Вы находитесь на странице: 1из 6

Journal of Molecular Liquids 137 (2008) 74 79 www.elsevier.

com/locate/molliq

Determination of the physico-chemical parameters and aggregation number of surfactant in micelles in binary alcoholwater mixtures
S. Javadian a,, H. Gharibi a,b , B. Sohrabi a , H. Bijanzadeh a , M.A. Safarpour c , R. Behjatmanesh-Ardakani d
c

Department of Chemistry, Tarbiat Modarres University, P.O. Box 14155-4838, Tehran, Iran Iranian Information and Documentation Center, IRANDOC, P.O. Box 13185-1371, Tehran, Iran Department of Chemistry, College of Chemistry, Iran University of Science and Technology, P.O. Box 16765-163, Tehran, Iran d Department of Chemistry, University of Payam Noor, Ardakan, Iran
b

Received 21 April 2006; accepted 10 April 2007 Available online 3 May 2007

Abstract The effect of ethanol on the micellization of the anionic surfactant, sodium dodecyl sulfate (SDS), was investigated using potentiometric and pulsed field gradientNMR spectroscopic techniques. Potentiometry studies showed that the critical micelle concentration (CMC) of surfactant decreases to a minimum value at around 10% ethanol (v/v) because of the co-surfactant effect. The mean values of the hydrodynamic radius, Rh, and aggregation number, NA, of micelles were determined by a combination of viscosity and self-diffusion coefficient measurements. The viscosity of the micellar solutions was approximately independent of ethanol concentration, indicating that the presence of the alcohol does not induce substantial changes in the micellar structure. The number of attachment of ethanol in the micellar core and the degree of dissociation of the counterion were evaluated using a potentiometric technique for alcohol concentrations between 5% and 40% v/v. In addition, an explicit expression is developed for the average size of micelles formed in aqueous solvents containing alcohol. 2007 Elsevier B.V. All rights reserved.
Keywords: Surfactant; Critical micelle concentration (CMC); Sodium dodecyl sulfate (SDS); Aggregation number; Micelle

1. Introduction Addition of a non-electrolyte additive to a surfactant solution greatly affects the nature of that solution. Many studies have examined the modulation of surfactant solutions by nonelectrolyte additives, in particular the effects of additives such as alcohols on the critical micelle concentration (CMC) and aggregation number of both anionic and cationic surfactants [17]. It is now generally believed that the major driving force for micellization in aqueous solutions is entropy. When an ionic surfactant molecule is placed in aqueous solution, it is thought that the water molecules surrounding the hydrocarbon portion of the

Corresponding author. Fax: +98 21 88009730. E-mail addresses: javadians@yahoo.com, javadian_s@modares.ac.ir (S. Javadian). 0167-7322/$ - see front matter 2007 Elsevier B.V. All rights reserved. doi:10.1016/j.molliq.2007.04.001

surfactant molecule take on a more structured configuration than is found in the bulk aqueous solution. Thus aggregation of the surfactant molecules into micelles will reduce the quantity of water molecules in highly structured environments, thereby increasing the entropy. One method of investigating the hydrophobic interaction between the hydrocarbon chains of a surfactant and solvent water molecules is to alter the structure of the water by modifying the solvent composition, and then observe the effect of this modification on micellar aggregation. The simplest way to alter the water structure is to add a co-solvent such as an alcohol, dioxane, or glycerol. At surfactant concentrations above the CMC, the micellar aggregation number, NA, is an important characteristic of a surfactant solution. Traditionally, several approaches have been used to determine this parameter, including small-angle neutron scattering (SANS) [8], fluorescence [9,10], and NMR selfdiffusion coefficient measurements [11,12]. The size, shape and

