Article pubs.acs.org/Langmuir

Light Scattering from Hydrophobe-Uptake Spherical Micelles near the Critical Micelle Concentration Ken Morishima and Takahiro Sato* Department of Macromolecular Science, Graduate School of Science, Osaka University, 1-1 Machikaneyama-cho, Toyonaka, Osaka 560-0043, Japan S Supporting Information *

ABSTRACT: We have investigated aqueous micellar solutions of mixtures of a surfactant (dodecylpyridinium chloride) and a hydrophobe (1dodecanol) near the critical micelle concentration (cmc), using simultaneous static and dynamic light scattering measurements. The static light scattering intensity for the aqueous solutions was separated into fast and slow relaxation components using dynamic light scattering results. The slow relaxation component gave us the information about the large scattering component. It turned out from this component that the amount of large colloidal particles of the hydrophobe was very tiny in the solution and hardly affects the association−dissociation equilibrium of the hydrophobe-uptake micelle. The free surfactant molecule and the hydrophobe-uptake spherical micelle in the solutions belong to the fast relaxation component. We have characterized the spherical micelle and also analyzed the association−dissociation equilibrium of the hydrophobe-uptake micelle up to near the cmc, using this scattering component extracted.



INTRODUCTION The hydrophobe-uptake micellization has been investigated for many systems,1−8 because hydrophobe uptake is one of the most important phenomena, which relates to extensive applications in detergents, paints, and drugs as well as potential applications to nanocarriers and nanoreactors. Many studies were interested in the micellization at surfactant concentrations much higher than the critical micelle concentration (cmc) and characterized the micellar structure in detail by scattering techniques and electron microscopy.9−13 However, near the cmc, it is difficult to apply these techniques to study the association−dissociation equilibrium behavior of the surfactants. Near the cmc, the micelle becomes less stable and sensitive to the solution temperature. Thus, it is difficult to prepare samples for cryo-transmission electron microscopy (cryoTEM) near the cmc. To the best of our knowledge, the micellization behavior near the cmc has been studied little by cryo-TEM. Moreover, the cmc for conventional surfactants is usually so low that the intensity of small-angle X-ray or neutron scattering is too weak to study the micellization behavior. Therefore, the most suitable technique to study the association−dissociation equilibrium near the cmc may be light scattering. However, this technique also has difficulty mentioned in the following. When a hydrophobic substance is added to an aqueous surfactant solution, two kinds of assemblies are formed in the solution. One is the small spherical micelle comprising the surfactant and a small amount of the hydrophobe included in © 2014 American Chemical Society

the hydrophobic core of the micelle. The other is a large colloidal particle consisting of the hydrophobe and stabilized by the surfactant. On one hand, when the amount of surfactant is much larger than the amount of the hydrophobe and the concentration of the surfactant is higher than the cmc, the assembly in the solution is the spherical micelle. On the other hand, when the amount of the surfactant is still larger than the amount of the hydrophobe but the surfactant concentration is near the cmc, large colloidal particles of the hydrophobe appear in the solution because of the difficulty of the spherical micelle formation. The formation of the large colloidal particles enormously increases light scattering intensity.1−4,13,14 Even if no hydrophobe is added to the solution, a tiny amount of impurity in the surfactant sample may act as a hydrophobe and light scattering intensity increases unexpectedly near the cmc.5,15 This strong scattering intensity from the large colloidal particles makes it difficult to analyze the association− dissociation equilibrium behavior of the surfactants using the light scattering data. Similar difficulties have often been encountered at light scattering studies on polymer solutions, which contain a small amount of large particles (sometimes dusts or aggregates). Kanao et al.16 proposed to escape such difficulties in polymer solutions by separating the static light scattering (SLS) intensity into small and large scattering components using dynamic light Received: June 25, 2014 Revised: September 10, 2014 Published: September 10, 2014 11513

dx.doi.org/10.1021/la502487x | Langmuir 2014, 30, 11513−11519

Langmuir

Article

scattering (DLS) data. Here, we apply their method to aqueous solutions containing mixtures of a surfactant and a hydrophobe near the cmc to study the association−dissociation equilibrium in the solutions using the scattering intensity for the small scattering component separated. In this study, we have chosen dodecylpyridinium chloride (DPC) as the surfactant and 1-dodecanol (DOH) as the hydrophobe. Although the DOH molecule has a hydroxyl group, its amphiphilicity is so weak that its aqueous solution undergoes a liquid−liquid phase separation. Thus, it behaves as a hydrophobe in aqueous solutions.



