You are on page 1of 8

Article

pubs.acs.org/Langmuir

Spontaneous Formation of Vesicles by Sodium 2‑Dodecylnicotinate


in Water
Aparna Roy, Monali Maiti, and Sumita Roy*
Department of Chemistry and Chemical Technology, Vidyasagar University, Paschim Medinipur-721 102, India
*
S Supporting Information

ABSTRACT: The surface activity and aggregation behavior of a


synthesized nicotinic acid based anionic surfactant, sodium 2-
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

dodecylnicotinate, were studied in aqueous solution. The self-


assembly formation was investigated by use of a number of
techniques, including surface tension and conductivity measure-
Downloaded via ESS INFLIBNET PCA 1 on August 21, 2023 at 09:38:27 (UTC).

ments, fluorescence spectroscopy, dynamic light scattering measure-


ment, gel permeation chromatography, and microscopy. The
amphiphile exhibits two breaks in the surface tension vs concentration plot, indicating stepwise aggregate formation and thus
producing two values of the aggregation concentration. Stepwise aggregation of the amphiphile was further confirmed by steady-
state fluorescence spectroscopy using pyrene as a probe molecule, and also the micropolarity of the aggregates was determined.
The rigidity of the microenvironment was estimated by determining steady-state fluorescence anisotropy using 1,6-diphenyl-
1,3,5-hexatriene as a fluorescence probe molecule. The average hydrodynamic radius and size distribution of the aggregate
suggest formation of larger aggregates in aqueous solution. The formation of vesicles in water was established by conductivity
measurement and a dye entrapment experiment. The entrapment of a small solute and the release capability have also been
examined to demonstrate these bilayers form enclosed vesicles. Transmission electron micrographs revealed the existence of
closed vesicles and closed tubules in aqueous solution. Therefore, for the first time, it has been observed that this simple single-
chain nicotinic acid based amphiphile spontaneously assembles to vesicles in aqueous solution.

■ INTRODUCTION
Amphiphilic molecules in aqueous solution self-assemble to
surfactant derived from nicotinic acid having a simple aliphatic
chain as a hydrophobic moiety. We became interested in the
form a variety of microstructures above a certain concentration nicotinic acid derived surfactant as it is well-known to all that
and have been a subject of intensive research during the past pyridine is an important model molecule in biology as many
four decades.1−3 Among all the self-assemblies, vesicular drugs and metabolites are pyridine based.12 Pyridine and
structures are most important because lipid bilayer vesicles or pyridinium salt both are used as ligands, and pyridine itself is a
liposomes have been widely studied for drug delivery and useful solvent in synthetic chemistry.13 “Niacin”, which is
material synthesis and as model biomembranes.4,5 Vesicles nothing but simple nicotinic acid, is widely known as vitamin
store, transport, or digest cellular products and waste materials. B3. N-Methylnicotinic acid (trigoneline) plays an important
They are involved in metabolism, buoyancy control,6 and biological role in controlling the G2 factor mechanism.14,15
enzyme storage. Because of their versatility, there are many Kalyanasundaram et al. have reported the critical micellar
reports on vesicle formation from natural amphiphiles (mainly concentration and critical micellar temperature for various 1-
phospholipids) and synthetic surfactants in the literature.7,8 alkyl-3-carbamoylpyridinium halide surfactants (N-alkylnicoti-
There are also instances of formation of vesicular aggregates by namides) in aqueous solution.16 The authors also studied the
amphiphiles having a large headgroup size and hydrogen micellazation properties of N-alkylnicotinic acid surfactants in
bonding. Grinberg and co-workers9 have reported that a triple- aqueous solution.17 Chevalier et al. have investigated the
headed derivative of vernonia oil assembles to vesicles in the micellar properties and mesophase structure of two isomers of
presence of cholesterol. Cantu et al.10 have shown that the N-alkylpyridiniocarboxylates with their carboxylate group at
vesicles made from ganglioside GM3 are stabilized in the position 3 or 4 on the pyridinium ring.18 Jiang and co-workers
presence of the larger headgroup ganglioside amphiphile GM1. have synthesized and studied the surface-active properties of a
A new class of nucleoamphiphiles assemble in water to form a novel class of gemini pyridinium surfactants having a 4-
nano- to micrometer left-handed helix with stacked bilayer-type methylene spacer group, and also they investigated the
aggregates induced by simply complexing chiral GMP or AMP interactions of these amphiphiles with polyacrylamide
(guanosine 5′-monophosphate or adenosine 5′-monophos- (PAM).19 Therefore, in this work, we have attempted to
phate) with a nonchiral monocationic amphiphile having two
hydrophobic chains of 12 and 14 carbons (dialkyldimethy- Received: December 13, 2011
lammonium bromide).11 In spite of all the reports, there is not Revised: August 1, 2012
a single paper on the formation of a vesicular structure by a Published: August 8, 2012

© 2012 American Chemical Society 12696 dx.doi.org/10.1021/la302484x | Langmuir 2012, 28, 12696−12703
Langmuir Article

synthesize and characterize the vesicles formed by a simple Table 1. Surface Properties of SDDNA
single-chain anionic surfactant derived from nicotinic acid in
property value property value
aqueous solution.


