+ All Categories
Home > Documents > Adsorption of Surfactants at the Solid-Liquid Interface: A ...

Adsorption of Surfactants at the Solid-Liquid Interface: A ...

Date post: 02-Dec-2021
Category:
Upload: others
View: 7 times
Download: 1 times
Share this document with a friend
71
Adsorption of Surfactants at the Solid-Liquid Interface: A Quartz Crystal Microbalance study Johan J.R. Stålgren Doctoral Thesis 2002 Department of Chemistry, Surface Chemistry Royal Institute of Technology Stockholm, Sweden
Transcript
Page 1: Adsorption of Surfactants at the Solid-Liquid Interface: A ...

Adsorption of Surfactants at the

Solid-Liquid Interface:

A Quartz Crystal Microbalance study

Johan J.R. Stålgren

Doctoral Thesis 2002

Department of Chemistry, Surface Chemistry

Royal Institute of Technology

Stockholm, Sweden

Page 2: Adsorption of Surfactants at the Solid-Liquid Interface: A ...

Akademisk avhandling som med tillstånd av Kungliga Tekniska Högskolan framlägges tilloffentlig granskning för avläggande av filosofie doktorsexamen, tisdagen den 29 januari,2002, kl. 09.00 i Kollegiesalen, Valhallvägen 79, KTH, Stockholm.

Address to the author:Johan J.R. Stålgren

Department of Chemistry, Surface ChemistryRoyal Institute of Technology

SE-100 44 StockholmSweden

ISSN 1650-0490ISBN 91-7283-238-XTRITA YTK-0201

Copyright 2002 by Johan J.R. Stålgren. All rights reserved. No part of this thesis may bereproduced without permission from the author.Other copyrighted material is used with permissionPaper I 2001 by the American Chemical SocietyPaper II 2001 by the Elsevier Science

Page 3: Adsorption of Surfactants at the Solid-Liquid Interface: A ...

Abstract

This thesis is concerned with the interfacial behaviour of surfactants at solid-liquid interfaces.

The main technique used for the adsorption measurements during this thesis work was the

Quartz Crystal Microbalance-Dissipation (QCM-DTM). This technique allows both the

adsorbed amount, as evaluated as a change in frequency (∆f), and the change in the dissipation

factor (∆D) that is a measure of the energy dissipated in the system, to be determined

simultaneously. Methods like null ellipsometry already exist, and they measure the amount

adsorbed to a planar, reflecting surface accurately, and the thickness of the adsorbed layer

may also be determined. The QCM-DTM technique has, however, some advantages. For

instance, the quartz crystal can be coated (physical/chemical) in a large number of ways. In

addition, simultaneous measurement of the dissipation factor allows another parameter to be

determined, this parameter is a measure of the interaction between the adsorbed surfactant

layers and the bulk solution. Further, opaque or even non-transparent solutions can be studied

with the QCM-DTM, which is not possible with the ellipsometer.

When the project began, an aim was to investigate the visco-elastic properties of

polyelectrolytes at different surfaces. This turned out to be more complex than we expected so

I decided to use a less complex systems in order to more fully understand the results. Hence,

the choice became to study surfactant adsorption, a topic which is well documented before by

several other techniques. The choice was based on the surfactants low molecular weight, and

the relatively simple distribution of a polar (hydrophilic) and non-polar (hydrophobic) part,

and a significant general knowledge about their interfacial behaviour.

The methodology for adsorption studies in liquid for the QCM-DTM was only in its infancy, so

parameters like temperature dependence, surface roughness, surface modification and

cleaning had to be kept under control or developed at the same time. Systematic surfactant

adsorption studies from liquids with the QCM technique do not exist. Hence, the aim of this

thesis was to achieve an understanding of the information provided by measured shifts in

frequency and dissipation factor for such systems, and from this draw conclusions about the

interfacial behaviour of both non-ionic and cationic surfactants. Further I aimed to learn how

valuable the QCM-DTM technique was for these systems and what pitfalls there are in

evaluating the results observed with this technique.

Page 4: Adsorption of Surfactants at the Solid-Liquid Interface: A ...

Sammanfattning

Den här avhandlingen handlar om det beteende som uppträder hos tensider, vid gränsytan

mellan fast fas och en vätska. Vid adsorptions mätningarna i denna avhandling har

huvudsakligen en teknik använts, en kvarts kristall mikrobalans (QCM-DTM). Denna teknik

tillåter både den adsorberade mängden, utvärderat genom skiftet i frekvens (∆f), samt

förändringen i dissipations faktorn (∆D), som är ett mått på systemets energi dämpning, att

uppmätas. Båda dessa faktorer kan uppmätas samtidigt. Metoder som noll ellipsometri

existerar redan och de mäter adsorptionen vid en plan, reflekterande yta. Denna teknik kan

också bestämma tjockleken hos det adsorberade skiktet. QCM-DTM tekniken har trots allt

några fördelar gentemot detta. Till exempel, kan kvarts kristallerna beläggas (fysiskt/kemiskt)

på flera olika sätt. En annan fördel är den samtidigt uppmätta dämpnings faktorn som tillåter

en parameter till att bestämmas. Dämpnings faktorn, är en parameter som gör att man kan

mäta styrkan hos interaktionerna mellan det adsorberade tensid skiktet och bulk lösningen.

Även lösningar som ej är genomskinliga kan studeras med QCM-DTM tekniken, detta är inte

möjligt med en teknik som till exempel ellipsometri.

När projektet började, var vårt mål att studera de viskoelastiska egenskaperna hos laddade

polymerer vid olika ytor. Detta visade sig vara mycket mer komplext än vad vi hade förväntat

oss, så vi bestämde oss för att använda ett mindre komplext system för att fullständigt förstå

våra resultat. Vårt val föll på olika tensid lösningar, ett ämne som är väl dokumenterat sedan

innan med flera olika tekniker. Valet grundade sig på tensidernas låga molekylära vikt samt

deras relativt enkla distribution av polära (hydrofila) och o-polära (hydrofoba) delar.

Dessutom fanns sedan innan en signifikant kunskapsbas om deras beteende vid gränsytan

mellan en fast fas och en vätska.

Eftersom metodologin för adsorptions studier i vätskor för QCM-DTM var bara i början på sin

utveckling behövdes faktorer som temperatur beroendet, ytråhet, ytmodifikation samt

rengöring kontrolleras samt utvecklas under tiden. Systematiska studier av tensid adsorption i

en vätska finns inte för QCM tekniken. Därav valet på avhandlingens innehåll, en ökad

förståelse av informationen för dessa system, genom studier av ändringen i frekvens samt

förändringen i dämpning. Samtidigt siktade jag på att lära mig hur värdefull QCM-DT M

tekniken var för dessa system, och vilka fallgropar det finns när man utvärderar resultaten från

denna teknik.

Page 5: Adsorption of Surfactants at the Solid-Liquid Interface: A ...

There are two kinds of scientists,

physicists and stamp collectors.

Ernest Rutherford (1871-1937)

Page 6: Adsorption of Surfactants at the Solid-Liquid Interface: A ...
Page 7: Adsorption of Surfactants at the Solid-Liquid Interface: A ...