R θ ,fast =

Materials. DPC was purchased from Sigma-Aldrich. Sodium tetraborate decahydrate (Borax) and DOH were purchased from Wako Pure Chemical Industries, Ltd. DPC was recrystallized from acetone 5 times. DOH was used without further purification. Water was purified with a Millipore Milli-Q system, which was used as the solvent of test solutions. Preparation of Test Solutions. Test solutions were prepared in the following manner. Aqueous Borax solution (25 mM) was used as an electrolyte solvent for all sample solutions. Purified DPC and neat DOH were mixed well at a weight ratio of DOH to DPC, cH/cR. Then, the DPC−DOH mixture was dissolved in the aqueous Borax solution. The concentrated original solution was prepared at cR = 5.0 × 10−2 g cm−3. Test solutions were prepared with dilution of the original solution by aqueous Borax solution. In this study, the mass concentration of DPC and DOH are represented by cR and cH, respectively. Then, the total mass concentration is represented by c (=cR + cH). The subscripts R and H indicate DPC and DOH, respectively. Specific Refractive Index Increment Measurements. Specific refractive index increments (∂n/∂c) for sample solutions were measured using a modified Schulz−Cantow-type differential refractometer (Shimazu) with 488 and 546 nm wavelength light at 25.0 °C. Values of ∂n/∂c at λ0 (wavelength) = 532 nm were obtained by interpolation of ∂n/∂c values at λ0 = 488 and 546 nm to be 0.172 cm3 g−1 at cH/cR = 0 and 0.169 cm3 g−1 at cH/cR = 0.1. We can obtain a value of ∂n/∂c at any cH/cR by

3

∫all A(τ ) dτ

Rθ ,

R θ ,slow =

∫slow A(τ ) dτ ∫all A(τ ) dτ

Rθ (2)

The light scattering power of the slow relaxation component is usually much stronger than that of the fast relaxation component. Therefore, if Rθ,fast and Rθ,slow are comparable, the weight fraction of the slow relaxation component may be very small. In such a case, Rθ,fast and Rθ,slow for dilute solutions are written as16,18

EXPERIMENTAL SECTION

c ⎛ ∂n ⎞ c ⎛ ∂n ⎞ ∂n = R⎜ ⎟ + H⎜ ⎟ ⎝ ⎠ ∂c c ∂c R c ⎝ ∂c ⎠H

∫fast A(τ ) dτ

⎞ ⎛ ∂n ⎞ −2 ⎛ K ′c 1 ⎜⎜ =⎜ ⎟ + 2A 2,fast c + ...⎟⎟ ⎝ ∂c ⎠fast ⎝ wfastM w,fast R θ ,fast ⎠

(3a)

⎡⎛ ∂n ⎞ 2 ⎤−1 K ′c = ⎢⎜ ⎟ wslowM w,slow Pz ,slow(k)⎥ R θ ,slow ⎣⎝ ∂c ⎠slow ⎦

(3b)

where ∂n/∂c, w, Mw, A2, and Pz(k) are the refractive index increment, the weight fraction (in the total scattering components), the weightaverage molar mass, the second virial coefficient, and the z-average particle scattering function, respectively, and subscripts fast and slow indicate the quantities for the fast and slow relaxation components, respectively. The optical constant K′ is defined as K′ =

4π 2n0 2 NAλ 0 4

(4)

with the refractive index of the solvent being n0 and the Avogadro constant being NA. The argument k of Pz,slow(k) is the magnitude of the scattering vector. In the above equations, the angular dependence of Rθ,fast was neglected on the assumption that the size of the fast relaxation component is much smaller than the wavelength of the incident light. The concentration dependence of Rθ,slow can be neglected in a good approximation.16 Using the bimodal A(τ), we can also calculate the first cumulants of the fast and slow relaxation components, Γfast and Γslow, respectively, by

Γfast =

∫fast τ −1A(τ ) dτ ∫fast A(τ ) dτ

,

Γslow =

∫slow τ −1A(τ ) dτ ∫slow A(τ ) dτ

(5)

These first cumulants give us the hydrodynamic radius RH of each component by

(1)

−1 kBT ⎛ Γslow ⎞ ⎜ lim 2 ⎟ 6πη0 ⎝k → 0 k ⎠ (6) where kBT is the Boltzmann constant multiplied by the absolute temperature and η0 is the solvent viscosity. In the above equation, the interparticle interaction effect on RH,slow was neglected on the assumption that the slow relaxation component appears only in very dilute solutions.