CAC (mM) 0.030, 0.0240,a 0.320, a0 (nm2 1.070
0.340a molecule−1)
RESULTS AND DISCUSSION
γCAC (mN m−1) 43, 36.600 ΔGa° (kJ mol−1) −27.540
We have performed the synthesis of a new nicotinic acid based ΠCAC (mN m−1) 33.900 ΔGad° (kJ mol−1) −49.550
surfactant named sodium 2-dodecylnicotinate, abbreviated as α 0.641 CAC/C20 35.660
SDDNA (Figure 1) according to the scheme given in the Γ2 × 106 (mol 1.540 pC20 5.0450
Supporting Information. The molecular structure of the m−2)
compound was identified by FT-IR, 1H NMR, and LC−MS a
Data are obtained from fluorescence measurements.
(see the Supporting Information).

the ratio CAC/C20, which is a measure of the tendency of the


surfactant to adsorb at the air−water interface relative to the
formation of the aggregates, is greater than those of
conventional hydrocarbon monomeric surfactants. The max-
imum surface excess concentration (Γ2) and minimum surface
Figure 1. Chemical structure of SDDNA. area per surfactant molecule at the air−water interface (a0)
were also estimated from the surface tension plot using the
Surface Tension Study. It is well-known that, among all Gibbs adsorption equation:33
the available methods, the surface tension method is the most Γ2 = −(1/2.303nRT )(dγ /(d log C)) (1)
useful and accurate method for determining the critical
aggregation concentration. Therefore, this method was used 18
a0 = 10 /(Γ2NA ) (2)
to determine the critical aggregation concentration (CAC) of
SDDNA. A plot of surface tension (γ) versus log- where dγ/(d log C) is the maximum slope, NA is Avogadro’s
(concentration, C) in aqueous solution (pH 6.70) of the number, T is the absolute temperature, and n is 2 for dilute
amphiphile is shown in Figure 2. solutions of monovalent ionic surfactants.34
For SDDNA, the a0 value thus obtained is in the range 1.0
nm2 ≤ a0 ≤ 1.1 nm2, which is indicative of the formation of
bilayer aggregates.25 The standard free energy of aggregation
per mole of surfactant was determined by using the equation22
ΔGa° = RT(1 + β) ln CAC (3)
where β = 1 − α, in which α is the degree of micellar ionization
which can be determined from the ratio of the slope of the
conductivity versus concentration lines above and below the
CAC (for the plot see the Supporting Information).
ΔGad° was evaluated from the equation35
ΔGad° = ΔGa° − πCAC/Γ2 (4)
Figure 2. γ vs log C of SDDNA in aqueous solution. where πCAC is the surface pressure (= γ0 − γCAC, γ0 being the
surface tension of pure water). The ΔGa° and ΔGad° values are
It is clear from the figure that there are two break points of also listed in Table 1. The large negative value of ΔGa° suggests
the plot γ vs log C corresponding to two CAC values. As the spontaneity of the aggregate formation in water.
number of points corresponding to the first break point is Microenvironment Study. To investigate the micro-
fewer, we thought this might be due to the presence of some environment of the self-assembly, fluorescence studies were
surface-active impurity in the compound. However, no performed using pyrene and 1,6-diphenyl-1,3,5-hexatriene
minimum was observed around the second break point of the (DPH) as an extrinsic fluorescence probe because both the
surface tension plot, which ruled out the presence of any probe molecules bind preferentially to the hydrophobic region
surface-active impurity in the amphiphile.20 Furthermore, the of the self-assemblies. We know that the ratio of the intensities
purity of the sample was confirmed by 1H NMR and LC−MS. corresponding to the first and third vibronic bands (I1/I3) is
This type of two break points due to post micellar aggregation sensitive to solvent polarity.36 Its value is maximum in water
has also been reported for other surfactants,20−27 mixed and decreases with a decrease in solvent polarity. Therefore, it
surfactants,28 and phenyl ring bearing cationic surfactants.29 has been widely used as a micropolarity probe for self-
The CAC values and all the interfacial properties are given in assemblies. The I1/I3 ratio was measured in the presence of
Table 1. various surfactant concentrations. A two-step change of the plot
The CAC and surface tension at the CAC (γCAC) values are I1/I3 vs log(concentration) of the surfactant was obtained
lower than those of the corresponding fatty acid soaps, which (Figure 3), which is consistent with the surface tension plot.
suggests that the amphiphile is a better surfactant than Two CACs and the polarity ratio corresponding to two
conventional soaps.30,31 The good surface activity is also CACs thus obtained are summarized in Table 2. The I1/I3 ratio
reflected by a larger pC20 (>3)32(where pC20 = negative above the first CAC is 1.77 (inset of Figure 3), which is very
logarithm of the surfactant concentration required to reduce close to that of water (1.86), indicating that the probe molecule
the surface tension of water by 20 units). On the other hand, is partially exposed to the bulk solvent, which is only possible if
12697 dx.doi.org/10.1021/la302484x | Langmuir 2012, 28, 12696−12703
Langmuir Article

