9
Colloids and Surfaces A: Physicochem. Eng. Aspects 450 (2014) 106–114 Contents lists available at ScienceDirect Colloids and Surfaces A: Physicochemical and Engineering Aspects j ourna l h om epa ge: www.elsevier.com/locate/colsurfa Triton X-100 micelles modulated by solubilized cinnamic acid analogues: The pH dependant micellar growth Vijay Patel a,, Debes Ray b , Vinod K. Aswal b , Pratap Bahadur a a Department of Chemistry, Veer Narmad South Gujarat University, Surat 395007, India b Solid State Physics Division, Bhabha Atomic Research Centre, Mumbai 400085, India h i g h l i g h t s Antioxidants from cinnamic acid family were solubilized in TX-100 micelles. Marked micellar growth/transition was seen in the presence of cinnamic acid. Micellar shape/size depends on pH and interaction of cinnamic acid with micelles. 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 Article history: Received 30 November 2013 Received in revised form 28 February 2014 Accepted 3 March 2014 Available online 12 March 2014 Keywords: Triton X-100 Cinnamic acid Micelles Scattering a b s t r a c t Aqueous micellar solutions of a nonionic surfactant p-tert-octylphenoxy polyethylene (9.5) ether, Triton X-100 (TX-100) in the presence of cinnamic acid and its hydroxyl derivatives viz. p-coumaric acid, caffeic acid, ferulic acid and sinapic acid were examined by cloud point (CP), viscosity, dynamic light scatter- ing (DLS), nuclear magnetic resonance (NMR) and small-angle neutron scattering (SANS) measurements. These solubilizates influence the CP, viscosity and micelle size of TX-100 depending on their hydropho- bicity. Cinnamic acid, a biomedically useful compound displayed maximum growth in the micelles and was therefore studied in detail. We found that an increase in the cinnamic acid concentration, a decrease in pH and the presence of sodium chloride all favour the micellar growth at low temperature. The location of cinnamic acid in micelles was elucidated from 1 H NMR. SANS study showed that ellipsoidal micelles transform into rod-like on solubilization of cinnamic acid which otherwise change to spherical ones in alkaline medium. © 2014 Elsevier B.V. All rights reserved. 1. Introduction Micellar growth and sphere-to-rod shape transitions in sur- factant solutions are observed in the presence of inorganic salts and organic additives [1,2]. Hydrophobic and weakly polar organic solutes have also been examined for their effect on micelles [3]. While highly polar substances like short chain alcohols [4,5] and Corresponding author. Tel.: +91 9979509888. E-mail addresses: [email protected] (V. Patel), [email protected] (D. Ray), [email protected] (V.K. Aswal), [email protected] (P. Bahadur). amides [6,7] disintegrate micelles; weakly polar and sparingly solu- ble substances such as medium chain alcohols [8,9], amines [10,11] and phenols [12–14] favour micellar growth. There are few reports on the effect of aromatic acids on the behaviour of surfactant solutions [15–17]. Acids with low aque- ous solubility act as solubilizates and favour micellar growth but their salts behave like hydrotropes [17]. The study of the pH effect on the surfactant systems containing solubilized anthranilic acid in Triton X-100 (TX-100) micelles is interesting in this context [18]. Hydrophobic organic anions interact with cationic surfactant which leads to the micellar growth developing high vis- cosity/viscoelasticity in the dilute solutions. The most extensively studied system has been CTAB-sodium salicylate [19,20]. http://dx.doi.org/10.1016/j.colsurfa.2014.03.015 0927-7757/© 2014 Elsevier B.V. All rights reserved.

Triton X-100 micelles modulated by solubilized cinnamic acid analogues: The pH dependant micellar growth

  • Upload
    pratap

  • View
    216

  • Download
    2

Embed Size (px)

Citation preview

Page 1: Triton X-100 micelles modulated by solubilized cinnamic acid analogues: The pH dependant micellar growth

Ta

Va

b

h

a

ARRAA

KTCMS

1

fasW

(

h0

Colloids and Surfaces A: Physicochem. Eng. Aspects 450 (2014) 106–114

Contents lists available at ScienceDirect

Colloids and Surfaces A: Physicochemical andEngineering Aspects

j ourna l h om epa ge: www.elsev ier .com/ locate /co lsur fa

riton X-100 micelles modulated by solubilized cinnamic acidnalogues: The pH dependant micellar growth

ijay Patela,∗, Debes Rayb, Vinod K. Aswalb, Pratap Bahadura

Department of Chemistry, Veer Narmad South Gujarat University, Surat 395007, IndiaSolid State Physics Division, Bhabha Atomic Research Centre, Mumbai 400085, India

i g h l i g h t s

Antioxidants from cinnamic acidfamily were solubilized in TX-100micelles.Marked micellar growth/transitionwas seen in the presence of cinnamicacid.Micellar shape/size depends on pHand interaction of cinnamic acid withmicelles.

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

r t i c l e i n f o

rticle history:eceived 30 November 2013eceived in revised form 28 February 2014ccepted 3 March 2014vailable online 12 March 2014

eywords:

a b s t r a c t

Aqueous micellar solutions of a nonionic surfactant p-tert-octylphenoxy polyethylene (9.5) ether, TritonX-100 (TX-100) in the presence of cinnamic acid and its hydroxyl derivatives viz. p-coumaric acid, caffeicacid, ferulic acid and sinapic acid were examined by cloud point (CP), viscosity, dynamic light scatter-ing (DLS), nuclear magnetic resonance (NMR) and small-angle neutron scattering (SANS) measurements.These solubilizates influence the CP, viscosity and micelle size of TX-100 depending on their hydropho-bicity. Cinnamic acid, a biomedically useful compound displayed maximum growth in the micelles and

riton X-100innamic acidicelles

cattering

was therefore studied in detail. We found that an increase in the cinnamic acid concentration, a decreasein pH and the presence of sodium chloride all favour the micellar growth at low temperature. The locationof cinnamic acid in micelles was elucidated from 1H NMR. SANS study showed that ellipsoidal micellestransform into rod-like on solubilization of cinnamic acid which otherwise change to spherical ones inalkaline medium.