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11.1. List of paper . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .31.2. Summary of papers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5

2. Surfaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 92.1. Gold . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .112.2. Self-Assembled Monolayers (SAMs) . . . . . . . . . . . . . . . . . . . . . . . . . 132.3. Silica . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 152.4. Silan-coated surfaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .16

3. Methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 183.1. Profilometer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .183.2. Contact Angle . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 203.3. X-ray Photoelectron Spectroscopy (XPS) . . . . . . . . . . . . . . . . . . . . . .223.4. The Quartz Crystal Microbalance-Dissipation (QCM-DTM) . . . . . . . . 23

4. Surfactants . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 344.1. Lipids . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. . . . . . . . . . . . . . . . . . . .344.2. Cationic surfactans . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 374.3. Non-ionic surfactants . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38

5. Results and Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 395.1. Adsorption of emulsion studied with the QCM-DTM . . . . . . . . . . . . . . 395.2. Polymer adsorption on phospholipid coated surfaces . . . . . . . . . . . . . 415.3. Adsorption of surfactants studied with the QCM-DTM . . . . . . . . . . . . .475.4. Counterion effects on sensed mass and energy dissipation . . . . . . . . . 495.5. Bound / trapped water . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .52

6. Concluding remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55

Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56

References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .57

Page 8: Adsorption of Surfactants at the Solid-Liquid Interface: A ...

0

Page 9: Adsorption of Surfactants at the Solid-Liquid Interface: A ...

1

1. Introduction

The Quartz Crystal Microbalance-Dissipation (QCM-DTM) technique is an ultra sensitive

weighing device based on the piezoelectric, electromechanical oscillator principle. It consists

of a thin single-crystal quartz disk, with one metal electrode deposited on each side. When the

electrodes are connected to an electric oscillator, the crystal can be made to oscillate in a very

stable manner at its resonance frequency, f. When a mass is adsorbed on one or both of the

electrodes, then this leads to a change in the resonance frequency of the quartz crystal, ∆f. If

the adsorbed mass is small compared to the mass of the quartz crystal and there is no slip or

deformation due to the oscillatory motion, then the resonance frequency decreases

proportionally to the mass of the adsorbed film according to the Sauerbrey relation. It is

possible to determine very small changes of the resonance frequency and hence very small

mass changes. This is possible since the QCM generally has very stable oscillations. In

addition to the adsorbed mass, simultaneous measurements of the change in dissipation factor,

(∆D), which is a measure of the energy dissipated in the system, is possible. Hence, this

parameter is a measure of the interaction between the adsorbed layer and the bulk solution.

This thesis is concerned with the interfacial behaviour of surfactants at solid-liquid interfaces.

Emphasis is placed on the adsorption / desorption of three different groups of surfactants;

cationic, non-ionic, and phospholipid surfactant.

To choose the surfactants to study was not easy, but it had to be surfactants with properties

which were well documented before by several different techniques. Further, to be able to

systematically vary the surfactant structure was regarded as important. I have used different

model surfaces to study the effect of the underlying substrate. The model surfaces also had to

be well characterized with several different techniques. The different model surfaces we

decided to work with were a metal (gold), silica, methylated silica, and several different self-

assembly monolayers on gold substrates. These surfaces had to be thoroughly evaluated to be

valuable for the QCM experiments, without adding more unknown parameters to the

interfacial study. In chapter 3 all the techniques used for evaluating both the surfaces and the

surfactants are discussed.

Systematic adsorption studies of surfactants from liquids using the QCM technique do not

exist. Hence, the aim of this thesis was to achieve an understanding of the information

Page 10: Adsorption of Surfactants at the Solid-Liquid Interface: A ...

2

provided by the measured shifts in frequency and dissipation factor for such system, and from

this draw conclusions about the interfacial behaviour of both non-ionic, cationic and

phospholipid surfactants. Last in the summary the main findings during these experiments,

and a hopefully valuable discussion of the different results obtained, is presented. More

details can be found in the manuscripts that constitute the second part of this thesis.

Page 11: Adsorption of Surfactants at the Solid-Liquid Interface: A ...

3

List of papers

This thesis consists of a summary and six papers. The papers are listed below and are in the

summary referred to by their Roman numerals (I to VI).

I Adsorption of Liposomes and Emulsions Studied with a Quartz Crystal

Microbalance.

Johan J.R. Stålgren, Per M. Claesson and Torbjörn Wärnheim

Advances in Colloid and Interface Science, 2001, 89-90, 383.

II Adsorption of a PEO-PPO-PEO Triblock Copolymer on Small Unilamellar

Vesicles: Equilibrium and Kinetic Properties and Correlation with

Membrane Permeability.

Markus Johnsson, Nill Bergstrand, Katarina Edwards and Johan J.R. Stålgren

Langmuir, 2001,17, 3902.

III Cationic and Non-ionic Surfactant Adsorption on Thiol Surfaces with

Controlled Wettability.

Katrin Boschkova and Johan J.R. Stålgren

Submitted to Langmuir.

IV A Correlation between Adsorbed Amount and Frictional Properties of Thin

Gemini Surfactant Films - CPP in Relation to Friction.

Katrin Boschkova, Adam Feiler, Bengt Kronberg and Johan J.R. Stålgren

Submitted to Langmuir.

V A Comparative Study of Surfactant Adsorption on Model Surfaces using

the Quartz Crystal Microbalance and the Ellipsometer.

Johan J.R. Stålgren, Jonny Eriksson and Katrin Boschkova

Submitted to Journal of Colloid and Interface Science.

Page 12: Adsorption of Surfactants at the Solid-Liquid Interface: A ...

4

VI Lubrication in Aqueous Solutions Using Cationic Surfactant –

a Study of Static and Dynamic Forces.

Katrin Boschkova, Bengt Kronberg, Johan J.R. Stålgren, Karin Persson, and

Monica Ratoi-Salagean

Accepted in Langmuir.

The papers are reproduced with permission from the publishers.

The author’s contribution to the papers is as follows:

I Major part of planning, experiments and evaluation.

II Part of planning, experiments and evaluation.

III Major part of planning and experiments, part of evaluation.

IV Part of planning, experiments and evaluation.

V Major part of planning, experiments and evaluation.

VI Part of planning, experiments and evaluation.

In all papers, I have been the main responsible for the QCM work and evaluation.

Page 13: Adsorption of Surfactants at the Solid-Liquid Interface: A ...

5

1.2 Summary of papers

Phospholipid adsorption at the solid-liquid interface.

The first two papers deal with adsorption of phospholipids at a gold surface, and effects of

additives. The results are important for comprehending the data obtained with the Quartz

Crystal Microbalance-Dissipation (QCM-DTM), in a useful way.

Paper I deals with adsorption from phospholipid liposome solutions (1.2%) and phospholipid

stabilised oil-in-water emulsions (20% purified soybean oil) with the same phospholipid

concentration. The main attention in the paper was given to the adsorption process at a gold

surface and the effect of repeated injections of the same solution. The second aim was to learn

how the dilution of the bulk solution affected the adsorbed layer and to determine what

remained on the surface after the dilution step was completed. The adsorption from the

liposome solution resulted in formation of a phospholipid bilayer with an additional and

incomplete outer layer of liposomes. The outer layer was removed by dilution leaving a

bilayer of phospholipids on the surface. The adsorption process observed from the

concentrated emulsion solution was considerably more complex. A slow spreading process

that also resulted in some expulsion of material from the interface followed the rapid initial

adsorption of emulsion droplets. After rinsing with water a phospholipid monolayer was

retained on the surface.