the formation of bilayer aggregates.21,40,41 The high anisotropy


value (r = 0.108) and high micropolarity index (I1/I3 = 1.77)
just above the first CAC indicates formation of a flat bilayer
structure above the first break point. The anisotropy value
gradually increases and remains constant up to 1.25 mM
(greater than the second CAC). On the other hand, the
micropolarity index gradually decreases to a very low value (I1/
I3 = 1.23). These two phenomena suggest that the flat bilayer
aggregates transformed into closed vesicles and tubules. After
an amphiphile concentration of 1.5 mM the fluorescence
anisotropy value of SDDNA decreases instantaneously,
indicating closed structures are malformed to open tubules
Figure 3. Polarity ratio I1/I3 vs log C. Inset: I1/I3 vs log C for the first and/or cylindrical micelles.
break point. Dynamic Light Scattering. The dynamic light scattering
method was employed to determine the hydrodynamic radius
Table 2. Self-Assembly Properties of SDDNA of the aggregates formed by SDDNA in aqueous solution
because it is an excellent technique for distinguishing bilayer
property value property value
structures such as vesicles, lamellae, and tubules from micelles
I1/I3 1.77, 1.23 Nagg × 10−4 22.92 on the basis of the size of the aggregates. The z-average
Rh (nm) 98.84 D × 1012 (m2 s−1) 2.24 hydrodynamic radius (Rh) of a 1 mM aqueous solution was
found to be 98.84 nm, which suggests the existence of larger
the probe molecule is solubilized in a nonspherical aggregate aggregates in solution (see the Supporting Information for the
such as a flat bilayer as indicated by the a0 value. The polarity size distribution). The apparent diffusion coefficient (Dapp) of
ratio above the second CAC is very low, which suggests that the the amphiphile was calculated using the Stokes−Einstein
probe molecule is solubilized in spherical aggregates. A similar equation:
type of result has also been reported by Dey et al.21,40 This is
due to the fact that pyrene molecules are nonpolar in nature, Dapp = kBT /6πηR h (5)
are generally soluble in the hydrocarbon layer, and encounter
only a few or no water molecules around them in the case of Rh and Dapp values thus obtained are listed in Table 2. The Dapp
spherical aggregates. Also the lower micropolarity value of the value is much smaller than that of normal spherical micelles
secondary aggregates compared to that of primary aggregates (∼10−10 m2 s−1).26 The mean aggregation number (Nagg) of the
indicates more ordering at the interface, which may be due to aggregates (assuming a bilayer vesicle) was also estimated by
the formation of a vesicular structure above the second CAC. use of the equation42
To further shed light on the structural change at
concentrations above the two break points, we measured the Nagg = 8πR h 2/a0 (6)
steady-state fluorescence anisotropy (r) using DPH as a
The large Rh and Nagg values with a very small (1/100 times)
fluorescence probe because it is a well-known membrane
diffusion coefficient compared to that of a normal spherical
fluidity probe and different scientists have used it to study many
micelle suggest that the aggregates formed by SDDNA are very
lipid bilayer membranes.37−39 r is an index of the equivalent
large.
microviscosity (microfluidity) of the vesicle core. Therefore, r
Conductivity Study. To demonstrate the transformation of
was measured at different concentrations above and below the
flat bilayer aggregates to a closed vesicle, the conductivity of a 1
second CAC. It has been observed that the anisotropy value
mM KCl solution was measured in the presence of a 1 mM
increases from 0.06 mM (above the first break point) and
SDDNA solution. It was observed that the sum of the
remains almost unchanged up to a concentration of 1.5 mM (r
conductivity values of a 1 mM KCl solution (160 μS cm−1) and
= 0.159). After this, the anisotropy value decreases and
a 1 mM SDDNA solution (77 μS cm−1) is 37 μS cm−1 greater
becomes almost constant at higher concentration. The r value
than that of 1 mM KCl containing a 1 mM (greater than the
for different concentrations of SDDNA is given in Table 3.
second CAC) SDDNA solution (200 μS cm−1). On the other
We are unable to measure the anisotropy below the first
hand, there was no decrease of conductivity of the 1 mM KCl
CAC because the intensity of DPH in the solutions was very
solution in the presence of 0.15 mM (less than the second
low. The value of anisotropy is greater than that of the lecithin
CAC) SDDNA. This study clearly indicates the existence of
liposomes (r ≈ 0.098) but less than that of sphingomyelin
vesicles after the second break point. A similar type of decrease
liposomes (r ≈ 0.247).37 The large value of r is due to tight
of the conductivity value of the KCl solution due to formation
packing of the hydrocarbon chains and perhaps is indicative of
of vesicles has been reported by Hoffmann43 and Dey et al.21
Actually the vesicles are able to entrap a part of the solvent and
Table 3. Anisotropy Values of SDDNA at Different the salt. As a result they prevent the entrapped charge carriers
Concentrations (K+ and Cl− ions) from contributing to the conductivity of the
solution, and thus, the conductivity of the vesicular solution
concn (mM) r concn (mM) r concn (mM) r
becomes lower than that of the conductivity of the salt solution.
0.06 0.108 1 0.163 2.25 0.123 The variation of the conductivity change (Δκ) with surfactant
0.08 0.137 1.25 0.162 2.50 0.117 concentration is shown in Figure 4.
0.20 0.156 1.50 0.159 2.75 0.108 The value of Δκ at first increases with the surfactant
0.40 0.160 1.75 0.144 3 0.107 concentration and then decreases. This is because of the fact
0.70 0.162 2 0.132 that the population of vesicles increases with the surfactant
12698 dx.doi.org/10.1021/la302484x | Langmuir 2012, 28, 12696−12703
Langmuir Article