© 2014 Elsevier B.V. All rights reserved.

. Introduction

Micellar growth and sphere-to-rod shape transitions in sur-

actant solutions are observed in the presence of inorganic saltsnd organic additives [1,2]. Hydrophobic and weakly polar organicolutes have also been examined for their effect on micelles [3].

hile highly polar substances like short chain alcohols [4,5] and

∗ Corresponding author. Tel.: +91 9979509888.E-mail addresses: [email protected] (V. Patel), [email protected]

D. Ray), [email protected] (V.K. Aswal), [email protected] (P. Bahadur).

ttp://dx.doi.org/10.1016/j.colsurfa.2014.03.015927-7757/© 2014 Elsevier B.V. All rights reserved.

amides [6,7] disintegrate micelles; weakly polar and sparingly solu-ble substances such as medium chain alcohols [8,9], amines [10,11]and phenols [12–14] favour micellar growth.

There are few reports on the effect of aromatic acids on thebehaviour of surfactant solutions [15–17]. Acids with low aque-ous solubility act as solubilizates and favour micellar growth buttheir salts behave like hydrotropes [17]. The study of the pHeffect on the surfactant systems containing solubilized anthranilicacid in Triton X-100 (TX-100) micelles is interesting in this

context [18]. Hydrophobic organic anions interact with cationicsurfactant which leads to the micellar growth developing high vis-cosity/viscoelasticity in the dilute solutions. The most extensivelystudied system has been CTAB-sodium salicylate [19,20].
Page 2: Triton X-100 micelles modulated by solubilized cinnamic acid analogues: The pH dependant micellar growth

hysico

fI∼mdaaaal

aopr[usanCdamwcia[tw

zIauaasidTN

2

2

(owa(

2

2

cUcat

V. Patel et al. / Colloids and Surfaces A: P

TX-100 is a polyoxyethylene based alkyl aryl type nonionic sur-actant, most extensively studied for its micellar behaviour [21–26].ts critical micellar concentration (CMC) and CP are ∼0.27 mM and67 ◦C, respectively [12]. TX-100 is commonly used for solubilizingembrane proteins [27] and in detergent formulations from heavy

uty industrial products to gentle detergents [28]. The effects ofromatic solutes like p-toluidine [11], phenols [12] and anthraniliccid [18] on TX-100 micelles have been documented. p-Toluidinend various phenolic compounds markedly alter the micellar char-cteristics strongly dependant on the pH which determines theocation of these solubilizates in the micelles.

Cinnamic acid (CA) occurs naturally in many spices (cinnamonnd cloves), cranberries and prunes and acts against pathogenicrganisms [29]. It is an FDA approved flavouring agent used inerfumes, pharmaceuticals and food industries. Cinnamic acid andelated polyphenols show antimicrobial and antifungal activity30]. Hydroxycinnamic acids viz. p-coumaric acid, caffeic acid, fer-lic acid and sinapic acid are well known antioxidants which onolubilization in surfactant solutions, can alter micelle size or shapend rheological behaviour. The solubility of CA and related polyphe-ols in water [31] and supercritical CO2 [32] has been reported.innamic acid possesses an aqueous solubility of 0.002 M withissociation constant (pKa) 4.55 [31]. Cinnamic acid is proton-ted at the pH conditions below the pKa value while in alkalineedium it becomes more soluble in water. The enhanced solubilityould make CA very applicable for industrial use viz. pharma-

eutical and perfume industry. Studies on the solubilization andnteraction of some phenolic antioxidants viz. p-hydroxybenzoiccid, syringic acid, sinapic acid and quercetin [33] and parabens34] in Pluronic® micelles have been reported. In the presence ofrans-ortho-methoxycinnamic acid, CTAB forms long and entangledormlike micelles resulting in highly viscous solutions [35].

To the best of our knowledge, there are no reports of solubili-ation and interaction of CA and its analogues in TX-100 micelles.n view of this, here we report the effect of solubilization of CAnd its hydroxyl derivatives viz. p-coumaric acid, caffeic acid, fer-lic acid and sinapic acid on the phase behaviour of TX-100 inqueous solution. Cinnamic acid displays the strongest interactionmong other solubilizates studied with TX-100 micelles. As a con-equence, we have studied the effect of CA on TX-100 micellesn detail at different temperature, pH and salt concentration con-itions using physical and scattering techniques (DLS and SANS).he location of cinnamic acid in micelles is revealed from 1HMR.

. Experimental

.1. Materials

Triton X-100 (TX-100), cinnamic acid (CA), p-coumaric acidPCA), caffeic acid (CFA), ferulic acid (FA) and sinapic acid (SA) werebtained from Sigma-Aldrich and used as received. Triply distilledater was used to prepare aqueous solutions. Deionized water from

Millipore Milli-Q system was used to prepare samples for DLS. D2O99%) was used for sample preparation for SANS and NMR.

.2. Methods

.2.1. ViscosityThe viscosity measurements were performed in a temperature

ontrolled water bath with stability of ±0.1 ◦C. Calibrated Cannon

bbelohde viscometers were used with size 25 and 150 havingonstants 0.001869 and 0.03462 cSt s−1, respectively [36]. The vari-tion in flow time was found to be ±5 s. The absolute viscosities ofhe solutions obtained are multiplied by viscometer constant to get

chem. Eng. Aspects 450 (2014) 106–114 107

kinematic viscosity in centi-stokes and then multiplied by densityof solvent (water) to give solution viscosities in centipoise. Theseviscosities of solutions were divided by viscosity of water to obtainthe relative viscosity [8].