Paper II is devoted to the adsorption of the triblock copolymer F127, poly(ethylene oxide)-

poly(propylene oxide)-poly(ethylene oxide), EO98PO67EO98, onto immobilized small

unilamellar vesicles (SUVs) of egg phosphatidylcholine (EPC). With the QCM-DTM technique

we first showed that SUVs of EPC adsorb on gold to form a monolayer of vesicles. This

supported monolayer of vesicles was then used to follow the adsorption of the F127 polymer

onto the lipid vesicle membrane surface. The adsorption of F127 was found to be a rapid

process and the measured polymer binding isotherm was fitted to a Freundlich type of

isotherm. The maximum, or plateau, adsorbed amount was determined to be of a magnitude

similar to that found for adsorption of F127 on hydrophobic surfaces. Furthermore, the

desorption of the triblock copolymers from the membrane surface was followed after rinsing

the SUV monolayer with pure buffer. It was found that the desorption process displayed

essentially the same rapid kinetics as the adsorption process, indicating a weak interaction

between the polymers and the lipid membrane. The determined polymer binding isotherm was

Page 14: Adsorption of Surfactants at the Solid-Liquid Interface: A ...

6

used to correlate the adsorbed amount of polymer with the polymer-induced leakage of

carboxy fluorescein (CF) from the SUVs. It was found that the membrane permeability was

increased severalfold already at low surface coverage, and that the maximum magnitude of

the CF release rate was obtained at, or close to, the F127 concentrations giving rise to

maximum adsorbed amount of polymer. In addition, the increased membrane permeability

induced by the triblock copolymers was compared with the effect of adding a conventional

ethylene oxide (EO)-surfactant, Triton X-100, to the SUVs. The result emphasizes the

dramatic effect of F127 on the bilayer permeability. Another interesting result was that the

stability of the liposomes used in this study was considerably higher compared to those

formed by the phospholipids mixtures employed in paper I.

Surfactant adsorption at model surfaces.

In papers III-VI we started to modify our surfaces with silica, methylated silica and several

different self-assembled monolayers (SAMs). The ionic surfactants used were the cationic,

DTAB (dodecyltrimethylammonium bromide), DDAB (didodecyldimethylammonium

bromide), and gemini surfactants having the same headgroup and chain length as DTAB but

with the additional feature that two headgroups were chemically connected with a spacer of

different length. The non-ionic surfactants used were the poly(ethylene oxide) monoalkyl

ethers C14EO6 and C12EO8(Octa-(ethylene oxide) mono n-dodecyl ether).

In paper III we showed that thiolated surfaces work very well as model substrates in

adsorption measurements using the QCM-DTM. Functionalised SAMs were prepared from

mixtures of hydrophobic, SH-C16 (thiohexadecane) and hydrophilic, SH-C16OH

(thiohexadecanol) terminated thiols, which allowed the interfacial energy of the surfaces to be

changed in a systematic way. The prepared thiol surfaces were used as substrates for

adsorption of a cationic, DTAB, and a non-ionic, C12EO8, surfactant. The experiments showed

that when the fraction of methyl groups at the surfaces was increased, the adsorption of both

DTAB and C12EO8 is increased. In particular, there is a transition from a micellar surfactant

layer to a surfactant monolayer at 25% to 50% surface coverage of SH-C16 groups with

monolayers being formed at higher coverage of SH-C16. In addition, the role of the counterion

in the adsorbed surfactant layer for the charged surfactant was discussed in terms of its

contribution to the mass and visco-elastic response determined by the quartz crystal

microbalance.

Page 15: Adsorption of Surfactants at the Solid-Liquid Interface: A ...

7

With Paper IV where we used the Gemini surfactants, we showed that by changing the length

of the spacer group from 3 to 12 a systematic change in the molecular packing at a gold

surface was obtained. Furthermore, the molecular packing was shown to correlate to the

frictional behavior of the surfactant film. An increasing length of the spacer group resulted in

lower, adsorbed amount and less good frictional properties. This is discussed in terms of the

critical packing parameter (CPP) of the surfactant and a relation between CPP and frictional

behavior is proposed. The results can be viewed upon either as controlled by the rigidity of

the surfactant layer or as a result of defects, holes, in the lubricating film. No correlation

between spacer length and viscoelasticity of the adsorbed surfactant layer was detected using

the QCM-DTM. This indicates that the resolution of the dissipation factor from QCM-DTM

measurements is not sufficient to describe the viscoelastic character of the thin surfactant

film. The degree of counterion-binding to charged surfactant films is a difficulty encountered

when converting the frequency response of the crystal to packing density. This problem is

again highlighted and discussed (see also paper III).

In Paper V we investigated the adsorption behaviour of hexa-ethylene oxide mono n-

tetradecyl ether (C14EO6), on different model surfaces. This investigation was conducted with

two different techniques, the QCM-DTM and the ellipsometer. The adsorbed amount of the

non-ionic surfactant was determined both at hydrophilic and hydrophobic surfaces. In

particular, the substrates employed were; hydrophilic silica, hydrophobized silica (using

dimethyldichlorosilane), hydrophobized gold surfaces (using 10-thiodecane and 16-

thiohexadecane). We showed that the frequency shift obtained from the QCM-DT M

experiments results in an overestimation of the adsorbed mass. This is attributed to two

different effects, viz, hydrodynamic coupling of water to the adsorbed surfactant layer and

secondly, trapped water within the adsorbed surfactant layer. Furthermore, from the

ellipsometry data the adsorbed layer thickness was determined. By combining the thickness

information and the dissipation parameter (obtained from the QCM-DTM experiments), we

again noted that the dissipation parameter was insufficient in describing the visco-elastic

character of thin surfactant films.

Paper VI is devoted to lubrication in aqueous surfactant systems where the surfactants adsorb

at surfaces in relative motion forming either a surfactant monolayer or a multi (liquid

crystalline) layer. The surfactants were of two kinds, viz., a double chain cationic surfactant,

Page 16: Adsorption of Surfactants at the Solid-Liquid Interface: A ...

8

didodecyldimethylammonium bromide, DDAB, and a single chain cationic surfactant,

dodecyltrimethylammonium bromide, DTAB. Excellent film forming capability was shown

for DDAB. We interpret this as being due to good packing of the surfactant molecules at the

surfaces, i.e. the inherent ability of these surfactant molecules to form liquid crystalline

structures at the surface, results in good load carrying capability. This is also reflected in the

bulk properties of the surfactants, where DDAB shows lamellar liquid crystalline phases at

concentrations much lower than DTAB, which does not show good lubrication properties.

The results were discussed in terms of film stability of a surfactant layer adsorbed at the

surface, which in turn is correlated to the critical packing parameter of the surfactant. The

systems were characterized using (i) the surface force apparatus determining the interaction

forces between the adsorbed layers at the surfaces, (ii) the EHD-rig (Elastohydrodynamic-rig)

determining film formation under shear. The adsorption kinetics and composition at the

surface were determined by QCM-DTM and X-ray photoelectron spectroscopy.

Page 17: Adsorption of Surfactants at the Solid-Liquid Interface: A ...

9

2. Surfaces

A controlled surface/environment is required when studying interfacial properties of

surfactant molecules. Otherwise it is impossible to interpret the experimental results.

