91
Interaction of sodium hyaluronate with a biocompatible cationic surfactant from lysine: a binding study Matej Bračič, a Per Hansson, b Lourdes Pérez, c Lidija F. Zemljič, a and Ksenija Kogej d* a Institute for the Engineering and Design of Materials, University of Maribor, Smetanova 17, 2000 Maribor, Slovenia. b Department of Pharmacy, Biomedical Centre, Uppsala University, SE-75123 Uppsala, Sweden c Department of Chemical and Surfactant technology, Instituto de Química Avanzada de Cataluña, CSIC, Jordi Girona 18-26, 08034 Barcelona, Spain 1 1 2 3 4 5 6 7 8 9 10 11 12 1 2

digital.csic.esdigital.csic.es/bitstream/10261/131661/1/2015_Langmuir.docx · Web viewAll aqueous NaHA solutions for potentiometric and fluorescence measurements were prepared by

Embed Size (px)

Citation preview

Page 1: digital.csic.esdigital.csic.es/bitstream/10261/131661/1/2015_Langmuir.docx · Web viewAll aqueous NaHA solutions for potentiometric and fluorescence measurements were prepared by

Interaction of sodium hyaluronate with a

biocompatible cationic surfactant from lysine: a

binding study

Matej Bračič,a Per Hansson,b Lourdes Pérez,c Lidija F. Zemljič,a and Ksenija Kogejd*

aInstitute for the Engineering and Design of Materials, University of Maribor, Smetanova 17,

2000 Maribor, Slovenia.

bDepartment of Pharmacy, Biomedical Centre, Uppsala University, SE-75123 Uppsala, Sweden

cDepartment of Chemical and Surfactant technology, Instituto de Química Avanzada de Cataluña,

CSIC, Jordi Girona 18-26, 08034 Barcelona, Spain

dDepartment of Chemistry and Biochemistry, Faculty of Chemistry and Chemical Technology,

University of Ljubljana, 1000 Ljubljana, Slovenia.

* [email protected]

Keywords: sodium hyaluronate, biocompatible cationic surfactant, fluorescence measurements,

binary-surfactant-mixture binding model; competitive adsorption.

1

1

2

3

4

5

6

7

8

9

10

11

12

13

14

15

16

12

Page 2: digital.csic.esdigital.csic.es/bitstream/10261/131661/1/2015_Langmuir.docx · Web viewAll aqueous NaHA solutions for potentiometric and fluorescence measurements were prepared by

Abstract 207 words

Mixtures of natural and biodegradable surfactants and ionic polysaccharides have attracted

considerable research interest in recent years because they prosper as antimicrobial materials for

medical applications. In the present work, interactions between the lysine derived biocompatible

cationic surfactant, abbreviated as MKM, and the sodium salt of hyaluronic acid (NaHA) are

investigated in aqueous media by potentiometric titrations using the surfactant sensitive electrode

(SSE) and pyrene based fluorescence spectroscopy. The critical micelle concentration (CMC) in

pure surfactant solutions and the critical association concentration (CAC) in the presence of

NaHA are determined in dependence on the added electrolyte (NaCl) concentration. The

equilibrium between the protonated (charged) and deprotonated (neutral) forms of MKM is

proposed to explain the anomalous binding isotherms observed in the presence of the

polyelectrolyte. The explanation is supported by theoretical model calculations of the mixed-

micelle equilibrium and the competitive binding of the two MKM forms to the surface of the

electrode membrane. It is suggested that the presence of even small amounts of the deprotonated

form can strongly influence the measured electrode response. Such ionic-nonionic surfactant

mixtures are a special case of mixed surfactant systems where the amount of the nonionic

component cannot be varied independently as is the case with some of the earlier studies.

2

17

18

19

20

21

22

23

24

25

26

27

28

29

30

31

32

33

34

35

34

Page 3: digital.csic.esdigital.csic.es/bitstream/10261/131661/1/2015_Langmuir.docx · Web viewAll aqueous NaHA solutions for potentiometric and fluorescence measurements were prepared by

1. Introduction

Aqueous mixtures of ionic surfactants and polyelectrolytes of opposite charge are of great

scientific and practical interest. They are used in a variety of technological fields in the form of

different advanced products, e.g. as household products (detergents, paints, and cosmetics),

coatings, wastewater treatment agents, pharmaceutical products, and many others.1,2 Their great

potential for the medical field use has been reported in the literature,3,4,5 e.g. in controlled drug

release or in direct treatment of acute inner organ injuries.3,4 In particular, natural and

biodegradable surfactants and ionic polysaccharides have attracted a lot of interest in recent

years.6,7,8,9 They prosper as alternative materials for medical applications where wound healing

properties and antimicrobial activity is needed, since a wide variety of common antimicrobial

chemical agents used in the medical field until now, such as metals and their salts, 10,11

iodophors,12 phenols and thiophenols,13 and antibiotics,14 is toxic to humans and harmful to the

environment.

When preparing aqueous mixtures for applications, special attention has to be paid to the

interactions between ionic surfactants and polyelectrolytes of opposite charge, since these can

alter the unique properties of each of the components in the mixture. Understanding of these

interactions is therefore crucial for optimizing the performance of the polyion-surfactant ion

mixtures. Within this frame, the purpose of this work is to perform a fundamental study of

interactions of a cationic surfactant derived from lysine with a completely ionized hyaluronic acid

(HA) in aqueous solutions. HA is a biocompatible and biodegradable polysaccharide consisting

of repeating β 1-4 D-glucuronic acid and β 1-3 N-acetyl-D-glucosamine. Due to its

biocompatibility and also to its excellent wound healing abilities, it is used in various

applications, the most notable being wound dressings, cosmetics, and pharmaceutical

products.15,16,17 The lysine-based cationic surfactant used in this work is an Nε-myristoyl-lysine

3

36

37

38

39

40

41

42

43

44

45

46

47

48

49

50

51

52

53

54

55

56

57

58

59

56

Page 4: digital.csic.esdigital.csic.es/bitstream/10261/131661/1/2015_Langmuir.docx · Web viewAll aqueous NaHA solutions for potentiometric and fluorescence measurements were prepared by

methyl ester, abbreviated as MKM.18 It is a biodegradable amino-acid based surfactant exhibiting

excellent antimicrobial properties and low toxicity.18 Amino acid based surfactants constitute an

important class of natural surface-active molecules and are of great interest in organic and

physical chemistry as well as in biology and medicine. Their physicochemical properties

resemble those of conventional cationic surfactants, but they are more environmentally friendly

and non-toxic to eyes and skin.6

The critical micelle concentration (CMC) and the critical association concentration (CAC)

values of MKM in pure solutions and in the presence of hyaluronate anion, respectively, were

determined by potentiometric titrations using the surfactant sensitive electrode (SSE) and by

fluorescence measurements using pyrene as a probe. All measurements were performed in

aqueous solutions without and with added NaCl. From the SSE data, binding isotherms where

constructed in order to study the binding behavior. The results are interpreted by taking into

account the acid-base equilibrium between the protonated and deprotonated forms of MKM. The

importance of this special feature of the studied system is highlighted by theoretical model

calculations of the micellar equilibrium and of the SSE signal in solutions in the absence and

presence of the polyelectrolyte.

2. Materials and methods

2.1. Materials

The sodium salt of HA, NaHA (molecular weight: Mw = 0.6-1.1 MDa; monomer unit molecular

weight: Mm = 401.3 g/mol; for structure see Scheme S1 in Supporting Information, SI) was

purchased from Lex, Slovenia. All aqueous NaHA solutions for potentiometric and fluorescence

measurements were prepared by suspending NaHA in triple-distilled water (or in aqueous NaCl

solution) and stirring for 30 min at ambient temperature in order to obtain clear solutions.

4

60

61

62

63

64

65

66

67

68

69

70

71

72

73

74

75

76

77

78

79

80

81

82

78

Page 5: digital.csic.esdigital.csic.es/bitstream/10261/131661/1/2015_Langmuir.docx · Web viewAll aqueous NaHA solutions for potentiometric and fluorescence measurements were prepared by

MKM (M = 407.0 g/mol; for structure see Scheme S1, SI) was synthesized as reported in

reference 18, where some biological and chemical properties of cationic surfactants derived from

lysine were also described. All aqueous MKM solutions were prepared by suspending MKM in

triple-distilled water (or in aqueous NaCl solution) and stirring for 15 min at around 40 °C. The

heating step is necessary because the dissolution of MKM in water at lower temperatures is slow

and its solubility moderate.

In view of the biodegradability of MKM,18 the stability of the surfactant in water was checked

by high pressure liquid chromatography (HPLC) and by nuclear magnetic resonance (NMR)

measurement (for instrumentation details see SI). 16 mg of MKM were dissolved in water and

stirred, first for 2 hours at 45°C and then for 18 hours at room temperature. Aliquots of 0.5 mL of

this solution were taken at the beginning (i.e. after 0 hours) and after 2, 4, and 20 hours. 4 HPLC

chromatograms were recorded at each time interval (t). These chromatograms were all identical;

therefore, only one is shown for the initial (t = 0 h) and final time (t = 20 h; c.f. Figure S1 in SI).

In Table S1, the height and area of the main peak observed at a retention time of around 13.5

minutes is reported at each t. After this treatment, the samples were lyophilized in order to

remove water, the obtained solid MKM was dissolved in deuterated methanol and an 1H NMR

spectrum was recorded. The 1H NMR spectra of MKM taken before (t = 0 h) and after (t = 20 h)

the described treatment in water are shown in Figure S2 in SI. It can be seen that also the NMR

spectra are identical. The HPLC and 1H NMR results thus show that no chemical degradation of

MKM is taking place in water at room temperature within the period of 20 hours. This time is

more than sufficient to perform the herein reported measurements without fear of MKM

degradation. Nevertheless, each measurement series was performed with freshly prepared MKM

solutions and immediately after their preparation.

2.2. Methods

5

83

84

85

86

87

88

89

90

91

92

93

94

95

96

97

98

99

100

101

102

103

104

105

106

910

Page 6: digital.csic.esdigital.csic.es/bitstream/10261/131661/1/2015_Langmuir.docx · Web viewAll aqueous NaHA solutions for potentiometric and fluorescence measurements were prepared by

SSE-potentiometric titrations. The potentiometric titrations using the SSE were used to

determine the CMC values of MKM and the binding isotherms of MKM binding by NaHA, both

as a function of NaCl concentration. The SSE was prepared following the procedure described in

the literature.19 The active part of the electrode is a poly(vinyl chloride) (PVC) membrane

containing so called carrier complex (CC) that is formed between a cationic and an anionic

surfactant in aqueous solutions. For CC preparation, a 0.4 M solution of sodium dodecyl sulphate

(SDS) was slowly added to a 0.005 M solution of MKM under continuous stirring until a 1:1

molar ratio between the surfactants was achieved. This resulted in precipitation of an insoluble

CC from water. The resulting white precipitate was repeatedly washed with water in order to

remove NaCl (the by-product of the reaction between MKM and SDS) and vacuum dried at 50-

60 °C. A solid membrane with the composition 23 wt. % of PVC, 76 wt. % of dioctyl phthalate

(DOP), and 1 wt. % of CC was then prepared in the following way: CC was dissolved in 5 mL of

tetrahydrofuran (THF) by heating, followed by a sequential addition of PVC and DOP. The clear

viscous THF solution was cast into a petri dish and the solvent was left to evaporate overnight at

room temperature, producing a thin solid membrane of approximately 1 mm thickness. A piece of

the PVC membrane was glued to the bottom of a hard 1 cm diameter PVC tube by using a dense

THF solution of PVC as an adhesive. The PVC tube was filled with the reference solution (1

104 M MKM in 0.01 M NaCl) and an Ag/AgCl electrode was inserted in order to provide

electrical contact. The saturated calomel electrode (SCE) was used as the reference electrode.

After inserting the electrodes into the solution, the potential difference (E) between the SSE and

SCE was recorded after each addition of the MKM titrant solution into a given volume of the

solvent using the pH meter MA 5740, Iskra. The stability criterion for taking a reading after each

addition was dE/dt = 0.1 mV/30 s. The solution was continuously stirred with a magnetic stirrer

and a constant temperature of 25 °C was maintained during the titration. The response of the SSE

6

107

108

109

110

111

112

113

114

115

116

117

118

119

120

121

122

123

124

125

126

127

128

129

130

1112

Page 7: digital.csic.esdigital.csic.es/bitstream/10261/131661/1/2015_Langmuir.docx · Web viewAll aqueous NaHA solutions for potentiometric and fluorescence measurements were prepared by

to surfactant ion concentration in water without added NaCl was linear in a wide MKM

concentration range (from around 5.0 x 10-6 to 4.9 x 10-4 M, the CMC value of MKM in water)

with a slope of 56 mV/decade, which is close to the theoretical value at 25 °C (59.2 mV/decade).