−1

RH,fast =

Here, (∂n/∂c)R is 0.172 cm g , which is the value at cH/cR = 0. (∂n/ ∂c)H could be determined by eq 1 using the experimental result at cH/ cR = 0.1 to be (∂n/∂c)H = 0.140 cm3 g−1. Light Scattering Measurements. Test solutions for light scattering measurements were optically cleaned by filtration through a 0.2 μm pore-size membrane filter before dilution of original solution. Simultaneous static and DLS measurements were carried out for aqueous Borax solutions of DPC−DOH mixtures using an ALV/SLS/ DLS-5000 light scattering instrument equipped with an ALV-5000 multiple τ digital correlator. Vertically polarized light with the wavelength λ0 = 532 nm emitted from a Nd:YAG laser was used as incident light. The light scattering system was calibrated using toluene as the reference material. The Rayleigh ratio of toluene Rtol for vertically polarized 532 nm light was taken to be 2.72 × 10−5 cm−1 at 25 °C.17 The excess Rayleigh ratio Rθ at scattering angle θ of each solution over that of the solvent was calculated from the scattering intensity of the solution Iθ,soln and solvent Iθ,solv by the standard procedure.16 The intensity autocorrelation function g(2)(t) obtained by DLS was analyzed by CONTIN to obtain the spectrum A(τ) of the relaxation time τ. All measurement was carried out at 25 °C. Analysis of Light Scattering Data. When the spectrum A(τ) obtained by DLS is bimodal, the solution contains two scattering components of largely different sizes. Using the bimodal A(τ), the excess Rayleigh ratio Rθ obtained by SLS is divided into the fast relaxation component Rθ,fast and the slow relaxation component Rθ,slow by

−1 Γ ⎞ kBT ⎛ ⎜ lim fast ⎟ , 6πη0 ⎝c , k → 0 k 2 ⎠

RH,slow =



RESULTS AND DISCUSSION Light Scattering Results. Figure 1 shows SLS results for aqueous Borax solutions of a DPC−DOH mixture of cH/cR = 0.1, with different total solute concentrations of c = cR + cH. While Rθ/K′c takes low values and exhibits no angular dependence at c ≥ 1.1 × 10−2 g/cm3, it increases rapidly and shows strong angular dependences at c ≤ 5.5 × 10−3 g/cm3. Because DPC is known to form a small spherical micelle in aqueous salt solution above the cmc, 19 the angular independence of the scattering function at c ≥ 1.1 × 10−2 g/ cm3 is normal but the strong angular dependences of Rθ/K′c demonstrate the existence of large colloidal particles, probably comprising DOH, in the solution with c ≤ 5.5 × 10−3 g/cm3. The anomalous light scattering starts occurring at the concentration close to the cmc for DPC in aqueous solutions 11514

dx.doi.org/10.1021/la502487x | Langmuir 2014, 30, 11513−11519

Langmuir

Article

slow relaxation components (Rθ,fast and Rθ,slow) and also estimated the first cumulants of the fast and slow relaxation components (Γfast and Γslow) using the bimodal A(τ), as shown in Figure 3. In what follows, we will discuss those fast and slow relaxation components separately. Occasionally, spectra for the slow relaxation component became multimodal, but we did not make further mode separation of the slow relaxation component, because the appearance of the multimodes was not systematic. It is noted that the CONTIN analysis is sometimes affected by noise of the scattering intensity. Slow Relaxation Component. Figure 4 shows the Guinier plots for the slow relaxation component in aqueous Borax solutions of the DPC−DOH mixture of cH/cR = 0.1 and DPC (cH/cR = 0) with different c. All of the scattering functions exhibit strong k dependences, indicating that the slow relaxation component has a large size comparable to the wavelength of the incident light. We assumed that the slow relaxation component in aqueous solutions of DPC−DOH mixtures is colloidal particles comprised of DOH, which are regarded as polydisperse spheres. Thus, in eq 3b, (∂n/∂c)slow = (∂n/∂c)H (=0.140 cm3 g−1) and Mw,slowPz,slow(k) is given by

Figure 1. Angular dependences of ln(Rθ/K′c) for aqueous Borax solutions of a DPC−DOH mixture of cH/cR = 0.1 with different total solute concentrations of c = cR + cH.

with a similar ionic strength,19,20 and such light scattering behavior was observed for many aqueous surfactant solutions.1−5,13,14 A similar behavior was observed for aqueous Borax solutions of DPC−DOH mixtures with cH/cR < 0.1 also near the cmc of aqueous DPC. As shown in Figure 2, similar strong angular dependences of Rθ/K′c are observed even for aqueous Borax solutions of DPC