Model surfaces, where chemical and structural properties can be controlled, are not readily

found and in most cases one has to prepare and characterize them oneself.

Surfaces can be characterized and classified in many different ways. The surface topology

allows classification into “rough” and “smooth” as quantified by e.g. the root mean square

roughness, Rq. Other classifications can be based on the chemical composition of the surface,

the surface energy, or the wetting properties. The latter classification is particularly suitable

when studying surfactant adsorption from aqueous media. The wetting properties can also

very conveniently be quantified by the contact angle, θ, of the liquid on the solid, i.e. cosθ =

(γSV- γSL) / γLV. The cosine of the contact angle is thus given by the difference in surface

energy between the solid-vapour (γSV) and solid-liquid (γSL) interface, normalized by the

liquid-vapour interfacial tension (γLV). Generally, high energy solids have by definition a high

value of γSV, and in most cases, a much lower interfacial tension against water due to the

hydrogen bonding capability and high dipolar moment of the water molecule. The contact

angle of water on such surfaces is low. Based on the contact angle one can classify a given

liquid on a given surface as completely wetting θ = 0, partly wetting (0 < θ ≤ 90°) or non-

wetting (θ ≥ 90°). When the liquid is water one normally talks about hydrophilic and

hydrophobic surfaces, but there is no general agreement about what contact angle the surface

is required to have in order to be classified as “hydrophilic” or “hydrophobic”. In this thesis

we use the general term “hydrophilic” for surfaces having a low contact angle, and

“hydrophobic” for surfaces with high contact angle. The quantitative measure of the wetting

behaviour is provided by the contact angle. We note that the contact angle is extremely

sensitive to the surface composition and sub monolayer adsorption of hydrophobic

compounds is easily detected. In fact, in many cases simple contact angle measurement is a

more sensitive probe of adsorption than sophisticated XPS (X-ray Photoelectron

Spectroscopy) analysis. However, of course, the contact angle does not give the same

chemical information as the XPS-spectra.

Page 18: Adsorption of Surfactants at the Solid-Liquid Interface: A ...

10

In this work we have varied the wetting properties from that of hydrophilic silica to that of

hydrophobic alkane thiol SAM on gold surfaces. Some of the properties of these surfaces are

described more extensively below. The wetting properties of the surfaces have been

characterized using contact angle measurement; the surface composition has been determined

employing x-ray photoelectron spectroscopy (XPS, ESCA). The topological character of the

surfaces has been determined by the profilometer for surface roughness effects and for some

surfaces, the scanning electron microscope (SEM) looking for eventual defects. Some data

can be found in table 1, where we have summarized the characteristics of our model surfaces.

Surface Ra [nm] Rq / Ra θθθθ

Silicaellipsometer 1.3 ± 0.1 1.5 ± 0.3 < 20°

Dimetyldichlorosilaneellipsometer 1.1 ± 0.2 1.5 ± 0.5 101° ± 1°

SilicaQCM 1.3 ± 0.2 1.5 ± 0.3 < 20°

DimetyldichlorosilaneQCM 1.0 ± 0.1 1.4 ± 0.3 101° ± 1°

GoldQCM 1.4 ± 0.1 6.0 ± 3 ≈ 30° ± 5°

ThiohexadecaneQCM 1.2 ± 0.1 3.0 ± 1.0 103° ± 3°

ThiodecaneQCM 1.2 ± 0.1 3.0 ± 1.0 91° ± 3°

ThiohexadecanolQCM 1.2 ± 0.1 3.0 ± 1.0 20° ± 2°

Table 1: Characteristics of our model surfaces, where silicaellipsometer means a silica surface for ellisometry studies

and silicaQCM means a quartz crystal coated with silica for measurements with the QCM.

Page 19: Adsorption of Surfactants at the Solid-Liquid Interface: A ...

11

2.1 Gold.

The electrodes on the quartz crystal are made of gold, and in many cases this has been a

suitable surface to conduct some of our adsorption studies with. The water contact angle on

the gold surface obtained after cleaning (see table 1) was low indicating a low degree of

contamination, and this surface will henceforth be classified as “ hydrophilic gold” to

distinguish it from a gold surface that has been exposed to air for a prolonged time and

appears hydrophobic due to adsorption of contaminants. Gold is a material that has a partly

filled electron band in its ground state. This means that it has both empty states and electrons

in the valence band. The electronegativity1 for gold is 2.41, which is very high for being a

metal. The electronic configuration for gold is:

1s22s22p63s23p63d104s24p64d105s25p64f145d106s1

The 6s electrons move around freely in the gold crystal, whereas the 5d electrons are tightly

bound to the nucleus. These 6s electrons play a major role for the chemical bonding to gold

atoms in the surface layer. In the ideal crystalline structure of gold, the atoms are located

themselves in a face centered cubic (fcc)1 lattice, which means that the unit cell consists of a

cube with one atom in each corner and one at the centre of each side. Each atom is thus in

contact with 12 others. The fcc structure provides maximum number of nearest neighbours

and is thus the preferred structure of crystalline materials of spherical molecules or individual

atoms. However, the bulk order has to end somewhere near the surface, for gold it prefers to

end in a structure called the (111) surface1 of an fcc crystal. It gives the closest packing of

atoms in the surface layer. Each surface atom has 6 neighbours on the surface. Gold are in

practice a polycrystalline material2, where a lot of small single crystals are joined together,

and the borders between the crystals are very far from the perfect (111) surface. At a

hydrophilic gold surface, there are a lot of unpaired electrons, and they are highly reactive2.

This leads to a fast contamination of a clean surface exposed to air. To keep the surface clean

you can either store it in ultra-high vacuum and never let it come in contact with

contaminations (solid-gas), or you could clean it in-situ under clean solvent conditions (solid-

liquid). Since we are doing all our experiments at the solid-liquid interface, we have chosen to

clean our surfaces in-situ, and keep the exposure to air contaminants to a minimum This is

very demanding since you have to keep all other surfaces in the experimental setup equally

Page 20: Adsorption of Surfactants at the Solid-Liquid Interface: A ...

12

clean and exposed to a liquid that is as pure as the liquid in contact with the cleaned gold

surface to prevent a contamination transport from the other surfaces to the clean (highly

reactive) gold surface. It exists several other techniques to clean surfaces, but most of them

are still dependent on the environment you do your cleaning and experiments in.

Page 21: Adsorption of Surfactants at the Solid-Liquid Interface: A ...