2.2.2. Cloud point (CP)The CPs of the solutions were determined by slowly heating the

sample solutions in thin 20 mL test tube immersed in a beaker hav-ing water. The solution temperature was increased slowly (heatingrate ∼1 ◦C min−1) with constant stirring using magnetic stirrerequipped with a heater. Temperature of first appearance of tur-bidity was taken as cloud point. The CP results were reproducibleup to ±0.5 ◦C.

2.2.3. Solubility experimentsUV–Vis double beam spectrophotometer (Thermo Fisher) was

used for solubilization experiments. It contains a matched pair ofstoppered quartz cells with optical path length of 1 cm. For analysis,wavelength used for CA, PCA, CFA, FA and SA are 272, 291, 321,320 and 319 nm, respectively, which correspond to their maximumabsorption. Dilute solutions of acids with concentrations rangingfrom 0.01 to 0.1 mM in methanol were used to prepare calibrationplot. It follows Beer–Lambert law (R2 = 0.9982). Sample solutionssaturated with acids were prepared by dissolving excess amount ofacids in water and 5% TX-100 and then stirred at 30 ◦C at 200 rpm for48 h. Excess amount of acid was removed by filtering the solutionswith millipore of size 0.45 �m. Filtered solution was diluted 10–100times with methanol. The very low amount of water permits thedirect use of calibration plot. The solubilities of acids in water andin 5% TX-100 were determined using the standard analytical shake-flask method.

2.2.4. Dynamic light scattering (DLS)The apparent hydrodynamic size of the micelles was obtained

from Zeta sizer Nano-ZS 4800 (Malvern Instruments, UK). The scat-tering angle was kept 90◦ fixed and the He–Ne laser was operatedat 633 nm. The Stokes–Einstein relationship was used to get thehydrodynamic size. Results are reproducible up to ±0.3 nm.

2.2.5. Small angle neutron scattering (SANS)SANS measurements were performed at the SANS diffractome-

ter at Guide Tube Laboratory, Dhruva reactor, BARC, Mumbai, India.The scattering data for a mean incident wavelength of 5.2 A with��/� = 15% were measured at 30 ◦C in the wave vector transfer(Q = 4�sin(�/2)/�, where � is scattering angle) range of 0.017 to0.35 A−1. The measured SANS data were corrected for the back-ground, the empty cell contributions and the transmission andnormalized to absolute cross-sectional unit using standard proto-cols.

2.2.5.1. SANS analysis. The differential scattering cross-section(d�/d�) per unit volume as a function of wave vector transfer Qand for a system of monodisperse particles in a medium can beexpressed as [37,38]

(d˙

)(Q ) = nV2(�p − �s)

2P(Q )S(Q ) + B, (1)

where n denotes the number density of particles, �p and �s are,respectively, the scattering length densities of particle and solvent

and V is the volume of the particle. P(Q) is the intraparticle structurefactor and S(Q) is the interparticle structure factor. B is a constantterm representing incoherent background, which is mainly due tothe hydrogen present in the sample.
Page 3: Triton X-100 micelles modulated by solubilized cinnamic acid analogues: The pH dependant micellar growth

108 V. Patel et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 450 (2014) 106–114

on X-1

af

P

F

tfS

(

p

wa

f

w(b

datlps

2

4Stpt

acts as a kosmotrope which reduces the solubility of surfactant inwater [46,47]. In 1.0 M NaCl, the CP of a TX-100 solution with solu-bilized CA decreases more sharply compared to only CA (Fig. 2(c)).

Table 1Solubility data of cinnamic acid and its analogues.

Organic acids Solubility (g/L) at 30 ◦C Po/w

Fig. 1. Chemical structures of Trit

The intraparticle structure factor P(Q) is decided by the shapend size of the particle and is the square of single-particle formactor F(Q) as determined by

(Q ) =⟨|F(Q )|2

⟩(2)

For a spherical particle of radius R, F(Q) is given by

(Q ) = 3

[sin(QR) − QR cos(QR)

(QR)3

]. (3)

S(Q) describes the interaction between the particles present inhe system and it is the Fourier transform of the pair correlationunction for the mass centres of the particles. For dilute systems,(Q) ∼1 and Eq. (1) subsequently reduces to

)(Q ) = nV2 ×

(�p − �s

)2 × P(Q ) + B. (4)

The polydispersity in size distribution of the particles is incor-orated using the following integration [39]:

d˝(Q ) =

∫d˙

d˝(Q, R) × f (R)dR + B, (5)

here f(R) is the particle size distribution and usually accounted by log–normal distribution as given by

(R) = 1√2�R

exp

[− 1

22

(ln

R

Rmed

)2]

, (6)

here Rmed is the median value and is the standard deviationpolydispersity) of the distribution. The mean radius (Rm) is giveny Rm = Rmed exp(2/2).

The data have been analyzed by comparing the scattering fromifferent models to the experimental data. Throughout the datanalysis, corrections were made for instrumental smearing, wherehe calculated scattering profiles smeared by the appropriate reso-ution function to compare with the measured data [40]. The fittedarameters in the analysis were optimized using nonlinear least-quare fitting program to the model scattering [41].

.2.6. Nuclear magnetic resonance (NMR)1H NMR experiments were performed on a Bruker AVANCE-II

00 MHz spectrometer. All measurements were performed at 30 ◦C.amples were prepared in D2O. The spectrum was calibrated by set-ing the HDO peak at a chemical shift of 4.65 ppm at 298 K. The HDOeak due to residual water was eliminated by solvent suppressionechniques.

00 and cinnamic acid analogues.

