Sie sind auf Seite 1von 12

Journal of Colloid and Interface Science 490 (2017) 762–773

Contents lists available at ScienceDirect

Journal of Colloid and Interface Science


journal homepage: www.elsevier.com/locate/jcis

Micelle-vesicle-micelle transition in aqueous solution of anionic


surfactant and cationic imidazolium surfactants: Alteration of the
location of different fluorophores
Rupam Dutta, Surajit Ghosh, Pavel Banerjee, Sangita Kundu, Nilmoni Sarkar ⇑
Department of Chemistry, Indian Institute of Technology, Kharagpur 721302, WB, India

g r a p h i c a l a b s t r a c t

a r t i c l e i n f o a b s t r a c t

Article history: The presence of different surfactants can alter the physicochemical behaviors of aqueous organized
Received 28 September 2016 assemblies. In this article, we have investigated the location of hydrophobic molecule (Coumarin 153,
Revised 5 December 2016 C153) and hydrophilic molecule (Rhodamine 6G perchlorate, R6G) during micelle-vesicle-micelle transi-
Accepted 5 December 2016
tion in aqueous medium in presence of anionic surfactant, sodium dodecylbenzenesulfonate (SDBS) and
Available online 8 December 2016
cationic imidazolium-based surfactant, 1-alkyl-3-methylimidazolium chloride (CnmimCl; n = 12, 16).
Initially, the physicochemical properties of anionic micellar solution of SDBS has been investigated in
Keywords:
presence of imidazolium-based surfactant, CnmimCl (n = 12, 16) in aqueous medium by visual observa-
Imidazolium-based surfactant
Zeta potential (f)
tion, turbidity measurement, zeta potential (f), dynamics light scattering (DLS), and transmission elec-
Rotational motion tron microscopy (TEM). Zeta potential (f) measurement clearly indicates that the incorporation
Translational diffusion efficiency of C16mimCl in SDBS micelle is better than the other one due to the involvement of strong
hydrophobic as well as electrostatic interaction between the two associated molecules. Turbidity and
DLS measurements clearly suggest the formation of vesicles over a wide range of concentration.
Finally, the rotational motion of C153 and R6G has also been monitored at different mole fractions of
CnmimCl in SDBS-CnmimCl (n = 12, 16) solution mixtures. The hydrophobic C153 molecules preferen-
tially located in the bilayer region of vesicle, whereas hydrophilic R6G can be solubilized at surface of
the bilayer, inner water pool or outer surface of vesicles. It is observed that rotational motion of R6G

⇑ Corresponding author.
E-mail address: nilmoni@chem.iitkgp.ernet.in (N. Sarkar).

http://dx.doi.org/10.1016/j.jcis.2016.12.009
0021-9797/Ó 2016 Elsevier Inc. All rights reserved.
R. Dutta et al. / Journal of Colloid and Interface Science 490 (2017) 762–773 763

is altered significantly in SDBS-CnmimCl solution mixtures in presence of different mole fractions of


CnmimCl. Additionally, the translational diffusion motion of R6G is monitored using fluorescence corre-
lation spectroscopy (FCS) techniques to get a complete scenario about the location and translational dif-
fusion of R6G.
Ó 2016 Elsevier Inc. All rights reserved.

1. Introduction [23]. Besides, these surfactants can form vesicular assemblies in


presence of other additives or oppositely charged surfactant mole-
On account of widespread application of surfactant-based self- cules [19,46–48]. Oppositely charged surfactant molecules pene-
assemblies, there is an increasing interest to investigate the struc- trate in the micellar assemblies which are formed from
tural aspects of their aggregates into the extensive variety of surfactants along with firm interactions with concerned molecules
nanostructures [1–4]. In aqueous medium, the physicochemical promote the facile approach to alter the nature of aggregation.
behavior of surfactant assemblies mainly depend upon the temper- In this manuscript, we are interested to investigate the location
ature, pH, external additives, ionic strength, nature of the surfac- dependent behavior of hydrophobic molecule C153 and hydrophi-
tants, counter ions, etc. [5–12,59]. Electrostatic, hydrophobic and lic molecule R6G during micelle-vesicle-micelle transition. We
steric interactions between head groups and alkyl part of the sur- have used sodium dodecylbenzenesulfonate (SDBS) as an anionic
factant molecules control the surfactant molecule containing self- surfactant which forms micelles in aqueous solution. Two different
assembled nanoaggregates formation as well as the micelle- imidazolium-based surfactants,1-dodecyl-3-methylimidazolium
vesicle-micelle transition [5,8,10]. The temperature or composition chloride(C12mimCl) and 1-hexadecyl-3-methylimidazolium chlo-
induced transition between micelles and vesicles in surfactant ride (C16mimCl) having variation in their alkyl chain length are
assemblies is an important phenomenon in colloid chemistry and used separately to observe the minute change in phase behavior
biological processes due to their application in drug delivery, of aqueous surfactant solution. Turbidity, dynamic light scattering
membrane protein reconstitution, and other useful purposes [13– (DLS), transmission electron microscopy (TEM) techniques and
17]. The morphology of aggregates of two oppositely charged sur- finally zeta potential (f) measurements are applied to characterize
factant molecules depend upon the geometry of surfactant mole- the micelle-vesicle-micelle transition in these oppositely charged
cules, pH of solution mixtures, ratio of mixing of surfactant surfactant assemblies. Additionally, to obtain further information
molecules, etc. [5,8,18–20]. As a consequence, a broad range of regarding the interaction pattern of probe molecules with surfac-
aggregates like micelles, mixed micelles, (disks, ellipsoidal, tant aggregates during aggregate transition, we have monitored
the rotational motion of C153 and R6G in SDBS-C16mimCl and
spheres, or wormlike), multilamellar or unilamellar vesicles are
SDBS-C12mimCl solution mixtures at different mole-fractions of
found to be present in aqueous surfactant mixtures [5,8,19,21].
CnmimCl (vC n mimCl ). As the appearance of larger vesicle aggregates
Recently, researchers are interested to investigate vesicle for-
mation in room temperature ionic liquid (RTILs) or using ionic liq- containing solutions is dense white. Therefore, we have prepared
uids [22–27]. The solvation capability of RTILs increase its small vesicles of SDBS-C16mimCl and SDBS-C12mimCl mixtures at
interaction with conventional surfactant molecules [24,25]. RTILs different vC n mimCl values by probe sonication. The formations of
are organic molten electrolytes that display unique physical char- small vesicles are also structurally characterized by DLS and TEM
acteristics including thermal stability, broad liquidus range, low measurements.
vapor pressure, etc. Due to these special features, RTILs have exten- Rotational motion of fluorophores in organized assemblies
sive use in catalysis, chemical synthesis as well as in separation mainly depends on its location inside organized assemblies along
and extraction processes [28–31]. To establish IL-modified aqueous with the nature of interaction with the surrounding environments
surfactant systems, solvophobic interaction between ionic liquid [49–51]. The solubilization site of probe molecules in surfactant
and amphiphilic (polymer, surfactant) molecules play the crucial assemblies is controlled by the surface charge of vesicle and chem-
role [32–36]. Pandey et al. showed the modification in aggregation ical nature of probe molecules [51]. In micelles, the hydrophobic
behavior of aqueous solution surfactants in presence of ILs [36– molecules generally solubilized in hydrophobic core of micelles
38]. Previously, we have also used ILs having chain length variation and hydrophilic molecules preferentially located at hydrophilic
of anion to modify the properties of zwitterionic micelles or catio- surfaces (interfaces) of micelle. Therefore, the microenvironments
nic micelles [33,34]. Generally, ILs, having short chain, behave like experienced by hydrophobic and hydrophilic molecules are com-
ordinary inorganic salts. But, long hydrophobic alkyl chain contain- pletely different in organized assemblies compared to the aqueous
ing ILs show amphiphilic nature, i.e., surface active properties [39– solution. In vesicles, the inner hydrophilic water pool is enclosed
41]. Basically, they are the members of an important series of ILs. by hydrophobic bilayer formed by surfactant molecules. Therefore,
However, in this present experimental condition they are not ILs. it would be important to study the rotational motion of C153 and
This study deals with the aggregation behavior of the oppositely R6G during micelle-vesicle-micelle transition. Sodium dodecylben-
charged surfactants in aqueous solutions and not about the ILs. zenesulphonate (SDBS) forms tiny micelles in aqueous solution
These type of molecules, having hydrophobic long alkyl chain (vC n mimCl ¼ 0:00). The interaction of imidazolium-based surfactants,
and a charged hydrophilic head group in either of their cationic (1-alkyl-3-methyl imidazolium chloride, CnmimCl, n = 12, 16) with
or anionic moiety enhances the formation of various supramolecu- aqueous SDBS solution results in formation of large vesicles which
lar nano-aggregates in aqueous medium [22,42,43]. The aggrega- subsequently converted into smaller ones by ultrasonic irradiation
tion phenomenon of imidazolium-based surfactants in aqueous using probe sonicator. Hydrophobic molecule, C153 is sparingly
as well as nonaqueous medium has been performed extensively water soluble and therefore, it is expected that it would be solubi-
by applying various techniques [22,42–45]. Recently, lized in the hydrophobic bilayer region of vesicles, inner core of
imidazolium-based surfactants are also used to form vesicle in micelles or mixed micelles at various vC n mimCl values. But due to
aqueous solution. Wang et al. demonstrated the concentration the hydrophilic nature of R6G, its location in SDBS-C16mimCl and
dependent micelle-vesicle transition in aqueous 1-alkyl-3- SDBS-C12mimCl aggregates depends upon the vC n mimCl values. This
methylimidazolium bromides [Cnmim]Br (n = 10, 12, 14) solution can be ascribed due to different solubilization possibilities
764 R. Dutta et al. / Journal of Colloid and Interface Science 490 (2017) 762–773

(micellar or mixed micellar surfaces, hydrophobic bilayer of vesi- kB T


dh ¼ ð1Þ
cles or water pool of vesicles) of R6G in SDBS-CnmimCl aggregates 3pgD
at different vC n mimCl values. To confirm this behavior of R6G, addi-
tionally, we have measured translational diffusion of R6G at differ- where kB, T, D and g denote Boltzmann constant, temperature, dif-
ent vC n mimCl values of SDBS-CnmimCl aggregates using fluorescence fusion coefficient and viscosity, respectively. The details of this
correlation spectroscopy(FCS) techniques. setup is already described in our earlier publication [51].