For the CMC determination, a 5 x 10-3 M MKM solution was added stepwise to 10 mL of water

or aqueous NaCl solution with a micro burette. The titrations were performed at the following

salt (NaCl) concentrations (csalt): csalt = 0, 0.01, 0.05, and 0.1 M.

Titrations for the binding studies were performed in the same manner as described above.

Instead of water or aqueous NaCl, a 5 x 10-4 M NaHA solution in the selected solvent was used as

the initial solution and a MKM solution in the same solvent together with the same volume of a 1

10-3 M NaHA solution, in order to keep the NaHA concentration in the cell constant during the

measurement, were added in a stepwise manner. The stability criterion for taking a reading after

each addition was dE/dt = 0.1 mV/180 s.

The apparent degree of MKM binding by NaHA (βapp) was calculated as βapp = (cst cs

f)/cp,

where cst and cs

f are the total and free surfactant (MKM) concentrations, respectively, which are

obtained from potentiometric curves20,21 measured in the absence and presence of NaHA, and cp is

the concentration of the polyion (NaHA) expressed in moles of monomer units per volume.

Fluorescence measurements. Fluorescence measurements were performed to determine the

surfactant’s CMC and its CAC in the presence of NaHA. Pyrene (Aldrich, optical grade) was

used as the external fluorescence probe to monitor the formation of MKM micelles. The pyrene

saturated aqueous solutions, without or with NaCl, were prepared as described previously22 and

were used for the preparation of the MKM and NaHA solutions. The emission spectra of pyrene

were recorded in a 1 cm quartz cuvette at a constant temperature of 25°C by using a

luminescence spectrometer LS 50 from Perkin-Elmer. The wavelength of the excitation

electromagnetic radiation was 330 nm. All spectra were recorded in the wavelength region 350-

7

131

132

133

134

135

136

137

138

139

140

141

142

143

144

145

146

147

148

149

150

151

152

153

154

1314

Page 8: digital.csic.esdigital.csic.es/bitstream/10261/131661/1/2015_Langmuir.docx · Web viewAll aqueous NaHA solutions for potentiometric and fluorescence measurements were prepared by

450 nm. The excitation and emission slit widths were 2.5 mm and the scan rate was 200 nm/min.

The stock MKM solution (c = 5 x 10-3 M) was added in small volume increments into a pyrene

saturated solution in the cuvette. The emission spectra were recorded after each addition until the

CMC (or CAC) of MKM was reached and exceeded. Nine scans were accumulated for each run

and spectra were fully corrected before the ratio of the fluorescence intensities of the first and

third vibronic bands, I1/I3, was calculated. The peaks appeared at approximately 373 and 384 nm.

All measurements were performed in water and in 0.01, 0.05, and 0.1 M NaCl.

2.3. Theory

2.3.1 Model of MKM self-assembly

Polyelectrolyte-free solutions of MKM are modelled using the Poisson-Boltzmann (PB)

approach.23,24 The protonated (S+) and deprotonated (S0) forms of MKM are considered to form

mixed micelles of a single aggregation number N (N=N +¿+N0¿, where N+¿¿ and N 0 are the

numbers of S+ and S0 per micelle, respectively). The chemical potential of surfactant species i in

micelles (μimic) is:23

μimic=μi

0 , mic+μimix+μi

surf+μiel. (1)

where μi0 , mic is the standard chemical potential of a hydrocarbon tail in the micellar core. The term

μimix is due to the entropy of mixing, which has two contributions: RTln xi from mixing S+ and S0

in the micelles and RTN

ln cmic from the translational entropy of the micelles. Here, xi is the mole

fraction of i in the micelle, R is the ideal gas constant and T is the absolute temperature. The term

μisurf (¿ai γ i, where a i is the area in the surface region of the micelle and γi is a proportionality

constant with units of surface tension) comes from the contact energy between hydrocarbon and

8

155

156

157

158

159

160

161

162

163

164

165

166

167

168

169

170

171

172

173

174

175

1516

Page 9: digital.csic.esdigital.csic.es/bitstream/10261/131661/1/2015_Langmuir.docx · Web viewAll aqueous NaHA solutions for potentiometric and fluorescence measurements were prepared by

water at the micelle surface and μiel is the contribution from the electrostatic free energy of

forming a micelle. For surfactant monomers in the aqueous region we have:

μiw=μ i

0 ,w+RTln c i (2)

with μi0 , w being the standard chemical potential of the surfactant monomer in water and c i its

concentration. By taking into account the conservation of mass and introducing an acid constant

Ka relating the protonated and deprotonated surfactant forms to the concentration of hydrogen

ions, the equilibrium of the system is determined by the following set of equations:

c i=x i cmic

1N e(∆ μ0+ μi

surf+ μiel )/RT (i = +, 0) (3)

cSt =N cmic+c+¿+c0¿ (4)

¿ (5)

Ka=c0¿¿ (6)

In eq. 3,∆ μ0 is equal to the difference μi0 , m−μi

0 , w. Eq. (5) is valid when dissociation of MKM is

the only source of hydrogen ions in the system. For the dilute micellar solutions considered here,

the activity of a dispersed ion is set equal to its concentration.

To model the self-assembly of MKM in polyelectrolyte solutions, the polyion-dressed micelle

(PDM) approach developed earlier was used.25,26,27 PDM is a complex between a surfactant

micelle and (part of) a polyion chain. The electrostatic free energy of a dilute solution is divided

into separate contributions from micelle-free regions (containing polyelectrolyte) and regions

containing PDMs. The latter, in turn, consists of an external free energy due to the net charge of

the complex and an internal free energy due to the interaction between the bare micelle and the

polyion charges in the complex. The external part is calculated using the PB model with the PDM

described as a uniformly charged sphere. Expressions for the contribution to the chemical

9

176

177

178

179

180

181

182

183

184

185

186

187

188

189

190

191

192

193

194

195

196

197

1718

Page 10: digital.csic.esdigital.csic.es/bitstream/10261/131661/1/2015_Langmuir.docx · Web viewAll aqueous NaHA solutions for potentiometric and fluorescence measurements were prepared by

potential (μiel ,ext) are the standard ones within the PB theory.23,24 The internal contributions to the

chemical potential of S+ and S0 in PDM can be written as25,26,27 (for details see SI, section S2):

μ iel ,∫¿

k BT=N

x+¿ lBh

4 r ¿¿¿¿¿ (i = +, 0) (7a)

g+¿=¿¿ (7b)

g0=−x+¿¿¿ (7c)

where k B is the Boltzmann constant, ε 0 εr is the permittivity of the medium, h is a model

parameter related to distribution of charges on the micelle (see below and SI, section S2), and lB

is the Bjerrum length: lB=e2

4 π kB T ε0 εr. The total chemical potential of surfactant i in PDM is:

μimic=μi

0 , mic+μimix+μi

surf+μiel ,ext+μi

el ,∫¿+f ∆ μ p¿. (8)

The first three terms on the right-hand side of eq. (8) have the same meaning as in eq. (1) and the

last term corresponds to the PB-model free energy change ∆ μp of transferring f polyion charged

segments per micelle charge from the free state to the micelle surface. The equilibrium

composition is obtained by combining eqs. (4) – (6) with the following equation:

c i=x i cmic

1N e¿ ¿ (i = +, 0) (9)

As a result, c+¿¿ and c0 are obtained as functions of cSt from which the degree of surfactant

binding to the polyelectrolyte (β), is calculated:

β=cSt −

c+¿−c0

c p¿ (10)

For comparison with experimental data an apparent degree of binding (app) is defined as:

βapp=cSt −

c+¿app

c p¿ (11)

10

198

199

200

201

202

203

204

205

206

207

208

209

210

211

212

213

214

215

216

1920

Page 11: digital.csic.esdigital.csic.es/bitstream/10261/131661/1/2015_Langmuir.docx · Web viewAll aqueous NaHA solutions for potentiometric and fluorescence measurements were prepared by

where c+¿app¿ is the surfactant concentration in a polyelectrolyte-free solution (below the CMC)

giving the same electrode signal (E) as that in the polyelectrolyte solution (see below).

2.3.2 Model of the SSE response

To calculate theoretically the SSE response, we assume that the main contribution to the

measured signal comes from the charge transfer between the solutions on each side of the

electrode membrane. In equilibrium models, the interference of an ion j of the same charge sign

and valence as the potential determining ion i is often handled by means of a selectivity

coefficient Kj.28 This can be interpreted as the equilibrium constant of the ion exchange reaction

in which i is replaced by j in the membrane. Under the condition that the total concentration of i

and j in the membrane is constant one obtains the familiar Nikolskii-equation for the measured

electromotive force:29

E=constant+ R TF

ln (ci+K j c j) (12)

with F being the Faraday constant. Eq. (12) is applicable to monovalent ions when the

composition of the solution on the reference side of the membrane is kept fixed. The form of the

equation remains the same when the overall membrane potential is assumed to be the sum of the

phase boundary potential and a membrane diffusion potential under steady state, but K j then

equals the product of the ion exchange constant and the ratio of the ion mobilities in the

membrane.29 The uncharged form of MKM (S0) does not contribute to the electric potential itself

but, when present in the membrane, it modulates the signal by affecting the chemical potential of

S+. Because of its dominating hydrophobic groups the concentration in the membrane is expected

to be quite high even at low concentrations in the solution. In principle, it can reduce the activity

of S+ in the membrane either by expelling it (competitive binding) or by lowering its activity

coefficient without replacing it. It may also affect the dynamics of the charge transfer process,

11

217

218

219

220

221

222

223

224

225

226

227

228

229

230

231

232

233

234

235

236

237

238

239

2122

Page 12: digital.csic.esdigital.csic.es/bitstream/10261/131661/1/2015_Langmuir.docx · Web viewAll aqueous NaHA solutions for potentiometric and fluorescence measurements were prepared by

e.g. by changing the mobility of S+ in the membrane. To the best of our knowledge there are no

reports in the literature of such types of interference by non-ionic surfactants. Since we are not in

a position to identifying the dominating potential generating mechanism we will assume that the

effect is captured by a generalized form of eq. (12):

E=E¿+ RTF

ln ¿ (13)

E¿ is a constant, c Na is the concentration of sodium ions in the solution, and K Na and K0 are

‘interference’ coefficients for sodium ions and S0, respectively. It will be demonstrated that eq.

(13) in combination with the above model of surfactant self-assembly captures the main

qualitative features observed experimentally. In SI (section S3) we describe the competitive

binding mechanism reproducing the form of eq. (13) to a constant near.

2.3.3 Model calculations

The non-linear PB equation is solved numerically by means of a computer program (PB cell)

written by Bengt Jönsson. The program also calculates the electrostatic contribution to the

chemical potentials in the thermodynamic model by Jönsson and Wennerström.24 The PDM is

modelled as a uniformly charged sphere with a radius 17 Å carrying N+¿ (1−f ) ¿ net charges. Instead

of attempting to determine the optimal value from the free energy expression we simply set f=1

for β+¿≤ 1¿, and f=1

β+¿¿ for β+¿>1 ¿, with β+¿¿ being the number of bound S+ per polyion charge.

The fully dissociated HA is modelled as a uniformly charged cylinder with a radius of 5 Å and a

linear charge density corresponding to one negative charge per 10 Å. Both, micelles and polyions

are in contact with infinitely large solution of monovalent salt at T = 298 K; the dielectric

constant (ε r) of the medium is set to ε r=78.5. When evaluating eqs. (3) and (9), we set

Δ μ0=−13.5 RT/mole (molarity scale), a typical value for a C12 surfactant,30 γ+¿=0.018¿ J/m2,24 and

12

240

241

242

243

244

245

246

247

248

249

250

251

252

253

254

255

256

257

258

259

260

261

2324

Page 13: digital.csic.esdigital.csic.es/bitstream/10261/131661/1/2015_Langmuir.docx · Web viewAll aqueous NaHA solutions for potentiometric and fluorescence measurements were prepared by

γ0=0.010 J/m2 (see the previous models of mixed aggregates of charged an uncharged

amphiphiles24). The dominating contribution to the free energy change of transferring polyion

chain segments to the micelle surface is considered to come from the entropy gain in releasing

the counterions. Hence, we put ∆ μp=−f μp−TS, where μp

−TS is the entropic part of the electrostatic

chemical potential of a polyion (per charge). PB calculations provide the following values of μp−TS

(in units of RT/mole): 0.55, 0.40, 0.31, and 0.26, in solutions containing 0.005, 0.01, 0.05 and 0.1

M salt respectively. In all calculations, we use pKa = 7.8 (suitable for MKM in the unimer form18),

N=56, and h=2.5. The latter is obtained from an optimization of the free energy of PDM for a

C12 surfactant by means of a procedure described in SI (section S2). In the calculation of the SSE

response, we use K Na=0.001 and K0=500. The ion concentrations in eq. (13) are converted to

activities by multiplying with activity coefficients (¿10−0.5√C salt /(1+√Csalt)). The correction has only a

marginal effect on the result.

