10 21.6 263.1 12.18
1.00 561 0.10 21.3 258.0 12.11
Fig. 6. SANS distribution for CTAB (0.050 M) micelles at different concen-
2.00 335 0.10 20.0 237.7 11.89
trations of phenol in water containing 0.01 M NaBr. The fit is shown after
3.00 94 0.10 19.8 87.0 4.39
convoluting with the resolution of the Dhruva instrument at 30 ◦ C.
914 J.P. Mata et al. / Journal of Colloid and Interface Science 299 (2006) 910–915

CTAB (50 mM) solution containing different phenol concen-


trations ((a) no phenol, (b) 0.75% phenol, (c) 1.00% phenol,
and (d) 2.00% phenol). From the spectra, it is evident that
α-CH2 β-CH2 protons of CTAB were shifted up-field on in-
creasing phenol concentration. On the other hand, methylene
and terminal methyl groups in CTAB were not affected by sol-
ubilizate molecules, suggesting that time average solubilization
locus of phenol is in the micelle–water interface. For the pro-
tons from the benzene ring, similar phenomena were observed,
but the extent of the shift was very small. In addition to the
up-field shifts, changes in peak shape were also observed at
solubilizate concentration near the peak in viscosity. At 0.75%
phenol, the viscosity of the solution increased noticeably and
peaks began to broaden as shown in Fig. 7. This suggests a
significant growth of the micelles and consequent increase of
rotational correlation time of the micelles at this concentra-
tion. At this solubilizate concentration, the adjacent two peaks
of β-CH2 and methylene protons were merged into one, and
they were not resolved anymore. Despite the shape change, the
peaks of the methylene and methyl groups were not shifted,
which suggests that phenol could not penetrate to the cores of
micelles even at high concentration. This is consistent with the
reports in literature. Kandori et al. reported that additives such
as phenol, which were solubilized in the palisade layer of dode-
cyltrimethylammonium bromide micelles, had a much greater
effect on the surfactant aggregation number and micellar shape
than benzene [38].

4. Conclusions

The change in the structure of CTAB micelles with addition


of phenol was monitored by viscosity, dynamic light scattering
and small angle neutron scattering. Addition of small amounts
of phenol increases the size of the aggregates due to solubi-
lization of phenol in the micelles and changes in the packing
of surfactants. However, high concentrations of phenol destabi-
lize the formation of micelles thereby decreasing the micelle
size. SANS studies reveal that the micelles can be modeled
as prolate ellipsoids with the semiminor axis equal to the ra-
dius of the spherical micelles without phenol. A typical peak
behavior is observed due to synergic condensation of phenol
to CTAB micelles. NMR studies show that phenol is located
Fig. 7. 1 H NMR spectra for aqueous CTAB (50 mM) with (a) no phenol,
at micelle–water interface in aqueous CTAB micellar solution.
(b) 0.75% phenol, (c) 1.00% phenol, (d) 2.00% phenol.
The solubilizate could not penetrate to the micellar core even at
high concentration, which could be attributed to the polar char-
CTAB micelles (150) is well reported in our previous stud- acteristics of the molecules.
ies. The results suggest that CTAB micelles undergo a prolate
ellipsoidal growth with the addition of phenol. The micelle ag-
Acknowledgment
gregation number increases initially and then decreases with
increasing phenol concentration. The fractional charge on the
We are thankful to Professor P. Bahadur, Department of
micelles decreased considerably on addition of phenol and the
Chemistry, VNSGU, Surat, for his guidance.
surface charge is insensitive to the concentration of phenol. The
observed variation in micelle aggregation number is consistent
References
with the results from DLS and viscosity measurements.
The locus of solubilization of phenol, and consequent [1] E. Jungerman, Cationic Surfactants, Dekker, New York, 1969.
changes in structure of the micelles, is inferred from NMR [2] J. Cross, E.J. Singer, Cationic Surfactants: Analytical and Biological Eval-
measurements. Fig. 7 shows 1 H NMR spectra for aqueous uation, Dekker, New York, 1994.
J.P. Mata et al. / Journal of Colloid and Interface Science 299 (2006) 910–915 915