2. Materials and methods 2.2.3. Transmission electron microscopy (TEM) measurement


The TEM measurements of surfactant aggregates were per-
2.1. Materials formed using the JEOL model JEM 2010 transmission electron
microscope which operates at 120 kV voltage.
Sodium dodecylbenzenesulphonate (SDBS) was obtained from For TEM measurements, 10 lL aqueous solution of SDBS-
Sigma-Aldrich and it might contain a mixture of various isomers CnmimCl (n = 12, 16) mixtures of desired mole fraction was posi-
[61–63]. Still, the isomer of the SDBS, which we have used in this tioned on carbon coated copper grid. Freshly prepared 0.1 wt%
study (shown in Scheme 1) is in higher amount compared to the aqueous uranyl acetate solution was used to observe the morphol-
other isomers and it influences more in all the experiments than ogy of vesicles.
the other isomers.1-alkyl-3-methyl imidazolium chloride (Cn-
mimCl, n = 12, 16) surfactants were purchased from Kanto Chemi- 2.2.4. UV–Vis absorption and emission measurements
cals (98% purity). Rhodamine 6G perchlorate (R6GClO4, laser The absorption spectra of C153 and R6G in different micelle and
grade), and coumarin153 (C153, laser grade) were purchased from vesicle solutions were monitored using Shimadzu (model no. UV-
Exciton. All these materials were used as received without further 2450) spectrophotometer. Simultaneously, the emission spectra
purifications. Milli-Q water was used for the preparation of surfac- were measured using Hitachi (model no. F-7000) spectrofluorime-
tant solutions. ter. To collect the emission spectra of C153 and R6G, the samples
The structures of all these materials were shown in Scheme 1. were excited at 410 nm and 440 nm, respectively.

2.2. Instrumentation
2.2.5. Time resolved anisotropy measurements
The time-resolved anisotropy decays were recorded using time
2.2.1. Turbidity measurement
correlated single photon counting (TCSPC) instrument from IBH (U.
Transmittance of SDBS-C16mimCl, and SDBS-C12mimCl solution
K.) In TCSPC set up, a picosecond diode laser (IBH, Nanoled) of
mixtures at different mole fractions were measured using Shi-
410 nm and 440 nm were used as excitation source. During aniso-
madzu (model no: UV2405) spectrophotometer at 600 nm due to
tropy measurement, a motorized polarizer was used on the emis-
absence of any absorption of the individual aqueous solution of
sion side using a Hamamatsu microchannel plate photomultiplier
SDBS and CnmimCl (n = 12, 16). Thus obtained transmittance val-
tube (3809U). The emission decays were collected at parallel Ik ðtÞ
ues were converted to turbidity.
and perpendicular I? ðtÞ polarizations at regular time intervals until
the desired peak difference between Ik ðtÞ and I? ðtÞ emission decays
2.2.2. Size distribution and zeta potential (f) measurements
was attained. Anisotropy decays were also analyzed using IBH DAS,
To determine the size of the aggregates Malvern Nano ZS instru-
version 6, decay analysis software.
ment with 4 mW He-Ne laser (k = 632.8 nm) and thermostated
sample holder was used. The same instrument was used for zeta
potential (f) measurements. During DLS measurements, the 2.2.6. Fluorescence correlation spectroscopy (FCS) measurements
collected scattering intensities were analyzed to calculate the In our present study to perform Fluorescence correlation spec-
hydrodynamic diameter (dh) of particles using following equation: troscopy (FCS) measurements we used DCS 120 Confocal Laser

O
S O Na
O
Sodium Dodecylbenzenesulfonate
(SDBS)
N
N
N Cl
N Cl

1-dodecyl-3-methylimidazolium chloride 1-hexadecyl-3-methylimidazolium chloride


(C12mimCl) (C16mimCl)

CF3 C2H5HN O NHC2H5


ClO4
H3C CH3
N O O COOC2H5

Coumarin 153 (C-153) Rhodamine 6G (R6G)

Scheme 1. Structures of Sodium dodecylbenzenesulfonate (SDBS), 1-dodecyl-3-methylimidazolium chloride (C12mimCl), 1-hexadecyl-3-methylimidazolium chloride (C16-
mimCl), Coumarin 153 (C 153) and Rhodamine 6G perchlorate (R6G ClO4).
R. Dutta et al. / Journal of Colloid and Interface Science 490 (2017) 762–773 765

Scanning Microscope (CLSM) system (Becker & Hickl DCS-120) tures were sonicated using probe sonicator and finally used for
with an inverted optical microscope of Zeiss (Carl Zeiss, Germany). spectroscopic measurements. During Spectroscopic measure-
For the diffusion of a single component system, a three dimen- ments, the corresponding fluorophore solution was allowed to
sional (3D) fitting model can be defined as make solvent free through evaporation and then the desired mole
    1 fractions of CnmimCl were added for the proper encapsulation of
1 s 1 s 2 the probe inside the micelle or vesicle aggregates.
GðsÞ ¼ 1þ 1þ 2 ð2Þ
N sD S sD
3. Results and discussion
GðsÞ is called autocorrelation function for simple solution con-
taining one diffusing molecule (excluding the contribution of tri-
3.1. Phase behavior study
plet state). In this equation, the average number of fluorescent
molecules in the detection volume is N, and the average time of
In aqueous solution, surfactant molecules form tiny aggregates
fluorescent molecules diffusing in the detection volume is sD . S is
called micelles of few nanometers in size. But, mixing of two oppo-
the structure parameter which is equal to the xxxyz where xz is the
sitely charged amphiphilic molecules in certain mole fractions,
longitudinal radius and xxy is the transversal or waist radius of alter their aggregation behavior. These changes are reflected in
the confocal volume. The relation between sD and the translational their phase behavior, which can be observed easily by naked eye.
diffusion coefficient is as follows: Initially with the addition of imidazolium-based surfactants in
aqueous SDBS solution, the turbidity of the solution increases shar-
x2xy
sD ¼ ð3Þ ply and it attains maxima around 0.45 mol fraction of
4D imidazolium-based surfactant. After vC n mimCl ¼ 0:45 the solution
The structure parameter (S) of the excitation volume was deter- becomes more turbid and sometimes precipitation appears which
mined using R6G in water as a reference sample of known diffusion indicates the formation of larger aggregates. At higher mole frac-
coefficient (Dt = 426 lm2 s1). The value of structure parameter is tion of imidazolium-based surfactant, i.e., vC n mimCl > 0.65, turbid-
found to be 5. This value has been fixed for the analysis of all data ity again decreases. These visual changes in phase behavior of
obtained in our systems. Fitting of the FCS data of R6G in water surfactant and imidazolium-based surfactants mixtures are moni-
resulted in the diffusion time of 73.6 ls which has been used to tored using UV–vis spectrophotometer by measuring the variation
calculate the transverse radius and the confocal volume of our of turbidity of the solution mixtures. The variation of turbidity at
set-up. The calculated transverse radius is 354 nm and the confo- different mole fractions of CnmimCl in SDBS-CnmimCl solutions is
cal volume is 1.39 fL. The details of this setup is already described shown in Fig. S1. Increase in mole fraction of CnmimCl from 0.20
in our earlier publication [51]. to 0.45 (vC n mimCl ¼ 0:20 to vC n mimCl ¼ 0:45), the solutions become
turbid indicating the formation of large vesicle aggregates. Besides,
2.3. Solution preparation further increase in mole fraction of CnmimCl, i.e., vC n mimCl values,
the solution becomes more turbid and appearance of white precip-
We have used SDBS as an anionic surfactant. This type of linear itate confirms the spontaneous aggregation of normal vesicles to
alkyl benzene sulfonate(ABS) surfactants generally consist of a larger ones. Actually, when the oppositely charged surfactants
complex mixture of different chain lengths and positional isomers. are close to their equimolar ratio [24], there occurs charge neutral-
Ma et al., have investigated the solubility, self-assembly and phase ization due to strong electrostatic attraction between the surfac-
behavior of the six positional isomers of SDBS and the obtained tants [25] and as a result phase separation takes place and
phase diagrams show different phase properties for different iso- precipitate settles out from aqueous solution. Precipitation indi-
mers [61]. They also showed that, with increase in the concentra- cates the growth of the vesicles as the vesicles become larger in
tion of surfactants, three of the six isomeric solutions of SDBS size [60]. It is well-established that oppositely charged surfactants
undergo transition from a micellar, to a lamellar phase through a can form different liquid crystalline phases due to the presence of
hexagonal phase, whereas the hexagonal phase is absent for the electrostatic and hydrophobic interactions [66,67]. There are vari-
transition of other three isomers of SDBS. This difference is due ous crystalline states like cubic, micellar, hexagonal, lamellar, etc.
to the variation of local molecular and global aggregate packing [66]. Bilayer (La) phase can be categorized into two phases, (a) pla-
constraints of the isomers. Klein and co-workers have also nar lamellar phase (Lal) and (b) closed bilayer structure i.e., vesicle
obtained similar kind of observations using molecular dynamics phase (Lav). However in this present study, we haven’t found any
study [62]. Thus, the mixture of isomers of SDBS may affect the lamellar phase during TEM measurements. There are experimental
physicochemical properties in aqueous solution. However, there evidences, which tell that vesicle phase forms spontaneously due
will be greater proportion as well as higher contribution of the iso- to thermodynamic stability [67]. Due to this reason, we are obtain-
mer of SDBS (shown in Scheme 1) which we are actually willing to ing vesicle phase only and no lamellar phase. At higher molar ratio
use in this present study. At first, water was filtered using syringe (vC n mimCl > 0:65), the sharp decrement in turbidity of the solution
filter (0.2 lm) to remove dust particles. Afterwards, 150 mM aque- indicates the transition of larger vesicles to smaller vesicles, mixed
ous solution of SDBS, C16mimCl, and C12mimCl were prepared sep- micelles or micelles. The vesicle region of these surfactant mix-
arately in different volumetric flasks. These solution mixtures were tures are quite broad and exists at low vC n mimCl values also. It is
stirred and kept overnight for stabilization. Then the various sets of interesting to note that at higher vC n mimCl , the decrease in turbidity
solution mixtures containing SDBS-C16mimCl, and SDBS-C12mimCl
is more rapid for SDBS-C16mimCl solution compared to SDBS-
of desired mole fraction [vC n mimCl = mole fraction of CnmimCl
C12mimCl solution indicating the transition from vesicle to micelle
vC n mimCl
(n = 12, 16) = v v ] were prepared using appropriate volume or mixed micelles. Due to the presence of strong hydrophobic
C n mimCl þ SDBS