3. Results and discussion

The synthesis and basic characterization of MKM was described recently.18 However, its

micellization process was followed only in aqueous solution without added salt by measuring the

self-diffusion coefficients at 25°C by NMR. It is well known that CMC of ionic surfactants

depends considerably on the added salt concentration. Therefore, we first report CMC values of

MKM in dependence on NaCl concentration (Section 3.1). Section 3.2 is devoted to the

discussion of experimental binding isotherms in NaHA solutions at various csalt values and

Section 3.3 to their model treatment.

3.1. Determination of CMC values

The response of the SSE electrode to the total concentration of MKM (i.e. the E vs. log cst

curve) in the absence of NaHA is presented in Figure 1 for all investigated csalt. The dependencies

13

262

263

264

265

266

267

268

269

270

271

272

273

274

275

276

277

278

279

280

281

282

283

284

2526

Page 14: digital.csic.esdigital.csic.es/bitstream/10261/131661/1/2015_Langmuir.docx · Web viewAll aqueous NaHA solutions for potentiometric and fluorescence measurements were prepared by

Figure 1. Plots of the electromotive force (E) as a function of the total MKM concentration (cst)

in aqueous solutions without and with added NaCl. Slopes of the linear parts of the curves below

the CMC are reported in the Figure. To highlight the difference between the curves those for

0.01, 0.05 and 0.10 M NaCl have been shifted vertically to start at the same value as the salt free

curve. The original curves are presented in SI (Figure S9).

are linear at low surfactant concentrations and display a break at a well-defined concentration as

expected for surfactant solutions below and above the CMC. The determined CMC values, taken

as the concentration at the break point, are reported in Table S2 (SI, section S4.1) and plotted in

Figure 2 as a function of the ionic strength I (= csalt + CMC) at the CMC. For the moment they

should be considered as apparent CMC values. The reason for this is that the electrode response

curves contain anomalies.

First of all, the slope below the break point is smaller than the theoretical value at 25°C (

dE /dlogcSt = 59 mM/decade) and depends on csalt (see values reported in Figure 1). This is not a

behavior expected for ionic surfactants. Normally, the addition of salt lowers the CMC whereas

14

285

286

287

288

289

290

291

292

293

294

295

296

297

298

299

300

2728

Page 15: digital.csic.esdigital.csic.es/bitstream/10261/131661/1/2015_Langmuir.docx · Web viewAll aqueous NaHA solutions for potentiometric and fluorescence measurements were prepared by

the slope of the E vs. log cst line below the CMC is not affected by csalt. Secondly, the slope above

the break point is negative and becomes more negative the higher the csalt. While the single-ion

activity of ionic surfactants is known to decrease above the CMC, an observation largely

attributable to a decreasing entropic penalty of binding counterions at a higher total surfactant

concentration, the effect is known to disappear when salt is added. In particular, the effect should

be absent at 0.1 M NaCl, where in contrast the experimental slope of the electrode response curve

above the CMC is the most negative. Because of the importance also for the construction of

surfactant binding isotherms in polyelectrolyte solutions, it is necessary to examine in detail the

origin of this ʻanomalousʼ behavior to see in what ways it affects the results. Therefore, the

potentiometric CMC values were compared with those obtained by the pyrene fluorescence

method. The formation of hydrophobic domains (such as micelles) is indicated by the change in

the pyrene polarity ratio, I1/I3, which is obtained from pyrene emission spectra22 (see examples of

such spectra reported in Figure S7 in SI). Plots of the I1/I3 ratio as a function of the MKM

concentration are shown in Figure 3 for all salt concentrations. The I1/I3 value is high below the

CMC (I1/I3 = 1.55-1.65) indicating a highly polar microenvironment of pyrene, it starts to

decrease as the micelles begin to form and finally stabilizes at a considerably lower value (I1/I3

1.1-1.2). The I1/I3 ratio around 1.2 indicates solubilization of pyrene in the hydrophobic interior

of surfactant micelles. The CMC values were determined at the point where the I1/I3 values

started to decrease and are plotted in Figure 2 together with the CMC data from potentiometric

measurements. It can be seen that potentiometric (SSE) and fluorimetric data are in good

agreement (see also the CMC values reported in Table S2 in SI). This indicates that the break

point in the SSE curves indeed coincides with the appearance of the hydrophobic domains in

solution. It is therefore interesting to point out that the CMC value in water (CMC = 4.9 10-4 M

by both methods) is approximately 3-times lower than the one determined previously by NMR

15

301

302

303

304

305

306

307

308

309

310

311

312

313

314

315

316

317

318

319

320

321

322

323

324

2930

Page 16: digital.csic.esdigital.csic.es/bitstream/10261/131661/1/2015_Langmuir.docx · Web viewAll aqueous NaHA solutions for potentiometric and fluorescence measurements were prepared by

Figure 2. Upper panel: CMC values of MKM in aqueous NaCl solutions at 25°C in dependence

on the logarithm of the ionic strength, I (= csalt + CMC), as determined by the SSE potentiometric

titrations (full triangles) and by pyrene fluorescence measurements (open triangles). The straight

line is the best fit to the experimental data and can be described by the equation

CMC=−3.28−1.25 log I (see text). Lower panel: CAC values of MKM in aqueous NaCl

solutions at 25°C in dependence on I as determined by the SSE potentiometric titrations.

measurements (CMC = 1.6 10-3 M18) and 4-6-times lower than the one obtained by

conductivity measurements (CMC = 2-3 10-3 M6,31). In reference 31, a considerably lower

CMC (= 2.3 10-5 M) in comparison with these values was determined by surface tension

measurements with the Wilhelmy plate. This finding was explained therein as a consequence of

16

325

326

327

328

329

330

331

332

333

334

335

336

3132

Page 17: digital.csic.esdigital.csic.es/bitstream/10261/131661/1/2015_Langmuir.docx · Web viewAll aqueous NaHA solutions for potentiometric and fluorescence measurements were prepared by

the multilayer adsorption of the surfactant’s amine groups on the platinum surface of the plate,

which supposedly affected the wetting properties of the plate and hence the measured surface

tension. However, the reasons for the above discrepancies in CMC values may also be that the

bulk methods (like NMR and conductivity) are less sensitive than the fluorescence and SSE

potentiometric ones for detecting the formation of hydrophobic domains at low surfactant

concentrations. NMR measures the self-diffusion coefficients of species in solution and low

solute concentrations (around 1 10-3 M) may present some limitations for the reliability of the

NMR determination. Another interpretation is that the values determined by the present methods

(and those obtained previously by surface tension31) are influenced by a small fraction of MKM

monomer present in the deprotonated form even at a neutral pH (note that pKa value of the

monomer MKM is 7.818). Thus, mixed micelles of the protonated (positively charged, S+) and

deprotonated (uncharged, S0) forms may exist in solution at concentrations low enough not to

Figure 3. The I1/I3 fluorescence ratio of pyrene in aqueous MKM solutions in the absence (water)

and in the presence of NaCl.

17

337

338

339

340

341

342

343

344

345

346

347

348

349

350

351

352

3334

Page 18: digital.csic.esdigital.csic.es/bitstream/10261/131661/1/2015_Langmuir.docx · Web viewAll aqueous NaHA solutions for potentiometric and fluorescence measurements were prepared by

markedly influence the concentration of the free MKM molecules, and therefore do not affect the

measured NMR self-diffusion coefficient or conductivity (note that the uncharged MKM has a

negligible effect on solution’s conductivity). Nevertheless, even at such low concentrations they

may very well be detectable by the pyrene method, which is very sensitive to the presence of

hydrophobic domains. It will be demonstrated below that the break point in the SSE response

curve should in fact correspond to the onset of mixed micelle formation of the two forms of

MKM, S+ and S0. Further support for this interpretation is provided by the plateau values of I1/I3

below the CMC. Usually, the I1/I3 values below the CMC increases with increasing salt

concentration because of a more polar environment when a larger number of ions from NaCl is

present in solution.22 The trend is just the opposite in Figure 3: the lowest I1/I3 ( 1.56) value is

detected in 0.1 M NaCl. This suggests that the concentration of S0 increases with increasing csalt.

The above results are in excellent agreement with recent studies of Fegyver et al.32,33 on the

impact of the nonionic surfactant additive on association between polyelectrolytes and oppositely

charged surfactants. Those studies revealed formation of mixed ionic-nonionic surfactant

micelles in polyelectrolyte-free solutions at very low surfactant concentrations by using pyrene

fluorescence spectroscopy, whereas the conductivity measurements detected micelles at orders of

magnitude higher surfactant concentrations. An important difference between the studies of

Fegyver et al.32,33 and ours is that the nonionic surfactant with a different chemical nature

compared to the ionic surfactant was used therein. The uncharged component could thus be added

to the mixed surfactant solution independently, for example so that its concentration was kept

constant. On the contrary, the concentration of nonionic surfactant (S0) is determined by the

inherent S+-S0 equilibrium in these solutions in the MKM case.

Let’s turn back to the ʻanomalousʼ behavior of MKM. Another indication of MKM’s

anomalous self-assembly comes from the variation of the apparent CMC with csalt (Figure 2,

18

353

354

355

356

357

358

359

360

361

362

363

364

365

366

367

368

369

370

371

372

373

374

375

376

3536

Page 19: digital.csic.esdigital.csic.es/bitstream/10261/131661/1/2015_Langmuir.docx · Web viewAll aqueous NaHA solutions for potentiometric and fluorescence measurements were prepared by

upper panel). The observation that CMC decreases with increasing csalt is a phenomenon typical

for ionic surfactants (see also Figure S8 in SI, where data for a conventional cationic and anionic

surfactant, cetylpyridinium chloride (CPC)34 and SDS,35 respectively, are included). The presence

of an electrolyte reduces the net repulsion between charged head groups of surfactant molecules

in the micelle and consequently micellization starts at a lower surfactant concentration.30

However, the trend indicated by the solid line in Figure 2 (upper panel) is not typical. This line is

described by an equation of the form CMC=−3.28−1.25logI , whereas for conventional ionic

surfactants like CPC or SDS the linear dependence is normally obtained on a double logarithmic

plot36,37 and is described by an equation of the form logCMC=−a−blogI (where constants a and

b have positive values). This is clearly demonstrated for CPC and SDS in Figure S8 in SI (lower

panel). As one can appreciate from Figure S8, the apparent CMC values of MKM do not fit such

functional relationship. Rather, they decrease less steeply with increasing I in comparison with

conventional surfactants, which may indicate that the micellar equilibrium is influenced by the

acid-base equilibrium.

Returning to the anomalous electrode responses displayed in Figure 1 that might also be

accounted for by proposing formation of MKM pre-micelles (dimers, trimers, etc., in equilibrium

with free monomers). It is easy to see that the formation of MKM pre-micelles would lower the

activity of the surfactant and reduce the slope of the curves below CMC. However, this cannot

account for the negative slopes above the CMC. As already pointed out the observed increase of

the effect with increasing csalt is not consistent with micellar equilibrium of regular cationic

surfactants. On the other hand it is possible to account for it if a fraction of MKM is deprotonated

but still interacts with the surfactant electrode membrane.