13

2.2 Self-Assembled Monolayers (SAMs).

The self-assembly of molecules at the solid-liquid interface has been an area of growing

interest. There are different combinations of surfaces and molecules that form SAMs and

many of them rely on the strong interaction between one part of the assembling molecules and

the surface, in addition to the interaction between the molecules within the monolayer. Long-

chain organosulphur compounds on noble metals, such as silver, platinum and in our case

gold, can be used for forming SAM coated surfaces with stable monomolecular (monolayer)

films. It is a versatile preparation technique, and SAM coated surfaces have served as model

systems in a number of applications, such as for biosensors, biomaterials, anti-corrosion

agents and lubrication1-2. In 1983 Nuzzo and Allara3 published the first observations of

organic disulphides that formed monolayers on gold from solutions as studied with infrared

spectroscopy. The general picture for the chemisorption of alkane thiols on gold is that the

thiol moiety adsorbs in a three-fold hollow site at the Au(111) lattice whereupon it loses its

hydrogen atom to become a thiolate. This gives an area of 21.4 Å2 per thiol molecule. SAM

coated surfaces have been studied with a number of techniques, including x-ray photoelectron

spectroscopy (XPS, ESCA), infrared spectroscopy, scanning tunnelling microscope (STM),

electrochemistry, ellipsometry, contact angles and various diffraction methods, see for

instance the review by Ulman1. The experimental findings4-5 strongly support the model

proposed for the structure of SAMs, and so do theoretical calculations6. However, the detailed

quantum mechanical processes are not completely understood. The structure of SAMs has

been determined by infrared reflection absorption spectroscopy (IRAS). The distance between

the sulphur atoms on the gold surface is slightly larger than the closest possible separation

between two alkyl chains, allowing an approximately 28-30° tilt of the chains in the layer to

increase van der Waals interaction between the chains, as confirmed both experimentally7 and

theoretically8.

Whereas the adsorption of molecules on the surface is fast (minutes) the self-assembling

process into an ordered monolayer is quite slow (hours)9. After a few minutes alkane thiol

molecules forming an almost fully covered layer have been adsorbed, but the order in the

layer is low. For one of the thiols used by us, the SH-(CH2)15CH3 thiol, different groups have

obtained very different kinetics for forming a well ordered SAM, ranging from seconds to

hours10. The concentration of the thiol in the liquid from which the SAM is formed is, of

course, important to take into account when investigating SAM kinetics. At low

Page 22: Adsorption of Surfactants at the Solid-Liquid Interface: A ...

14

concentrations a diffusion limited adsorption kinetics has been observed4, whereas at high

concentrations the time limiting step in the process is the actual surface attachment1.

The stability of alkane thiol SAMs when immersed in a solution containing the SAM

molecules is excellent. It is a stationary system, i.e. there is a continuous exchange of

molecules from the monolayer at the surface to the bulk solution. If the bulk solution contains

another alkane thiol, the surface monolayer will be exchanged by the new alkane thiol from

the bulk solution. This exchange could take from hours to days depending on which SAMs

that are involved10. Even though molecules in the SAM can be exchanged for other SAMs, the

desorption of the SAM in contact with a SAM-free solution is very slow due to the interaction

with the surface and within the tightly packed layer. It is very easy to prepare mixed SAMs by

just mixing the adsorbing species in the solution (see paper III). The surface composition will

be highly dependent on properties like chain length, terminal functionality and solubility11.

The preparation procedure of thin monolayers of alkanethiols on surfaces is relatively easy.

One dissolves the film forming molecules in an appropriate solvent (in our case ethanol) to a

rather low concentration, for our SH-C16 a 1mM concentration is enough. A previously

cleaned gold surface is then immersed in the solution for approximately 24 hours. When the

surface is withdrawn from the solution it is rinsed and then put into pure solvent for

approximately a week in order to dissolve any loosely bound, physisorbed, molecules that

might be attached to the monolayer.

Page 23: Adsorption of Surfactants at the Solid-Liquid Interface: A ...

15

2.3 Silica.

The gold coated quartz crystals used in the QCM can be further modified with an evaporated

100 nm thick layer of SiO2. The roughness and contact angle is well defined, as described in

table 1. Silica has been widely used the last decade as a model for a hydrophilic surface. The

chemistry of silica is rather complex, and a more detailed description can be found

elsewhere1. SiO2 surfaces consist of two very different surface groups, relatively hydrophobic

siloxane (Si2O), and more hydrophilic silanol groups (SiOH). The silanol group is

amphoteric1, which means that it can act both as a base and an acid. Hence, when the silica

surface is exposed to water solution the surface charge is determined by the density of silanol

groups on the surface and both the ionic strength of the solution and its pH. This forced us to

perform all our experiments with SiO2 under controlled pH and electrolyte concentration, and

preferably keeping them constant.

In order to increase the number of silanol groups at the surface (making it more hydrophilic)

and to remove eventual contaminations, we treated the surface with surfactants (Hellman

ExTM) followed by plasma cleaning as described in detail in paper III.

Yaminsky et al2 have an explanation to the instability of SiO2 surfaces in water. The surface

decomposition of SiO2, into polysilicic acids may result in the formation of a diffuse silica gel

layer. This gel is probably the main reason for instability of our SiO2 surfaces in water, which

are shown as a small drift in the frequency, but not in the dissipation factor. This is contrary

to ellipsometric studies using silica surfaces where instabilities could be seen at the gas-solid

interface but not at the liquid-solid interface3.

Page 24: Adsorption of Surfactants at the Solid-Liquid Interface: A ...

16

2.4 Silan coated silica.

The silanol groups present at the surface of silica crystals allow a surface modification by

reactions with different types of silanes. Surfaces can for instance be made more or less

hydrophobic by reaction with different alkylchlorosilanes1-2. For all our measurements (QCM

and ellipsometry) we used a dimethyldichlorosilane to modify our silica crystals/wafers in

order to obtain a small surface roughness and a reproducible hydrophobicity. This silane does

not form large hydrophobic islands as probed by profilometry, whereas this was found to

occur when for instance dimethyloctylchlorosilane (DMOS) was used. When surfaces coated

with DMOS were used in surface force experiments, a long-range attractive force was

observed2. This may be a direct consequence of these silane islands present on the surface.

When two surfaces of DMOS are brought together for the first time, capillary condensation

immediately starts3 and drops of DMOS are formed. The reason for the long-range attraction

seen in the surface forces experiment was suggest to be is the coalescence of these drops

between the surfaces. The dimethyldichlorosilane does not show this behaviour, probably due

to the strong cross-linking between the dimethyldichlorosilanes in the hydrophobic layer and

no hydrophobic islands are formed. We note however, that a long-range attraction has also

been observed for silica surfaces coated with dimethyldichlorosilane6 despite that we do not

see any islands of silane on such surfaces. The reason may be that air-bubbles are attached to

the surface and it is the coalescence of these bubbles that gives rise to the attractive force.

This mechanism was first suggested by Parker et al. for other silane coated surfaces7. For a

further discussion on long-range attractive forces between non-polar surfaces in water we

refer the reader to a recent review9. It has been shown that correctly prepared

dimethyldichlorosilane coated surface are surprisingly stabile over several days, as long as

they are kept in Milli-Q water or in clean organic solvents4. Such good stability has not been

found for DMOS, which likely is due to the lack of crosslinking in the hydrophobic layer,

leading to a hydrolyse of the silanol-silane bond in water. This indicates that the silanes are

kept at the surface partly by the low solubility, and that they are only partly stabilized by

chemical reaction with silanol groups5.

The preparation of the silane surface is a process that easily can go wrong, because of that a

precise experimental protocol needs to be used. The method described below has shown to

have the highest success rate.

Page 25: Adsorption of Surfactants at the Solid-Liquid Interface: A ...

17

The SiO2 surfaces were cleaned in a mixture of 25% NH4OH, 30% H2O2 and H2O (1:1:5, by

volume) at 80° C for 10 min, followed by cleaning in a mixture of 25% HCl, 30% H2O2 and

H2O (1:1:5, by volume) at 80° C for 10 min. In between and after the cleaning procedures by

the two mixtures, the substrates were rinsed in water. The substrates were then immediately

put into a reactor and exposed to vapours of dimethyldichlorosilane for 24 hours. Afterwards

these substrates were rinsed in toluene, ethanol and water, followed by heating to 200°C for 1

hour. All substrates were stored in ethanol until use.