of each stock solution. The solution mixtures were then stirred and interaction between the surfactant and alkyl chain of C16mimCl,
thermostated at 298 K for 2–3 days for stabilization. the system reorganizes accordingly, so that surfactant molecules
Finally, these solution mixtures were used to characterize the can easily incorporate in imidazolium-based surfactant micelles
aggregate transition between micelle-vesicle-micelle by visual which induces the formation of mixed micelles with much smaller
observation, dynamic light scattering (DLS) measurement, zeta sizes. In SDBS-C16mimCl aggregates, at higher vC 16 mimCl values, the
potential study, TEM images, and eventually these solution mix- system contains cation rich mixed micelles composed of SDBS
766 R. Dutta et al. / Journal of Colloid and Interface Science 490 (2017) 762–773

and C16mimCl. Vesicle formation from oppositely charged surfac- 3.3. Transmission electron microscopy (TEM) studies
tants are well reported [14,15]. Several groups have also reported
vesicle formation involving of one conventional surfactant and The formation of vesicles in mixtures of SDBS-CnmimCl (n = 12,
the another oppositely charged imidazolium-based surfactant 16) are further characterized by the analysis of TEM images. The
[19,24,25]. Moreover, two oppositely charged imidazolium-based TEM micrographs of aqueous solutions of SDBS-C12mimCl, and
surfactants can also form vesicles [26]. The driving force of vesicle SDBS-C16mimCl at vC n mimCl = 0.50 are shown in Fig. 1 (a and b for
formation is the electrostatic and hydrophobic interactions SDBS-C16mimCl and c–e for SDBS-C12mimCl, respectively). The
between the two oppositely charged surfactants. The differences TEM images clearly indicate the formation of spherical vesicular
between a conventional surfactant and an imidazolium surfactant aggregates for each oppositely charged surfactant-mixtures. The
of same chain length appears due to the variation of charge density average diameters of these spherical aggregates are more than
and geometric packing of the head groups of the surfactants. 300 nm. The TEM images of SDBS-C16mimCl and SDBS-
Therefore, the phase behaviors of conventional surfactant and C12mimCl solution mixtures at these certain mole fractions also
imidazolium-based surfactants are also very distinct [24]. Positive suggest a broad range of distribution in size of the vesicles. The
charge delocalization is also higher for imidazolium cations com- synergistic interaction between the cationic imidazolium moieties
pared to alkyl substituted ammonium cations and thus the charge with sulfate group of surfactant and also hydrophobic interaction
density of the head groups of the surfactants cause the solubility between alkyl parts of the concerned pairs are responsible for
mismatch [64]. Kumar and co-workers have explained the advan- the formation of these stable vesicles.
tages of imidazolium cations that they can prevent precipitation
by maintaining the electrostatic interactions and thus increase 3.4. Zeta potential (f) measurements
the concentration range of vesicular aggregates [65].
Zeta potential (f) signifies the surface charge of the colloidal
3.2. Dynamic light scattering (DLS) measurements particles in aqueous solution. It is the electrical potential difference
between stationary layer of fluid related to micellar aggregates and
To observe the structural modification of SDBS micelles by dispersion medium which stretches out from the particle surface
imidazolium-based surfactants, dynamic light scattering (DLS) due to thermal motion of solvent molecules. Better information
experiment has been performed at different vC n mimCl values. The about potential stability of the aggregates can also be obtained
from f values.
SDBS molecules form tiny micelles in aqueous solution as obtained
In general, the surface of SDBS micelles is negatively charged
by DLS measurement. But variation in size of solution mixtures at
and as a result it shows the f value about 36.5 ± 2.0 mV. Initially,
different vC n mimCl values show regular increase with increase in
on the addition of imidazolium-based surfactant, f value decreases
vCn mimCl values and at higher vCn mimCl values the size of aggregates
and in SDBS rich region the overall surface charge of vesicles is
again decreases. Therefore, the observed variations in DLS profiles
negative. Now, on further addition of imidazolium-based surfac-
indicate micelle-vesicle-micelle transition in SDBS-CnmimCl solu- tants, surface charge increases and becomes positive in CnmimCl
tion mixtures. The intensity-size distribution profiles in SDBS-
rich aggregates. The variation of zeta potential values (Fig. S4) of
C16mimCl and SDBS-C12mimCl aggregates at certain mole fraction SDBS-C16mimCl, and SDBS-C12mimCl solution mixtures also pro-
of imidazolium-based surfactants are shown in Figs. S2 and S3
vide an indication about the incorporation efficiency of different
(Supporting Information), respectively. The broad distribution of imidazolium-based surfactants into micellar aggregates. The
DLS profiles indicate that the aggregate sizes are more than
increase in f value is more rapid in SDBS-C16mimCl solution com-
300 nm, although small vesicles, and mixed micelles are also pre- pared to SDBS-C12mimCl solutions. The strong interaction of long
sent along with the larger aggregates. It is interesting that at higher
alkyl chain containing imidazolium cation with the sulfate anion
vC16 mimCl values, i.e., at vC16 mimCl = 0.80, the DLS profiles of SDBS- of SDBS is responsible for the incorporation of C16mimCl into SDBS
C16mimCl indicate that the system mainly composed of mixed micellar assemblies effectively. The observed trend in variation of
micelles whereas in SDBS-C12mimCl solution mixtures, at zeta potential values clearly suggests micelle-vesicle-micelle tran-
vC12 mimCl = 0.80, mainly vesicular aggregates are present in solution. sition in aqueous SDBS-CnmimCl solution.
At vC n mimCl > 0.90, the systems contain mainly micellar phase. But,
at vC n mimCl = 1.00 an intense peak (minor population) is observed 3.5. Transformation of large vesicles to small ones: Preparation and
in the larger diameter range, which indicates the presence of larger characterizations
aggregates. As the light scattering intensity (Iscatter) is directly pro-
portional to the sixth power of the diameter of particles (d6), in the 3.5.1. Preparation of small vesicles by ultrasonication
larger diameter range (i.e., at minor population) an intense peak is The mixing of two oppositely charged SDBS and CnmimCl solu-
also noticed. The DLS measurements along with turbidity study tions lead to the large aggregates formation and the solution
clearly indicate the micelle-vesicle-micelle transition in aqueous becomes turbid. Therefore, to prepare the small aggregates, these
SDBS-C16mimCl and SDBS-C12mimCl aggregates with the variation mixtures are sonicated using ultrasonicator (Processor SONOPROS
in vC n mimCl values. PR-250 MP, Oscar Ultrasonics Pvt. Ltd. India) having titanium

(a) (b) (c) (d) (e)