Figure 4. Δκ vs surfactant concentration. Figure 5. Gel filtration profile of the separation of the dye entrapped
(small peaks) in a vesicle of SDDNA from the corresponding free dye
(methylene blue).
concentration and more and more K+ and Cl− ions are being
entrapped inside the aqueous core. As a result the Δκ value
increases. The decrease in the Δκ value at higher surfactant widely used due to their simplicity and convenience.50 A
concentrations is most probably due to the transformation of number of probes such as pyranine,51 calcein,52 carboxyfluor-
closed vesicles to open tubules and/or cylindrical micelles, escein,53 riboflavin,46,47 etc. have been used to estimate
which causes release of the entrapped water with salt. membrane permeability. Here we used riboflavin as a probe
Entrapment of Methylene Blue in Vesicles. Further- molecule because it is a neutral water-soluble molecule which is
more, to confirm the formation of vesicles from this newly strongly fluorescent in the neutral form and becomes
synthesized amphiphile, we employed dye entrapment studies nonfluorescent upon deprotonation (pKa of riboflavin ∼10.2).
to check whether the aggregates contain closed inner aqueous It can also withstand a low ionic strength. This probe has also
compartments, because micelle-forming amphiphiles in water been successfully used by different scientists54−56 for the
have no capacity to entrap hydrophilic water-soluble molecules determination of transmembrane pH gradients with unrelated
(dye). Walde and Namani44 have used arsenazo III as a water- vesicle-forming surfactants of diverse surface charges. The
soluble dye to establish the aqueous core of decanoic acid/ fluorescence intensity due to the membrane-bound and
dodecylbezenesulfonate vesicles. A similar type of experiment entrapped riboflavin was measured at 514 nm (λex = 374
was also performed by Dey et al.40 to confirm the inner nm). This includes the dye molecules that were adsorbed on
aqueous compartments of vesicles formed by N-[4-(n- the outer membrane surfaces and the dye molecules that were
dodecyloxy)benzoyl]-L-aminoacidates in aqueous solution entrapped within the inner water pools. When the pH of the
using methyl orange as a dye molecule. Gokel et al.45 showed vesicle dispersion was increased from 6.8 to 10.2, the
that the aggregates of steroidal azacrown derivatives in aqueous fluorescence intensity of the riboflavin decreased initially
media possessed entrapment capacities. Riboflavin was used as “instantaneously” to about 60 ± 5% of the original value at
a dye molecule by Bhattacharya and co-workers to establish the pH 6.8. This characteristic of immediate capacity (40 ± 5%)
vesicular structure formed by cationic mixed-chain surfactants46 loss in the fluorescence intensity is due to the deprotonation of
in 1995 and hybrid bolaphile/amphiphile ion pairs47 in 1998. riboflavin molecules that are bound at the outer surfaces of the
They have also used methylene blue (MB) as a water-soluble vesicles. This result (decrease of the fluorescence intensity of
dye to establish inner aqueous compartments of some riboflavin due to the conversion of its neutral form to the
synthesized click adducts48 and sugar-based amphiphiles.49 deprotonated form upon pH adjustment from 6.8 to 10.2)
Therefore, to investigate the entrapment capacity of the established that it contained an internal aqueous compartment.
aggregate generated from the amphiphile SDDNA, we chose The residual fluorescence intensity disappeared as a function of
MB as a dye molecule. To achieve our goal, we prepared a 1 time (Figure 6) since instantaneous ionization of the riboflavin
mM stock solution of SDDNA in water containing 0.1 mM molecules occurred at the exposed exovesiculer surfaces as the
MB. The resulting solution (2 mL) was then loaded into a gel pH was increased.
filtration column (preequilabrated Sepharose 4B) and eluted Since the fluorescence intensity disappeared as a function of
with triply distilled water. Translucent vesicular suspensions time, we calculated the half-time of the process (t1/2 ≈ 1.52
eluted right after the void volume. The gel filtration was
continued until free MB was gel filtered. The separate fractions
of each were collected (2 mL each), and the amount of dye
present was evaluated by the spectrophotometric method. The
profile of gel filtration was obtained by plotting the absorbance
at 665 nm of all the fractions obtained from the gel filtration
column vs the elution volume (Figure 5). It was observed that
there was a small initial portion containing vesicles entrapping
approximately 1.79% of the total dye followed by a large peak
which was due to the free, unentrapped dye.
Kinetics of Transmembrane Permeation. To shed light
on whether the bilayer membranes of the vesicles are
permeable, a kinetic study was performed fluorometrically
because of all the methods available to study the membrane Figure 6. OH− permeability profile of riboflavin-entrapped vesicular
permeability, methods utilizing the fluorescence probe are SDDNA.

12699 dx.doi.org/10.1021/la302484x | Langmuir 2012, 28, 12696−12703


Langmuir Article

Figure 7. Negatively stained (with 2% aqueous uranyl acetate solution) TEM micrographs of (A−C) 0.7 mM (inset of (C), vesicle with wall
thickness of 3.18 nm; the bar represents 50 nm), (D) 1 mM, (E) 1.25 mM, (F, G) 1.50 mM (inset of (G), tubule with closed end), and (H, I) 2 mM
(inset of (H), open tubules) SDDNA in water.

min, kperm ≈ 7.58 × 10−3 s−1). Taking the bilayer thickness of


the amphiphile as 3.18 nm (obtained from a TEM micrograph,
Figure 7C) and assuming the time-dependent loss of
fluorescence intensity by exovesicular OH− as a permeation-
limited deprotonation of riboflavin entrapped in the internal
aqueous compartment, we calculated P ≈ 1.53 × 10−9 cm/s for
the permeation constant of OH− toward vesicles of SDDNA at
pH 10.2.
TEM. The transmission electron microscopy technique was
used to visualize the actual morphology of the aggregates in
dilute and concentrated solutions of the amphiphile. TEM
pictures of negatively stained specimens prepared from different
aqueous SDDNA solutions are shown in Figure 7. The
transmission electron micrographs of a 0.7 mM (just above 2
times the second CAC) aqueous solution of SDDNA (Figure
7A−C) clearly exhibit formation of closed vesicles in an
aqueous solution of the amphiphile having an internal diameter
in the range of 30−100 nm. The inner and outer diameters of
the vesicle are 79.92 and 86.28 nm, respectively. Therefore, the
wall thickness of the lamella is about 3.18 nm (inset of Figure
Figure 8. Schematic representation of bilayer formation of SDDNA.
7C).
As the hydrocarbon chain of the amphiphile has a length of At a concentration of 1 mM, the size of the vesicles increases
1.74 nm (obtained from the energy-minimized structure via the to ∼200 nm. The size of the vesicles is enhanced to ∼325 nm at
Hartree−Fock method by Gauss View 3.0), each lamella should a 1.25 mM amphiphile concentration. At a concentration of
have a thickness of about 3.48 nm. However, the length of the 1.50 mM, not only does the size of the vesicle become very
wall thickness is smaller than twice the length of the large (∼750 nm, Figure 7F) but also the vesicles connect to
hydrocarbon chain but much larger than the length of a single each other to form elongated aggregates and closed tubules
hydrophobic tail. This implies that the aggregate has ((Figure 7G). However, the micrographs (Figure 7H,I)
interdigitated hydrocarbon tails in the bilayer vesicles (Figure corresponding to 2.0 mM SDDNA solutions show the existence
8). of open tubules20,40 and cylindrical micelles.57 Therefore, the
12700 dx.doi.org/10.1021/la302484x | Langmuir 2012, 28, 12696−12703
Langmuir Article