S. Javadian et al. / Journal of Molecular Liquids 137 (2008) 7479

75

dispersity of micelles in a solution are sensitive to numerous internal (e.g., hydrophobic structure and head group type) and external (e.g., temperature, pressure, pH, and electrolyte concentration) parameters. Recently, water/alcohol/surfactant mixtures have been extensively used in the study of chemical equilibrium and reaction rates [1317]. In the present study, we used the ion selective electrode (ISE) technique to determine the characteristics of sodium dodecyl sulfate (SDS) in solvents comprised of a mixture of ethanol (EtOH) and water. In addition, the sizes and aggregation numbers of the micelles in these solutions were determined by measuring the self-diffusion coefficient and viscosity as a function of solvent composition and total surfactant concentration. Finally, we developed an explicit expression for the growth of SDS micelles in the presence of alcohol. 2. Experimental section 2.1. Materials EtOH and SDS were obtained from Merck (purity N 99%) and were used as received. D2O (99.95% purity) and deuterated ethanol were obtained from Fluka Co. 2.2. Methods 2.2.1. EMF measurement An anionic surfactant-selective electrode and a sodium ion selective electrode were used to measure the concentrations of monomeric anionic surfactant and counterion. The anionic surfactant-selective electrode was fabricated as described previously [18,19]. The experimental errors in the potentiometry measurements were 0.2 mV. For each solution, the surfactant ion (SD) and counterion (Na+) activities were determined using the following two cells: Surfactant electrodej test solutionj reference electrode
calomel

2.2.3. Viscosity The viscosities of the mixed surfactant solutions were measured using semi-micro-Cannon Ubbelohde capillary viscometers immersed in a water bath of temperature 25 0.1 C. The time required for the surfactant solution to flow through the capillary was measured with an accuracy of 0.01 s. The viscosity of each sample was measured at least five times. The standard deviation in these measurements was less than 1.5% and the relative standard deviation or coefficient of variation was less than 0.5. The viscosity was measured as a function of total surfactant concentration for solvents with various EtOH/ water ratios. 2.2.4. Density The densities of SDS in the various water/ethanol mixtures were measured using a 25 ml pycnometer at 25 C. To determine the density and viscosity data for the surfactant/D2O system, and to consider the isotope effect, the density and viscosity data for the surfactant/H2O system were multiplied by the ratios D2O/H2O or D2O/H2O. Substitution of D for H in the solvent could potentially alter the structural properties of the micellar aggregates. In fact, D2O is thought to be slightly more structured than H2O [22]. However, Berr showed that these differences are sufficiently small that they can be neglected [23]. 3. Empirical expression for micelle size in aqueous solvent containing alcohol Through experiments on a series of sodium alkyl sulfates, Huisman obtained the following empirical relationship amongst the molecular weight of the surfactant monomer, Mw, the aggregation number, NA, and the CMC: log NA k1 log CMC log Mw k2 3

Cell1 and Sodium ion electrodej test solutionj reference electrode


calomel

where k1 and k2 are constants. Fitting data at a fixed temperature for a homologous series of n-alkyl sodium sulfates yields k1 = 0.336 and k2 = 3.465 [24]. This relationship has been generalized to above the CMC by replacing the CMC with the concentration of a free monomer, m1 [25]: log NA k1 log m1 log Mw k2 4

Cell2 All experiments were performed at 25 C. 2.2.2. NMR measurements NMR self-diffusion studies were performed on a Bruker 500 NMR spectrometer at 25 C. The pulsed field gradient technique developed by Stejskal and Tanner was used [20]. In this technique, micelle self-diffusion is monitored based on signals from trace amounts of tetramethylsilane (Me4Si) added to the solution, which are assumed to be completely solubilized in the micellar phase. The Core program was used to calculate the self-diffusion coefficients [21].

Eqs. (3) and (4) are independent of the alkyl chain length. According to the charged phase separation model in aqueous solution [26]: nDS mNA DSn Nam nm Thus: n log m1 m log m2 Constant 6 5

where m1 and m2 are the free concentrations of surfactant molecules and counterions, respectively. Also, based on a theoretical consideration of the thermodynamics of micelle formation by ionic surfactants, Hall suggested the following equation [27]: log m1 k3 b log m2 7

76

S. Javadian et al. / Journal of Molecular Liquids 137 (2008) 7479

Table 1 Variation of micellization parameters of SDS (CMC, and N) and self-diffusion coefficient of monomer of SDS with ethanol concentrations at 25 C Percentage of ethanol (% v/v) 0 10 20 30 40
a b c d

Combining Eq. (4) with Eqs. (16) and (8) then yields: log NA K V g Vlog aCT bm1 17

CMC (EMF) (mmol/l)

CMC (NMR) (mmol/l) 8.3 6.3 7.2 14.1

D 1010 (m2/s) 5.0 (4.8 d) 3.5 2.0 1.9 1.84

0.84(0.90 a) 8.7(8.5 b) 0.69(0.63 a) 0.47(0.53 a) 0.38(0.47 a) 0.28(0.3 a) Ref. [29]. Ref. [39]. Refs. [40,41]. Ref. [42]. 5.9 7.4 17.0 42.0