Fig. 1. TEM images of SDBS-C16mimCl (a and b) and SDBS-C12mimCl (c–e) solution at vCn mimCl = 0.50.
R. Dutta et al. / Journal of Colloid and Interface Science 490 (2017) 762–773 767

probe (frequency 20 ± 3 kHz). The solution is then diluted with In the next section, the location and dynamics of a hydrophobic
Milli-Q water (1:1 ratio) and centrifuged to remove the titanium molecule (coumarin 153, C153) and a hydrophilic molecule (rho-
particles (if present). These solutions are suitable for spectroscopic damine 6G perchlorate, R6GClO4) have been investigated in
measurement. We have chosen seven different mole fractions SDBS-C16mimCl and SDBS-C12mimCl solution mixtures at different
(vC n mimCl = 0.00, 0.20, 0.35, 0.55, 0.65, 0.80 and 1.00) of SDBS- vC n mimCl values to observe how the location of dye molecule alter
C12mimCl and SDBS-C16mimCl mixtures for spectroscopic mea- during micelle-vesicle-micelle transition.
surements. We have chosen these vC n mimCl values where initial
two vC n mimCl values (vC n mimCl = 0.20 and 0.35), the surface charge of 3.6. Binding or location of hydrophobic and hydrophilic probe
SDBS-CnmimCl aggregates are negative. But at higher vC n mimCl val- molecules (C153 and R6G) in SDBS-C16mimCl and SDBS-C12mimCl
ues (vC n mimCl = 0.55, 0.65 and 0.80), the surface charge of SDBS- aggregates
CnmimCl aggregates becomes positive. The photographs of SDBS-
C16mimCl and SDBS-C12mimCl solution mixtures are shown in
Fig. S5 (Supporting Information). 3.6.1. Steady-state absorption and emission studies
The steady state absorption and emission spectra of C153 and
R6G were monitored in SDBS-C16mimCl and SDBS-C12mimCl
3.5.2. Characterizations of small vesicles by dynamic light scattering aggregates at all the above mentioned mole fractions of CnmimCl
(DLS) and transmission electron microscopy (TEM) to study the microenvironmental change around the probe mole-
Finally, DLS experiment is carried out to get the idea about the cules upon micelle-vesicle-micelle transition. The variation of
size of SDBS-C16mimCl and SDBS-C12mimCl aggregates. The absorption and emission spectra of R6G and C153 in SDBS, C16-
intensity-size distribution profiles of SDBS-C16mimCl and SDBS- mimCl and C12mimCl micellar solutions are shown in Figs. S8
C12mimCl aggregates are shown in Figs. S6 and S7, respectively. and S9, respectively. In water, C153 shows an emission maximum
DLS profiles indicate that the sizes of SDBS-C12mimCl aggregates at 549 nm. In neat SDBS solution, C153 exhibits intense emission
are around 80–100 nm range at all compositions (i.e., maxima around 540 nm whereas neat C16mimCl and C12mimCl
vC12 mimCl = 0.20, 0.35, 0.65, 0.80). At vC 12 mimCl = 0.55, the solution solution, it shows emission maxima around 536 nm and
mixture of SDBS-C12mimCl mixture is hazy (phase separation 537 nm respectively. Therefore, the large blue-shifted emission
occurs on standing), therefore, DLS measurement cannot be per- maxima of C153 in micellar solution compared to water indicate
formed at this particular mole fraction. But in SDBS-C16mimCl mix- that the probe molecules are solubilized in micellar media of sur-
ture, diameter of aggregates at vC 16 mimCl = 0.20, 0.35, 0.55, 0.65 is factants. But, the variation of emission spectra of C153) in SDBS-
around 80–100 nm whereas at higher mole fraction C16mimCl and SDBS-C12mimCl solution shown in Figs. S10 and
(vC 16 mimCl = 0.80) the diameter of the aggregate is less, around S11 (Supporting Information) respectively, at different vC n mimCl val-
10 nm. The DLS measurement along with visual observation sug- ues suggest that probe molecules are preferentially located in the
gest that SDBS-C16mimCl mixture at higher mole fraction of bilayers instead of water pool of the vesicles. The close inspection
imidazolium-based surfactant forms cation rich mixed micelles. of above mentioned spectral profiles clearly suggest that variations
To confirm the vesicle formation, TEM measurements were per- of emission intensity are more prominent at the blue end of the
formed for both the solution mixtures. TEM micrographs are emission spectra of oppositely charged surfactant mixtures at dif-
depicted in Figs. 2 and 3 for SDBS-C16mimCl and SDBS-C12mimCl ferent vC n mimCl values. These observations provide a clear idea that
solution, respectively at vC n mimCl = 0.20, and vC n mimCl = 0.65. The for- the micropolarity experienced by the fluorophores is also depends
mation of spherical vesicles with distinct bilayer can be clearly on the variation of SDBS and CnmimCl concentration.
visualized from TEM images and which also nicely correlates with Interestingly, the red shifts in absorption and emission maxima
DLS study. (Fig. S8) of R6G in micellar assemblies compared to water indicate

(a) (b) (c)

(d) (e) (f)

Fig. 2. TEM images of SDBS-C16mimCl solution mixture at vC16 mimCl = 0.20 (a–c) and vC16 mimCl = 0.65 (d–f).
768 R. Dutta et al. / Journal of Colloid and Interface Science 490 (2017) 762–773

(a) (b) (c)

(d) (e) (f)

Fig. 3. TEM images of SDBS-C12mimCl solution mixture at vC12 mimCl = 0.20 (a–c) and vC12 mimCl = 0.65 (d–f).

the binding of it to surface of the micelles. At lower mole fraction of the horizontally and vertically polarized components of the emis-
imidazolium-based surfactant (i.e., vC n mimCl = 0.20, 0.35), slight red sion intensity (HH and HV), each for the same period of time.
shift in absorption and emission spectra of R6G is observed in The ratio of IHV and IHH expresses the G value of the system; i.e.,
SDBS-C16mimCl and SDBS-C12mimCl aggregates shown in G = IHV/IHH and for our TCSPC set up, G value is 0.6.
Figs. S12 and S13 (Supporting Information), respectively. This The anisotropy decays of C153 and R6G in micellar solution
observation suggests that R6G molecules bind to the surface of (SDBS, C16mimCl, and C12mimCl), are shown in Fig. 4. For all solu-
the SDBS-CnmimCl aggregates. Zeta potential measurements indi- tions, the anisotropy decays of C153and R6G are biexponential in
cate that surface charge of the aggregates at the above mentioned nature and the corresponding decay parameters are shown in
mole fraction is negative, so, electrostatic interaction with R6G Tables 1and 2, respectively. The average rotational time constant
allows it to bind to the surface efficiently. Again, at (sr ) can be calculated utilizing the following equation:
vCn mimCl ¼ 0:55 or 0:65, the surface charge of the aggregates
becomes positive and also increased hydrophobicity allows R6G sr ¼ a1 s1 þ a2 s2 ð5Þ
molecules to migrate into the aqueous pool of the vesicles. As a
Anisotropy decays of C153 in water is single exponential in nat-
result, blue shift is observed in absorption and emission spectral
ure with a time constant of 100 ps [19]. Though, in micelles and
profiles of R6G. But at higher mole fraction of imidazolium-based
vesicles, the observed anisotropy decays are biexponential with
surfactant, although the surface charge of aggregates is high, again
longer rotational time constants. In confined environment of
red shift in absorption and emission spectra suggests binding of
micelles or vesicles, the rotational motion of probe molecules
R6G in the mixed micelles or vesicle surface. This observation also
becomes hindered and as a result rotational relaxation time of
suggests that in cation rich mixed micelles or vesicles, the
C153 increases. The average rotational relaxation time of C153 in
hydrophobicity of the vesicle bilayer is less than the anion rich
aqueous SDBS solution is 1.10 ns with components of 2.14 ns
vesicles. Therefore, in conclusion, it can be considered that surface
(39%) and 0.43 ns (61%). Again the rotational time constants of
charge and the hydrophobicity of the aggregates are the controlling
C153 in C16mimCl and C12mimCl solution are 0.72 ns (with com-
factors for the motion or binding of R6G molecule in SDBS-
ponents of 1.76 ns (44%) and 0.28 ns (56%)) and 0.57 ns (with
CnmimCl aggregates.
components of 1.13 ns (31%) and 0.27 ns (69%)), respectively.
Therefore, the chain length of imidazolium-based surfactant also
3.6.2. Time-resolved anisotropy studies influences the anisotropy of C153 molecules. Moreover, it is found
The time-resolved anisotropy measurements provide better that, with the increment in chain length of imidazolium-based sur-
information regarding the location of C153 and R6G during factant, the contribution of slow component increases significantly.
micelle-vesicle-micelle transition in mixed surfactant aggregates. In SDBS-C16mimCl aggregates at vC 16 mimCl ¼ 0:20, the average rota-
To calculate the time-resolved anisotropy (r(t)) the following equa- tional time increases to 1.76 ns. But at higher vC n mimCl values, it
tion is used: again decreases. Similar phenomenon is observed for SDBS-
C12mimCl aggregates. Interestingly, the decreases in rotational
Ik ðtÞ  GI? ðtÞ
rðtÞ ¼ ð4Þ time constant from vC n mimCl ¼ 0:65 to 0:80, is more prominent for
Ik ðtÞ þ 2GI? ðtÞ
SDBS-C16mimCl solution compared to SDBS-C12mimCl aggregates.
Ik ðtÞ and I? ðtÞ represent the emission decays which are polar- The rotational time constant decreases 34% for SDBS-C16mimCl
ized in parallel and perpendicular direction with respect to the aggregates; whereas for SDBS-C12mimCl aggregates only 24%
direction of polarization of the excitation light, respectively. G is decrease is observed. This scenario is easily understandable by
the experimental correction factor and does not depend to the considering the visual observation, DLS measurement of SDBS-
property of the sample. G factor can be measured by exciting the CnmimCl solution at vC n mimCl = 0.80. The observation suggests that
sample using the horizontally polarized light and then measuring at this mole fraction SDBS-C16mimCl mixture remains as mixed
R. Dutta et al. / Journal of Colloid and Interface Science 490 (2017) 762–773 769

0.4 0.4
R6G SDBS C 153 SDBS
C16mimCl
C16mimCl
0.3 C12mimCl 0.3
C12mimCl

r (t)

r (t)
0.2 0.2

0.1 0.1

0.0 0.0
0 3 6 9 12 15 0 3 6 9 12 15
Time (ns) Time (ns)

Fig. 4. Anisotropy decays of R6G and C 153 in SDBS, C16mimCl and C12mimCl.