3.2. Binding isotherms

19

377

378

379

380

381

382

383

384

385

386

387

388

389

390

391

392

393

394

395

396

397

398

399

3738

Page 20: digital.csic.esdigital.csic.es/bitstream/10261/131661/1/2015_Langmuir.docx · Web viewAll aqueous NaHA solutions for potentiometric and fluorescence measurements were prepared by

The degree of binding of MKM to NaHA was determined by the use of SSE. The SSE

potentiometric titration curves (E vs. log cs) obtained in pure MKM solution and in the presence

of NaHA are presented in Figure 4 for the case of 0.01 M NaCl. Other curves are shown in Figure

S9 in SI. The curve obtained in the presence of NaHA almost fits the calibration line

(measurement in pure MKM) for cs < 6 10-5 mol/L and starts to deviate from it at higher

surfactant concentrations due to MKM binding by NaHA. In the region where binding is strong

and cooperative the free surfactant concentration in mixed solution with the polyelectrolyte

usually almost does not change. Because the potential of the SSE depends on the free (monomer)

MKM concentration in solution, also the value of the measured E should be more or less

constant. In the MKM case, however, E in this region clearly decreases with increasing MKM

Figure 4. SSE potentiometric titration curves for pure MKM and for MKM in the presence of

NaHA (cp = 5 10-4 mol/L) in 0.01 M NaCl; the values of Δcs and csf, needed for the calculation

of the degree of binding, are indicated (see text).

20

400

401

402

403

404

405

406

407

408

409

410

411

412

413

414

3940

Page 21: digital.csic.esdigital.csic.es/bitstream/10261/131661/1/2015_Langmuir.docx · Web viewAll aqueous NaHA solutions for potentiometric and fluorescence measurements were prepared by

concentration (c.f. the negative slope of the curve in the range 6 10-5 M < cst < 5 10-4 M).

This result suggests that the free MKM concentration even decreases in the region of cooperative

binding, which ultimately leads to negative slopes of the experimental surfactant binding

isotherms (see Figure 5). This is generally not observed in surfactant/polyelectrolyte mixtures.

More extended discussion related to this observation is given below.

Figure 5. Binding isotherms (degree of binding app as a function of csf) for MKM binding by

NaHA (cp = 5 10-4 mol/L) in water (open circles) and in solutions with added NaCl: csalt = 0.01

(diamonds), 0.05 (squares), and 0.1 mol/L (triangles). The arrows indicate CMC values of MKM

determined by potentiometry: 1 - water, 2 - 0.01 M NaCl, 3 - 0.05 M NaCl, and 4 - 0.1 M NaCl.

Normally, the difference between the curves in Figure 4 should give the information on the

amount of MKM bound by the polyion (designated as cs) and enable the calculation of app (=

(cst - cs

f)/cp = cs/cp) and construction of binding isotherms. Apparent binding isotherms obtained

in this way are shown in Figure 5 for all salt concentrations. Some features are characteristic for

cooperative binding of surfactants by polymers.20 Binding is negligible (app 0) at MKM

21

415

416

417

418

419

420

421

422

423

424

425

426

427

428

429

430

4142

Page 22: digital.csic.esdigital.csic.es/bitstream/10261/131661/1/2015_Langmuir.docx · Web viewAll aqueous NaHA solutions for potentiometric and fluorescence measurements were prepared by

concentrations lower than a certain value: for example, in water app is virtually zero for csf below

4 10-5 mol/L. At higher surfactant concentrations, app increases rapidly (i.e. csf does not change

appreciably), indicating strongly cooperative interaction, and finally levels of at some point

indicating that cooperative binding to the polyelectrolyte is no longer possible.

The free surfactant concentration at the point of the steep increase in app is referred to as the

critical association concentration (CAC) or ‘CMC’ in the presence of the polyion. The CAC

values determined from the isotherms in Figure 5 are reported in Table S2 in SI and plotted in

Figure 2 versus I. They increase with increasing I, contrary to CMC values, a finding related to

the screening effect of simple electrolyte on attractive interactions between polyions and

oppositely charged surfactant cations. Strong interactions are indicated by a pronounced

reduction of CAC in comparison with the CMC values (CMC values are indicated by arrows in

Figure 5). In the NaHA-MKM system, the largest lowering of CAC (around 1 order of

magnitude) is observed in solutions with no added NaCl.

At the highest NaCl concentration (csalt = 0.1 mol/L), no levelling off of the isotherm is

indicated; rather, app continuously increases to values above 1. The reason for this observation is

the coincidence of the CMC and the CAC values of MKM in 0.1 M NaCl (CMCCAC = 1 × 10-4

mol/L; c.f. Table S2 and arrow 4 in Figure 5). The free micelle formation starts instantly after (or

simultaneously with) the saturation of the polyion by MKM, with SSE registering both processes.

The slopes of the experimental binding isotherms in Figure 5 deserve special attention due to

their pronounced negative value in the region of cooperative binding, which is not typical for the

binding of conventional surfactants by polymers. The “back swing” of the isotherms even

increases with increasing salt concentration and seems to indicate that the free MKM

concentration decreases as the cooperative binding of MKM to NaHA is taking place.

22

431

432

433

434

435

436

437

438

439

440

441

442

443

444

445

446

447

448

449

450

451

452

453

4344

Page 23: digital.csic.esdigital.csic.es/bitstream/10261/131661/1/2015_Langmuir.docx · Web viewAll aqueous NaHA solutions for potentiometric and fluorescence measurements were prepared by

Possible explanations for negative slopes in the NaHA-MKM system could be the following:

(1) the presence of a non-ionic surfactant (with different chemical structure as MKM) that is not

detected by the SSE but takes part in the aggregate formation at the polyion together with MKM;

this option was excluded based on the chemical analysis of the solid MKM; (2) polyion

properties (its charge density) change significantly during the binding process. In order to get an

increasing negative slope, the charge density should increase upon surfactant binding, which

seems highly unrealistic in the case of cationic surfactant binding to an anionic polyelectrolyte;

(3) some other equilibrium process, such as the above mentioned acid-base equilibrium plays an

important role. In the following, we present results of model calculations favoring the latter

explanation.

3.3. Results of theoretical modelling

To investigate theoretically the importance of the equilibrium between the protonated (S+) and

deprotonated (S0) MKM on the results we used the model described in Section 2.3 to calculate

the concentrations of free surfactant monomers (C+¿¿ andC0¿ in the solution in equilibrium with

micelles as functions of the total surfactant concentration. The result obtained for the systems

containing 0.1 M salt in the absence of the polyelectrolyte is presented in Figure 6. Shown in the

figure is also the concentration of hydrogen ions ([H+]), the mole fraction of S+ in the micelles (

x+¿¿) and the total fraction of surfactant in the micelles (N Cmic /Cst ). As can be seen, the fraction

in micelles starts to increase fairly abruptly when the total surfactant concentration is about 10-4

M. Below that concentration, which is interpreted as the critical micelle concentration (CMC),

essentially all surfactant exists in the form of free monomers. Below CMC, the concentrations of

H+ and free S0 are the same since (by neglecting the dissociation of water) all H+ in the system

derives from the reaction S+¿→ S 0+H+¿¿ ¿. Because of the rather high pKa, they are present at much

23

454

455

456

457

458

459

460

461

462

463

464

465

466

467

468

469

470

471

472

473

474

475

476

4546

Page 24: digital.csic.esdigital.csic.es/bitstream/10261/131661/1/2015_Langmuir.docx · Web viewAll aqueous NaHA solutions for potentiometric and fluorescence measurements were prepared by

lower concentration than S+. Nevertheless, at the CMC the micelles contain about equal amounts

of S+ and S0 (x+¿¿ = 0.47), which is explained by the fact that micelles of low charge density are

favored by the electrostatic free energy. Above CMC the concentration of H+ increases faster

with increasing surfactant concentration than below CMC showing that, despite the fact that the

ratio C0/C+ of free monomers decreases faster than below CMC (in agreement with Eq. 6), the

total concentration of S0 in the system increases in this range. Again this is an effect of the

electrostatic free energy favoring low charge of the micelles. Note that this is not in conflict with

the result that x+¿¿ increases and C0 decreases above CMC since the fraction partitioned to

micelles increases for both S+ and S0.

Figure 6. Theoretically calculated concentrations of free S+ (C+), S0 (C0), H+ ([H+], left axis), and

mole fraction of S+ (x+) and total fraction of surfactant in micelles (N Cmic /Cst ) (right axis) as a

function of total surfactant concentration at 0.1 M salt in polyelectrolyte-free solutions.

The result in Figure 6 of the largest significance for the interpretation of the data from the

electrode measurements in Figure 1 is that C0 starts to decrease immediately above the CMC. C0

24

477

478

479

480

481

482

483

484

485

486

487

488

489

490

491

492

4748

Page 25: digital.csic.esdigital.csic.es/bitstream/10261/131661/1/2015_Langmuir.docx · Web viewAll aqueous NaHA solutions for potentiometric and fluorescence measurements were prepared by

decreases in part because incorporating uncharged surfactants in charged micelles lowers the

electrostatic free energy of the micelles (and thus the electrostatic interactions give a negative

contribution to the chemical potential of S0 in micelles), and in part because it is favored by the

entropy of mixing in the micelles. Before dealing with how the decrease affects the response of

the SSE, we note that the onset of micelle formation appears to have a much less dramatic effect

on C+ than on C0. This can explain the discrepancy between the experimental CMC values

reported here and those obtained from measuring bulk properties (NMR self-diffusion and

conductivity) in salt-free solutions. This is further discussed in SI, section S6. The theoretically

calculated CMC values depend, of course, on the standard free energy of transferring a

hydrocarbon tail from water to the micelle core (Δ μ0). The latter value was adjusted to give a

CMC in agreement with experiments at 0.1 M salt. However, the value (13.5 RT/mole) is of the

same order of magnitude as values used earlier for surfactants of comparable hydrophobicity in

connection with the PB theory. E.g., the values 11.4 and 14.4 RT/mole (molarity scale) were

assigned to dodecyl (C12TAB) and tetradecyltrimethyl ammonium bromide (C14TAB),

respectively, in an earlier study.38 By taking into account the contribution from the methylene

groups in the spacer in the MKM head group, the value used is fully reasonable. This value and

the values of the other parameters are used also for the lower salt concentrations. The following

CMC values are obtained at 0.0005, 0.01, 0.05 and 0.1 M salt, respectively: 2.2×10-4, 1.7×10-4,

1.1×10-4 and 9.4×10-5 M. In agreement with the results in Figure 2, the CMC decreases with

increasing salt concentration but the trend is weaker than in the experiments. The agreement

could have been improved by using, e.g., the cell model at the lowest salt concentrations, by

optimizing the aggregation number instead of using a fixed value, or by adjusting the value

25

493

494

495

496

497

498

499

500

501

502

503

504

505

506

507

508

509

510

511

512

513

514

4950

Page 26: digital.csic.esdigital.csic.es/bitstream/10261/131661/1/2015_Langmuir.docx · Web viewAll aqueous NaHA solutions for potentiometric and fluorescence measurements were prepared by

assigned toγ0. However, since our purpose is to demonstrate qualitative features rather than to

seek quantitative agreement we choose not to embark on that task.

We return now to the concentration of free surfactant and how it affects the SSE signal.

According to Figure 6, the concentration of free S+ increases monotonically as a function of the

total concentration, a result expected for solutions of mixed micelles of the present type. It is

indeed difficult to imagine a mechanism that would decrease C+ above the CMC when salt is

present in excess. Thus, we conclude that the lowering of the SSE signal above CMC (Fig. 1)

does not reflect a lowering of the activity of S+ in the solution. The only other charged species

displaying a variant activity in the solution is H+, which is expected to interfere with the S+-

response in the same way as sodium ions. However, since the activity of it increases above the

CMC (and more rapidly than below) it cannot give rise to a reduction of the membrane potential.

Furthermore, it is present at very low concentrations and should be outcompeted by the sodium

ions. The remaining component displaying a varying activity is S0, which is not expected to give

rise to a membrane potential itself. However, by partitioning to the membrane it can affect the

response indirectly by lowering the local activity of S+ or, alternatively, increasing its mobility.

Figure 7 shows the electrode response functions calculated by means of eq. (13) using the data

behind Figure 6 and the corresponding ones for the other salt concentrations. To facilitate

comparison with the experimental data in Figure 1, the curves have been adjusted slightly in the

vertical direction. We emphasize that the adjustment has no influence on the shape of the curves.

The model captures on a qualitative level the two most intriguing features observable in Fig. 1,

namely the deviation from the Nernstian slope at low surfactant concentrations (< CMC) and the

turnover to negative slopes at higher concentrations (> CMC). The quantitative agreement with

experiments decreases with decreasing salt concentration. In particular, the slope below the CMC

approaches the Nernstian value (59 mV/decade) slower with decreasing salt concentrations than

26

515

516

517

518

519

520

521

522

523

524

525

526

527

528

529

530

531

532

533

534

535

536

537

538

5152

Page 27: digital.csic.esdigital.csic.es/bitstream/10261/131661/1/2015_Langmuir.docx · Web viewAll aqueous NaHA solutions for potentiometric and fluorescence measurements were prepared by

Figure 7. Theoretically calculated electrode response (E) as a function of the total surfactant

concentration (cst) in the surfactant solutions at different salt concentrations (in mol L -1) as

indicated. The dashed line: Nernstian response function (slope at 25°C: 59 mV/decade).

in the experiments, leading also to lower maximum E values. At low surfactant concentrations,

the interference by Na+ is responsible for the slight bending of the curves (mainly at the higher

salt concentrations). Apart from that, the slope deviates from the Nernstian below CMC because

C0 increases slower than C+.