M w,slow Pz ,slow(k) =

∫ w(M)MPsphere,M(k) dM

(7)

where w(M) is the weight fraction of the sphere with the molar mass M. Psphere,M(k) is the particle scattering function for the sphere with M given by ⎡ sin(kR ) − kR cos(kR ) ⎤2 M M M ⎥ Psphere, M(k) = ⎢3 3 (kRM ) ⎦ ⎣

(8)

with the radius of the sphere RM calculated using the specific volume υH̅ of DOH (=1.2 cm3/g) by

⎛ 3MυH̅ ⎞1/3 RM = ⎜ ⎟ ⎝ 4πNA ⎠

(9)

We further assume that the size distribution of the DOH spheres is represented by the log-normal distribution defined by w(M ) dM = Figure 2. Angular dependences of ln(Rθ/K′c) for aqueous Borax solutions of DPC (cH/cR = 0) with different total solute concentrations of c = cR + cH.

1 exp( −x 2) dx π

(10)

with x≡

(cH/cR = 0) with c = cR ≤ 3.0 × 10−3 g/cm3. The large scattering component in DPC solutions may consist of a tiny amount of some hydrophobic impurity, including in the DPC sample. Figure 3 shows the concentration dependence of the relaxation time spectra A(τ) of the intensity autocorrelation function g(2)(t) for aqueous Borax solutions of DPC and the DPC−DOH mixture of cH/cR = 0.1 at scattering angle θ = 60°. While A(τ) is unimodal at the highest concentration (c = 2.75 × 10−2 g cm−3), it becomes bimodal with decreasing c. At low concentrations (c ≤ 5.5 × 10−3 g cm−3), A(τ) becomes unimodal again. Similar c dependences of A(τ) were also observed for the solutions having other cH/cR. According to the procedure mentioned in the Experimental Section, we divided Rθ, as shown in Figure 1, into the fast and

ln(M / M w,slow M n,slow ) 2 ln(M w,slow /M n,slow )

(11)

As shown by solid curves in Figure 4b, the above equations can fit the experimental Rθ,slow/K′c satisfactorily and we can determine wslow, Mw,slow, and Mw,slow/Mn,slow by the fitting. Although colloidal particles in aqueous solutions of DPC with cH/cR = 0 do not comprise DOH but probably some hydrophobic impurity in the DPC sample, we made the same fitting by assuming that the specific volume of the impurity is close to that of DOH. Table S1 of the Supporting Information summarizes the fitting results, along with the weight-average radius of the DOH sphere Rw,slow, calculated by 1/9 ⎛ 3M w,slow υH̅ ⎞1/3⎛ M n,slow ⎞ ⎟⎟ R w,slow = ⎜ ⎟ ⎜⎜ ⎝ 4πNA ⎠ ⎝ M w,slow ⎠

11515

(12)

dx.doi.org/10.1021/la502487x | Langmuir 2014, 30, 11513−11519

Langmuir

Article

Figure 3. Relaxation time spectra of the intensity autocorrelation function at scattering angle θ = 60° for aqueous Borax solutions of (a) DPC−DOH mixture of cH/cR = 0.1 and (b) DPC (cH/cR = 0) with different c.

Figure 4. Angular dependences of ln(Rθ,slow/K′c) for aqueous Borax solutions of (a) DPC−DOH mixture of cH/cR = 0.1 and (b) DPC (cH/cR = 0) with different c.

and also the hydrodynamic radius RH,slow of the slow relaxation component. It can be seen from Table S1 of the Supporting Information that the weight-average molar masses Mw,slow are very high, the dispersities Mw,slow/Mn,slow are considerably high, and the weight fractions wslow are very small in all of the solutions, where both fast relaxation component is observable by DLS (cf. Figure 3). Such tiny amounts of the scattering component are detectable only by light scattering, because the light scattering power is proportional to Mw,slow. In aqueous Borax solutions of DPC (cH/cR = 0), wslow values are at most in the order of 10−5, which indicates that high purity of the DPC sample was used. The anomalous light scattering behavior in aqueous surfactant solutions near the cmc can occur even for highly purified surfactant samples. In other words, light scattering is a very good detector for hydrophobic impurity in surfactant samples. One of reviewers of this paper pointed out that the appearance of the slow relaxation component in our system near the cmc resembles the transition from the spherical micelle to the bilayer vesicle in aqueous solutions of mixtures of two surfactants at approaching the cmc of the major surfactant component with higher cmc. There are two kinds of mixtures studied previously for the micelle−bilayer transition: one is the mixture of single- and double-tailed surfactants,21,22 and the other is the mixture of oppositely charged surfactants.23,24 However, DOH is not a surfactant but a hydrophobe not