stable range of the vesicles is 0.7−1.25 mM. These results are in the saturated aqueous probe solution. The samples were excited at 335
good agreement with the conductivity and steady-state nm, and the emission spectrum was recorded between 350 and 550
fluorescence anisotropy measurements. nm with excitation and emission slit widths both set at 1 nm. Each


spectrum was blank subtracted and was corrected for lamp intensity
variation during measurement. The steady-state fluorescence aniso-
CONCLUSION tropy (r) of DPH was obtained in the same instrument equipped with
In summary, we have synthesized a single-chain nicotinic acid filter polarizers that use the L format configuration. DPH was excited
based surfactant which is a useful surface-active agent and at 350 nm and the fluorescence intensity was measured at 450 nm with
band passes of 2.5 and 5 nm, respectively. The fluorescence anisotropy
spontaneously forms vesicular structures in water. To our
value was calculated by the equation
knowledge, this is the first report of formation of vesicles by a
nicotinic acid based amphiphile. Surface tension measurements r = (I − GI⊥)/(I − 2GI⊥) (1)
at different surfactant concentrations indicate stepwise
aggregate formation and thus produce two values of where I∥ and I⊥ are the fluorescence intensities polarized parallel
aggregation concentration. Stepwise aggregation of the (excitation and emission polarizers at 0°) and perpendicular
amphiphile has been further confirmed by the steady-state (excitation polarizer at 0° and emission polarizer at 90°) to the
fluorescence spectroscopy method. The microenvironments of excitation light. G is the instrumentation correction factor (G = i90−0/
the bilayer self-assemblies are nonpolar in nature and are highly i90−90).
ordered. DLS measurement suggests the existence of larger For anisotropy measurements a 2 mM solution of the probe was
aggregates in aqueous solution. Gel permeation chromatog- prepared in a 20% methanol−water mixture. The final concentration
of the probe was adjusted to 2 μM by addition of the appropriate
raphy and conductivity study indicate formation of closed
amount of stock solution.
vesicles in solution. Kinetic study by the fluorometric method Conductivity Measurements. Conductivity measurements were
suggests the bilayer membrane of the vesicles is permeable. performed with a digital conductivity meter (Systronics, model 304)
TEM pictures revealed the existence of closed bilayer vesicles with a cell constant of 1.05 cm−1. A 1 × 10−3 M stock solution of the
and tubules in aqueous solution. Thus, the amphiphile may surfactant was prepared. The conductivity was measured at different
have potential use in surfactant chemistry and the pharma- concentrations by adding a subsequent volume of stock solution to a
ceutical industry as a drug delivery vehicle. The amphiphile can beaker containing a known volume of water. For each measurement,
also be used as a useful ligand to form complexes. the solution was equilibrated for 5 min to obtain a constant


conductivity value. For the entrapment experiment, a 1 mM KCl
solution was prepared in triply distilled water. For each concentration,
EXPERIMENTAL SECTION a weighed amount of the amphiphile was dissolved in triply distilled
General Procedures. 2,5-Dibromopyridine, pyrene, DPH, uranyl water and the exact same amount of the amphiphile was dissolved in
acetate, and Sepharose 4B were purchased from Aldrich Chemicals, the prepared 1 mM KCl solution. All the solutions (KCl solution,
and 1-bromododecane, riboflavin, and methylene blue were purchased amphiphile in water, and amphiphile in KCl solution) were left for 4 h
from SRL. Pyrene, DPH, and riboflavin were used after recrystalliza- for equilibration before measurement. After equilibration, the
tion from ethanol. Other chemicals were procured locally and used as conductivity of all the solutions was measured in a digital conductivity
they were obtained. All organic solvents were purchased from Merck meter.
and were distilled and dried as required. Aqueous solutions were made Dye Entrapment Studies. For the dye entrapment study, a 1 mM
by triply distilled water. 1H NMR spectra were recorded using a stock solution of the amphiphile was prepared in triply distilled water
Bruker 400 MHz instrument. LC−MS spectra were taken in a Waters containing 0.1 mM methylene blue (λmax = 665 nm), which is a water-
2996 machine. UV−vis absorption spectra were obtained with a soluble dye, and the solution was left for 4 h for spontaneous vesicle
Shimadzu (model 1601) spectrophotometer. The pH measurement formation. A 2 mL volume of the resulting solution was then loaded
was done with a pH meter, model pH LI613 (ELICO), using a glass into a column packed with a pre-equilibrated Sepharose 4B matrix (25
electrode. All measurements were carried out at room temperature cm height and 1.2 cm diameter) and eluted with triply distilled water.
(∼30 °C) unless otherwise mentioned. The fractions containing vesicles were pulled out. The eluent was
Preparation of Aggregates. A weighed amount of amphiphile collected in 2 mL fractions. The absorbance for all the fractions
was dissolved in a certain amount of triply distilled water and shaken containing solutions was determined at 665 nm and plotted against the
by hand to form a homogeneous mixture. The mixture was left for 4 h elution volume.
and afforded approximately spherical and tubular aggregates as Kinetics of Transmembrane Permeation. An aliquot of 1 mL
evidenced from transmission electron microscopy. (pH 6.8) was taken in a fluorescence cuvette, and the time-dependent
Surface Tension Measurement. Surface tension measurement decrease of the fluorescence intensity at 514 nm (λex = 374 nm,
was carried out with a surface tensiometer (manual) using the Du excitation and emission slit widths of 5 and 10 nm, respectively) at
Nuoy ring detachment method. Before each experiment, the ∼30 °C was followed using a Hitachi F-7000 spectrophotometer upon
platinum−iridium ring was thoroughly cleaned with ethanol−HCl adjustment of the pH from 6.8 to 10.2 by the addition of an aliquot of
solution and burned in an oxidizing flame by use of a Bunsen burner. 48 μL of a 0.035 M KOH solution. Since riboflavin fluorescence
The instrument was calibrated, and the accuracy was checked by decreases upon ionization in an alkaline medium (pKa ≈ 10.2), the
measuring the surface tension of triply distilled water. A 1 × 10−3 M time-dependent quenching of the fluorescence intensity at 514 nm due
stock solution of the surfactant was prepared. The surface tension (γ) to riboflavin under an imposed transmembrane pH gradient of 3.4 pH
was measured at different concentrations by adding a subsequent units was taken as a measure of OH− permeation across the bilayer
volume of stock solution to a beaker containing a known volume of vesicles. The rate constant was obtained from the monoexponential
water until the value almost reached saturation. For each time-dependent portion of the loss of the fluorescence intensity. The
concentration, the solution was equilibrated for 15 min, and the rate constant value represents an average of three independent
average of three readings was taken. The CAC was then obtained from experiments.
the break point of the plot of γ versus log C. DLS Studies. The size of the aggregates was determined from DLS
Fluorescence Measurements. The steady-state fluorescence measurement by using a Zetasizer Nano S (ZEN 1600) light scattering
spectra of the pyrene probe (∼2 × 10−7 M) were measured on a apparatus (Malvern Instruments, Westborough, MA) with a He−Ne
Hitachi F-7000 spectrophotometer optical system equipped with a 150 laser (632.8 nm, 4 mW). The Nano S (ZEN 1600) instrument
W Xe lamp. For the pyrene probe, a saturated solution of pyrene was incorporates noninvasive backscatter (NIBS) optics with a detection
prepared in water. A weighed amount of surfactant was then added to angle of 173°. For DLS measurements, an aqueous solution of the