The model of polyion-dressed micelles (PDM) was used to calculate the concentration of

micelle-bound and free surfactant in solutions containing polyelectrolyte and simple salt. The free

polyelectrolyte chain is modelled using parameters suitable for fully dissociated HA as described

in section 2.3.3. All other parameters are the same as used to model the polyelectrolyte-free

system. When the degree of S+ binding to the polyelectrolyte ¿ is below unity, the concentrations

of all species and x+ vary with the total surfactant concentration qualitatively in the same way as

in the polyelectrolyte-free solutions. When β+¿¿ is larger than unity, C+ starts to increase faster

with increasing total surfactant concentration due to the entropic penalty of binding surfactant

27

539

540

541

542

543

544

545

546

547

548

549

550

551

552

553

554

555

5354

Page 28: digital.csic.esdigital.csic.es/bitstream/10261/131661/1/2015_Langmuir.docx · Web viewAll aqueous NaHA solutions for potentiometric and fluorescence measurements were prepared by

counterions to PDM. This is accompanied by an increase in C0 and a decrease in x+. The same

qualitative behavior is observed at all salt concentrations investigated. The results are used to

calculate the electrode response functions in the same way as in the polyelectrolyte-free case.

Figure 8 shows the result obtained for the 0.01 M salt case (solid line). For surfactant

concentrations below the CAC, the response is the same as in polyelectrolyte-free solutions

(dashed line). Above the CAC, the signal first drops and then increases again as expected from

the variation of C0 just described. A comparison with Fig. 4 shows that the calculated curve

captures the main features observed experimentally. The theoretically calculated curves for the

other salt concentrations are presented together with the corresponding experimental curves in SI

(Figure S9). It can be mentioned here that the experimental curves in

Figure 8. Theoretically calculated electrode response (E vs. log cS) as a function of the total

surfactant concentration, cst, in the presence and absence of the polyelectrolyte at 0.01 M salt.

E* = 257 mV.

Fig. 1 plotted without vertical adjustment display a shift towards lower E with increasing salt

concentration (at 0.01, 0.05, and 0.1 M salt the shift is ca. 25, 25 and 35 mV, respectively, from

28

556

557

558

559

560

561

562

563

564

565

566

567

568

569

570

571

5556

Page 29: digital.csic.esdigital.csic.es/bitstream/10261/131661/1/2015_Langmuir.docx · Web viewAll aqueous NaHA solutions for potentiometric and fluorescence measurements were prepared by

the salt free curve). The reason for this salt effect, which is not captured by eq. (13), is not

understood.

Figure 9 shows theoretically calculated binding isotherms for all salt concentrations studied.

Shown are both the apparent binding isotherms (βapp=f ¿; curves to the left) calculated from eq.

(11) and the actual binding isotherms (β=f (C+¿¿); curves to the right) calculated from eq. (10).

Note that the former ones are constructed from the curves in Fig. 8 (and the corresponding ones

for the other salt concentrations) in the same way as the experimental binding isotherms in Fig. 5

were constructed from the data in Figs. 4 and S7. In agreement with experiments, the slope of the

apparent binding isotherms calculated from theory is negative in a large binding range above

CAC and changes rather abruptly to positive values at high degrees of

Figure 9. Theoretically calculated binding isotherms. Left branch: Apparent degree of surfactant

binding, app, plotted vs. the apparent free concentration of S+. Right branch: true degree of

binding, , of MKM plotted vs. its free monomer concentration. In both cases, free monomer

concentration is designated as csf.

binding. In contrast, in the binding isotherms showing the actual degree of binding (“β”) the

slope is never negative. The apparent binding isotherm for the highest salt concentration does not

29

572

573

574

575

576

577

578

579

580

581

582

583

584

585

586

587

588

5758

Page 30: digital.csic.esdigital.csic.es/bitstream/10261/131661/1/2015_Langmuir.docx · Web viewAll aqueous NaHA solutions for potentiometric and fluorescence measurements were prepared by

‘swing back’ as much as in the experiment but the effect is most pronounced for this salt

concentration and the curve even intersects the other ones, in agreement with experiments. The

crossover to positive slope takes place at a higher binding level in the theoretically calculated

isotherms for the two lowest salt concentrations than in the experiments. This is directly an effect

of setting f=1 for β+¿≤ 1¿ in the PDM model, an approximation introduced to simplify the

calculations. Recall also that in the present model the crossover is a consequence of the entropic

penalty of binding surfactant counterions when the net charge of the PDM becomes positive (

β+¿>1¿). The mechanism is thus different from that in the site binding model39 often used to

analyze binding isotherms.

We conclude that according to the theoretical model calculations the data in Fig. 5 do not

represent true binding isotherms for MKM. The in-model explanation is that they are constructed

with the assumption that the measured electrode potential is a function of the activity of the

protonated form of MKM only, when it is in fact a function of both the protonated and

deprotonated forms. The largest consequence of that is that the reduction of the signal above the

CAC is interpreted as a lowering of the free concentration of (protonated) MKM, whereas it is

only the free concentration of the deprotonated form that is reduced. This is the origin of the

negative slope of the binding isotherms. In contrast, the protonated fraction, which is much larger

in magnitude, continues to increase above the CAC. The explanation is strongly supported by the

overall good qualitative agreement between the theoretically calculated electrode response

functions and experiments, both in the present and absences of the polyelectrolyte.

Conclusions

30

589

590

591

592

593

594

595

596

597

598

599

600

601

602

603

604

605

606

607

608

609

610

5960

Page 31: digital.csic.esdigital.csic.es/bitstream/10261/131661/1/2015_Langmuir.docx · Web viewAll aqueous NaHA solutions for potentiometric and fluorescence measurements were prepared by

The micellization of a recently synthesized biocompatible lysine-based surfactant MKM having

a C12 hydrocarbon chain and its interaction with the sodium salt of hyaluronic acid (NaHA) were

studied in aqueous solutions at different NaCl concentrations by the surfactant-selective electrode

(SSE)-based potentiometric titrations and by pyrene-based fluorescence measurements. The

surfactant showed several anomalies. First, its CMC values did not fit the usual double

logarithmic relationship in dependence on the ionic strength as known for conventional

surfactants but followed a semi-logarithmic relation of the form CMC=−a−b logI (I is the ionic

strength). Second, the slopes of the SSE response curves decreased with increasing I below the

CMC and turned to negative above the CMC, which resulted in pronounced negative slopes of

the experimental binding isotherms in the region of cooperative binding. The ‘back swing’ of the

isotherms even increased with increasing salt concentration. These observations were explained

by proposing equilibrium between the protonated (charged) and deprotonated (uncharged) form

of MKM in solution. All important features of the experimental isotherms were captured by a

theoretical model that considered the possibility of competitive binding of the two forms to the

electrode membrane. In particular, it was shown that the measured electrode potential is a

function of the activity of both the protonated and deprotonated MKM, which brings around the

anomalous behavior as detected by SSE.

One important implication of the result in the present study is that determination of binding

isotherms for pH sensitive surfactants by means of surfactant selective electrodes may be

difficult. The same should apply to mixtures of ionic and non-ionic amphiphiles in general. Such

example may be an anomalous binding isotherm obtained from the electrode measurements in

mixtures of a cationic surfactant and a non-ionic lipid as reported by Kadi et al.40 Additional

examples of electrode measurements that resulted in negative slopes of binding isotherms were

31

611

612

613

614

615

616

617

618

619

620

621

622

623

624

625

626

627

628

629

630

631

632

633

6162

Page 32: digital.csic.esdigital.csic.es/bitstream/10261/131661/1/2015_Langmuir.docx · Web viewAll aqueous NaHA solutions for potentiometric and fluorescence measurements were prepared by

reported also for mixtures of anionic surfactants and neutral polymers41 or cationic surfactant and

an anionic polyelectrolyte.42

Acknowledgements

This work was partially supported by the Slovenian Research Agency, ARRS, through the

Physical Chemistry program P1-0201, program P2 0118, and project L2-4060. LP generously

thanks the Spanish Plan National I+D+I MAT2012-38047-C02-02.

Supporting Information. Structures of MKM and HA, additional information about the PDM

model and a simple model of the SSE response, examples of pyrene emission spectra,

experimentally determined CMC and CAC values, dependence of CMC on ionic strength, SSE

potentiometric titration curves, and discussion about CMC values obtained by NMR. This

material is available free of charge as Supporting Information on the ACS Publications website.

References

1. Fukui, H.; Satake, I.; Hayakawa, K. Thermodynamic Effects of Alcohol Additives on the

Cooperative Binding of Sodium Dodecyl Sulfate to a Cationic Polymer. Langmuir 2002, 18 (11),

4465-4470.

2. Chen, Y.; Lapitsky, Y. Interactions of anionic surfactants with cationic polyelectrolyte

gels: Competitive binding and application in separation processes. Colloids and Surfaces A:

Physicochemical and Engineering Aspects 2010, 372 (1–3), 196-203.

32

634

635

636

637

638

639

640

641

642

643

644

645

646

647

648

649

650

651

652

653

654

6364

Page 33: digital.csic.esdigital.csic.es/bitstream/10261/131661/1/2015_Langmuir.docx · Web viewAll aqueous NaHA solutions for potentiometric and fluorescence measurements were prepared by

3. Lu, K. W.; Taeusch, W. H.; Robertson, B.; Goerke, J.; Clements, J. A. Polymer-

Surfactant Treatment of Meconium-induced Acute Lung Injury. Am. J. Respir. Crit. Care Med.

2000, 162 (2), 623-628.

4. Panyam, J.; Chavanpatil, M. D. Polymer-surfactant nanoparticles for sustained release of

compounds. United States Patent2011.

5. Khdair, A.; Gerard, B.; Handa, H.; Mao, G.; Shekhar, M. P. V.; Panyam, J.

Surfactant−Polymer Nanoparticles Enhance the Effectiveness of Anticancer Photodynamic

Therapy. Molecular Pharmaceutics 2008, 5 (5), 795-807.

6. Colomer, A.; Pinazo, A.; Manresa, M. A.; Vinardell, M. P.; Mitjans, M.; Infante, M. R.;

Perez, L. Cationic Surfactants Derived from Lysine: Effects of Their Structure and Charge Type

on Antimicrobial and Hemolytic Activities. J Med Chem 2011, 54 (4), 989-1002.

7. Bongaerts, J. H. H.; Cooper-White, J. J.; Stokes, J. R. Low Biofouling Chitosan-

Hyaluronic Acid Multilayers with Ultra-Low Friction Coefficients. Biomacromolecules 2009, 10

(5), 1287-1294.

8. Vindigni, V.; Cortivo, R.; Iacobellis, L.; Abatangelo, G.; Zavan, B. Hyaluronan Benzyl

Ester as a Scaffold for Tissue Engineering. International Journal of Molecular Science 2009, 10.

9. Luppi, B.; Bigucci, F.; Mercolini, L.; Musenga, A.; Sorrenti, M.; Catenacci, L.; Zecchi,

V. Novel mucoadhesive nasal inserts based on chitosan/hyaluronate polyelectrolyte complexes

for peptide and protein delivery. Journal of Pharmacy and Pharmacology 2009, 61 (2), 151-157.

10. Shateri Khalil-Abad, M.; Yazdanshenas, M.; Nateghi, M. Effect of cationization on

adsorption of silver nanoparticles on cotton surfaces and its antibacterial activity. Cellulose 2009,

16 (6), 1147-1157.

33

655

656

657

658

659

660

661

662

663

664

665

666

667

668

669

670

671

672

673

674

675

676

6566

Page 34: digital.csic.esdigital.csic.es/bitstream/10261/131661/1/2015_Langmuir.docx · Web viewAll aqueous NaHA solutions for potentiometric and fluorescence measurements were prepared by

11. Yadav, A.; Prasad, V.; Kathe, A. A.; Raj, S.; Yadav, D.; Sundaramoorthy, C.;

Vigenshwaran, N. Functional finishing in cotton fabrics using zinc oxide nanoparticles. Bulletin

of material science 2006, 29 (6), 641-645.