dispersed as a micelle in water by itself. Therefore, the appearance of the slow relaxation component in our system near the cmc may have nothing to do with the micelle−bilayer transition reported previously for aqueous solutions containing two surfactants. The reviewer also gave us the comment that the slow relaxation component may not be the uniform density sphere but a vesicle. Thus, we tried to fit the Rθ,slow/K′c data by the scattering function for the vesicle (cf. Supporting Information). The fitting was also good, and it was difficult to distinguish the uniform density sphere from the vesicle by light scattering (we have failed to distinguish them also from the ratio of the radius of gyration and hydrodynamic radius; cf. Supporting Information). Sato et al.25 distinguished spherical particles formed by a thermosensitive block copolymer in hot water from the vesicle, by small-angle X-ray scattering (SAXS). However, the present low-molar-mass surfactant solutions close to the cmc were too dilute to obtain enough scattering intensity by SAXS (using the beamline 40B2 of SPring-8, Hyogo, Japan). Therefore, we cannot specify the shape of the slow relaxation component at present. As shown in Table S1 of the Supporting Information (in the parentheses), even if we assume that the slow relaxation component is the vesicle, the weight fraction wslow of the slow relaxation component is also very small for solutions with a detectable fast relaxation component. Therefore, in the 11516

dx.doi.org/10.1021/la502487x | Langmuir 2014, 30, 11513−11519

Langmuir

Article

considered as the fast relaxation component; it should be noted that we do not define the fast relaxation component as the micellar component. The micelle Rm can be characterized by the aggregation number m of R and the molar mass Mm,H of H included in the micelle. Here, we assumed that the Rm is a spherical micelle. According to the micellization theory,26 the spherical micelle formed by low mass surfactants has some optimum aggregation number, which is independent of the concentration. Therefore, the value of m is assumed to be independent of the concentration. The molar mass Mm of the spherical micelle is given by

following discussion of the fast relaxation component, we will neglect the amounts of DPC and DOH in the slow relaxation component. Fast Relaxation Component. Figure 5 shows the Guinier plots for the fast relaxation component in aqueous Borax

Mm = mMR + Mm ,H

(13)

where MR is the molar mass of R, and the refractive index increment (∂n/∂c)m of the micelle is given by Mm ,H ⎛ ∂n ⎞ ⎛ ∂n ⎞ mMR ⎛ ∂n ⎞ ⎜ ⎟ = ⎜ ⎟ + ⎜ ⎟ ⎝ ∂c ⎠m Mm ⎝ ∂c ⎠R Mm ⎝ ∂c ⎠H

(14)

The hydrophobic core of the micelle Rm comprises H and the hydrophobic tail of R. Thus, the volume of the core Vcore is written as Figure 5. Angular dependences of ln(Rθ,fast/K′c) for solutions with cH/ cR = 0.1.

Vcore = υH̅ (Mm ,H /NA ) + υRhm

(15)

where υRh is the volume of the hydrophobic tail of R. The area aR on the interface of the core occupied by each R molecule is calculated by

solutions of the DPC−DOH mixture of cH/cR = 0.1 and DPC with different c. All of the scattering functions exhibit no angular dependences, indicating that the fast relaxation component has a size much smaller than the wavelength of the incident light. For each c, we have estimated R0,fast at zero k by averaging Rθ,fast at all measured k. Figure 6 shows the

maR = (3 4π Vcore)2/3

(16)

Because the total amount of R on the interface of colloidal particles is negligibly small (cf. wslow in Table S1 of the Supporting Information), the molar concentrations of the micelle [Rm] and the free molecule [R] are related by c R /MR = m[R m] + [R]

(17)

Moreover, using the association constant Km, [Rm] and [R] are related by

K m = [R m]/[R ]m

(18)

For given values of m and Km, we can calculate [Rm] and [R] from eqs 17 and 18. Because the amount of DOH in the colloidal particle component is also negligibly small in comparison to the total H in the solution, Mm,H is calculated from [Rm] by Mm ,H = c H/[R m]

(19)

The weight-average molar mass of the fast relaxation component Mw,fast is calculated by Figure 6. Fitting results of R0,fast/K′c by eq 3a along with eqs 13, 14, and 17−20, using m, Km, and A2,fast as adjustable parameters, and calculated aR by eqs 15 and 16 using the adjustable parameters determined.