3. Results and discussion

3.1. Solubility of cinnamic acid analogues in TX-100 micelles

The CA analogues find numerous industrial as well as biologicalapplications/activity viz. antioxidant, antitubercular, antidiabetic,antimicrobial, antiviral and anticancer agents [30,42]. These acidshave limited aqueous solubility and get solubilized in TritonX-100 micelles. The surfactant solution behaviour depends onhydrophobicity, quantity solubilized and location of these acids inmicelles. Depending upon the octanol–water partition coefficients(Po/w), these solubilizates partition in micelles and accordinglyalter the micellar characteristics. All the studied solubilizatespossess a common aromatic ring, –COOH group, double bondand vary in functional groups like –OH, –OCH3 which play akey role in determining solubility and location in the micelles(Fig. 1).

The solubility data are presented for cinnamic acid analogues(Table 1). It clearly shows that CFA has the maximum aqueoussolubility due to two –OH group while CA possesses maximum sol-ubility in 5% TX-100 compared to other solubilizates studied due tothe highest Po/w. SA has the lowest aqueous and micellar solubilitydue to the presence of two –OCH3 groups and lower Po/w. With thehighest Po/w, CA penetrates in the TX-100 micelles and displays thestrongest interaction. The solubilization of CA in TX-100 micellesleads to an increase in solution viscosity, indicating micellar growthwith a corresponding decrease in CPs (Fig. 2). TX-100 forms rodlikemicelles with progressive solubilization of CA (discussed later inSANS part of MS). Here, CA acts as a water structure maker andtherefore, dehydration of TX-100 occurs which results in loweringits CP (Fig. 2(c)). Addition of inorganic salt like sodium chloride alsoleads to dehydration of micelles [44,45]. It is well known that NaCl

Water (Exp.) Water* (L) 5% TX-100

CA 0.29 0.31 7.75 2.41PCA 0.97 1.02 4.86 1.87CFA 1.25 1.23 3.16 1.42FA 0.88 0.92 4.28 1.64SA 0.19 0.22 2.65 0.99

* Data are taken from Ref. [31].Po/w Data are taken from Ref. [43] (RSC website).

Page 4: Triton X-100 micelles modulated by solubilized cinnamic acid analogues: The pH dependant micellar growth

V. Patel et al. / Colloids and Surfaces A: Physico

30

40

50

60

70

0

2

4

6

8

10

0 10 20 30 40 500

20

40

60

80

CP,

°Cc

a

b

rel

Dh,

nm

[Cinnamic aci d], mM

Fig. 2. (a) Micelle hydrodynamic diameter (Dh), (b) relative viscosity and (c) cloudpoint of 5% TX-100 with CA in the (�) presence and (©) absence of 1.0 M NaCl at3

Hiid(

hiilhohDgsih

h1Tuotis

0 ◦C.

ere, the lowering of the cloud point is dominated by dehydrat-ng effects of CA and NaCl which favours the micellar growth andt is further supported by an increase in apparent hydrodynamiciameter (Dh) of TX-100 micelles acquired from DLS measurementsFig. 2(a)).

The presence of –OH groups in PCA and CFA makes them moreydrophilic and hence it is difficult for their molecules to penetrate

nto the micelle. This is probably the reason for their lower solubil-ty in TX-100 micelles (Table 1). Consequently, PCA and CFA have aower solubility compared to CA in TX-100. Now, FA has a slightlyigher solubility than CFA due to the presence of –OCH3 in placef second –OH group. p-Coumaric acid, CFA and FA due to theirydrophilic nature do not have a significant effect on CP, rel andh of TX-100 (figure not shown). Sinapic acid has one extra –OCH3roup compared to FA. Hence, it becomes more hydrophobic andhould have a higher solubility in the micelle but an opposite effects observed (Table 1). This may be due to its large structure i.e. stericindrance and low Po/w.

Incorporation of solubilizates to TX-100 micelles inducesydrophobicity in the system. The size distribution plots for 5% TX-00 in the presence of cinnamic acid analogues are shown in Fig. 3.he Dh of 5% TX-100 micelle is 10.6 nm which increases with sol-bilization of all acids. The order of Dh of micelles in the presence

f acid is as follows: CA > PCA > FA > CFA > SA. This clearly reflectshat solubilization of CA (highest Po/w) has a maximum modulat-ng effect on the apparent micelle size (73.7 nm) indicating thetrongest interaction with TX-100 micelles [12].

chem. Eng. Aspects 450 (2014) 106–114 109

3.2. Effect of temperature

Dehydration of nonionic micelles can be brought up by the addi-tion of salt, increasing temperature or solubilizing the hydrophobicsolutes [8,12,38,25,48]. Nonionic surfactant solutions often showa micellar transition near the phase separation temperature (CP).Solubilization of CA lowers it to room temperature. Fig. 4 showsthe effect of temperature on an aqueous solution of TX-100 in theabsence and presence of CA. This indicates that the cloud point isdecreased to 53, 39 and 32 ◦C (Fig. 2) by the addition of 20, 40 and50 mM CA, respectively. The increase in the hydrodynamic size andviscosity confirms the micellar growth.