12. Singhal, J. P.; Ray, A. R. Adsorption of iodine on nylon-6. Trends in biomaterials and

artificial organs 2002, 16 (1), 46-51.

13. Fillat, A.; Gallardo, O.; Vidal, T.; Pastor, F. I. J.; Díaz, P.; Roncero, M. B. Enzymatic

grafting of natural phenols to flax fibres: Development of antimicrobial properties. Carbohydrate

Polymers 2012, 87 (1), 146-152.

14. Qian, L.; Guan, Y.; Ziaee, Z.; He, B.; Zheng, A.; Xiao, H. Rendering cellulose fibers

antimicrobial using cationic β-cyclodextrin-based polymers included with antibiotics. Cellulose

2009, 16 (2), 309-317.

15. Brown, M. B.; Jones, S. A. Hyaluronic acid: a unique topical vehicle for the localized

delivery of drugs to the skin. Journal of the European Academy of Dermatology and Venereology

2005, 19 (3), 308-318.

16. Dumitriu, S. Polysaccharides: Structural Diversity and Functional Versatility. Marcel

Dekker: New York, 1998.

17. Lee, K. Y.; Jeong, L.; Kang, Y. O.; Lee, S. J.; Park, W. H. Electrospinning of

polysaccharides for regenerative medicine. Adv Drug Deliver Rev 2009, 61 (12), 1020-1032.

18. Perez, L.; Pinazo, A.; Teresa García, M.; Lozano, M.; Manresa, A.; Angelet, M.; Pilar

Vinardell, M.; Mitjans, M.; Pons, R.; Rosa Infante, M. Cationic surfactants from lysine:

Synthesis, micellization and biological evaluation. Eur J Med Chem 2009, 44 (5), 1884-1892.

19. Hayakawa, K.; Kwak, J. C. T. Surfactant-polyelectrolyte interactions. 1. Binding of

dodecyltrimethylammonium ions by sodium dextransulfate and sodium poly(styrenesulfonate) in

34

677

678

679

680

681

682

683

684

685

686

687

688

689

690

691

692

693

694

695

696

697

698

699

6768

Page 35: digital.csic.esdigital.csic.es/bitstream/10261/131661/1/2015_Langmuir.docx · Web viewAll aqueous NaHA solutions for potentiometric and fluorescence measurements were prepared by

aqueous solution in the presence of sodium chloride. The Journal of Physical Chemistry 1982, 86

(19), 3866-3870.

20. Hayakawa, K.; Kwak, J. C. T. Study of surfactant-polyelectrolyte interactions. 2. Effect of

multivalent counterions on the binding of dodecyltrimethylammonium ions by sodium dextran

sulfate and sodium poly(styrene sulfonate) in aqueous solution. The Journal of Physical

Chemistry 1983, 87 (3), 506-509.

21. Kogej, K. Association and structure formation in oppositely charged polyelectrolyte-

surfactant mixtures. Advances in Colloid and Interface Science 2010, 158 (1-2), 68-83.

22. Kogej, K.; Škerjanc, J. Fluorescence and Conductivity Studies of Polyelectrolyte-Induced

Aggregation of Alkyltrimethylammonium Bromides†. Langmuir 1999, 15 (12), 4251-4258.

23. Gunnarsson, G.; Jönsson, B.; Wennerström, H. Surfactant Association into Micelles. An

Electrostatic Approach. Journal of Physical Chemistry 1980, 84, 3114-3121.

24. Jönsson, B.; Wennerström, H. Phase Equilibria in a Three-Component Water-Soap-

Alcohol System. A Thermodynamic Model. Journal of Physical Chemistry 1987, 91, 338-352.

25. Hansson, P. Self-assembly of ionic surfactants in polyelectrolyte solutions: A model for

mixtures of opposite charge. Langmuir 2001, 17, 4167.

26. Gernandt, J.; Hansson, P. Hysteresis in the surfactant-induced volume transition of

hydrogels. Journal of Physical Chemistry B 2015, 119 (1717), 1725.

27. Hansson, P. Phase behavior of aqueous polyion-surfactant ion complex salts: A

theoretical analysis. Journal of Colloid and Interface Science 2009, 332, 183-193.

28. Buck, R. P.; Lindner, E. Recommendations for nomenclature of ion-selective electrodes.

Pure & Applied Chemistry 1994, 66 (12), 2527-2536.

29. Bobacka, J.; Ivaska, A.; Lewenstam, A. Potentiometric Ion Sensors. Chemical Reviews

2008, 108, 329-351.

35

700

701

702

703

704

705

706

707

708

709

710

711

712

713

714

715

716

717

718

719

720

721

722

723

6970

Page 36: digital.csic.esdigital.csic.es/bitstream/10261/131661/1/2015_Langmuir.docx · Web viewAll aqueous NaHA solutions for potentiometric and fluorescence measurements were prepared by

30. Evans, F. D.; Wennerström, H. T. The Colloidal Domain: Where Physics, Chemistry,

Biology, and Technology Meet. 2 ed.; Wiley-VCH Verlag GmbH: New York, 1999; p 672.

31. Mezei, A.; Perez, L.; Pinazo, A.; Comelles, F.; Infante, M. R.; Pons, R. Self assembly of

pH-sensitive cationic lysine based surfactants. Langmuir 2012, 28, 16761-16771.

32. Fegyver, E.; Meszaros, R. The Impact of Nonionic surfactant Additives on the

Nonequilibrium Association between Oppositely charged Polyelectrolytes and Ionic Surfactants,

Soft Matter 2014, 10, 1953-1962.

33. Fegyver, E.; Meszaros, R. Fine-Tunin the Nonequilibrium Behavior of Oppositely

Charged Macromolecule/Surfactant Mixtures via the Addition of Nonionic Amphiphiles,

Langmuir 2014, 30, 15114-15126.

34. Bakshi, M. S.; Kaur, I. Benzyl and pyridinium head groups controlled surfactant-polymer

aggregates of mixed cationic micelles and anionic polyelectrolytes. Colloid and polymer science

2004, 282, 476-485.

35. Bauer, S.; Arpa-Sancet, M. P.; Finlay, J. A.; Callow, M. E.; Callow, J. A.; Rosenhahn, A.

Adhesion of Marine Fouling Organisms on Hydrophilic and Amphiphilic Polysaccharides.

Langmuir 2013, 29 (12), 4039-4047.

36. Rosen, M. J. Surfactants and interfacial phenomena; Wiley: Michigan, 1978.

37. Hunter, R. J.; White, L. R. Foundations of colloid science; Clarendon Press: Oxford,

1987.

38. Hansson, P.; Jönsson, B.; Ström, C.; Söderman, O. Determination of micellar aggregation

numbers in dilute surfactant systems with the fluorescence quenching method. Journal of

Physical Chemistry B 2000, 104, 3496-3506.

36

724

725

726

727

728

729

730

731

732

733

734

735

736

737

738

739

740

741

742

743

744

745

7172

Page 37: digital.csic.esdigital.csic.es/bitstream/10261/131661/1/2015_Langmuir.docx · Web viewAll aqueous NaHA solutions for potentiometric and fluorescence measurements were prepared by

39. Satake, I.; Yang, J. T. Interaction of sodium decyl sulfate with poly(L-ornithine) and

poly(L-lysine) in aqueous solution. Biopolymers 1976, 15, 2263-2275.

40. Kadi, M.; Hansson, P.; Almgren, M. Determination of Isotherms for Binding of

Surfactants to Vesicles Using a Surfactant-Selective Electrode. Journal of Physical Chemistry B

2004, 108, 7344-7351.

41. Takisawa, N.; Brown, P.; Bloor, D.; Hall, D. G. Chemical relaxation and equilibrium

studies associated with binding of anionic surfactants to neutral polymers. Journal of the

Chemical Society, Faraday Transactions 1: Physical Chemistry in Condensed Phases 1989, 85

(8), 2099-2112.

40. Thalberg, K.; Van Stam, J.; Lindblad, C.; Almgren, M.; Lindman, B. Time-resolved

fluorescence and self-diffusion studies in systems of a cationic surfactant and an anionic

polyelectrolyte. The Journal of Physical Chemistry 1991, 95 (22), 8975-8982.

37

746

747

748

749

750

751

752

753

754

755

756

757

7374

Page 38: digital.csic.esdigital.csic.es/bitstream/10261/131661/1/2015_Langmuir.docx · Web viewAll aqueous NaHA solutions for potentiometric and fluorescence measurements were prepared by

Table of contents graphics for the paper

Interaction of hyaluronic acid with a biocompatible cationic surfactant from lysine: a binding study

by Matej Bračič, Per Hansson, Lourdes Perez, Lidija F. Zemljič, and Ksenija Kogej

38

758

759

760761

762

763

764

765

766

767

768

769

770

771

772

773

774

775

776

777

778

7576

Page 39: digital.csic.esdigital.csic.es/bitstream/10261/131661/1/2015_Langmuir.docx · Web viewAll aqueous NaHA solutions for potentiometric and fluorescence measurements were prepared by

Supporting information for the paper

Interaction of sodium hyaluronate with a

biocompatible cationic surfactant from lysine: a

binding study

Matej Bračič a Per Hansson,b Lourdes Pérez,c Lidija Fras Zemljič,a* and Ksenija Kogejd*

a Institute for the Engineering and Design of Materials, University of Maribor, Smetanova 17,

2000 Maribor, Slovenia.

b Department of Pharmacy, Biomedical Centre, Uppsala University, SE-75123 Uppsala, Sweden

c Department of Chemical and Surfactant technology, Instituto de Química Avanzada de

Cataluña, CSIC, Jordi Girona 18-26, 08034 Barcelona, Spain

d Department of Chemistry and Biochemistry, Faculty of Chemistry and Chemical Technology,

University of Ljubljana, 1000 Ljubljana, Slovenia.

* [email protected]

39

779

780

781

782

783

784

785

786

787

788

789

790

791

792

793

794

795

7778

Page 40: digital.csic.esdigital.csic.es/bitstream/10261/131661/1/2015_Langmuir.docx · Web viewAll aqueous NaHA solutions for potentiometric and fluorescence measurements were prepared by

S1. Materials

Scheme S1. Chemical structures of a) MKM1 and b) NaHA2.

S1.1. Additional characterization of MKM

HPLC analysisof MKM were carried out using an analytical HPLC instrument, model Elite

LaChrom, using a UV-VIS detector at 215 nm. A Lichrospher 100 CN (propylciano) 5 μm, 250 x

4 mm, column was used. A gradient elution profile was employed from the initial solvent

composition of A/B 75:25 (by volume), changing during 24 min to a final composition of 5:95.

Solvent A was 0.1% (v/v) trifluoroacetic acid (TFA) in H2O and solvent B was 0.085% TFA in

H2O/CH3CN 1:4. The flow rate through the column was 1.0 mL min -1. The 1H NMR spectra were

recorded on a Varian spectrometer at 400 MHz using the deuterium signal of the solvent as the

lock. All measurements were performed in deuterated methanol using 5 mm tubes.

40

796

797

798

799

800

801

802

803

804

805

806

807

808

809

810

7980

Page 41: digital.csic.esdigital.csic.es/bitstream/10261/131661/1/2015_Langmuir.docx · Web viewAll aqueous NaHA solutions for potentiometric and fluorescence measurements were prepared by

Figure S1. The HPLC chromatograms (UV signal as a function of the retension time) taken at t = 0 and 20 hours, as indicated in the panels.

Table S1. Retention time, area, and height of the main peak in the HPLC chromatogram of MKM in water

after t = 0, 2, 4, and 20 hours of stirring the solution.

t / hours retention time (min) area (a.u.) height (a.u.)

0 13.490 9812177 404641

2 13.493 10748325 422992

4 13.520 11030196 428674

20 13.567 9856165 401192

41

Minutes0 5 10 15

mA

U

0

20

40

60

80

100

120

140

160

180

200

mA

U

0

20

40

60

80

100

120

140

160

180

200

13,4

90

UVmkmT0

Retention Time

Minutes0 5 10 15

mA

U

0

20

40

60

80

100

120

140

160

180

200

mA

U

0

20

40

60

80

100

120

140

160

180

200

13,5

67

UVmkmT20

Retention Timet = 0 hours t = 20 hours

811812813814815

816

817

818

819

820

821

8182

Page 42: digital.csic.esdigital.csic.es/bitstream/10261/131661/1/2015_Langmuir.docx · Web viewAll aqueous NaHA solutions for potentiometric and fluorescence measurements were prepared by

Figure S2. The 1H NMR spectra of MKM in deuterated methanol taken at t = 0 and 20 hours, as indicated in the panels.