Page 26: Adsorption of Surfactants at the Solid-Liquid Interface: A ...

18

3. Methods

When positioned at one of the interfaces between chemistry and physics, called surface

science, you relatively soon realize that surface chemists are lacking in the knowledge of how

to characterize physical (mechanical), properties and surface physics are lacking in

knowledge about the basic chemical properties. Hence, in this chapter a number of analytical

methods that can be used to study thin surface layers and surface structures are described. I

will mainly discuss the Quartz Crystal Microbalance-Dissipation (QCM-DTM) technique.

However, also some of the other techniques that I have been using, and which I consider

being the most valuable ones for the characterization of my model surfaces, will be briefly

presented.

3.1 Profilometer.

The surface roughness analysis was carried out using a Zygo View 5010TM, which is a non-

contacting technique (see figure 1 for a schematic illustration). It is a precision vertical

scanning transducer and a camera put together to generate a three dimensional interferogram

of the model surface. This is processed by the software (Metro Pro PCTM) in the computer and

transformed using frequency domain analysis to give a quantitative 3-D image.

The vertical resolution is 1 Å, independent of microscope magnification and the lateral

resolution is at best 0.3 µm. The model surfaces are characterized using the average surface

roughness, Ra, which is the average deviation of all points from a plane fit to the test surface.

The standard deviation of the profile heights Rq (rms) is also given. For a gaussian surface the

ratio between Ra and Rq is close to 1.3. Both Ra and Rq (rms) can be found in table 1 (see

chapter 2) for all our model surfaces. Ten measurements were made on random spots on each

model surface. The measurement area was 0.18 mm * 0.13 mm, which gives an area of

0.0234 mm2. All measurements were carried out in ambient atmosphere.

Page 27: Adsorption of Surfactants at the Solid-Liquid Interface: A ...

19

Sample

ReferenceSurface

PZTStack

InterferenceMicroscopeObjective

LightSource

Camera

Figure 1: The Zygo View 5010TM, a precision vertical scanning transducer and a camera put together.

Page 28: Adsorption of Surfactants at the Solid-Liquid Interface: A ...

20

3.2 Contact angle.

Contact angle measurements were conducted with a Fibro DAT 1100 system. This instrument

is used for fast absorption and wetting studies for contact angles above 20°. The application of

the droplet from the syringe onto the test surface was computer controlled; giving a controlled

drop volume (4 µl). The syringe used was a Teflon syringe in order to avoid any liquid

remaining onto the tip. The spreading process was recorded using a CCD camera connected to

an image analyser. The images were analysed with respect to base width and height in terms

of contact angle and drop volume. The drop volume starts to decrease due to evaporation after

10 s of spreading time. These data are discarded in the evaluation of contact angles. As this

simple method is a sensitive measure of the interfacial properties it is widely used to

characterize surfaces. The surfaces tension γ [J/m2] are only one of many properties, but also

estimates of surface roughness and chemical heterogeneity can be obtained from spreading

experiment. It is a simple and reliable method to use as a quality control of the self-assembly

monolayer formation process and general cleanliness of for instance silica and gold or other

hydrophilic materials (in general contaminations are of hydrophobic nature).

The contact angel α is related to the involved surface tensions, in this case there are three, the

solid-vapour, γSV , solid-liquid, γSL , and finally liquid-vapour, γLV , surface tension. This

relation is described by the Young’s expression (see equation 1).

cos( )αγ γ

γ=

−SV SL

LV

(1)

Whereas the Young-Dupre equation (see equation 2) relates to the adhesion energy per unit

areas of the solid (S) and liquid (L) adhering in the medium gas/liquid in our case vapour (V),

∆WSLV [J/m2] .

γ αLV SLVW( cos )1+ = ∆ (2)

Different chemical or structural components on a surface produces a heterogeneous surface.

Cassie suggested a way to calculate the contact angle of a heterogeneous multicomponent

model surface (see equation 3).

Page 29: Adsorption of Surfactants at the Solid-Liquid Interface: A ...

21

cos cosα χ α= Σ i i (3)

where, χi is the fraction of the i:th component on the surface and cosαi is its contact angle on

that type of surface1. In 1989, Israelachvili2, derived a revised version for the contact angle on

a heterogeneous surface (see equation 4).

( cos ) ( cos )1 12 2+ = +α χ αΣ i i (4)

Israelachvili’s assumptions are that the work of adhesion/cohesion is proportional to the

square root of the interaction forces involved, and again adding the works of

adhesion/cohesion to give the overall work of adhesion/cohesion. This equation claims to

account for the heterogeneity that is likely to occur in patches of molecular/atomic

dimensions.

Page 30: Adsorption of Surfactants at the Solid-Liquid Interface: A ...

22

3.3 X-ray Photoelectron Spectroscopy (XPS).

XPS is in the chemistry society often called, Electron Spectroscopy for Chemical Analysis

(ESCA). It is described in detail in references 1-5. Photons with energy hv from an X-ray

source are irradiated at the surface under study and adsorbed by the atoms. As a direct

consequence of this irradiation, emissions of electrons with lower binding energy, the

ionisation energy (Eb), than the energy of the incoming photons will occur. As a consequence

of the law of energy conservation, the emitted photoelectron obtains a kinetic energy (EK) that

is characteristic for the type of atom, the shell of the electron and its chemical environment.

The photoelectrons are separated by their kinetic energy before they reach the detector. By the

uniqueness of the kinetic energies of the photoelectrons emitted from the atoms an elemental

analysis can be conducted. In Albert Einstein’s equation (see equation 1) for the photoelectric

effect all this is described, for his work in this area Einstein got the Nobel Prize in 1921.

E hv EK b= − − φ (1)

where, φ is a correction for the spectrometer work function.

The photoelectrons, having kinetic energy up to around 1500 eV (if the AlKα electrode is

used), do not move more than a few nanometers in the solid material until they collide and

loose all or part of their kinetic energy. The average distance that photoelectrons move within

the solid material before they collide inelastically is mainly a function of the density of the

material and the kinetic energy. The inelastic mean free path λ(EK) describes this process. A

fraction of 1/e, about 37 %, of the photoelectrons move the distance λ before being scattered,

about 5 % moves as far as 3λ before they get scattered inelastically. The mean free path is

often referred to as sampling escape depth or sampling depth. It is around 0.5-2 nm for a

metal and 1.5-4 nm for oxides, these values are typical mean free paths2 for photoelectrons

having a kinetic energy of 1000 eV. Hence, XPS is truly a surface sensitive technique for

chemical analysis.

The equipment employed is a Kratos, AXIS HS X-ray photoelectron spectrometer (Kratos

Analytical, Manchester, UK). The X-ray emitted in our case comes from an MgKα (1253.6

eV) source. There are other sources available, one of them is the energetic AlKα (1486.6 eV)

source that produces more energetic photoelectrons, and due to this in some cases is more

useful.

Page 31: Adsorption of Surfactants at the Solid-Liquid Interface: A ...