3.3. Effect of pH

3.3.1. Viscosity, cloud point and DLSCinnamic acid shows enhanced antimicrobial/antioxidant activ-

ity in acidic media [49]. Hence, microstructure changes in TX-100micelles with solubilized CA are of prime importance and arestudied in detail. Fig. 5 shows the CP, rel and micelle Dh of 5%TX-100 solutions in the presence of 40 mM CA (fixed) at varyingpH at 30 ◦C. Being a nonionic surfactant, the micellar behaviourof TX-100 is unaffected by pH [12]. The presence of pH respon-sive –COOH group makes CA sensitive to pH. Our SANS data showthat originally (5% TX-100 + 40 mM CA) micelles are rodlike. Beinghydrophobic, CA molecules penetrate into the TX-100 micelles byhydrophobic interaction and accordingly, induce micellar growthwhich is supported by increase in both the viscosity as well asthe apparent hydrodynamic size (Fig. 3). Originally, pH of the 5%TX-100 solution having 40 mM solubilized CA is 3.6. Now, theionization of CA is opposed at the pH <3.6 and protonation ofCA molecules induces more hydrophobicity and eventually micel-lar growth is observed due to its penetration within the micelle.Protonation of CA induces the highest hydrophobicity and con-sequently maximum micelle growth is observed. The solutionviscosity attains the maximum with a reverse trend for CP. Furtherlowering of pH does not involve any change in micelle param-eters and a plateau is observed. The pKa of CA is 4.55 [31]. AtpH > 4.55, negative charge is developed on CA and hence its solubil-ity in water is increased and hydrophobic interaction is decreasedand thus CA is pushed out towards the micelle interface. As aresult, the viscosity of the solution is decreased with a simultane-ous increase in the CP. Above pH > 6.0 almost all the CA moleculesare converted in to cinnamate ions and as a result, no markedchange in the viscosity and corresponding CP has been observed.Here, the change in pH towards alkaline media pulls CA towardsthe micellar interface and rod-like micelles are transformed intospherical-like geometry. This has been discussed in the later partof the manuscript. Furthermore, the size distribution profiles ofmicelles obtained from DLS measurement can give a better ideaabout the pH-induced change in hydrodynamic size of micelles(Fig. 6).

3.3.2. SANSThe progressive solubilization of cinnamic acid leads to morpho-

logical changes in TX-100 micelles. At higher level of solubilization,the large increase in viscosity and Dh suggests the formation ofbigger size micelles. SANS measurements have been performed tomonitor CA-induced morphological changes in TX-100 micelles andare shown in Fig. 7. SANS data show the effect of CA varied over aconcentration range from 0 to 40 mM with fixed concentration (5%)of TX-100. There is a strong build up of scattering intensity in the

low-Q region with the increase in the CA concentration whereasthey overlap at higher Q region irrespective of the CA concentra-tion. This is an indication of the morphological changes in micelles.The analysis shows that originally TX-100 (5%) micelles are
Page 5: Triton X-100 micelles modulated by solubilized cinnamic acid analogues: The pH dependant micellar growth

110 V. Patel et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 450 (2014) 106–114

Fig. 3. Size distribution curves for 5% TX-100 in the presence of CA, PCA, CFA, FA and SA at 30 ◦C.

30 35 40 45 500

20

40

60

80

100

0

2

4

6

8

10

Dh,

nm

Temperatur e, °C

a

0 mM 20 mM 40 mM

0 mM 20 mM 40 mM

b

rel

Fig. 4. Effect of temperature on (a) micelle hydrodynamic diameter (Dh) and (b)viscosity of 5% TX-100 solutions in the absence and presence of CA.

0 2 4 6 8 10 12 140

10

20

30

401

2

3

4

5

6

730

45

60

75

90 c

a

Dh,

nm

pH

b

rel

CP

°C

Fig. 5. Effect of pH on (a) hydrodynamic size, (b) relative viscosity and (c) cloudpoint of 5% TX-100 micelles in the presence of 40 mM CA at 30 ◦C.

Page 6: Triton X-100 micelles modulated by solubilized cinnamic acid analogues: The pH dependant micellar growth

V. Patel et al. / Colloids and Surfaces A: Physico

Fig. 6. Size distribution curves for 5% TX-100 + 40 mM CA at varying pH at 30 ◦C.

Fig. 7. Effect of concentration and pH on SANS pattern of 5% TX-100 in the presenceof CA at 30 ◦C. Inset shows the plots in a linear scale. The solid lines represent thefitted data.

chem. Eng. Aspects 450 (2014) 106–114 111

ellipsoidal with axial ratio of ∼3 [50]. Mahajan et al. [51] showedthat Nagg of 3% TX-100 is 378 and approximately same semi-minoraxis (20 A). However, a slight change was observed in the semi-major axis which may be due to the compact packing of TX-100micelles. In the presence of a small amount of CA (10 mM) the axialratio slightly increases and becomes 3.8, presenting the ellipsoidalshaped micelles (Table 2). We conclude that the micelle shape isnot much affected by lower [CA]. With increase in [CA], the axialratio increases and at 30 mM CA, the axial ratio is ∼4.7 depictingstill ellipsoidal micelles. Also a marked increase in the aggrega-tion number is observed as a result of the concentration gradient ofCA. At 40 mM CA, uniaxial micellar growth was observed; micelleshave an axial ratio ∼10 which presents a rod-like structure with aradius 20.3 A and length 206.8 A. SANS results for 40 mM at pH = 3.6show typical slope of about (−1) on log–log scale in low-Q regionwhich corresponds to the presence of rod-like micelles [52,53].These fitted parameters have been obtained by the model fittingas well as by the normalization of scattering data in the absoluteunit.

A change in the pH of the solution has a striking effect on themicellar transition and aggregation characteristics [11,54]. Lower-ing of the pH leads CA into its nonionized form (hydrophobic) whichgets solubilized in the micelles and consequently induces micellargrowth. Now, a change in the pH of the system towards the alkalinemedium has a prominent effect on micelle shape, size and aggre-gation characteristics [54]. Fig. 7 shows that at pH = 9.0, the overallscattered intensity in the low Q-region decreased drastically (lowerthan pure TX-100) which indicates decrease in micellar size. This isdue to the development of negative charge on CA which increasesits solubility in water and consequently CA is expelled out of themicelle. The absence of slope (−1) and a flat scattering pattern sug-gest the transformation from rod-like to spherical micelles at pH9. Spherical micelles have a radius of 23.1 A and an aggregationnumber of 142. Our SANS results are also supported by viscosityand DLS results. Micellar parameters obtained from SANS data aresummarized in Table 2.