69.1 (65.9 c) 76.5 36.4 10.9

where = k1 and K = k1Ki log Mw + k2. K and depend on the alcohol concentration. The concentrations m1 and m2 can be estimated from EMF measurements of cells (1) and (2) using the following equations:
=RT eF E1 E1 V m1 g1
-

18

and m2 where ln g1 p X z2 iA I p ; I m i z2 i 1 I 20
=RT eF E2 E2 V g2
-

19

where k3 is a constant and is the degree of attachment of the counterion. Substituting the conventional pseudo-phase ion exchange mass balance relationship [25,28]: m2 aCT m1 m1 8

where = 1 is the apparent degree of counterion dissociation, into Eq. (7), we obtain: log m1 k3 b log faCT m1 m1 g Then, combining Eq. (9) with Eq. (4) yields: log NA k1 fk3 b log aCT bm1 g log Mw k2 which can be written as: log NA K g log aCT bm1 11 10 9

where = k1 and K = k1k3 log Mw + k2. For alcoholwater mixtures, on the other hand, the mixed micelles are formed according to the following process [1,29]: nDS mNa pROH DSn Nam ROHp nm with n log m1 m log m2 p log aAlc constant: Eq. (13) can be written in the form [1]: log m1 k4 p m log aAlc log m2 n n 14 13 12

First, m1 and m2 are estimated by assuming 1 = 1, and then these estimates are used to evaluate 1 from Eq. (20), where E1 1 2 and E2 are the EMF values of cells (1) and (2), E0 and E0 are the standard potentials of cells (1) and (2), 1 and 2 are the activity coefficients of the ionic surfactant and counterion, respectively, F is Faraday's constant, R is the gas constant, and T is the absolute temperature. The values of m1 and m2 can be calculated iteratively using Eqs. (18)(20). Ki and can be determined from the intercept and slope, respectively, of a plot of log m1 versus log m2 [Eq. (16)]. The values for alcoholwater mixtures with alcohol contents of 0 40% v/v are shown in Table 1. According to Eq. (15), the variation of Ki with log aAlc should give a straight line of slope (p / n), which is defined as the degree of alcohol attachment. From a plot of Ki versus log aAlc (Fig. 1), we found that the degree of ethanol attachment is 1.50. The hydrodynamic radius of a micelle can be calculated by the StokesEinstein equation [30,31]: Rh kB T 6pgD 21

where kB is the Boltzmann constant, T is the absolute temperature, D is the diffusion coefficient and is the viscosity of the medium, which is taken as that of the ternary surfactant/

where p / n and m / n = are the number of attachment of alcohol and counterion, respectively. Given that in each series of experiments the activity of the alcohol remains constant, Eq. (7) is equivalent to Eq. (10) with Ki k4 p log aAlc n Then, Eq. (14) can be written as: log m1 Ki b log m2 15

16

Fig. 1. log Ki versus log aAlc in water/SDS/ethanol systems.

S. Javadian et al. / Journal of Molecular Liquids 137 (2008) 7479

77

Fig. 2. Plot of EMF versus log(CT) for SDS in () water, () 10% EtOH v/v at 25 C using a Br reference electrode.

Fig. 4. The self-diffusion coefficient (D) of the SDS as a function of total concentration in different mixtures of D2O\EtOH at 25.0 C.

water/alcohol. The Rh values obtained using Eq. (21) can be used to compute the aggregation number of the micelles. Assuming a spherical micelle and neglecting its intrinsic instability, the aggregation number can be computed using the following relation [32,33]:
NA
4 3 3 pRh