23

3.4 The Quartz Crystal Microbalance-Dissipation (QCM-DTM).

The Quartz Crystal Microbalance (QCM) is by no means a new technique1-2. In vacuum

physics for instance it has existed for decades, providing film thickness measurements for

metal film deposition3. However, lately numerous advancements have been made in the

measurement of the frequency factor of the QCM. Among these is the new Quartz Crystal

Microbalance-Dissipation (QCM-DTM) instrument from Q-Sense, Gothenburg, Sweden (see

figure 2), which we have used for our experiments4. This new setup has two advantages. First,

an improved resolution of the frequency factor in aqueous solutions. In fact, it is hard to find

any comparable non-vacuum QCM setup5-6. Secondly, this instrument also measures the so-

called dissipation factor, which is a measure of the damping of the crystal as will be discussed

later. Hence, the QCM technique has only recently become a potentially useful tool for the

surface scientists concerned with “wet” surface chemistry. The possibility to monitor the

interfacial processes quantitatively in real time opens up new windows of opportunity. The

QCM principle is based on evaluating a change in frequency of the oscillating crystal, ∆f

[Hz], due to the change occurring on or adjacent to its electrodes. This frequency change is

most often interpreted as being due to the change in surface mass loading. In this work we

have used the QCM as a “probe”, in order to characterize physical changes at interfaces

occurring as a result of surfactant addition7-8. Such effects may arise due to adsorption or

different phase changes9-11. This chapter will describe the basic operational parameters of the

QCM, and the focus will be on its operation in liquids, even though its use is more widely

documented for the gas phase system. We note that not so many studies have explored the use

of the QCM for studying solid-liquid interfaces in surfactant solutions12-14. However, a much

more extensive literature on surfactant films deposited onto surfaces via the air-solution

interface with monolayers of insoluble surfactants is available. A discussion of this extensive

literature is beyond the scope of this thesis. The interested reader is referred to reference 12.

Page 32: Adsorption of Surfactants at the Solid-Liquid Interface: A ...

24

Figure 2: The Quartz Crystal Microbalance-Dissipation (QCM-DTM).

The Quartz crystal.

Most QCMs consist of a thin wafer of piezoelectric material, usually quartz, sandwiched

between a pair of thin metal electrodes (see figure 3), usually gold. Quartz is a piezoelectric

crystalline form of silicon dioxide (SiO2). For oscillating systems the α-quartz is the preferred

choice because of its thermodynamic stability at temperatures up to 846° K. The other form,

the β-quartz, is metastable at room temperature and it is not piezoelectric15. Piezoelectricity1 is

literally “pressure electricity”, the prefix piezo- being derived from the Greek word “to press”.

The direct piezoelectric effect refers to the electric polarisation of certain materials by

mechanical stress. The converse effect refers to the deformation of the same material by an

electric field. Electrostriction is a property of all dielectric materials; it means that when they

are placed in an electric field they deform. The difference between piezoelectric materials and

purely electrostriction materials is that the piezoelectric deformation is much larger than the

ordinary electrostriction deformation and the piezoelectric deformation is reversible. As an

example, a rod of quartz is cut in such a way that an applied field causes an elongation of the

rod. Reversing the direction of the field will cause the rod to contract in a piezoelectric

material, whereas in a non-piezoelectric material, whatever deformation is caused will be

independent of the direction of the field.

The piezoelectric effect being reversible gives that it is also anisotropic, which means that the

mechanical deformation and the electric field (see below) in the material depends on the

Page 33: Adsorption of Surfactants at the Solid-Liquid Interface: A ...

25

direction within the material. Such materials cannot have a centre of symmetry; with a centre

of symmetry the reversal of an applied field would have no effect on the materials internal

structure. Lord Kelvin gave the first explanation of the origin of the piezoelectric effect1 in

terms of molecular structure; the assumption he made is useful as a qualitative and heuristic

guide to understand piezoelectricity.

The electrode.

The electric field is in almost all quartz crystals applied via electrodes deposited at the quartz

surface in a key hole pattern, as shown in figure 3. In general gold electrodes give a

considerably more chemically stable surface compared to other electrode materials16 such as

silver (Ag) and aluminium (Al), which both tend to oxidize in aqueous solutions. Although it

has been suggested that the gold electrodes of the QCM may also be subject to minor

oxidation. This is certainly the case when they are treated with UV/ozone (AuO3, is the

product from the UV/ozone treatment17). For the gold and silver electrodes a thin adherent

layer (2.5-5 nm) made of chromium (Cr) or titanium (Ti) are used to improve the adhesion of

the electrodes to the quartz crystal. The disadvantage of this thin adherent layer is the

increased stress in the electrodes, which can influence the output frequency.

Figure 3: The Quartz crystal, with its gold electrode in a characteristic “keyhole” pattern.

Page 34: Adsorption of Surfactants at the Solid-Liquid Interface: A ...

26

Frequency.

A quartz crystal used in the QCM is normally cut at an angle θ ≈ 35° from the ZX-plane, this

is known as an AT-cut18, this cut angle makes the quartz crystal less sensitive to temperature

drifts as compared to the large temperature drifts in the original X-cut quartz crystal (quartz

crystals cut normal to the x-axis) where the temperature drifts are as large as 30 ppm / °K. In

the AT-cut case the temperature drift could be as small as 2 ppm. / °K. Another improvement

from the original X-cut is that when applying an electrical field across an AT-cut crystal, a

shear strain will be induced instead of the induced strain in the thickness direction in the X-

cut case.

Consequently, an alternating electric field onto an AT-cut quartz crystal will induce shear

waves. Of all the vibrational modes that may exist in a quartz crystal, only those that can be

driven by an alternating electrical field are relevant in the context of this thesis. Mechanical

resonance begins when the thickness of the quartz crystal contains an integral number n of

half wavelengths of the extensional wave or longitudinal waves. The quartz crystal’s surfaces

will be the anti-nodes of vibration from a standing wave within the plate. When n is even the

vibrational modes of the two surfaces are in phase (destructive), and in anti-phase

(constructive) when n is odd (n=1 being the fundamental mode, n = 3 is called the first

overtone, and so forth). The resonance frequency condition is (equation 1):

fnv

tq

=2

(1)

Where v is the velocity of the extensional waves, (v/f) is the wavelength, tq is the thickness of

the quartz crystal and n is an odd integer (1,3,5,…).

Sauerbrey1 9 was the first to show that any mass, ∆ m , deposited on one or both of the

electrodes of a QCM crystal, induces a shift in the frequency, ∆ f, that is proportional to the

added mass. If the mass is deposited evenly over the electrode(s), and ∆ f is much smaller

than f, then the frequency shift versus mass relationship is:

∆∆ ∆ ∆

mt f

nf

f

nf

C f

nC

t

fq q q q q q= − = − = − ⇒ =

ρ ρ ν ρ

0 02

02(2)

Page 35: Adsorption of Surfactants at the Solid-Liquid Interface: A ...