Table 1
Anisotropy Decay Parameters of C-153 in SDBS, C16mimCl, C12mimCl, SDBS-C16mimCl and SDBS-C12mimCl surfactant mixture at different mole fractions.

System vCn mimCl aslow sslow afast sfast hsr ia


(n = 12, 16) (ns) (ns) (ns)

Neat SDBS 0.00 0.39 2.14 0.61 0.43 1.10


Neat C16mimCl 1.00 0.22 1.76 0.78 0.43 0.72
Neat C12mimCl 1.00 0.18 1.48 0.82 0.37 0.57
SDBS-C16mimCl 0.20 0.54 2.81 0.46 0.52 1.76
0.35 0.53 2.80 0.47 0.53 1.73
0.55 0.50 2.75 0.50 0.52 1.64
0.65 0.48 2.53 0.52 0.34 1.39
0.80 0.46 1.67 0.54 0.32 0.92
SDBS-C12mimCl 0.20 0.43 2.64 0.57 0.41 1.37
0.35 0.42 2.42 0.58 0.38 1.24
0.65 0.40 2.44 0.60 0.39 1.21
0.80 0.31 2.16 0.69 0.37 0.92
a
Experimental error 5%.

Table 2
Anisotropy Decay Parameters of R6G in SDBS, C16mimCl, C12mimCl, SDBS-C16mimCl and SDBS-C12mimCl surfactant mixture at different mole fractions.

System vCn mimCl aslow sslow afast sfast hsr ia


(n = 12, 16) (ns) (ns) (ns)

Neat SDBS 0.00 0.84 3.82 0.16 0.77 3.33


Neat C16mimCl 1.00 0.61 2.77 0.39 0.62 1.93
Neat C12mimCl 1.00 0.69 1.51 0.31 0.40 1.17
SDBS-C16mimCl 0.20 0.82 6.10 0.18 0.26 5.05
0.35 0.80 5.88 0.20 0.21 4.75
0.55 0.24 4.15 0.76 0.19 1.14
0.65 0.14 3.95 0.86 0.20 0.73
0.80 0.49 2.95 0.51 0.33 1.61
SDBS-C12mimCl 0.20 0.78 5.39 0.22 2.55 4.77
0.35 0.77 5.88 0.23 0.36 4.61
0.65 0.23 4.04 0.77 0.21 1.09
0.80 0.43 2.07 0.57 0.23 1.02
a
Experimental error 5%.

micelles, but, SDBS-C12mimCl solution exists as vesicles. As vesi- and 0.77 ns (14%)), 1.93 ns (with components of 2.77 ns
cles provide more restriction on the rotational motion of (61%) and 0.62 ns (39%)) and 1.17 ns (with components of
hydrophobic molecule, C153 than micelles or mixed micelles, 1.51 ns (69%) and 0.40 ns (31%)) respectively. The cationic
therefore, the anisotropy decays become faster in SDBS-C16mimCl R6G molecules strongly bind to the surface of negatively charged
aggregates than SDBS-C12mimCl aggregates between SDBS micellar aggregates, consequently, the rotational time con-
vCn mimCl ¼ 0:65 and vC n mimCl ¼ 0:80. The observed variation of aniso- stant of R6G is higher in SDBS micelle compared to the cationic
tropy values of C153 at different vC n mimCl value suggests micelle- micellar solution of C16mimCl and C12mimCl. Moreover, between
vesicle-micelle transition in SDBS-C16mimCl and SDBS-C12mimCl C16mimCl and C12mimCl micellar solution, the anisotropy decay
solution. of R6G becomes slower in C16mimCl micelle as it offers better
The anisotropy decays of R6G in micellar solution of SDBS, C16- restriction on rotational motion of R6G due to longer alkyl chain
mimCl and C12mimCl exhibit biexponential nature with average length. The rotational time constant of R6G increases significantly
time constant of 3.33 ns (with components of 3.82 ns (84%) at vC n mimCl ¼ 0:20 and vC n mimCl ¼ 0:35 for both SDBS-C16mimCl and
770 R. Dutta et al. / Journal of Colloid and Interface Science 490 (2017) 762–773

SDBS-C12mimCl aggregates. In aqueous medium the anisotropy picture about the location of probe molecules in SDBS-C16mimCl
decay of R6G is single exponential, having a time constant around and SDBS-C12mimCl aggregates during micelle-vesicle-micelle
0.26 ns [51]. In vesicular solution, the fast component of aniso- transition.
tropy decay of R6G similar to time constant in aqueous phase
and the slow component corresponds to the rotation of R6G
3.6.3. Diffusion of R6G in SDBS-C16mimCl and SDBS-C12mimCl
entrapped in vesicle bilayer. As the surface charge of SDBS-
aggregates
CnmimCl is negative (as indicated by zeta potential measurement),
In order to get better insight about the micelle-vesicle-micelle
R6G efficiently binds to the vesicle bilayer which results in
transition in SDBS-C16mimCl and SDBS-C12mimCl solution mix-
increased rotational time constant. At vC n mimCl ¼ 0:55 or 0:65, the
tures, the translational diffusion of R6G has been monitored using
surface charge of the aggregates becomes positive and the
fluorescence correlation spectroscopic (FCS) technique. In FCS
hydrophobicity of vesicle bilayer is also higher compared to
technique, the fluctuation of fluorescence intensity of dye molecule
micelle. Therefore, R6G molecules may migrate towards the inner
is monitored in a small observation volume (1 fL) and eventually
water pool or outer aqueous phase from the vesicle bilayer. As
these fluctuations in emission intensity are correlated to generate
depicted in Table 2, the anisotropy decays of R6G at
the autocorrelation function. This correlation function can provide
vCn mimCl ¼ 0:55 or 0:65 in both aggregates of SDBS-C16mimCl and information regarding different molecular processes e.g., reaction
SDBS-C12mimCl is also biexponential. But the relative contribution kinetics, translational diffusion, conformation fluctuation, etc.
of fast component of anisotropy decay increases significantly at [52]. Therefore, this methodology has been applied extensively to
vCn mimCl ¼ 0:55 or 0:65 compared to vCn mimCl ¼ 0:20 or 0:35. There- monitor bimolecular reaction, conformational dynamics of biomo-
fore, the variation of anisotropy values and their relative contribu- lecules or translational diffusion in different organized assemblies
tion suggest that majority of the R6G molecules located in micelle- and ionic liquids [53–58].
water interface or outer aqueous phase of vesicles at FCS is an efficient technique to monitor the mobility of fluo-
vCn mimCl ¼ 0:55 or 0:65. But at in SDBS-C16mimCl solution mixture, rophore in various complex environments and the morphology,
at vC n mimCl ¼ 0:80, the solution mixtures mainly contains mixed structural heterogeneity of organized assemblies can influence
micelles, therefore, R6G molecules binds in hydrophobic surface the diffusion coefficient (Dt) of dye molecule [51,58]. In SDBS-
of mixed micelles, rotational time constant again increases. In CnmimCl aggregates, absorption-emission spectra and time
SDBS-C12mimCl solution mixtures, the systems mainly exist as resolved anisotropy measurements clearly indicate that motion,
vesicles and it is expected that R6G mainly located in inner core i.e., location of R6G can provide better visualization about
of vesicles or outer aqueous phase. As a result, the rotational time micelle-vesicle-micelle transition in surfactant assemblies than
constant of R6G decreases. Anisotropy decays of R6G and C 153 in C153. As R6G is a hydrophilic cationic dye molecule; therefore,
SDBS-C12mimCl and SDBS-C16mimCl mixtures at different mole the motion of R6G strongly depends on hydrophobicity along with
fractions are shown in Figs. 5 and 6 respectively. So, the average the surface charge of bilayer. As being hydrophobic in nature, C153
rotational time constant values of R6G can provide the actual preferentially resides in the bilayer of vesicles and experiences

0.4 0.4
SDBS-C16mimCl χ C16mimCl =0.20 SDBS-C12mimCl χ C mimCl =0.20
12
χ C16mimCl =0.35 χ C mimCl =0.35
R6G χ C16mimCl =0.55 R6G 12

0.3 0.3 χ C mimCl =0.65


χ C16mimCl =0.65 12

χ C16mimCl =0.80 χ C mimCl =0.80


12
r (t)

r (t)

0.2 0.2

0.1 0.1

0.0 0.0
0 3 6 9 12 15 0 3 6 9 12 15
Time (ns) Time (ns)

Fig. 5. Anisotropy decays of R6G in SDBS-C12mimCl and SDBS-C16mimCl mixtures at different mole fractions.