12701 dx.doi.org/10.1021/la302484x | Langmuir 2012, 28, 12696−12703


Langmuir Article

amphiphile was prepared in triply distilled water. The solution was (10) Cantu, L.; Corti, M.; Del Favero, E.; Maurer, N. Spontaneous
filtered directly into the scattering cell through a Whatman syringe vesicle formation: transition between single-and bicomponent system.
filter (sterile and endotoxin free, 0.2 μm). Before measurement, the Colloid Polym. Sci. 1995, 98, 197−200.
scattering cell was rinsed with the filtered solution and the sample (11) Aime, C.; Manet, S.; Satoh, T.; Ihara, H.; Park, K. Y.; Godde, F.;
solution was equilibrated for 5−10 min in the DLS optical system. The Oda, R. Self-assembly of nucleoamphiphiles: investigating nucleosides
data acquisition was carried out for 10 min. effect and the mechanism of micrometric helix formation. Langmuir
Transmission Electron Microscopy Studies. Transmission 2007, 23, 12875−12885.
electron microscopic measurements were performed with a high- (12) Berg, E. R.; Green, D. D.; Moliva, A. D. C.; Bjerke, B. T.; Gealy,
resolution transmission electron microscope (JEOL-JEM 2100, Japan) M. W.; Ulness, D. J. Ion-pair interaction in pyridinium carboxylate
operating at 200 keV. The transmission electron microscopic
solutions. J. Phys. Chem. A 2008, 112, 833−838.
measurements were performed using the aqueous solutions of the
(13) Krygowski, T. M.; Szatylowicz, H.; Zachara, J. E. How H-
amphiphile after equilibration for 4 h. For all the cases the surfactant
solutions were filtered by use of a Whatman syringe filter (sterile and bonding modifies molecular structure and π-electron delocalization in
endotoxin free, 0.2 μm). A drop of the surfactant solution was put on a the ring of pyridine/pyridinium derivatives involved in H-bond
copper grid coated with carbon and allowed to soak for 1−2 min; the complexation. J. Org. Chem. 2005, 70, 8859−8865.
surface solvent on the grid was removed by tapping in a filter paper (14) Pimental, G., Ed. Opportunities in Chemistry; National Academy
followed by staining with 2% aqueous uranyl acetate solution. The of Sciences Report; National Academy of Sciences: Washington, DC,
specimens were then dried in desiccators until measurement. 1987.


(15) Dictionary of Organic Compounds, 5th ed.; Chapman and Hall:
ASSOCIATED CONTENT London, 1982; Vol. 4.
(16) Kalyanasundaram, K.; Colassis, T.; Humphry-Baker, R.;
* Supporting Information
S Savarino, P.; Barni, E.; Pellizzetti, E.; Graetzel, M. Micellar properties
Synthesis scheme, all spectral data, conductivity as a function of of nicotinamide-based surfactants. J. Colloid Interface Sci. 1989, 132,
the surfactant concentration, and cumulant distribution of the 469−474.
DLS study. This material is available free of charge via the (17) Wiederkehr, N. A.; Kalyanasundaram, K.; Grätzel, M.; Viscardi,
Internet at http://pubs.acs.org. G.; Savarino, P.; Barni, E. Micellization properties of zwitterionic


surfactants derived from nicotinic acid in aqueous solutions. Langmuir
AUTHOR INFORMATION 1991, 7, 23−29.
(18) Amrhar, J.; Chevalier, Y.; Gallot, B.; Perchec, P. L.; Auvray, X.;
Corresponding Author Petipas, C. Solution properties of zwitterionic pyridinio surfactants.
*E-mail: rsumita4@yahoo.co.in. Langmuir 1994, 10, 3435−3441.
Notes (19) Zhou, L.; Jiang, X.; Li, Y; Chen, Z.; Hu, X. Synthesis and
The authors declare no competing financial interest. properties of a novel class of gemini pyridinium surfactants. Langmuir