M w,fast =

[R]MR 2 + [R m]Mm 2 [R]MR + [R m]Mm

(20)

and the weight fraction of the fast relaxation component wfast can be approximated to be unity, because of negligibly small wslow. Although the refractive index increment of the free R is slightly different from (∂n/∂c)m, we approximate (∂n/∂c)fast in eq 3a by (∂n/∂c)m (eq 14), because the contribution of the free R to Rθ,fast is small. Thus, we can calculate K′c/Rθ,fast by eq 3a along with eqs 13, 14, and 17−20, using m, Km, and A2,fast as adjustable parameters. In the upper panel of Figure 6, solid curves indicate fitting results of R0,fast/K′c in the Guinier plot by choosing suitable values for m, Km, and A2,fast, which are listed in Table 1. The

concentration dependences of ln(R0,fast/K′c) for four different cH/cR (the upper panel). For all cH/cR, ln(R0,fast/K′c) exhibits a maximum at concentrations slightly higher than the cmc. There are three scattering components in the aqueous Borax solution containing DPC and DOH; the dissociated surfactant molecule R, the spherical micelle Rm, and the colloidal particle. While the last component is regarded as the slow relaxation component, the first and second scattering components are 11517

dx.doi.org/10.1021/la502487x | Langmuir 2014, 30, 11513−11519

Langmuir

Article

If the spherical micelle Rm is regarded as the hard sphere, RH,m and its second virial coefficient A2,m may be approximately calculated by

Table 1. Characteristics of the Spherical Micelle Rm cH/cR

m

log(Km) (M1−m)

A2,fast (×10−3, cm3 g−2 mol)

RH,m (nm)

0.1 0.05 0.001 0

44 40 35 40

101 78.5 65.5 71.0

1.13 1.40 1.60 1.75

2.2 1.6 1.7 1.5

RH, m

⎛ 3υH̅ Mm ⎞1/3 =⎜ ⎟ , ⎝ 4πNA ⎠

A 2, m =

4υH̅ Mm

(22)

At the fit of R0,fast/K′c shown in Figure 6, the second virial coefficient is important only in a high c region, where the scattering from Rm is predominant in R0,fast, so that A2,fast determined by the fitting may be approximately equal to A2,m. Figure 8 compares experimental RH,m and A2,fast with eq 22. We

fitting is satisfactorily good for all cH/cR, demonstrating the appropriateness of the spherical model for the fast relaxation component. With increasing cH/cR, m and Km seem to increase and A2,fast decreases, although the order is reversed for m and Km at cH/cR = 0 and 0.001. Figure 6 also shows the area aR on the interface of the hydrophobic core occupied by each R molecule (in the lower panel), calculated by eqs 15 and 16 using m and Km determined. At cH/cR = 0 and 0.001, aR is constant in the c range examined by light scattering, but at cH/cR = 0.05 and 0.1, it increases rapidly with approaching c, where the fast relaxation component disappears. The decrease in the area density of the hydrophilic moiety of R leads to penalty of the interfacial energy, and this instability makes the spherical micelle destruct and the hydrophobe release. Figure 7 shows the concentration dependences of the first cumulant of the fast relaxation component. As explained the

Figure 8. Comparison of experimental RH,m and A2,fast to those for the hard sphere calculated by eq 22.

Figure 7. Concentration dependences of (Γfastk−2)k = with four different cH/cR.

0

have used υ̅H for DOH (=1.2 cm3/g) to obtain the solid curves in the figure, but the results did not so much change even if we use υ̅H for n-dodecane (=1.3 cm3/g). The agreement between the experiment and theory for RH,m is reasonably good, demonstrating that DPC forms the spherical micelle containing DOH inside the core in aqueous Borax solution. However, the magnitude of A2,fast is higher than A2,m calculated by eq 22, although they show a similar concentration dependence (∝1/ Mm). The latter disagreement is owing to the electrostatic repulsion between micelles, which is not considered in eq 22.