VSDS nh;SDS VD2 O bVNa nh;con VD2 O p n VAlc

22

where VSDS, VNa+ and VAlc are the molecular volumes of SDS, sodium ion and ethanol, respectively, and nh,SDS and nh,con are the hydration numbers of head group of SDS and Na+ in micelle, respectively, which were taken as VSDS = 0.362 nm3 [34], VNa+ = 0.003 nm3 [34], VD2O = 0.030 nm3 [34], VAlc = 0.067 nm3 [35], nh, SDS = 4 [34] and nh,con = 6 [34]. 4. Results and discussion Fig. 2 shows a typical plot of E1 versus the logarithm of the total concentration of SDS (CT). At low surfactant concentration, E1 is directly proportional to log(CT). The Nernstian slopes of the electrodes in these experiments were within the acceptable range of 5961 mV/decade. A definite break in the slope of the plot is considered to correspond to the CMC. Fig. 3 1/2 shows typical plots of EMF against log(m1/2 1 m2 ), where m1 and m2 are the ionic surfactant monomer and counterion concentrations respectively. In these plots, Nernstian behavior is observed at low surfactant concentrations, and a sharp break corresponding to the CMC occurs with increasing surfactant

concentration. At surfactant concentrations above the CMC, the slope decreases because counterions attach to the micelles. According to Fig. 4, the self-diffusion coefficient for SDS is approximately constant up to the CMC of the surfactant. At SDS concentrations greater than the CMC the self-diffusion coefficient decreases with increasing surfactant concentration for ethanol concentration between 0% and 30%. In 40% ethanol/water (v/v %), no micelle was detected up to 45 mM. Typical plots of the variation of the relative viscosity, r = / o (where and o are the viscosities of the solution and the solvent (water), respectively) of the micellar solution versus ethanol concentration for several different SDS concentrations are shown in Fig. 5. It is clear from this figure that ethanol has only a marginal effect on the solvent viscosity, which remains approximately constant, particularly for ethanol concentrations up to 10%. Given that the solvent viscosity is the primary determinant of micellar structure [36] our results indicate that addition of ethanol to the aqueous solvent has little effect on the structure of the SDS micelles, and thus that the micelles maintain a spherical shape. The results are found in good agreement with the literature reports [36]. The relative viscosity increases slightly as the ethanol content is increased beyond 10% due to increasing dispersity of the system. Fig. 6 shows the variation of the self-diffusion coefficient of the SDS micelles as a function of the alcohol concentration. The self-diffusion coefficient goes through a minimum as the concentration of alcohol is increased from 0% to 40%. The initial decrease in the self-diffusion coefficient is due to incorporation of the alcohol into the micellar phase. Penetration of alcohol into the Stern layer and

1/2 Fig. 3. Plot of EMF versus log(m1/2 1 m2 ) for SDS in () water, () 10% EtOH v/v at 25 C using a Na+ reference electrode.

Fig. 5. Relative viscosity (r) as a function of alcohol concentration for different total concentrations of SDS at 25.0 C.

78

S. Javadian et al. / Journal of Molecular Liquids 137 (2008) 7479

Fig. 6. Micelle self-diffusion coefficients versus alcohol concentration for different total concentrations of SDS at 25.0 C.

Fig. 8. Logarithm of the aggregation number of SDS micelles versus log ( CT m1) for different alcohol concentrations. The straight line is a linear least-squares fit to Eq. (17).

micellar core leads to an increase in the area per head group, and thus to a decrease in the charge density at the micellar surface. The increase in the self-diffusion coefficient at higher alcohol concentration can be mainly attributed to a decrease in the size of the micelles at higher alcohol contents. As alcohol is added to a surfactant micellar solution, it disrupts the three dimensional network structure of the water; ultimately, equilibrium is reached between the micellar and solvent phases. The presence of alcohol also lowers the dielectric constant of the solvent and the degree of structure. Additionally, the low dielectric constant of the alcohol/water medium favors the formation of smaller aggregates and consequently the charged head groups are further apart. Moreover, these smaller aggregates will likely have a large percentage of alcohol molecules associated with them, oriented at the hydrocarbonsolvent interface. If still more alcohol is added, these phenomena lead to the production of smaller and smaller aggregates until ultimately, an alcohol concentration is reached at which aggregation is unnecessary: the solvent structure has been destroyed so effectively that the hydrophobic effect no longer dominates (Fig. 7). The results in Table 1 show that, at low alcohol concentration, the CMC decreases with addition of more alcohol, whereas at higher alcohol concentration, it increases. On the other hand, the aggregation number passes through a maximum as the alcohol concentration is increased (Fig. 7). These results are consistent with previous results [1,7,29]. The addition of ethanol is expected to destroy the water structure and hence to diminish the hydrophobic effect, which would lead to a continuous increase in the CMC with increasing alcohol concentration. However, the forma-