2007, 23, 11404−11408.
(20) Ghosh, A.; Dey, J. Physicochemical characterization and tube-
ACKNOWLEDGMENTS
like structure formation of a novel amino acid-based zwitterionic
This work was supported by a grant from the Department of amphiphile N-(2-hydroxydodecyl)-L-valine in water. J. Phys. Chem. B
Science and Technology (SR/FT/CS-026/2008). A.R. thanks 2008, 112, 6629−6635.
the University Grants Commission (Grant F.17-130/98(SA-I)) (21) Roy, S.; Dey, J. Spontaneously formed vesicles of sodium N-(1-
and M.M. thanks the Council of Scientific and Industrial acrylamidoundecanoyl)glycinate and L-alaninate in water. Langmuir
Research (Grant 09/599(0044)/2011-EMR-I) for research 2005, 21, 10362−10369.
fellowships. We thank Dr. Sajal Kanti Mal for his valuable (22) Bhattacharya, S.; Halder, J. Thermodynamics of micellization of
suggestion for synthesis of the amphiphile and Dr. S. Patra, multiheaded single-chain cationic surfactants. Langmuir 2004, 20,
Government College of Engineering and Ceramic Technology, 7940−7947.
Kolkata, for his assistance with the DLS measurement. (23) Tanford, C.; Nozakk, Y.; Rohde, M. F. Size and shape of


globular micelles formed in aqueous solution by n-alkyl polyoxy-
ethylene ethers. J. Phys. Chem. 1977, 81, 1555−1560.
REFERENCES (24) Ghosh, H. N.; Palit, D. K.; Sapre, A. V.; Ramarao, K. V. S.;
(1) Fendler, J. H. Membrane Mimetic Chemistry; Wiley: New York, Mittal, J. P. Dual sites of solvation for electrons produced by
1983. photoionisation in aqueous micellar solutions. Chem. Phys. Lett. 1993,
(2) Hoffman, H. Hundred years of colloid science fascinating 203, 5−11.
phenomena in surfactant solutions. Ber. Bunsen-Ges. Phys Chem. 1994, (25) Won, Y. −Y.; Brannan, A. K.; Davis, H. T.; Bates, F. S.
98, 1433−1455. Cryogenic transmission electron microscopy (cryo-TEM) of micelles
(3) Laughlin, R. G. The Aqueous Phase Behavior of Surfactants; and vesicles formed in water by poly(ethylene oxide)-based block
Academic Press: London, 1994.
copolymers. J. Phys. Chem. B 2002, 106, 3354−3364.
(4) Collier, J. H.; Messersmith, P. B. Phospholipid strategies in
(26) Hassan, P. A.; Raghavan, S. R.; Kaler, E. W. Microstructural
biomineralization and biomatarials research. Annu. Rev. Mater. Res.
changes in SDS micelles induced by hydrotropic salt. Langmuir 2002,
2001, 31, 237−263.
(5) Torchilin, V. P. Recent advances with liposomes as 18, 2543−2548.
pharmaceutical carriers. Nat. Rev. Drug Discovery 2005, 4, 145−160. (27) Roy, S.; Dey, J. Self-organization properties and microstructures
(6) Walsby, A. E. Gas vesicles. Microbiol. Rev. 1994, 58, 94−144. of sodium N-(11-acrylamidoundecanoyl)-L-valinate and -L-threoninate
(7) Fendler, J. H. Membrane Mimetic Chemistry; Wiley: New York, in water. Bull. Chem. Soc. Jpn. 2006, 79, 59−66.
1982; Chapter 6. (28) (a) Yan, Y.; Huang, J.; Li, Z.; Han, F.; Ma, J. Aggregates
(8) Kunitake, T., Mittal, K. L., Bothorel, P., Eds. Surfactants in transition depending on the concentration in the cationic
Solution; Plenum: New York, 1986; Vol. 5, p 727. bolaamphiphile/SDS mixed systems. Langmuir 2003, 19, 972−974.
(9) Grinberg, S.; Linder, C.; Kolot, V.; Waner, T.; Wiesmen, Z.; (b) Han, F.; He, X.; Huang, J.; Li, Z.; Wang, y.; Fu, H. Surface
Shaubi, E.; Heldmen, E. Novel cationic amphiphilic derivatives from properties and aggregates in the mixed systems of bolaamphiphiles and
vernonia oil: synthesis and self-aggregation into bilayer vesicles, their oppositely charged conventional surfactants. J. Phys. Chem. B
nanoparticles, and DNA complexants. Langmuir 2005, 21, 7638−7645. 2004, 108, 5256−5262.