3.3.3. NMRChanges in the microenvironment of the surfactant solution can

easily be monitored by the analysis of 1H NMR spectrum. Cinnamicacid-induced microstructural changes can be quantified from theobserved chemical shifts of TX-100 protons (Fig. 8). The 1H NMRspectra and chemical shifts of TX-100 and CA protons are shownin Fig. 8(A). TX-100 protons are highly sensitive to the change inthe polarity of their microenvironment. In the presence of 20 mMcinnamic acid, all TX-100 protons show an upfield shift indicat-ing that these protons experience more hydrophobic environment(Fig. 8(B)). The decrease in the intensity of the NMR peaks of cin-namic acid protons clearly indicates that they are shielded whereasthe broadening of NMR peaks indicates the micellar growth [55,56].At a higher level of solubilization (40 mM), NMR peaks of the aro-matic protons of cinnamic acid (C4 and C5) completely merged withT8 proton peak while the other protons show upfield shifts. Herethe T3, T2 and T1 protons of the alkyl chain of TX-100 form the coreof the micelle and are located at 1.59, 1.22 and 0.65 ppm which in40 mM CA shifts to 1.25, 0.87 and 0.31 ppm, respectively. At thesame time, the vinyl protons of CA also experience large upfieldshifts. The shell protons of TX-100 are affected more compared tothe core and aromatic protons by the presence of cinnamic acidwhich indicates that majority of CA molecules may reside at thecore–shell interface at higher level of solubilization.

Now, a change in the pH of the solution has a significant effect

on the NMR spectrum also. Fig. 8(B) shows the 1H NMR spectrum of5% TX-100 in the presence of 40 mM cinnamic acid at pH ∼ 9. All theprotons exhibit a downfield shift and the intensity of the aromaticprotons of CA increases suggesting that they are deshielded and
Page 7: Triton X-100 micelles modulated by solubilized cinnamic acid analogues: The pH dependant micellar growth

112 V. Patel et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 450 (2014) 106–114

Table 2Micellar parameters for 5% TX-100 in the presence of CA at 30 ◦C.

[CA] (mM) Semi-major axis a (Å) Semi-minor axis b (Å) Axial ratio Nagg Shape

0 63.5 19.8 3.2 287 Ellipsoidal10 76.0 20.0 3.8 351 Ellipsoidal30 95.6 20.3 4.7 454 Ellipsoidal40 (pH = 3.6) 206.8 20.3 10.2 737 Rod-like

40 (pH = 9.0) 23.1 23.1 1.0 142 Spherical

F fect of

eCmt

ig. 8. (A) Structures and 1H NMR spectra of (a) TX-100 and (b) cinnamic acid (B) Ef

xperience a relatively more hydrophilic environment. As a resultA molecules are pushed out from core–shell interface towardsicelle interface and hence rod-like micelles are transformed in

o spherical micelles.

concentration and pH on TX-100 protons in the presence of cinnamic acid at 30 ◦C.

4. Conclusion

The influence of the solubilization of phenolic antioxidantsfrom cinnamic acid family viz. CA, PCA, CFA, FA and SA on the

Page 8: Triton X-100 micelles modulated by solubilized cinnamic acid analogues: The pH dependant micellar growth

hysico

abuos

wroacmCoaiinsp

A

(

R

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[[

[

[

[

[

[

[

[

[

[

[

[

V. Patel et al. / Colloids and Surfaces A: P

ggregation behaviour of TX-100 micelles in aqueous medium haseen investigated by viscosity, DLS, SANS and NMR studies. The sol-bility data for CA and its analogues in TX-100 micelles follow therder CA > PCA > FA > CFA > SA, a trend expected from their chemicaltructures.

Cinnamic acid alters the micellar characteristics of TX-100hich is strongly dependant on pH as shown from DLS and SANS

esults and supported by the CP and viscosity data. The behaviourf CA as a hydrophobic solubilizate in acidic pH and a hydrotrope inlkaline pH dictates its location in micelles and as a consequence,hanges the micellar size and shape. SANS data shows that theicellar aggregation number increases with the concentration of

A. The formation of rodlike micelles on progressive solubilizationf cinnamic acid and their transformation into spherical micelles inlkaline medium has been confirmed from SANS. NMR experimentsndicate that majority of CA molecules may reside at core–shellnterface of TX-100 micelles at higher level of solubilization. As cin-amic acid and related compounds are biologically important, thistudy might be useful for their surfactant based formulations inharmaceutical and food industry.

cknowledgement

The authors thank UGC New Delhi for the financial assistanceProject No: 37-527/(2009) SR).

eferences

[1] P. Alexandridis, T.A. Hatton, Poly(ethylene oxide) poly(propylene oxide)poly(ethylene oxide) block copolymer surfactants in aqueous solutions and atinterfaces: thermodynamics, structure, dynamics, and modeling, Colloids Surf.,A 96 (1995) 1–46.

[2] P.A. Hassan, J.V. Yakhmi, Growth of cationic micelles in the presence of organicadditives, Langmuir 16 (2000) 7187–7191.

[3] R.A. Rahem, The influence of hydrophobic counterions on micellar growth ofionic surfactants, Adv. Colloid Interface Sci. 141 (2008) 24–36.

[4] T. Gu, P.A. Galera-Gomez, The effect of different alcohols and other polar organicadditives on the cloud point of Triton X-100 in water, Colloids Surf., A 147(1999) 365–370.

[5] P. Parekh, K. Singh, D.G. Marangoni, P. Bahadur, Effect of alcohols on aqueousmicellar solutions of PEO–PPO–PEO copolymers: a dynamic light scattering and1H NMR study, J. Mol. Liq. 165 (2012) 49–54.