tion of mixed micelles will stabilize the micellar phase by decreasing the surface charge density and hence head grouphead group repulsions. These two major effects with increasing alcohol concentration oppose each other, causing the CMC to pass through a minimum and the aggregation number to pass through a maximum. Above 40% v/v ethanol, no micelles form [37] because the solvent structure has been disrupted to such an extent that it can accommodate the hydrocarbon tails of the surfactant without any more unfavorable disruption. Moreover, at high alcohol concentrations the driving force for micellization is diminished to the point where surfactant ions or ion pairs remain free in solution. Evidence of decreased aggregation numbers in the presence of high concentrations of co-solvents has also been obtained by Muller in 19F NMR studies of a partially fluorinated analogue of SDS [38,39]. We now consider the ability of the expression developed in this paper, Eq. (17), to predict the average size of micelles formed in aqueous solutions containing alcohol. Fig. 8 shows the variation in aggregation number as a function of the arguments of Eq. (17). Note that, in the absence of alcohol, there is a good linear dependency between log NA and log(CT m1) that conforms to the following expression (Fig. 8): log NA 0:669 log aCT bm1 1:351 R2 0:997 23

In addition, in the presence of 10%30% alcohol, the following linear relation was observed between log NA and log(CT m1): 10%: log NA 1:083 log aCT bm1 0:789 R2 0:944 20%: log NA 0:846 log aCT bm1 0:618 R2 0:989 30%: log NA 0:811 log aCT bm1 0:095 R2 0:901 26 25 24

Fig. 7. Aggregation number of SDS as a function of alcohol concentration in different total concentrations of SDS at 25 C.

At higher alcohol concentrations (N 10% ethanol), the aggregation number is decreased due to the co-solvent effect.

S. Javadian et al. / Journal of Molecular Liquids 137 (2008) 7479

79

It is generally accepted that disruption of the structured water around the hydrophobic group favors demicellization. 5. Conclusion To study micelle growth at a total surfactant concentration above the CMC, PFGNMR spectroscopy and viscometry were used to measure the average aggregation number and the size or hydrodynamic radius. The solvent viscosity remained approximately constant with increasing ethanol content, indicating that adding ethanol to the aqueous medium does not affect the micellar structure. The micelle self-diffusion coefficient was observed to go through a minimum as the alcohol content was increased from 0 to 40% v/v. The initial decrease in the selfdiffusion coefficient is due to an increase in the micellar size (the co-surfactant effect), while the increase in the self-diffusion coefficient at higher alcohol concentration is due to an increase in dispersity resulting from destruction of the water structure (the co-solvent effect). Upon addition of ethanol to aqueous SDS solutions, the CMC decreases to a minimum and the aggregation number increases to a maximum at around 10% ethanol. A linear equation was derived to predict the size of SDS micelles as a function of surfactant concentration and degree of dissociation of the counterion in the presence of alcohol. This equation provides a good description of the variation of the aggregation number with surfactant and counterion concentrations. Notations CMC Critical micelle concentration CT Total concentration D Self-diffusion coefficient E1,2 EMF of cells 1, 2 E1,2 Standard potentials of cells 1, 2 F Faraday constant m1 monomer concentration of SDS m2 Counterion concentration Mw Molecular weight R Gas constant Rh Hydrodynamic radius of micelle T Absolute temperature V Molecular volume z Charge Viscosity Density Mean activity coefficient 1 Activity coefficient of SDS 2 Activity coefficient of counterion