42

t = 0 hours

t = 20 hours

822

823

824

825826827

8384

Page 43: digital.csic.esdigital.csic.es/bitstream/10261/131661/1/2015_Langmuir.docx · Web viewAll aqueous NaHA solutions for potentiometric and fluorescence measurements were prepared by

S2. Additional information about the model of polyion-dressed micelle (PDM)

Electrostatic free energy. The free energy expression underlying the PDM model derives from a

recent model of complexes between surfactant micelles and oppositely charged flexible polyion

chains.3, 4, 5, 6 The aggregation number N of a micelle with a spherical hydrocarbon core is related

to the core radius r, the volume per surfactant tail v, and the area per surfactant in the head group

region a:

N= 4 π r3

3v=4 π r2

a=36 π v2

a3 (S:1)

When v is the same for all components, (S:1) is valid also for mixed micelles. The polyion and

small ions are solubilized in the aqueous regions between the micelles (see Figure S1). To

describe the interactions in concentrated solutions each micelle is positioned in a spherical cell of

radius R=L+r, where L is the thickness of the aqueous layer surrounding the micelle in the cell.

R is directly related to the molar concentration of micelles Cmic:

NvCmic N A=( rR )

3

(S:2)

with NA being the Avogadro number.

43

828

829

830

831

832

833

834

835

836

837

838

839

840

841

842

8586

Page 44: digital.csic.esdigital.csic.es/bitstream/10261/131661/1/2015_Langmuir.docx · Web viewAll aqueous NaHA solutions for potentiometric and fluorescence measurements were prepared by

Figure S3. Schematic picture of micelle – polyion complex phase indicating how the radius of the bare micelle (r), the effective double layer thickness (d), the thickness of the aqueous layer (L), and the radius of the cell (R) are defined in the electrostatic model. The electrostatic free energy is expressed as a function of L and the number of micelles Nmic

(equal to the number of cells):

Gel=2 πNmic σ2 r3

ε0 εr( d (L)r+d (L) ) (S:3)

d (L)=Lc (1−e−L/Lc) (S:4)

Lc=hr√Z

(S:5)

Eq. (S:3) has the form of the energy of a spherical capacitor. The same type of expression or

similar has been used to model complex formation between micelles and polyelectrolytes with

results in good agreement with experiments.3, 4, 5, 6 The theoretical basis for it rests on a

comparison with an analogous expression for the electrostatic energy between planar surfaces

separated by counterions.7 In the strong electrostatic coupling regime the electrostatic force

between the surfaces is independent of the valence of the counterions and their position in the gap

between the planes. The behavior is described well by the strong coupling (SC) theory by

Moreira and Netz,8 in which the energy expression has the form of the energy of a capacitor.7 The

SC theory works fairly good also at considerably lower coupling strength as long as the

separation between planes is smaller than the average distance between the counterions. This is

attributed to the correlation between the counterions: In order to avoid each other the counterions

to both surfaces tend to form one correlated layer in the gap between the planes (rather than one

at each surface).9 As a result, the average distance between counterions does not change much

when the distance between the planes changes, and therefore the coulomb energy due to

counterion-counterion interactions is nearly constant. This means that the variation of the energy

only depends on the plate – plate and plate – counterion interactions. The net effect is an

44

843844845846

847

848

849

850

851

852

853

854

855

856

857

858

859

860

861

862

863

864

865

866

8788

Page 45: digital.csic.esdigital.csic.es/bitstream/10261/131661/1/2015_Langmuir.docx · Web viewAll aqueous NaHA solutions for potentiometric and fluorescence measurements were prepared by

electrostatic energy increasing linearly with increasing plate-plate distance,9 similar to a

capacitor. This is the origin of the electrostatic ion-ion correlation attraction between like charged

surfaces. The applicability range of the SC theory can be extended to lower electrostatic coupling

strengths by replacing the constant force with an exponential force law.7 Comparison with Monte

Carlo simulations shows that the decay length of the interaction (defined as 2Lc) is of order the

average distance between the counterions in the compressed monolayer.7, 9 For monovalent

counterions this means 2 Lc∝√a, where a is the area per surface charge. The interpretation placed

on this is that upon increasing the distance between the planes the attraction vanishes when the

monolayer of counterions separates into two layers (one at each surface), i.e. when the electric

double layers at each surface no longer overlap. Lc can thus be interpreted as an effective

thickness of the electric double layer at a single surface. In agreement with that, the energy (per

plate) in the modified SC theory approaches that of a capacitor with separation Lc in the limit of

infinite separation.7 More specifically, the model assumes that the energy between two planes at a

distance 2L is twice that that of a capacitor with separation d, where d is a function of L given by

eq. (S:4). The form of the function is mainly intuitive but comparison with MC simulations

shows that it works fairly well in a wide range of coupling strengths.7 Eqs. (S:3) – (S:4) are based

on the assumption that the capacitor analogy is transferrable to systems of charged spheres.6 The

energy (per sphere) is equal to that of a spherical capacitor (eq. S:3) with separation d as shown

in Figure S1, where d is a function (eq. S:4) of the thickness L of the aqueous layer surrounding

each sphere. In further analogy with the case of planes separated by monovalent counterions, Lc

is related to the area a+¿¿ per charge on the (positively charged) sphere, but since for spheres

√a+¿∝r√Z

¿, we use eq. (S:5) with h being a constant.

45

867

868

869

870

871

872

873

874

875

876

877

878

879

880

881

882

883

884

885

886

887

888

8990

Page 46: digital.csic.esdigital.csic.es/bitstream/10261/131661/1/2015_Langmuir.docx · Web viewAll aqueous NaHA solutions for potentiometric and fluorescence measurements were prepared by

The usefulness of the model is that it allows us to quantify both the correlation induced attraction

between like charged micelles in the concentrated regime and the energy of the electric double

layer surrounding single micelles in the dilute regime. The last aspect is the central one in the

present work. Upon dilution the dense complex phase disintegrates into single PDM:s or single

PDM:S and free polyelectrolytes depending on the polyion/surfactant ratio; see Figure S2. Eq.

(S:3) assumes that small ions are evenly distributed in the aqueous regions (i.e., no counterion

‘condensation’ at the polyion and the micelle), which is a reasonable approximation in the

concentrated regime. In the dilute regime counterion binding is treated on the Poisson-Boltzmann

(PB) level. The electrostatic free energy expression (S:3) is replaced by

Gel=G polel +Gmic

el ,ext+Gmicel ,∫¿¿ (S:6)

The first term in (S:6) is the electrostatic free energy of the free polyion chains and the second

term is the electrostatic free energy due to the net charge of PDM ( f ≠ 1). Both are evaluated by

means of the PB-equation as described in the paper. The last term is the “internal” electrostatic

(free) energy of the PDM:s which is the limiting form of eq. (S:3) as L goes to infinity:

Gmic

el ,∫¿=2 πNmicσ

2 r3

ε 0 εr( Lc

r+Lc)¿ (S:7)

46

889

890

891

892

893

894

895

896

897

898

899

900

901

902

903

904

905

906

9192

Page 47: digital.csic.esdigital.csic.es/bitstream/10261/131661/1/2015_Langmuir.docx · Web viewAll aqueous NaHA solutions for potentiometric and fluorescence measurements were prepared by

Figure S4. Schematic picture of a single PDM (left) and a polyion chain (right) surrounded by simple ions in the dilute regime. Dashed line has the same meaning as in Figure S1. In the PB-cell model the PDM and the polyion is described as a uniformly charged sphere and cylinder, respectively. In the dilute solutions considered in the present work the spheres and cylinders are treated as in contact with a common bulk electrolyte solution, meaning that the radii of the cells (solid red lines) are infinitely large.

For mixed micelles of a cationic and a nonionic surfactant

Nmic=N+ ¿mic+N0

mic

N ¿ (S:8)

σ=Nx+¿e

4 π r2 ¿ (S:9)

Lc=hr

√N x+¿¿ (S:10)

with N imic being the number of i molecules in the micellar sub-phase and h a model parameter

related to distribution of charges on the micelle. With eqs. (S:8) – (S:10), the definition of the

Bjerrum length and the equalities Z=N x+¿ ¿, and Nmic N=N +¿mic+N0mic ¿, (S:7) can be rewritten as:

Gmicel ,∫¿=kBT ¿¿ (S:11)

Surface free energy. The surface free energy of the micelles due to the contact between the

hydrocarbon core and water is:10

Gmicsurf=a¿ (S:12)

Chemical potential of surfactant in PDM. The electrostatic contribution to the chemical potential

of the two surfactants in the PDM is obtained in the usual way by taking the derivatives of eq.

(S:6) with respect to N+¿mic¿ and N 0mic, respectively. The contribution from Gmic

el , ext is the expressions

obtained within the PB-theory for ionic and non-ionic surfactants in mixed micelles of a given net

47

907908909910911912913

914

915

916

917

918

919

920

921

922923

924

925

926

927

928

929

930

931

932

933

9394

Page 48: digital.csic.esdigital.csic.es/bitstream/10261/131661/1/2015_Langmuir.docx · Web viewAll aqueous NaHA solutions for potentiometric and fluorescence measurements were prepared by

charge.10 For S+ there is a contribution also from Gpolel since, for each micelle formed, f free

polyion charges are removed from the solution per charged surfactant. This is also evaluated

using the PB-theory for a cylindrical polyion.10 However, we include only the entropic part

associated with the release of the counterion bound to the free polyion. Thus, in eq. (8) and (9)

we put ∆ μp=− f μp−TS; see section 2.3.3. The argument comes from a previous study6 suggesting

that the energetic part is nearly equal to a standard free energy of a polyion charged group

positioned in the PDM double layer (an excess energy due to the interaction between neighboring

charges on the polyion). Since we have chosen not

1

2

3

4

5

6

7

0 20 40 60 80 100 120

(Gm

ic /

k BT) /

(Nm

icN

)

N

tot

surf

el

a

0

50

100

150

200

250

1 1.5 2 2.5 3 3.5 4

Nop

t

h

b

48

934

935

936

937

938

939

940

941

942

9596

Page 49: digital.csic.esdigital.csic.es/bitstream/10261/131661/1/2015_Langmuir.docx · Web viewAll aqueous NaHA solutions for potentiometric and fluorescence measurements were prepared by

Figure S5. (a) Total free energy (Gmicel ,∫¿+Gmic

surf ¿) per surfactant in PDM (tot) and individual contributions from the surface (surf) and electrostatic free energy (el) as a function of N calculated from theory with h=2.5; x+=1. (b) Optimal aggregation number (Nopt) as a function of h (x+=1).

to include the latter in Gmicel ,∫¿¿ it is most correct not to include the energetic part in G pol

el when

taking the derivative with respect to N imic. The contribution from Gmic

el ,∫¿¿ to the chemical potential

is given by eq. (7), and the contribution from Gmicsurf is just μi

surf=a γi.

Optimal aggregation number and determination of h. The optimal aggregation number is

obtained by minimizing the sum Gmicel ,∫¿+Gmic

surf ¿ with respect to N keeping N+¿mic¿ and N0mic fixed. The

result for pure micelles of a cationic ionic surfactant (x+¿=1¿) with 12 carbons in the hydrocarbon

chain (v = 351 Å3)9 in aqueous solution at 298 K (lB = 7.14 Å) is shown in Figure S3. For the

surface free energy we have used γ+¿¿=0.018 J/m2 a value often used in models of ionic

surfactants.10, 11 Fig. S3a shows the individual and total free energy contributions as a function of

N for h = 2.5, for which the optimal aggregation number is 62; Fig. S3b shows optimized N-

values as a function of h. Aggregation numbers for C12-tailed ionic surfactants in dilute solution

reported in the literature are typically around 60. For example, for dodecyltrimethylammonium

bromide values between 55 and 65 have been reported both for the pure surfactant12 and in

complexes with hydrophilic polyelectrolytes,3 motivating setting h = 2.5 in the model calculations

for the surfactant under study. As a check, insertion of this value in (S:5) gives Lc = 5.5 Å. Since

2Lc is of the order of the distance between the charged head groups at the micelle surface (

8 Å) and 5.5 Å is a reasonable value of the thickness of the polyion layer surrounding

a micelle (see, e.g., MC simulations on sphere + polyion13) the h-value used is consistent with the

conditions underlying the model. The maximum aggregation number for a C12 surfactant

49

943944945946947

948

949

950

951

952

953

954

955

956

957

958

959

960

961

962

963

964

965

966

967

9798

Page 50: digital.csic.esdigital.csic.es/bitstream/10261/131661/1/2015_Langmuir.docx · Web viewAll aqueous NaHA solutions for potentiometric and fluorescence measurements were prepared by

consistent with a spherical hydrocarbon core is 56. However, in practice the micelle radius is

allowed to exceed the length of a fully stretched tail by ca. 10%,10 and so the pure micelles (x+=1)

should still have an average spherical shape. For the mixed micelles the optimal value provided

by the model increases as the fraction of the non-ionic component increases, and exceeds that

conceivable for a sphere already for x+≈0.9, as shown in Figure S3. This is not unexpected

considering that the non-ionic component lowers the net repulsion between the head groups.