12702 dx.doi.org/10.1021/la302484x | Langmuir 2012, 28, 12696−12703


Langmuir Article

(29) De, S.; Aswal, V. K.; Ramakrishnan, S. Phenyl-ring-bearing amphiphiles. Evidence of vesicle formation from single-chain
cationic surfactants: effect of ring location on the micellar structure. amphiphiles bearing a disaccharide headgroup. Langmuir 2000, 16,
Langmuir 2010, 26 (23), 17882−17889. 87−97.
(30) Pegiadou, S.; Perez, L.; Infante, M. R. Synthesis, characterization (50) Brunner, J.; Graham, D. E.; Hauser, H.; Semenza, G. Ion and
and surface properties of 1-N-L -tryptophan-glycerol-ether surfactants. sugar permeabilities of lecithin bilayers: comparison of curved and
J. Surfactants Deterg. 2000, 3, 517−524. planar bilayers. J. Membr. Biol. 1980, 57, 133−141.
(31) Ono, D.; Masuyama, A.; Nakatsuji, Y.; Okahara, M.; Yamamara, (51) Kano, K.; Fendler, J. H. Pyranine as a sensitive pH probe for
S.; Takeda, T. Preparation, surface-active properties and acid liposome interiors and surfaces. pH gradients across phospholipid
decomposition profiles of a new “soap” bearing a 1,3-dioxolane ring. vesicles. Biochim. Biophys. Acta 1978, 509, 289−299.
J. Am. Oil Chem. Soc. 1993, 70, 29−36. (52) Matzuzaki, A.; Harada, S.; Fujii, K.; Miyajima, K. Physicochem-
(32) Rosen, M. J. Surfactants and Interfacial Phenomena, 4th ed.; ical determinants for the interactions of magainins 1 and 2 with acidic
Wiley Interscience: New York, 2004. lipid bilayers. Biochim. Biophys. Acta 1991, 1063, 162−170.
(33) Chu, B.; Zhou, Z. Physical chemistry of polyalkylene block (53) (a) Neumann, R.; Ringsdorf, H.; Patton, E. V.; O’Brien, D. F.
copolymer surfactants. In Nonionic Surfactants; Nace, V. M., Ed.; Preparation and characterization of long chain amino acid and peptide
Surfactant Science Series, Vol. 60; Marcel Dekker: New York, 1996; p vesicle membranes. Biochim. Biophys. Acta 1987, 898, 338−348.
67. (b) Weinstein, J. N.; Yoshikami, S.; Henkart, P.; Blumenthal, R.;
(34) Rosen, M. J. Surfactants and Interfacial Properties, 3rd ed.; Wiley Hajins, W. A. Liposome-cell interaction: transfer and intracellular
Interscience: New York, 2004; pp 60−64. release of a trapped fluorescent marker. Science 1977, 195, 489−492.
(35) Mukherjee, P. The nature of the association equilibria and (54) Bhattacharya, S.; De, S.; George, S. K. Synthesis and vesicle
hydrophobic bonding in aqueous solutions of association colloids. Adv. formation from novel pseudoglyceryl dimeric lipids. Evidence of
Colloid Interface Sci. 1967, 1, 242−275. formation of widely different membrane organizations with exceptional
(36) Kalyanasundaram, K.; Thomas, J. K. Environmental effects on thermotropic properties. Chem. Commun. 1997, 2287−2288.
vibronic band intensities in pyrene monomer fluorescence and their (55) Kaler, E. W.; Murthy, A. K.; Rodrigueg, B. E.; Zasadzinski, J. A.
application in studies of micellar systems. J. Am. Chem. Soc. 1977, 99, N. Spontaneous vesicle formation in aqueous mixtures of single-tailed
2039−2044. surfactants. Science 1989, 245, 1371−1374.
(37) Shinitzky, M.; Barenholz, Y. Dynamics of the hydrocarbon layer (56) Kunitake, T.; Okahata, Y.; Yasunami, S. Substrate entrapment
in liposomes of lecithin and sphingomyelin containing dicetylphos- and formation of pH gradient in ammonium bilayer vesicles. Chem.
phate. J. Biol. Chem. 1974, 249, 2652−2657. Lett. 1981, 1397−1400.
(38) Shinitzky, M.; Dianoux, A. C.; Itler, C.; Weber, G. Micro- (57) Han, D.; Li, X.; Hong, S.; Jinnai, H.; Liu, G. Morphological
viscosity and order in the hydrocarbon region of micelles and transition of triblock copolymer cylindrical micelles responding to
membranes determined with fluorescent probes. I. Synthetic micelles. solvent change. Soft Matter 2012, 8, 2144−2151.
Biochemistry 1971, 10, 2106−2113.
(39) Shinitzky, M.; Yuli, I. Lipid fluidity at the submacroscopic level:
determination by fluorescence polarization. Chem. Phys. Lipids 1982,
30, 261−282.
(40) Mohanty, A.; Dey, J. Effect of the headgroup structure on the
aggregation behavior and stability of self-assemblies of sodium N-[4-
(n-dodecyloxy)benzoyl]-L-aminoacidates in water. Langmuir 2007, 23,
1033−1040.
(41) Ghosh, A.; Dey, J. Effect of hydrogen bonding on the
physicochemical properties and bilayer self-assembly formation of N-
(2-hydroxydodecyl)-L-alanine in aqueous solution. Langmuir 2008, 24,
6018−6026.
(42) Christov, N. C.; Denkov, N. D.; Kralchevsky, P. A.; Broze, G.;
Mehreteab, A. Kinetics of triglyceride solubilization by micellar
solutions of nonionic surfactant and triblock copolymer. 1. Empty
and swollen micelles. Langmuir 2002, 18, 7880−7886.
(43) Horbaschek, K.; Hoffmann, H.; Thunig, C. Formation and
properties of lamellar phases in systems of cationic surfactants and
hydroxy-naphthoate. J. Colloid Interface Sci. 1998, 206, 439−456.
(44) Namani, T.; Walde, P. From decanoate micelles to decanoic
acid/dodecylbenzenesulfonate vesicles. Langmuir 2005, 21, 6210−
6219.
(45) De Wall, S. L.; Wang, K.; Berger, D. R.; Watanabe, S.;
Hernandez, J. C.; Gokel, G. W. Azacrown ethers as amphiphile
headgroups: formation of stable aggregates from two- and three-armed
lariat ethers. J. Org. Chem. 1997, 62, 6784−6791.
(46) Bhattacharya, S.; Haldar, S. Synthesis, thermotropic behavior,
and permeability properties of vesicular membranes composed of
cationic mixed-chain surfactants. Langmuir 1995, 11, 4748−4757.
(47) Bhattacharya, S.; De, S.; Subramanian, M. Synthesis and vesicle
formation from hybrid bolaphile/amphiphile ion-pairs. Evidence of
membrane property modulation by molecular design. J. Org. Chem.
1998, 63, 7640−7651.
(48) Bhattacharya, S.; Biswas, J. Vesicle and stable monolayer
formation from simple “click” chemistry adducts in water. Langmuir
2011, 27, 1581−1591.
(49) Bhattacharya, S.; Ghanashyam Acharya, N. S. Vesicle and
tubular microstructure formation from synthetic sugar-linked

12703 dx.doi.org/10.1021/la302484x | Langmuir 2012, 28, 12696−12703

You might also like