[6] B. Bharatiya, C. Guo, J.H. Ma, O. Kubota, K. Nakashima, P. Bahadur, Urea-induced demicellization of Pluronic L64 in water, Colloid Polym. Sci. 287 (2009)63–71.

[7] C.C. Ruiz, F.G. Sanchez, Effect of urea on aggregation behaviour of Triton-X 100micellar solutions: a photophysical study, J. Colloid Interface Sci. 165 (1994)110–115.

[8] V. Patel, J. Dey, R. Ganguly, S. Kumar, S. Nath, V.K. Aswal, P. Bahadur, Solubi-lization of hydrophobic alcohols in aqueous Pluronic solutions: investigatingthe role of dehydration of micellar core in tuning the restructuring and growthof Pluronic micelles, Soft Matter 9 (2013) 7583–7591.

[9] P. Parekh, P. Bahadur, Dynamic light scattering studies on the effect of 1-alkanols, alkanediols and alkoxyethanols on the micelles of a moderatelyhydrophobic/hydrophilic PEO–PPO–PEO triblock copolymer, J. SurfactantsDeterg. 14 (2011) 425–432.

10] B. Bharatiya, G. Ghosh, V.K. Aswal, P. Bahadur, Effect of effect of n-hexanol andn-hexylamine on the micellar solutions of Pluronic F127 and P123 in water and1 M NaCl, J. Dispersion Sci. Technol. 31 (2010) 660–667.

11] K. Singh, N. Dharaiya, G. Marangoni, P. Bahadur, Dissimilar effects of solubilizedp-toluidine on the shape of micelles of differently charged surfactants, ColloidsSurf., A 436 (2013) 521–529.

12] N. Dharaiya, P. Bahadur, Phenol induced growth in Triton X-100 micelles: effectof pH and phenols’ hydrophobicity, Colloids Surf., A 410 (2012) 81–90.

13] J.P. Mata, V.K. Aswal, P.A. Hassan, P. Bahadur, A phenol-induced structuraltransition in aqueous cetyltrimethylammonium bromide solution, J. ColloidInterface Sci. 299 (2006) 910–915.

14] P. Sabatino, A. Szczygiel, D. Sinnaeve, M. Hakimhashemi, H. Saveyn, J.C. Martins,P.V. Meeren, NMR study of the influence of pH on phenol sorption in cationicCTAB micellar solutions, Colloids Surf., A 370 (2010) 42–48.

15] W.S. Higazy, F.Z. Mahmoud, A.A. Taha, S.D. Christian, Effects on the micellar sol-

ubilization of organic compounds by surfactant micelles. I. Length of carboxylicside chains in aromatic acids, J. Solution Chem. 17 (1988) 191–202.

16] S.S. Shah, K. Naeem, S.W.H. Shah, H. Hussain, Solubilization of short chainphenylalkanoic acids by a cationic surfactant, cetyltrimethylammonium bro-mide, Colloids Surf., A 148 (1999) 299–304.

[

[

chem. Eng. Aspects 450 (2014) 106–114 113

17] K. Naeem, S.W.H. Shah, B. Naseem, S.S. Shah, Interactions of short chain pheny-lalkanoic acids within ionic surfactant micelles in aqueous media, J. Serb. Chem.Soc. 77 (2012) 201–210.

18] G. Verma, V.K. Aswal, S.K. Kulshreshtha, P.A. Hassan, E.W. Kaler, Adsorbedanthranilic acid molecules cause charge reversal of nonionic micelles, Langmuir24 (2008) 683–687.

19] N.K. Pokhriyal, J.V. Joshi, P.S. Goyal, Viscoelastic behaviour of cetyltrimethylammonium bromide/sodium salicylate/water system: effect of solubilisationof different polarity oils, Colloids Surf., A 218 (2003) 201–212.

20] V. Hartmann, R. Cressely, Influence of sodium salicylate on the rheologi-cal behaviour of an aqueous CTAB solution, Colloids Surf., A 121 (1997)151–162.

21] C.J. Biaselle, D.B. Millar, Studies on Triton X-100 detergent micelles, Biophys.Chem. 3 (1975) 355–361.

22] H.H. Paradies, Shape and size of a nonionic surfactant micelle. Triton X-100 inaqueous solution, J. Phys. Chem. 84 (1980) 599–607.

23] R.J. Robson, E.A. Dennis, The size, shape, and hydration of nonionic surfactantmicelles. Triton X-100, J. Phys. Chem. 81 (1977) 1075–1078.

24] W. Brown, R. Rymden, J.V. Stam, M. Almgren, G. Svensk, Static and dynamicproperties of non-ionic amphiphile micelles: Triton X-100 in aqueous solution,J. Phys. Chem. 93 (1989) 2512–2519.

25] K. Streletzky, G.D.J. Phillies, Temperature dependence of Triton X-100 micellesize and hydration, Langmuir 11 (1995) 42–47.

26] M. Kumbhakar, T. Goel, T. Mukherjee, H. Pal, Role of micellar size and hydrationon salvation dynamics: a temperature dependent study in Triton-X-100 andBrij-35 micelles, J. Phys. Chem. B 108 (2004) 19246–19254.

27] G.C. Kresheck, J. Hwang, Phase separation properties of several aqueous alkyldimethylphosphine oxide/phospholipid mixtures and their potential use forprotein purification, Chem. Phys. Lipids 76 (1995) 193–199.

28] E.W. Flick, Advanced Cleaning Product Formulations, 2, Nomad Press, 2008, pp.159.

29] J.A. Hoskins, The occurrence, metabolism and toxicity of cinnamic acid andrelated compounds, J. Appl. Toxicol. 4 (1984) 283–292.

30] P. Sharma, Cinnamic acid derivatives: a new chapter of various pharmacologicalactivities, J. Chem. Pharm. Res. 3 (2011) 403–423.