60

70

80

90

100

110

0.75 0.8 0.85 0.9 0.95 1 1.05

Nop

t

x+

Figure S6. Optimal aggregation number (Nopt) as a function of x+; h =2.5.

S3. Simple model of surfactant selective electrode response

In this section we derive an electrode response function based on a simple model of how the

charged and uncharged forms of MKM interact in the electrode membrane. The selectivity of the

electrode for cationic surfactants over inorganic ions indicates that the surfactant has the highest

permeability through the membrane. This can be explained in part by the hydrophobic tail of the

surfactant and in part by the presence of (fixed) anionic groups in the plastic membrane

(explaining why the performance of the present type of electrode is usually excellent for cationic

50

968

969

970

971

972

973

974975

976

977

978

979

980

981

982

983

99100

Page 51: digital.csic.esdigital.csic.es/bitstream/10261/131661/1/2015_Langmuir.docx · Web viewAll aqueous NaHA solutions for potentiometric and fluorescence measurements were prepared by

but poor for anionic surfactants). The membrane is assumed to contain a concentration c−¿ ¿ of

negative charges fixed to the polymer matrix. S+ and Na+ dissolved in the membrane are restricted

to the regions immediately surrounding the negative charges, the latter functioning as sites always

associated with exactly one cation. S0 is allowed to reside at a site together with Na+ but not with

S+ due to steric hindrance. S0 at these sites are in equilibrium with those distributed elsewhere in

the membrane (and in the solution). Sites occupied by S+, Na+, and simultaneously by Na+ and S0

are denoted S+¿¿, Na+¿¿, and Na+¿S 0

¿, respectively. The correspondning concentrations of site are

denoted c+¿¿, c Na, and c Na0, respectively. For the ion exchange between cations in the solution and

the membrane we have:

S+¿+Na+¿⇆Na+¿+S+¿¿¿ ¿¿ (S:13)

K Na=K Na

'

K+¿=c+¿ cNa

cNa c+¿¿¿¿

(S:14)

where K Na is the equilibrium constant and K+¿=exp¿¿ and K Na' =exp {−(μNa

0 −μNa0 , w )/RT }. For the

binding of S0 to a site we have:

S0+Na+¿⇆Na+¿ S0

¿¿ (S:15)

K 0' =

cNa0

c Na c0 (S:16)

where the equilibrium constant K 0' =exp {−( μNa 0

0 −μNa0 −μ0

0 ,w ) /RT }. Provided that the equilibrium

distruibution of S+ between the membrane and the solution is given by the relationship

μ+¿0,w+RTln c+¿=μ+¿0+RTln

c+¿c−¿+F ∆∅ ¿¿ ¿

¿ ¿ (S:17)

51

984

985

986

987

988

989

990

991

992

993

994

995

996

997

998

999

1000

1001

101102

Page 52: digital.csic.esdigital.csic.es/bitstream/10261/131661/1/2015_Langmuir.docx · Web viewAll aqueous NaHA solutions for potentiometric and fluorescence measurements were prepared by

where c−¿=c+¿+cNa+cNa0 ¿¿, the electrostatic potential difference between the membrane and the solution

∆∅ can be written

∆∅= RTF

lnK+¿+

RTF ln c

+¿ c−¿

c+¿=RTF

lnK+¿+ RT

F ln¿ ¿¿¿¿ ¿ (S:18)

After using (S:14) and (S:16) we arrive at an expression resembling eq. (13):

∆∅=const .+ RTF

ln¿ (S:19)

where K 0' '=K Na cNa K0

' . Double prime is used to signify that K0' ' is a conditional equilibrium

constant dependent on the salt concentration. The net result of the mechanism is that S0 helps Na+

in replacing S+ in the membrane, which can also be seen by adding together (S:13) and (S:15):

S+¿+Na+¿ +S0⇆ Na+¿S0

+S+¿¿¿ ¿¿. The salt dependence of K0' ' appears not to be in agreement with

experiments.

S4.1 Experimentally determined CMC and CAC values

52

1002

1003

1004

1005

1006

1007

1008

1009

1010

1011

1012

1013

1014

1015

1016

1017

1018

1019

1020

103104

Page 53: digital.csic.esdigital.csic.es/bitstream/10261/131661/1/2015_Langmuir.docx · Web viewAll aqueous NaHA solutions for potentiometric and fluorescence measurements were prepared by

Figure S7. Examples of fluorescence spectra of pyrene in aqueous MKM solutions without (H2O) and with added NaCl (csalt = 0.05 M). Numbers 1-6 correspond to increasing MKM concentration from below (1-3) to above the CMC (4-6).

Table S2. CMC values in pure MKM and CAC values in mixed NaHA-MKM solutions

determined by the SSE based potentiometric titrations (CMC(SSE) and CAC(SSE)) and by

fluorescence measurements (CMC(FM)) at various salt (NaCl) concentrations, csalt.

csalt (mol/L) CMC(SSE) (mol/L) CMC(FM) (mol/L) CAC (SSE) (mol/L)

0 4.9 10-4 4.9 10-4 3.9 10-5

0.01 3.1 10-4 2.5 10-4 7.6 10-5

0.05 1.6 10-4 8.3 10-5 1.1 10-4

0.1 1.1 10-4 7.4 10-5 1.1 10-4

53

1021

102210231024

1025

1026

1027

1028

105106

Page 54: digital.csic.esdigital.csic.es/bitstream/10261/131661/1/2015_Langmuir.docx · Web viewAll aqueous NaHA solutions for potentiometric and fluorescence measurements were prepared by

0

5

10

CM

C (m

M)

MKM: SSE MKM: fluorescence SDS: Rosen CPC: Kogej

10-3 10-2 10-1

0.1

1

10

CM

C (m

M)

I (M)

Figure S8. The dependence of the critical micelle concentration, CMC, on the ionic strength, I (= csalt + CMC) in aqueous MKM solutions. For comparison, CMC data for SDS 14 and CPC15 are included. Above: linear axis for CMC and logarithmic axis for I. Below: a double logarithmic plot.

54

1029

1030

1031103210331034

107108

Page 55: digital.csic.esdigital.csic.es/bitstream/10261/131661/1/2015_Langmuir.docx · Web viewAll aqueous NaHA solutions for potentiometric and fluorescence measurements were prepared by

S5.

Figure S9. Left panel: SSE potentiometric titration curves for pure MKM (full squares) and for MKM in the presence of NaHA (cp = 5 10-4 mol/L; open circles) in aqueous NaCl solutions: the NaCl concentration (csalt) is indicated in the Figures. Right panel: Theoretical titration curves calculated as described in Section 3.2. The curves have been adjusted in the vertical direction by adding a constant to the calculated E to coincide with the experimental data at the lowest surfactant concentration. E*/mV: 236 (0.1 M), 248 (0.05 M), 257 (0.01 M), 283 (0.0005 M).

55

1035

1036

103710381039104010411042

109110

Page 56: digital.csic.esdigital.csic.es/bitstream/10261/131661/1/2015_Langmuir.docx · Web viewAll aqueous NaHA solutions for potentiometric and fluorescence measurements were prepared by

S6. Discussion about CMC values determined by NMR

The results from the theoretical calculations show that the fraction of surfactant in micelles

decreases with decreasing salt concentration and with that the extent to which C+ is affected by

the micellar equilibrium. To check the consequences of that for the possibility to determine the

CMC using the NMR self-diffusion method we used the data in Fig. 6 (at 0.1 M salt) and the

corresponding data for 0.0005 M salt case to calculate the expected observable self-diffusion

coefficients (Dobs). The result showed that an accurate estimate should be possible at the higher

salt concentration but not at the lower salt concentration where the variation of Dobs in the

vicinity of CMC is too small on an absolute scale. In the calculations we used the relationship

Dobs= p Dmic+(1− p)D1 11 where p, Dmic (= 1×10-10 m2/s) and D1 (= 4×10-10 m2/s) represent the

fraction of surfactant in micelles, and the self-diffusion coefficient for micelles and monomers,

respectively.

References

1. Perez, L.; Pinazo, A.; Teresa García, M.; Lozano, M.; Manresa, A.; Angelet, M.; Pilar

Vinardell, M.; Mitjans, M.; Pons, R.; Rosa Infante, M. Cationic surfactants from lysine:

Synthesis, micellization and biological evaluation. Eur J Med Chem 2009, 44 (5), 1884-1892.

2. Dumitriu, S. Polysaccharides: Structural Diversity and Functional Versatility. Marcel

Dekker: New York, 1998.

56

1043

1044

1045

1046

1047

1048

1049

1050

1051

1052

1053

1054

1055

1056

1057

1058

1059

1060

1061

1062

1063

1064

111112

Page 57: digital.csic.esdigital.csic.es/bitstream/10261/131661/1/2015_Langmuir.docx · Web viewAll aqueous NaHA solutions for potentiometric and fluorescence measurements were prepared by

3. Hansson, P. Self-assembly of ionic surfactants in polyelectrolyte solutions: A model for

mixtures of opposite charge. Langmuir 2001, 17, 4167.

4. Hansson, P. Phase behavior of aqueous polyion-surfactant ion complex salts: A

theoretical analysis. Journal of Colloid and Interface Science 2009, 332, 183-193.

5. Gernandt, J.; Hansson, P. Hysteresis in the surfactant-induced volume transition of

hydrogels. Journal of Physical Chemistry B 2015, 119 (1717), 1725.

6. Hansson, P. Surfactant self-assembly in oppositely charged polymer networks. Theory. .

Journal of Physical Chemistry B 2009, 113, 12903-12915.

7. Hansson, P.; Månsson, R.; Bysell, H.; Malmsten., M. Peptide-microgel interactions in the

strong coupling regime. Journal of Physical Chemistry B 2012, 116, 10964-10975.

8. Moreira, A. G.; Netz, R. R. Binding of Similarly Charged Plates with Counterions Only.

Phys Rev Lett 2001, 87, 0708301-078304.

9. Naji, A.; Jungblut, S.; Moreira, A. G.; Netz, R. R. Electrostatic interactions in strongly

coupled soft matter. Physica A 2005, 352, 131-170.

10. Jönsson, B.; Wennerström, H. Phase Equilibria in a Three-Component Water-Soap-

Alcohol System. A Thermodynamic Model. Journal of Physical Chemistry 1987, 91, 338-352.

11. Evans, F. D.; Wennerström, H. T. The Colloidal Domain: Where Physics, Chemistry,

Biology, and Technology Meet. 2 ed.; Wiley-VCH Verlag GmbH: New York, 1999; p 672.

12. Hansson, P.; Jönsson, B.; Ström, C.; Söderman, O. Determination of micellar aggregation

numbers in dilute surfactant systems with the fluorescence quenching method. Journal of

Physical Chemistry B 2000, 104, 3496-3506.

13. T.Wallin; Linse, P. Monte Carlo Simulations of Polyelectrolytes at Charged Micelles. 3.

Effects of Surfactant Tail Length. Journal of physical Chemistry B 1997, 101, 5506-5513.

14. Rosen, M. J. Surfactants and Interfacial Phenomena. 2 ed.; John Wiley: New York, 1989.

57

1065

1066

1067

1068

1069

1070

1071

1072

1073

1074

1075

1076

1077

1078

1079

1080

1081

1082

1083

1084

1085

1086

1087

1088

113114

Page 58: digital.csic.esdigital.csic.es/bitstream/10261/131661/1/2015_Langmuir.docx · Web viewAll aqueous NaHA solutions for potentiometric and fluorescence measurements were prepared by

15. Škerjanc, J.; Kogej, K.; Vesnaver, G. Polyelectrolyte-Surfactant Interactions. Enthalpy of

Binding of Dodecyl- and Cetylpyridinium Cations to Poly(styrenesu1fonate) Anion. Journal of

Physical Chemistry 1988, 92, 6382-6385.

58

1089

1090

1091

1092

1093

115116