[3] P.M. Holland, D.N. Rubingh (Eds.), Cationic Surfactants: Physical Chem- [22] T. Nash, J. Colloid Sci. 13 (1958) 134.
istry, Dekker, New York, 1991. [23] T. Nash, J. Colloid Sci. 14 (1959) 59.
[4] J.M. Richmond, Cationic Surfactants: Organic Chemistry, Dekker, New [24] N.C. Verma, B.S. Valaulikar, C. Manohar, J. Surf. Sci. Technol. 3 (1987)
York, 1990. 19.
[5] T. Imae, S. Ikeda, Colloid Polym. Sci. 265 (1987) 1090. [25] V.K. Aswal, P.S. Goyal, Curr. Sci. 79 (2000) 947.
[6] R. Zielinski, S. Ikeda, H. Nomura, S. Kato, J. Chem. Soc. Faraday Trans. [26] T. Tominaga, T.B. Stem, P. Solyom, J. Colloid Interface Sci. 35 (1971)
I 85 (1989) 1619. 519.
[7] M. Sasaki, T. Imae, S. Ikeda, Langmuir 5 (1989) 211. [27] S.H. Chen, T.L. Lin, in: D.L. Price, K. Skold (Eds.), Methods of Experi-
[8] H. Hirata, Y. Sakaiguchi, J. Colloid Interface Sci. 121 (1988) 300. mental Physics, vol. 23B, Academic Press, New York, 1987, p. 489.
[9] (a) S. Ozeki, S. Ikeda, J. Colloid Interface Sci. 77 (1980) 219; [28] J.N. Israelachvili, Intermolecular and Surface Forces, Academic Press,
(b) S. Ozeki, S. Ikeda, J. Colloid Interface Sci. 87 (1982) 424. New York, 1992.
[10] T. Imae, S. Ikeda, J. Phys. Chem. 90 (1986) 5216. [29] S.H. Chen, E.Y. Sheu, J. Kalus, H. Hoffmann, J. Appl. Crystallogr. 21
[11] D. Nguyen, G.L. Bertrand, J. Phys. Chem. 96 (1992) 1994. (1988) 751.
[12] M. Cates, S.J. Candau, J. Phys. Condens. Matter 2 (1990) 6889. [30] J.B. Hayter, J. Penfold, Colloid Polym. Sci. 261 (1983) 1022.
[13] H. Rehage, H. Hoffman, Mol. Phys. 74 (1991) 933. [31] J.S. Pedersen, D. Posselt, K. Mortensen, J. Appl. Crystallogr. 23 (1990)
[14] V.K. Aswal, P.S. Goyal, Phys. Rev. 61 (2000) 2947. 321.
[15] Z. Weican, L. Ganzuo, M. Jianhai, S. Qiang, Z. Liqiang, L. Haojun, [32] P.R. Bevington, Data Reduction and Error Analysis for Physical Sciences,
W. Chi, Chin. Sci. Bull. 45 (2000) 1854. McGraw–Hill, New York, 1969.
[16] S. Backlund, G. Bostrom, H. Hoiland, M. Ruths, in: Proc. 5th Conf. Col- [33] A. Stradner, O. Glatter, P. Schurtenberger, Langmuir 16 (2000) 5351.
loid Chem., Hungary, 1988. [34] B.J. Kim, S.S. Im, S.G. Oh, Langmuir 17 (2001) 565.
[17] Ch.D. Prasad, H.N. Singh, Colloid Surf. 50 (1990) 37. [35] R. Zana (Ed.), Surfactant Solutions: New Methods of Investigation,
[18] Ch.D. Prasad, H.N. Singh, Colloid Surf. 59 (1991) 27. Dekker, New York/Basel, 1987, p. 295.
[19] D. Attwood, A.T. Florence, in: Surfactant Systems, Their Chemistry, Phar- [36] C.A. Bunton, C.P. Cowell, J. Colloid Interface Sci. 122 (1988) 154.
macy and Biology, Chapman & Hall, London, 1983, p. 304. [37] J.H. Fendler, E.J. Fendler, G.A. Infante, P.S. Shih, L.K. Patterson, J. Am.
[20] E. Angelescu, D.M. Popescu, Kolloid Z. 51 (1930) 336. Chem. Soc. 97 (1975) 89.
[21] E. Angelescu, T. Manolescu, Kolloid Z. 94 (1941) 319. [38] K. Kandori, J.M. Robert, R.S. Schechter, J. Phys. Chem. 93 (1989) 1506.

Vous aimerez peut-être aussi