0.4 0.4
χ C16mimCl =0.20 χ C12mimCl =0.20
C 153 C 153
SDBS-C16mimCl χ C16mimCl =0.35 SDBS-C12mimCl χ C12mimCl =0.35
0.3 χ C16mimCl =0.55 0.3 χ C12mimCl =0.65
χ C16mimCl =0.65
χ C12mimCl =0.80
χ C16mimCl =0.80
r (t)
r (t)

0.2 0.2

0.1 0.1

0.0 0.0
0 3 6 9 12 15 0 3 6 9 12 15
Time (ns) Time (ns)

Fig. 6. Anisotropy decays of C 153 in SDBS-C12mimCl and SDBS-C16mimCl mixtures at different mole fractions.
R. Dutta et al. / Journal of Colloid and Interface Science 490 (2017) 762–773 771

similar microenvironment in vesicle assemblies at all vC n mimCl val- aggregates, at vC n mimCl ¼ 0:80, due to formation of mainly mixed
ues. Therefore, the diffusion of R6G molecule is monitored during micelles containing SDBS and C16mimCl, the relative contribution
micelle-vesicle-micelle transition in aqueous SDBS-C16mimCl and of diffusion coefficient increases significantly compared to
SDBS-C12mimCl solution mixtures at different vC n mimCl values. In vC n mimCl ¼ 0:55 or vC n mimCl ¼ 0:65. Similarly, in SDBS-C12mimCl
vesicle, the aqueous core is surrounded by surfactant bilayer dis- aggregates, due to alteration in the location of R6G molecules at
persed in aqueous medium; whereas the hydrophobic tail part of different vC n mimCl values variation in diffusion time is also observed.
surfactant molecules is sequestered into the hydrophobic core in In SDBS-C16mimCl and SDBS-C12mimCl aggregates, at different
surfactant micelle. Therefore, the basic interest of the present vC n mimCl values the translational diffusion motion of R6G clearly
study is to observe how R6G molecules oscillate between the sur- depicts micelle-vesicle-micelle transition in surfactant assemblies.
face of micelle, vesicle or aqueous core during micelle-vesicle- The diffusion motion of R6G in SDBS-C12mimCl and SDBS-
micelle transition. C12mimCl aggregates clearly suggests that the location of R6G is
The FCS traces of R6G in water, SDBS-C16mimCl and SDBS- strongly influenced during micelle-vesicle-micelle transition. It is
C12mimCl aggregates at different vC n mimCl values are presented in found that the diffusion motion ofR6G is 18 times slower in anio-
Fig. 7. The FCS trace of R6G in water is fitted using one- nic SDBS micelles. But in cationic C16mimCl and C12mimCl
component fitting equation with diffusion coefficient (Dt) of micelles, the diffusion times of R6G are 9 times and 7 times
426 lm2 s1. But in SDBS-C16mimCl and SDBS-C12mimCl aggre- slower than in water. As R6G is a cationic dye, the diffusion motion
gates, two component fitting equation is used to fit the experimen- becomes slower in anionic SDBS micelles than in cationic C16-
tally obtained FCS curves. During FCS curve fitting by two mimCl or C12mimCl micelles. However, in SDBS-CnmimCl (n = 12,
component fitting equation, one component i.e., Dt value of R6G 16) aggregates, variation in Dt values and their relative contribu-
in water is fixed and evaluated the other component. This other tion signifies that with change in vC n mimCl values, location of R6G
component corresponds to diffusion of R6G entrapped in micelle molecules alter spontaneously, i.e., R6G molecules oscillate
or vesicle aggregates. The calculated diffusion coefficients in these between aqueous phase and surfactant containing aggregates.
systems are given in Table S1 (Supporting Information).
When cationic R6G molecules bind to anionic SDBS micelles,
the diffusion motion ascribed to the motion of that self- 4. Conclusions
assembly; therefore, the slow diffusion of R6G molecules are
observed in SDBS micelles compared to the diffusion of R6G in bulk There are several literature reports on the formation of
water. This clearly indicates efficient binding of the R6G within the micelles/mixed micelles and vesicles from oppositely charged sur-
SDBS micelles. Similarly, in micellar solution of C12mimCl and C16- factants through fine regulation of the mixing ratio of surfactants
mimCl, we have observed slow diffusion of R6G compared to i.e., the anionic/cationic charge ratio [1–3,24,25]. In the previous
water. Moreover, between C16mimCl and C12mimCl micellar solu- studies, our group has also reported micelle-vesicle-micelle transi-
tion, the Dt value of R6G is lower in C16mimCl micellar solution. tion from cationic/anionic surfactants and/or imidazolium-based
The C16mimCl contains more number of alkyl chain, therefore, surfactants. Furthermore, they have monitored the rotational
the translational motion of R6G encapsulated in C16mimCl micelle motion of hydrophobic dye C153 during the aggregate transition
is slower than C12mimCl micelle. In vesicular solution at [19,26]. Cholesterol/surfactant mixture was also used to prepare
vesicles and subsequently the translational diffusion of hydrophilic
vCn mimCl ¼ 0:20 and vCn mimCl ¼ 0:35, the diffusion time of R6G
probe R6G was studied during this vesicular transition [51].
increases significantly compared to the diffusion time of R6G in
Recalling these observations, we further extend the work in
SDBS micelle in both SDBS-C16mimCl and SDBS-C12mimCl
present study. In this article, the aggregate transition has been
aggregates. Due to the negative surface charge, R6G
investigated from micelles to vesicles and then again micelles in
molecules strongly bind to the surface of vesicle bilayer
mixed system containing anionic surfactant SDBS and cationic
and therefore, their translational motion is very slow. But at
imidazolium-based surfactants, C16mimCl and C12mimCl, respec-
vCn mimCl ¼ 0:55 and vC n mimCl ¼ 0:65, the diffusion of R6G becomes
tively. Moreover, we have also investigated how the location of a
faster compared to SDBS micelle solution in SDBS-C16mimCl aggre-
hydrophilic dye R6G and a hydrophobic dye C153 alter during
gates. It is observed that the relative contribution of that compo-
micelle-vesicle-micelle transition using fluorescence spectroscopic
nent decreases significantly at vC n mimCl ¼ 0:55 and vC n mimCl ¼ 0:65 techniques. Steady state absorption-emission and time resolved
compared to vC n mimCl ¼ 0:20 and vC n mimCl ¼ 0:35. This observation anisotropy measurements indicate that the location of C153
suggests that at these mole fractions (vC n mimCl ¼ 0:55 and remains almost unaltered in vesicular aggregates at different
vCn mimCl ¼ 0:65) small number of R6G molecules preferentially vC n mimCl values whereas the location of cationic R6G alters. As
solubilized into the vesicle aggregates. But in SDBS-C16mimCl C153 molecule is neutral and hydrophobic, the rotational motion

Water Water
1.0 (a) SDBS-C16mimCl
χ C16 mimCl=0.00
1.0 (b) SDBS-C12mimCl
χ C12 mimCl=0.00
χ C16 mimCl=0.20 χ C12 mimCl=0.20
0.8 0.8
χ C16 mimCl=0.35 χ C12 mimCl=0.35
χ C16 mimCl=0.55 χ C12 mimCl=0.65
0.6 0.6
G (τ)

G (τ)

χ C16 mimCl=0.65 χ C12 mimCl=0.80


χ C16 mimCl=0.80 χ C12 mimCl=1.00
0.4 0.4
χ C16 mimCl=1.00

0.2 0.2

0.0 0.0
1
10 10 2
10 3
10 4
105
10 6
101 102 103 104 105 106
Time (μs) Time (μs)

Fig. 7. FCS traces of R6G in water, SDBS-C16mimCl and SDBS-C12mimCl solution mixtures at different mole fractions.
772 R. Dutta et al. / Journal of Colloid and Interface Science 490 (2017) 762–773