31] F.L. Mota, A.J. Queimada, S.P. Pinho, E.A. Macedo, Aqueous solubility of somenatural phenolic compounds, Ind. Eng. Chem. Res. 47 (2008) 5182–5189.

32] Y.-P. Chen, Y.-M. Chen, M. Tang, Solubilities of cinnamic acid, phenoxyaceticacid and 4-methoxyphenylacetic acid in supercritical carbon dioxide, FluidPhase Equilib. 275 (2008) 33–38.

33] A. Parmar, K. Singh, A. Bahadur, G. Marangoni, P. Bahadur, Interaction and sol-ubilization of some phenolic antioxidants in Pluronic® micelles, Colloids Surf.,B 86 (2011) 319–326.

34] M. Khimani, R. Ganguly, V.K. Aswal, S. Nath, P. Bahadur, Solubilization ofparabens in aqueous Pluronic solutions: investigating the micellar growth andinteraction as a function of paraben composition, J. Phys. Chem. B 116 (2012)14943–14950.

35] A.M. Ketner, R. Kumar, T.S. Davies, P.W. Elder, S.R. Raghavan, A simple class ofphotorheological fluids: surfactant solutions with viscosity tunable by light, J.Am. Chem. Soc. 129 (2007) 1553–1559.

36] ASTM Standard D 445-12 and D 446-12 2012 and the references within.37] J.B. Hayter, J. Penfold, Determination of micelle structure and charge by neutron

small-angle scattering, Colloid Polym. Sci. 261 (1983) 1022–1030.38] E.W. Kaler, Small-angle scattering from colloidal dispersions, J. Appl. Cryst. 21

(1988) 729–736.39] J.S. Pedersen, Analysis of small-angle scattering data from colloids and poly-

mer solutions: modeling and least-squares fitting, Adv. Colloid Interface Sci.70 (1997) 171–210.

40] J.S. Pedersen, C. Riekel, Resolution function and flux at the sample for small-angle X-ray scattering calculated in position–angle–wavelength space, J. Appl.Cryst. 24 (1991) 893–909.

41] P.R. Bevington, Data Reduction and Error Analysis for Physical Sciences,McGraw-Hill, New York, NY, 1969.

42] P. De, M. Baltas, F. Bedos-Belval, Cinnamic acid derivatives as anticanceragents—a review, Curr. Med. Chem. 18 (2011) 1672–1703.

43] Chem Spider-The free Chemical Database from RSC (http://www.chemspider.com/).

44] J.A. Molina-Bolivar, J. Aguiar, C.C. Ruiz, Growth and hydration of Triton X-100micelles in monovalent alkali salts: a light scattering study, J. Phys. Chem. B106 (2002) 870–877.

45] R.K. Mahajan, K.K. Vohra, N. Kaur, V.K. Aswal, Organic additives and electrolytesas cloud point modifiers in octylphenol ethoxylate solutions, J. SurfactantsDeterg. 11 (2008) 243–250.

46] Y. Zhang, P.S. Cremer, Interactions between macromolecules and ions: theHofmeister series, Curr. Opin. Chem. Biol. 10 (2006) 658–663.

47] Y. Zhang, S. Furyk, D.E. Bergbreiter, P.S. Cremer, Specific ion effects on the watersolubility of macromolecules: PNIPAM and the Hofmeister series, J. Am. Chem.Soc. 127 (2005) 14505–14510.

48] P. Parekh, R. Ganguly, V.K. Aswal, P. Bahadur, Room temperature sphere-to-rodgrowth of Pluronic® P85 micelles induced by salicylic acid, Soft Matter 8 (2012)5864–5872.

49] M. Friedman, H.S. Jurgens, Effect of pH on the stability of plant phenolic com-pounds, J. Agric. Food Chem. 48 (2000) 2101–2110.

50] P.S. Goyal, S.V.G. Menon, B.A. Dasannacharya, P. Thiyagarajan, SANS study ofmicellar structure and interactions in Triton X-100 solutions, Physica B 213(214) (1995) 610–612.

Page 9: Triton X-100 micelles modulated by solubilized cinnamic acid analogues: The pH dependant micellar growth

1 hysico

[

[

[

[

[

14 V. Patel et al. / Colloids and Surfaces A: P

51] R.K. Mahajan, K.K. Vohra, V.K. Aswal, Small angle neutron scattering measure-ments of aggregation behaviour of mixed micelles of conventional surfactantswith triblock polymer L64, Colloids Surf., A 326 (2008) 48–52.

52] V. Patel, S. Chavda, V.K. Aswal, P. Bahadur, Effect of a hydrophilic PEO–PPO–PEO

copolymer on cetyltrimethylammonium tosylate solutions in water, J. Surfac-tants Deterg. 15 (2012) 377–385.

53] R. Kumar, G.C. Kalur, L. Ziserman, D. Danino, S.R. Raghavan, Wormlike micellesof a C22-tailed zwitterionic betaine surfactant: from viscoelastic solutions toelastic gels, Langmuir 23 (2007) 12849–12856.

[

chem. Eng. Aspects 450 (2014) 106–114

54] G. Verma, V.K. Aswal, P.A. Hassan, pH-Responsive self-assembly in an aqueousmixture of surfactant and hydrophobic amino acid mimic, Soft Matter 5 (2009)2919–2927.

55] J.H. Ma, C. Guo, Y.-L. Tang, H.-Z. Liu, 1H NMR spectroscopic investigations on

the micellization and gelation of PEO–PPO–PEO block copolymers in aqueoussolutions, Langmuir 23 (2007) 9596–9605.

56] T. Nivaggioli, B. Tsao, P. Alexandridis, T.A. Hatton, Microviscosity in Pluronicand Tetronic poly(ethylene oxide)–poly(propylene oxide) block copolymermicelles, Langmuir 11 (1995) 119–126.