[2] N. Alizadeh, H. Gharibi, M. Shamsipur, Bull. Chem. Soc. Jpn. 68 (1995) 1. [3] D.J. Shaw, Introduction to Colloid and Surface Chemistry4th ed., Butterworth Heinemann, 1993. [4] S.P. Moulik, Md.E. Haque, P.K. Jana, A.R. Das, J. Phys. Chem. 100 (1996) 701. [5] M.S. Bakshi, Bull. Chem. Soc. Jpn. 69 (1996) 2723. [6] M. Manabe, A. Tokunaga, H. Kawamura, M. Shiomi, K. Hiramatsu, Colloid Polym. Sci. 280 (2002) 929. [7] M.A. Safarpour, A.A. Rafati, H. Gharibi, J. Chin. Chem. Soc. 46 (1999) 983. [8] E. Capontti, D. Chillura Martino, M.A. Floriono, R. Triolo, Langmuir (1997) 3277. [9] P.C. Griffiths, N. Hirst, A. Paul, S.M. King, R.K. Heenan, R. Farley, Langmuir 20 (2004) 6904. [10] M. Almgren, S. Swarup, J. Colloid Interface Sci. 91 (1983) 256. [11] H. Gharibi, B. Sohrabi, S. Javadian, S.M. Hashemianzadeh, Colloids Surf., A Physicochem. Eng. Asp. 244 (2004) 187. [12] B. Lindman, J. Phys. Chem. 87 (1983) 1377. [13] R. Zana, Adv. Colloid Interface Sci. 57 (1995) 1. [14] D. Atwood, V. Mosquera, J. Rodriguez, M. Garcia, M. Suarez, Colloid Polym. Sci. 272 (1994) 584. [15] M. Shamsipur, N. Alizadeh, H. Gharibi, Bull. Chem. Soc. Jpn. 69 (1996) 311. [16] J. Armestrong, B. Chowdhry, J. Mitchell, A. Beezer, S. Leharne, J. Phys. Chem. 100 (1996) 1738. [17] S. Cinelli, G. Onori, A. Santucci, Colloids Surf., A Physicochem. Eng. Asp. 160 (1999) 3. [18] H. Gharibi, R. Palepu, D.M. Bloor, D.G. Hall, E. Wyn-Jones, Langmuir 8 (1992) 782 (and the references there in). [19] Davidson, Ph. D., Thesis, University of Aberdeen, 1993. [20] E.O. Stejskal, J.E. Tanner, J. Chem. Phys. 42 (1965) 288. [21] P. Stilbs, K. Paulsen, P.C. Griffiths, J. Phys. Chem. 100 (1996) 8180. [22] G. Nemethy, H.A. Scheraga, J. Chem. Phys. 41 (1964) 680. [23] S.S. Berr, J. Phys. Chem. 91 (1987) 4760. [24] H.F. Huisman, Proc. K. Ned. Akad. Wet. B 67 (1964) 367. [25] F.H. Quina, M.P. Nasser, J.B.S. Bonilha, B.S. Bales, J. Phys. Chem. 99 (1995) 17028. [26] M. Kashinuma, T. Sasaki, Bull. Chem. Soc. Jpn. 48 (1975) 2755. [27] D.G. Hall, J. Chem. Soc., Faraday Trans. 1 (177) (1981) 1121. [28] F.H. Quina, H. Chaimovich, J. Phys. Chem. 83 (1979) 1844. [29] A.A. Rafati, H. Gharibi, M. Rezaie-Sameti, J. Mol. Liq. 111 (2004) 109. [30] D.F. Evans, B.W. Ninham, J. Phys. Chem. 87 (1983) 5025. [31] A. Eintein, Investigation in the Theory of Brownian Movement, Dover, New York, 2004, p. 109. [32] P. Stilbs, Prog. NMR Spectr. 19 (1987) 1. [33] F. Bockstahl, G. Dupltre, J. Phys. Chem. B 105 (2001) 13. [34] S.S. Berr, E. Caponett, J.S. Johnson, R.R.M. Jones, L. Magid, J. Phys. Chem. 90 (1986) 5766. [35] D. Ciccarelli, L. Costantino, G. Derrico, L. Paaduano, V. Vitagliani, Langmuir 14 (1998) 7130. [36] Din K-ud, S. Kumar Kirti, P.S. Goyal, Langmuir 12 (1996) 1490. [37] N. Muller, T.W. Johnson, J. Phys. Chem. 73 (1969) 2042. [38] N. Muller, H. Simsohn, J. Phys. Chem. 75 (1971) 952. [39] S.S. Shah, N.U. Jamroz, Q.M. Sharif, Colloids Surf., A Physicochem. Eng. Asp. 178 (2001) 199. [40] Y. Moroi, N. Yoshida, Langmuir 13 (1997) 3909. [41] Y. Moroi, K. Matsuoka, Bull. Chem. Soc. Jpn. 67 (1994) 2057. [42] A.M. Misselyn-Bauduin, A. Thibaut, J. Grandjean, G. Broze, R. Jrme, J. Colloid Interface Sci. 238 (2001) 1.

References
[1] H. Gharibi, B.M. Razavizadeh, A.A. Rafati, Colloids Surf., A Physicochem. Eng. Asp. 136 (1998) 123.

Вам также может понравиться