indicates that it is located in the hydrophobic region of micelles or [5] M. Tian, Y. Fan, G. Ji, Y. Wang, Spontaneous aggregate transition in mixtures of
a cationic gemini surfactant with a double-chain cationic surfactant, Langmuir
vesicles. But due to the cationic nature of R6G, hydrophobicity and
28 (2012) 12005–12014.
surface charge of vesicle bilayer influence the location of R6G [6] T. Lu, Y. Lan, C. Liu, J. Huang, Y. Wang, Surface properties, aggregation behavior
molecules during micelle-vesicle-micelle transition. Finally, the and micellization thermodynamics of a class of gemini surfactants with ethyl
translational diffusion motion of R6G in SDBS-C16mimCl and ammonium headgroups, J. Colloid Interface Sci. 377 (2012) 222–230.
[7] M. Benrraou, B.L. Bales, R. Zana, Effect of the nature of the counterion on the
SDBS-C12mimCl aggregates at different vC n mimCl values has been properties of anionic surfactants. 1. Cmc, ionization degree at the cmc and
monitored using FCS study to confirm the alteration of location aggregation number of micelles of sodium, cesium, tetramethylammonium,
tetraethylammonium, tetrapropylammonium, and tetrabutylammonium
during micelle-vesicle-micelle transition and it supports the previ-
dodecyl sulfates, J. Phys. Chem. B 107 (2003) 13432–13440.
ously obtained results. [8] M. Tian, L. Zhu, D. Yu, Y. Wang, S. Sun, Y. Wang, Aggregate transitions in
Though there is a significant amount of work concerning the mixtures of anionic sulfonate gemini surfactant with cationic ammonium
micelle to vesicle transition and the study of translational diffusion single-chain surfactant, J. Phys. Chem. B 117 (2013) 433–440.
[9] H. Yin, S. Lei, S. Zhu, J. Huang, J. Ye, Micelle-to-vesicle transition induced by
of hydrophilic as well as hydrophobic probe inside the aggregates organic additives in catanionic surfactant systems, Chem. Eur. J. 12 (2006)
[54,58], here we have mainly focused on the effect of chain length 2825–2835.
of imidazolium-based surfactants in the formation of micellar or [10] L. Jiang, K. Wang, M. Deng, Y. Wang, J. Huang, Bile salt-induced vesicle-to-
micelle transition in catanionic surfactant systems: steric and electrostatic
vesicular aggregates. It is interesting to note that at vC n mimCl = 0.80 interactions, Langmuir 24 (2008) 4600–4606.
SDBS-C16mimCl solution mixture forms mixed micelles, whereas [11] L. Ferrer-Tasies, E. Moreno-Calvo, M. Cano-Sarabia, M. Aguilella-Arzo, A.
SDBS-C12mimCl solution mixture mainly exists as vesicles. As a Angelova, S. Lesieur, S. Ricart, J. Faraudo, N. Ventosa, J. Veciana, Quatsomes:
vesicles formed by self-assembly of sterols and quaternary ammonium
result, the location of R6G alters in these two aggregates which surfactants, Langmuir 29 (2013) 6519–6528.
are exhibited in the anisotropy and translational diffusion mea- [12] H. Yin, J. Huang, Y. Lin, Y. Zhang, S. Qiu, J. Ye, Heating-induced micelle to
surements. Due to greater surface activity and directional polariz- vesicle transition in the cationic-anionic surfactant systems: comprehensive
study and understanding, J. Phys. Chem. B 109 (2005) 4104–4110.
ability, imidazolium-based surfactants can form highly stable and [13] J.H. Fendler, Surfactant vesicles as membrane mimetic agents:
ordered supramolecular assemblies which can be used as reactor characterization and utilization, Acc. Chem. Res. 13 (1980) 7–13.
for nanoparticle synthesis. This type of composition dependent [14] I.I. Yaacob, A.C. Nunes, A. Bose, D.O. Shah, Synthesis and characterisation of
magnetic nanoparticles in spontaneously generated vesicles, J. Colloid
micelle to vesicle aggregate transition has potential application Interface Sci. 168 (1994) 289–301.
in drug delivery [14] and the vesicles can be used as membrane [15] E.J. Danoff, X. Wang, S.H. Tung, N.A. Sinkov, A.M. Kemme, S.R. Raghavan, D.S.
mimetic agent [13]. On the basis of the above experiment, as a English, Surfactant vesicles for high-efficiency capture and separation of
charged organic solutes, Langmuir 23 (2007) 8965–8971.
future plan we can use a long chain containing charged fluorophore [16] M. Dolder, A. Engel, M. Zulauff, The micelle to vesicle transition of lipids and
along with an oppositely charged surfactant to prepare different detergents in the presence of a membrane protein: towards a rationale for 2D
dye-surfactant assemblies [60]. crystallization, FEBS Lett. 382 (1996) 203–208.
[17] G.P. Kumar, P. Rajeshwarrao, Nonionic surfactant vesicular systems for
effective drug delivery—an overview, Acta Pharm. Sin. B 1 (2011) 208–219.
[18] J.B. Engberts, J. Kevelam, Formation and stability of micelles and vesicles, Curr.
Acknowledgments Opin. Colloid Interface Sci. 1 (1996) 779–789.
[19] S. Ghosh, C. Ghatak, C. Banerjee, S. Mandal, J. Kuchlyan, N. Sarkar, Spontaneous
N.S. gratefully acknowledges SERB, Department of Science and transition of micelle – vesicle – micelle in a mixture of cationic surfactant and
anionic surfactant-like ionic liquid: a pure nonlipid small unilamellar vesicular
Technology (DST), Council of Scientific and Industrial Research
template used for solvent and rotational relaxation study, Langmuir 29 (2013)
(CSIR), Government of India for providing generous research grant. 10066–10076.
R.D., S.G., P.B and S.K. are thankful to CSIR for providing their [20] Y.I. González, H. Nakanishi, M. Stjerndahl, E.W. Kaler, Influence of pH on the
research fellowships. micelle-to-vesicle transition in aqueous mixtures of sodium dodecyl
benzenesulfonate with histidine, J. Phys. Chem. B 109 (2005) 11675–11682.
[21] M. Kepczynski, J. Bednar, D. Kuzmicz, P. Wydro, M. Nowakowska, Spontaneous
formation of densely stacked multilamellar vesicles in dioctadecyldimethy-
Appendix A. Supplementary material lammonium bromide/oleosiloxane mixtures, Langmuir 26 (2010) 1551–1556.
[22] K.S. Rao, P.S. Gehlot, T.J. Trivedi, A. Kumar, Self-assembly of new surface active
ionic liquids based on Aerosol-OT in aqueous media, J. Colloid Interface Sci.
Diffusion co-efficient (Dt) of R6G in water and SDBS-CnmimCl 428 (2014) 267–275.
aggregates, variation of turbidity in aqueous SDBS-CnmimCl solu- [23] H. Wang, L. Zhang, J. Wang, Z. Li, S. Zhang, The first evidence for unilamellar
tions at different vC n mimCl , intensity size distribution of SDBS- vesicle formation of ionic liquids in aqueous solutions, Chem. Commun. 49
(2013) 5222–5224.
CnmimCl aggregates, variation of zeta potential (f) of SDBS solution [24] J. Yuan, X. Bai, M. Zhao, L. Zheng, C12mimBr ionic liquid/SDS vesicle formation
with addition of CnmimCl, visual observations of SDBS-CnmimCl and use as template for the synthesis of hollow silica spheres, Langmuir 26
solution mixtures, intensity-size distribution profiles of SDBS- (2010) 11726–11731.
[25] M. Zhao, J. Yuan, L. Zheng, Spontaneous formation of vesicles by N-dodecyl-N-
CnmimCl aggregates ,normalized absorption and emission spectra methylpyrrolidinium bromide (C12MPB) ionic liquid and sodium dodecyl
of R6G and C153 have been provided in the supporting informa- sulfate (SDS) in aqueous solution, Colloids Surf. A 407 (2012) 116–120.
tion. Supplementary data associated with this article can be found, [26] S. Mandal, J. Kuchlyan, D. Banik, S. Ghosh, C. Banerjee, V. Khorwal, N. Sarkar,
Ultrafast FRET to study spontaneous micelle-to-vesicle transitions in an
in the online version, at http://dx.doi.org/10.1016/j.jcis.2016.12. aqueous mixed surface-active ionic-liquid system, ChemPhysChem 15 (2014)
009. 3544–3553.
[27] K.S. Rao, S. So, A. Kumar, Vesicles and reverse vesicles of an ionic liquid in ionic
liquids, Chem. Commun. 49 (2013) 8111–8113.
References [28] T. Welton, Room-temperature ionic liquids. Solvents for synthesis and
catalysis, Chem. Rev. 99 (1999) 2071–2083.
[29] B. Tang, W. Bi, M. Tian, K.H. Row, Application of ionic liquid for extraction and
[1] E.W. Kaler, A.K. Murthy, B.E. Rodriguez, J.A.N. Zasadzinski, Spontaneous vesicle
separation of bioactive compounds from plants, J. Chromatogr. B 904 (2012)
formation in aqueous mixtures of single-tailed surfactants, Science 245 (1989)
1–21.
1371–1374.
[30] F.V. Rantwijk, R.A. Sheldon, Biocatalysis in ionic liquids, Chem. Rev. 107 (2007)
[2] L.M. Bergström, S. Skoglund, K. Edwards, J. Eriksson, I. Grillo, Spontaneous
2757–2785.
transformations between surfactant bilayers of different topologies observed
[31] J.P. Hallett, T. Welton, Room-temperature ionic liquids: solvents for synthesis
in mixtures of sodium octyl sulfate and hexadecyltrimethylammonium
and catalysis. 2, Chem. Rev. 111 (2011) 3508–3576.
bromide, Langmuir 30 (2014) 3928–3938.
[32] S. Dey, A. Adhikari, D.K. Das, D.K. Sasmal, K. Bhattacharyya, Femtosecond
[3] E.W. Kaler, K.L. Herrington, A.K. Murthy, J.A.N. Zasadzinski, Phase behavior and
solvation dynamics in a micron-sized aggregate of an ionic liquid and P123
structures of mixtures of anionic and cationic surfactants, J. Phys. Chem. 96
triblock copolymer, J. Phys. Chem. B 113 (2009) 959–965.
(1992) 6698–6707.
[33] V.G. Rao, C. Ghatak, S. Ghosh, S. Mandal, N. Sarkar, The chameleon-like nature
[4] C. Tondre, C. Caillet, Properties of the amphiphilic films in mixed cationic/
of zwitterionic micelles: the effect of ionic liquid addition on the properties of
anionic vesicles: a comprehensive view from a literature analysis, Adv. Colloid
aqueous sulfobetaine micelles, ChemPhysChem 13 (2012) 1893–1901.
Interface Sci. 93 (2001) 115–134.
R. Dutta et al. / Journal of Colloid and Interface Science 490 (2017) 762–773 773