for solutions

CONCLUSION We have investigated the light scattering peak phenomenon in hydrophobe-uptake micellar solutions near the cmc by separating the scattering intensity into the fast and slow relaxation components using DLS results. There are three scattering components in the solutions: the free surfactant molecule R, the spherical micelle Rm consisting of m surfactant molecules and a small amount of the hydrophobe in the hydrophobic core, and the large colloidal particle of the hydrophobe stabilized by the surfactant. The fast relaxation component corresponds to the scattering from the first and second scattering components R and Rm, and we were able to characterize the spherical micelle Rm and also analyze the association−dissociation equilibrium of the hydrophobe-uptake micelle even near the cmc, using this scattering component extracted. The size and aggregation number m of Rm increase only slightly with increasing the mixing ratio cH/cR and depend little upon the total solute concentration c (=cR + cH) at fixed cH/cR, except near the cmc. The slow relaxation component gave us the information about the large colloidal particle component. Although we could not judge whether this component is the uniform density

reason below, we have determined the hydrodynamic radius RH,fast by extrapolating Γfast/k2 to the infinite dilution using the data of c > 0.02 g/cm3. The results of RH,fast were 2.1, 1.6, 1.7, and 1.5 nm at cH/cR = 0.1, 0.05, 0.001, and 0, respectively. The hydrodynamic radius RH,fast is given by RH,fast−1 =

[R]MR RH,R −1 + [R m]MmRH, m−1 [R]MR + [R m]Mm

(21)

where RH,R and RH,m are the hydrodynamic radii of the free R and the micelle Rm, respectively. As mentioned above, when c > 0.02 g/cm3, aR and then m of Rm are almost constant for all cH/ cR, so that we can extrapolate Γfast/k2 to the infinite dilution without changing RH,m. In Figure 7, we have made such extrapolations to obtain RH,fast. We can also roughly estimate RH,R of the free DPC; the root-mean-square end-to-end distance of n-dodecane is ca. 1 nm, and the size of the pyridine ring is about 0.3 nm, so that RH,R ∼ 1.5 nm. Using RH,fast and RH,R along with m and Km determined above, we can calculate RH,m from eq 21. The results are also listed in Table 1. 11518

dx.doi.org/10.1021/la502487x | Langmuir 2014, 30, 11513−11519

Langmuir

Article

dodecyl ethers with uptake of n-dodecane. J. Phys. Chem. B 2008, 112, 4648−4655. (13) Ochi, M.; Matsue, S.; Einaga, Y. Change in characteristics of wormlike micelles of hexaoxyethylene tetradecyl and decyl ethers (C14E6 and C10E6) with uptake of n-dodecane. Polym. J. 2008, 40, 442−449. (14) Paillet, S.; Grassl, B.; Desbrieres, J. Rapid and quantitative determination of critical micelle concentration by automatic continuous mixing and static light scattering. Anal. Chim. Acta 2009, 636, 236−241. (15) Sorci, G. A.; Walker, T. D. Phenomenon observed at the onset of micellization using static light scattering. Langmuir 2005, 21, 803− 806. (16) Kanao, M.; Matsuda, Y.; Sato, T. Characterization of polymer solutions containing a small amount of aggregates by static and dynamic light scattering. Macromolecules 2003, 36, 2093−2102. (17) Dezelic, G.; Vevra, J. Angular dependence of light scattering in pure liquids. Croat. Chem. Acta 1966, 38, 35−47. (18) Hashidzume, A.; Kawaguchi, A.; Tagawa, A.; Hyoda, K.; Sato, T. Synthesis and structural analysis of self-associating amphiphilic statistical copolymers in aqueous media. Macromolecules 2006, 39, 1135−1143. (19) Fujio, K.; Ikeda, S. Size of micelles of 1-dodecylpyridinium chloride in aqueous NaCl solutions. Bull. Chem. Soc. Jpn. 1992, 65, 1406−1410. (20) van Os, N. M.; Haak, J. R.; Rupert, L. A. M. Physico-Chemical Properties of Selected Anionic, Cationic and Nonionic Surfactants; Elsevier Science: Amsterdam, Netherlands, 1993. (21) Grillo, I.; Penfold, J. Self-assembly of mixed anionic and nonionic surfactants in aqueous solution. Langmuir 2011, 27, 7453− 7463. (22) Bergström, L. M.; Skoglund, S.; Danerlöv, K.; Garamus, V. M.; Pedersen, J. S. The growth of micelles, and the transition to bilayers, in mixtures of a single-chain and a double-chain cationic surfactant investigated with small-angle neutron scattering. Soft Matter 2011, 7, 10935−10944. (23) Kaler, E. W.; Herrington, K. L.; Murthy, A. K.; Zasadzinski, J. A. N. Phase behavior and structures of mixtures of anionic and cationic surfactants. J. Phys. Chem. 1992, 96, 6698−6707. (24) Bergström, L. M.; Skoglund, S.; Edwards, K.; Eriksson, J.; Grillo, I. Self-assembly in mixtures of an anionic and a cationic surfactant: A comparison between small-angle neutron scattering and cryo-transmission electron microscopy. Langmuir 2013, 29, 11834−11848. (25) Sato, T.; Tanaka, K.; Toyokura, A.; Mori, R.; Takahashi, R.; Terao, K.; Yusa, S. Self-association of a thermosensitive amphiphilic block copolymer poly(N-isopropylacrylamide)-b-poly(N-vinyl-2-pyrrolidone) in aqueous solution upon heating. Macromolecules 2013, 46, 226−235. (26) Israelachvili, J. N. Intermolecular and Surface Forces, 2nd ed.; Academic Press: New York, 1992.