[34] V.G. Rao, C. Ghatak, S. Ghosh, R. Pramanik, S. Sarkar, S. Mandal, N. Sarkar, Ionic [50] K.S. Mali, G.B. Dutt, T. Mukherjee, Rotational diffusion of organic solutes in
liquid-induced changes in properties of aqueous cetyltrimethylammonium surfactant-block copolymer micelles: role of electrostatic interactions and
bromide: a comparative study of two protic ionic liquids with different anions, micellar hydration, J. Phys. Chem. B 111 (2007) 5878–5884.
J. Phys. Chem. B 115 (2011) 3828–3837. [51] S. Ghosh, A. Roy, D. Banik, N. Kundu, J. Kuchlyan, A. Dhir, N. Sarkar, How does
[35] L. Zheng, C. Guo, J. Wang, X. Liang, S. Chen, J. Ma, B. Yang, Y. Jiang, H. Liu, Effect the surface charge of ionic surfactant and cholesterol forming vesicles control
of ionic liquids on the aggregation behavior of PEO-PPO-PEO block copolymers rotational and translational motion of rhodamine 6G perchlorate (R6G ClO4)?,
in aqueous solution, J. Phys. Chem. B 111 (2007) 1327–1333. Langmuir 31 (2015) 2310–2320
[36] R. Rai, G.A. Baker, K. Behera, P. Mohanty, N.D. Kurur, S. Pandey, Ionic liquid- [52] W. Becker, The bh TCSPC Handbook, fourth ed., Becker & Hickl GmbH, 2010.
induced unprecedented size enhancement of aggregates within aqueous [53] N. Pal, S.D. Verma, M.K. Singh, S. Sen, Fluorescence correlation spectroscopy:
sodium dodecylbenzene sulfonate, Langmuir 26 (2010) 17821–17826. an efficient tool for measuring size, size-distribution and polydispersity of
[37] K. Behera, H. Om, S. Pandey, Modifying properties of aqueous microemulsion droplets in solution, Anal. Chem. 83 (2011) 7736–7744.
cetyltrimethylammonium bromide with external additives: ionic liquid 1- [54] S. Ghosh, A. Adhikari, S. Sen Mojumdar, K. Bhattacharyya, A fluorescence
hexyl-3-methylimidazolium bromide versus cosurfactant n- correlation spectroscopy study of the diffusion of an organic dye in the gel
hexyltrimethylammonium bromide, J. Phys. Chem. B 113 (2009) 786–793. phase and fluid phase of a single lipid vesicle, J. Phys. Chem. B 114 (2010)
[38] K. Behera, S. Pandey, Interaction between ionic liquid and zwitterionic 5736–5741.
surfactant: a comparative study of two ionic liquids with different anions, J. [55] U. Haupts, S. Maiti, P. Schwille, W.W. Webb, Dynamics of fluorescence
Colloid Interface Sci. 331 (2009) 196–205. fluctuations in green fluorescent protein observed by fluorescence correlation
[39] O.A.E. Seoud, P.A.R. Pires, T. Abdel-Moghny, E.L. Bastos, Synthesis and micellar spectroscopy, Proc. Natl. Acad. Sci. U.S.A. 95 (1998) 13573–13578.
properties of surface-active ionic liquids: 1-alkyl-3-methylimidazolium [56] D.K. Sasmal, T. Mondal, S. Sen Mojumdar, A. Choudhury, R. Banerjee, K.
chlorides, J. Colloid Interface Sci. 313 (2007) 296–304. Bhattacharyya, An FCS study of unfolding and refolding of CPM-labeled human
[40] P.D. Galgano, O.A.E. Seoud, Micellar properties of surface active ionic liquids: a serum albumin: role of ionic liquid, J. Phys. Chem. B 115 (2011) 13075–13083.
comparison of 1-hexadecyl-3-methylimidazolium chloride with structurally [57] T. Torres, M. Levitus, Conformational dynamics: a new FCS-FRET approach, J.
related cationic surfactants, J. Colloid Interface Sci. 345 (2010) 1–11. Phys. Chem. B 111 (2007) 7392–7400.
[41] M. Blesic, M.H. Marques, N.V. Plechkova, K.R. Seddon, L.P.N. Rebelo, A. Lopes, [58] S. Dey, U. Mandal, S. Sen Mojumdar, A.K. Mandal, K. Bhattacharyya, Diffusion
Self-aggregation of ionic liquids: micelle formation in aqueous solution, Green of organic dyes in immobilized and free catanionic vesicles, J. Phys. Chem. B
Chem. 9 (2007) 481–490. 114 (2010) 15506–15511.
[42] O. Nacham, A. Martín-Pérez, D.J. Steyer, M.J. Trujillo-Rodríguez, J.L. Anderson, [59] X. Li, Y. Yang, J. Eastoe, J. Dong, Rich self-assembly behavior from a simple
V. Pino, A.M. Afonso, Interfacial and aggregation behavior of dicationic and amphiphile, ChemPhysChem 11 (2010) 3074–3077.
tricationic ionic liquid-based surfactants in aqueous solution, Colloids Surf. A [60] J. Shen, X. Xin, T. Liu, S. Wang, Y. Yang, X. Luan, G. Xu, S. Yuan, Ionic self-
469 (2015) 224–234. assembly of giant vesicle as smart microcarrier and microreactor, Langmuir 32
[43] P. Brown, C.P. Butts, J. Eastoe, D. Fermin, I. Grillo, H.C. Lee, D. Parker, D. Plana, R. (2016) 9548–9556.
M. Richardson, Anionic surfactant ionic liquids with 1-butyl-3-methyl- [61] J.G. Ma, B.J. Boyd, C.J. Drummond, Positional isomers of linear sodium dodecyl
imidazolium cations: characterization and application, Langmuir 28 (2012) benzene sulfonate: solubility, self-assembly, and air/water interfacial activity,
2502–2509. Langmuir 22 (2006) 8646–8654.
[44] Q. Feng, H. Wang, S. Zhang, J. Wang, Aggregation behavior of 1-dodecyl-3- [62] X. He, W. Shinoda, R. DeVane, K.L. Anderson, M.L. Klein, Paramaterization of a
methylimidazolium bromide ionic liquid in non-aqueous solvents, Colloids coarse-grained model for linear alkylbenzene sulfonate surfactants and
Surf. A 367 (2010) 7–11. molecular dynamics studies of their self-assembly in aqueous solution,
[45] B. Dong, N. Li, L. Zheng, L. Yu, T. Inoue, Surface adsorption and micelle Chem. Phys. Lett. 487 (2010) 71–76.
formation of surface active ionic liquids in aqueous solution, Langmuir 23 [63] L. He, A.S. Malcolm, M. Dimitrijev, S.A. Onaizi, H.H. Shen, S.A. Holt, A.F. Dexter,
(2007) 4178–4182. R.K. Thomas, A.P.J. Middelberg, Cooperative tuneable interactions between a
[46] Y. Gu, L. Shi, X. Cheng, F. Lu, L. Zheng, Aggregation behavior of 1-dodecyl-3- designed peptide biosurfactant and positional isomers of SDOBS at the air-
methylimidazolium bromide in aqueous solution: effect of ionic liquids with water interface, Langmuir 25 (2009) 4021–4026.
aromatic anions, Langmuir 29 (2013) 6213–6220. [64] B.F.B. Silva, E.F. Marques, U. Olsson, R. Pons, Headgroup effects on the unusual
[47] S. Mandal, J. Kuchlyan, S. Ghosh, C. Banerjee, N. Kundu, D. Banik, N. Sarkar, lamellar-lamellar coexistence and vesicle-to-micelle transition of salt-free
Vesicles formed in aqueous mixtures of cholesterol and imidazolium surface catanionic amphiphiles, Langmuir 26 (2010) 3058–3066.
active ionic liquid: a comparison with common cationic surfactant by water [65] K.S. Rao, T. Singh, A. Kumar, Aqueous-mixed ionic liquid system: phase
dynamics, J. Phys. Chem. B 118 (2014) 5913–5923. transitions and synthesis of gold nanocrystals, Langmuir 27 (2011) 9261–
[48] S. Chabba, S. Kumar, V.K. Aswal, T.S. Kang, R.K. Mahajan, Interfacial and 9269.
aggregation behavior of aqueous mixtures of imidazolium based surface active [66] K.J. Tangso, S. Lindberg, P.G. Hartley, R. Knott, P. Spicer, B.J. Boyd, Formation of
ionic liquids and anionic surfactant sodium dodecylbenzenesulfonate, Colloids liquid-crystalline structures in the bile salt-chitosan system and triggered
Surf. A 472 (2015) 9–20. release from lamellar phase bile salt-chitosan capsules, ACS Appl. Mater.
[49] G.B. Dutt, Do ionic and hydrophobic probes sense similar microenvironment in Interfaces 6 (2014) 12363–12371.
Triton X-100 nonionic reverse micelles?, J Chem. Phys. 129 (2008) 014501(1)– [67] Z. Yuan, S. Dong, W. Liu, J. Hao, Transition from vesicle phase to lamellar phase
014501(6). in salt-free catanionic surfactant solution, Langmuir 25 (2009) 8974–8981.

Das könnte Ihnen auch gefallen