sphere or vesicle from the light scattering data alone, it turned out that the amount of this component is very tiny in the solution. Because of its small amount, the droplet component hardly affects the association−dissociation equilibrium of the hydrophobe-uptake micelle.



ASSOCIATED CONTENT

S Supporting Information *

Characterization of the slow relaxation component. This material is available free of charge via the Internet at http:// pubs.acs.org.



AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected]. Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS This work was supported in part by a Grant-in-Aid for Scientific Research (26620184) from the Japan Society for the Promotion of Science.



REFERENCES

(1) Phillips, J. N.; Mysels, K. J. Light scattering by aqueous solutions of sodium lauryl sulfate. J. Phys. Chem. 1955, 59, 325−330. (2) Princen, L. H.; Mysels, K. D. Some effects of lauryl alcohol on light scattering by sodium lauryl sulfate. J. Phys. Chem. 1959, 63, 1781−1782. (3) Zhu, Z.; Reed, W. F. Enhanced surfactant supramicellar assembly by hydrophobic dopants. Langmuir 2013, 29, 10376−10382. (4) Corti, M.; Degiorgio, V. Intensity-correlation study of the effect of dodecanol on sodium dodecyl sulfate aqueous solutions near the critical micelle concentration. Chem. Phys. Lett. 1977, 49, 141−144. (5) Konev, S. A.; Nikitin, V. V.; Olefirenko, G. I.; Yudin, I. K.; Zhuravleva, E. V. Static and dynamic light scattering study of aqueous micellar solutions formed by polydisperse surfactant. Mol. Cryst. Liq. Cryst. 1991, 200, 19−27. (6) Matsumura, S.; Kataoka, K. Preclinical and clinical studies of anticancer agent-incorporating polymer micelles. Cancer Sci. 2009, 100, 572−579. (7) Kataoka, K.; Harada, A.; Nagasaki, Y. Block copolymer micelles for drug delivery: Design, characterization and biological significance. Adv. Drug Delivery Rev. 2001, 47, 113−131. (8) Yang, Y. Q.; Zheng, L. S.; Guo, X. D.; Qian, Y.; Zhang, L. J. pHsensitive micelles self-assembled from amphiphilic copolymer brush for delivery of poorly water-soluble drugs. Biomacromolecules 2011, 12, 116−122. (9) Sanada, Y.; Akiba, I.; Sakurai, K.; Shiraishi, K.; Yokoyama, M.; Mylonas, E.; Ohta, N.; Yagi, N.; Shinohara, Y.; Amemiya, Y. Hydrophobic molecules in filtrating into the poly(ethylene glycol) domain of the core/shell interface of a polymeric micelle: Evidence obtained with anomalous small-angle X-ray scattering. J. Am. Chem. Soc. 2013, 135, 2574−2582. (10) Bae, Y.; Nishiyama, N.; Fukushima, S.; Koyama, H.; Matsumura, Y.; Kataoka, K. Preparation and biological characterization of polymeric micelle drug carriers with intracellular pH-triggered drug release property: Tumor permeability, controlled subcellular drug distribution, and enhanced in vivo antitumor efficacy. Bioconjugate Chem. 2005, 16, 122−130. (11) Wassenius, H.; Nyden, M.; Vincent, B. NMR diffusion studies of translational properties of oil inside core−shell latex particles. J. Colloid Interface Sci. 2003, 264, 538−547. (12) Miyake, M.; Asano, A.; Einaga, Y. Size change of the wormlike micelles of pentaoxyethylene, hexaoxyethylene, and heptaoxyethylene 11519

dx.doi.org/10.1021/la502487x | Langmuir 2014, 30, 11513−11519

Light scattering from hydrophobe-uptake spherical micelles near the critical micelle concentration.

We have investigated aqueous micellar solutions of mixtures of a surfactant (dodecylpyridinium chloride) and a hydrophobe (1-dodecanol) near the criti...
1MB Sizes 1 Downloads 5 Views