27

Where ρ q and ν q are the specific density and the shear-wave velocity in quartz, respectively,

tq is the thickness of the quartz crystal, f0 the fundamental resonant frequency and n is the

shear wave number. With ρ q = 2648 kg/m3, ν q = 3340 m/s, tq = 0.33 mm, and f0 = 5 MHz, C

is 17.7 ng cm-2 Hz-1. For the relation to be valid, Sauerbrey assumed that the added mass

should be much smaller than the mass of the quartz crystal, and it should be rigidly attached to

the electrode(s), with no slip or inelastic deformation in the added mass due to the oscillatory

motion. Pulker later confirmed equation 2 by experimental data up to mass loadings (madsorbed /

mcrystal ) of approximately 2 %. There are various models or converting the frequency shift to

mass loadings up to approximately 5 %, and they all behave similarly21. Another property,

probably the most important to have under control, is that the surface area should be smooth.

The QCM surface area is approximately the same as the projected geometrical area for low Ra

values, and this roughness effect will be discussed later.

The use of QCM in liquid media.

In 1980, Nomura showed that a quartz crystal could be completely immersed in a liquid and

still be excited to stable oscillations22, after this theories had to be worked out. In 1985

Kanazawa and Gordon published a theory23 on the QCM behaviour in the liquid phase. They

were totally unaware of the theory that Stockbridge24 had published already in 1966. This

paper was concerned with the gas pressure effect on the QCM oscillations, and it turned out to

be exactly the same as the Kanazawa and Gordon theory. The relationship derived describes

the change in oscillation frequency of the quartz crystal in contact with a fluid in terms of

material parameters of the fluid and the quartz. This relationship is shown below (equation 3).

∆ff

t nq q

f f= − 0

2 ρ πρ η (3)

Where ρ f and ηf are the specific density and the absolute viscosity of the film, respectively, tq

is the thickness of the quartz crystal, f0 the fundamental resonant frequency of the dry crystal,

ρq is the specific density of quartz and n is the shear wave number. With ρ q = 2648 kg/m3, tq

= 0.33 mm, and f0 = 5 MHz. This relation is obtained from a simple physical model, which

Page 36: Adsorption of Surfactants at the Solid-Liquid Interface: A ...

28

couples the shear wave in the quartz crystal, to a damped shear wave in the fluid. The shear

wave extension or, as it is more commonly called, the decay length, is given by equation 4.

δπη

ρ=

4 f

n ff(4)

where, δ is the decay length of the shear wave, ηf is the absolute viscosity of the film, fn the

resonant frequency of the dry crystal in mode n and ρq is the specific density of quartz. The

decay length is the distance into the liquid where the amplitude of the shear wave has fallen

by a factor of e, and for a 5 MHz quartz crystal oscillating in water this decay length is

approximately 250 nm at 20° C.

Roughness properties.

Martin, Frye and Wessendorf examined the frequency response of smooth (low surface

roughness, Ra < 10 nm) and textured surfaces (high surface roughness, Ra > 100 nm) on

quartz crystals in liquids in 199425. Smooth quartz crystals, which viscously entrain a layer of

contacting liquid, exhibited a response that depends on the square root of the product of liquid

density and viscosity. Textured-surface quartz crystals, which also trap liquid in surface

crevices, pores, etc., exhibit an additional response that depends linearly on liquid density

alone. The resulting modification to the Stockbridge and Kanazawa equation 4, is shown in

equation 5.

∆ ∆ ∆f f ff

t n

f

ttv t

q q

f fn

q qf f= + = − −0

2 ρ πρ η

ρρ (5)

Where ∆fv is the induced frequency shift due to the liquids viscosity and density over a

uniform crystal. ∆ft is the induced frequency shift due to trapped liquid with an average

thickness of tf , ρ f and ηf are the specific density and the absolute viscosity of the film,

respectively, tq is the thickness of the quartz crystal, fn the resonant frequency of the dry

crystal, ρq is the specific density of quartz and n is the shear wave number. The liquid

Page 37: Adsorption of Surfactants at the Solid-Liquid Interface: A ...

29

entrained by the oscillating smooth surface is described as viscously coupled. This liquid does

not move synchronously with the surface, but undergoes a progressive phase lag with

increasingly distance from the surface. The textured-surface also traps a quantity of fluid in

excess of that viscously entrained by a smooth surface. The perpendicular character of the

texture-surface constrains this trapped liquid to move synchronously with the surface, rather

than undergoing a progressive phase lag. This trapped liquid can be viewed as an added mass,

contributing to an areal mass density, ρtf, where ρ is the absolute density of the liquid and tf is

the effective thickness of the perpendicular features of the surface.

The frequency shift measured in an adsorption experiment is relative to the frequency of the

crystal immersed in water. Under the conditions we have used the instrument, i.e. relatively

low solute concentrations, no measurable effects due to changes in bulk viscosity or density is

expected. Hence, the measured frequency shift in our experiments is due to changes occurring

close to the surface. Most importantly adsorption including bound water.

Martins addition to Gordon’s and Kanazawa’s equation is valid under the condition that the

effective thickness h, of trapped liquid is small compared to the liquid decay length δ. In such

a case the relative response due to liquid trapping is small, compared to the frequency shift

due to viscous entrainment and may be neglected. This defines a criterion for hydrodynamic

smoothness29: h << δ. If h is comparable to or larger than δ, then a significant additional

response arises due to liquid trapping. Martins definition for a smooth surface are that the

roughness should be less than 10 nm, but he was working with frequency shifts (∆f ) in the

kHz region. We are working within a lot more sensitive regime, the frequency shifts in our

experiments are in the Hz region, where this effect is significant, this is in addition to the

viscous entrainment.

Longitudinal waves.

The generation of longitudinal waves in liquids has been largely ignored, except in a few

reports30-33. The occurrence of the longitudinal wave component is usually demonstrated by a

movable plate, parallel to the quartz crystal, and starting from a remote distance (much larger

than the decay of the shear wave). The presence of longitudinal waves can be observed so far

away as centimetres34-35. There presence results in a periodicity of the resonant frequency,

inductance, capacitance, and resistance as the distance between the quartz crystal and the

movable plate is varied36-38. The wavelength, of the standing longitudinal waves that are

Page 38: Adsorption of Surfactants at the Solid-Liquid Interface: A ...

30

observed agree with that expected for the resonant frequency of the quartz crystal and the

properties of the fluid medium, and the periodicity of the quantities mentioned above is

approximately half the longitudinal wavelength. For a 5 MHz quartz crystal in a water

medium at 20° C the periodicity is approximately 150 µm.

Pc

f= =

λ2 2 0

(6)

where, P is the periodicity [m] , λ is the wave length [m], c is the phase velocity of the waves

in water, c = 1465 m/s at 20° C and f0 is the fundamental frequency [Hz].

Various measurements of the standing wave frequency and amplitude have revealed effects of

crystal contour, liquid properties, interface reflection coefficients, and the radial dependence

on the standing wave amplitude. Clearly it is important to design the experiments so that

contributions from such longitudinal waves are avoided, mainly taking the phase velocity of

the waves in the medium into consideration.

Amplitude of vibration.

From classical driven oscillator theory, it is clear that the values of the shear wave amplitude

depends on both the drive voltage applied to the quartz crystal and the quality factor, Q (the

inverse of the dissipation factor, D), of the system. This has to be remembered than

comparing data from different authors and theoretical calculations. In view of recent

theoretical work by Kanazawa39, it is now possible to compare experimental measurements of

the vibration amplitude itself to detailed calculations for quartz. Kanazawa calculates an

amplitude of 133 nm using a peak drive voltage of 1.0 V, and a Q of 100 000 (D = 10*10-6).

Just as with classical oscillator theory, the amplitude of vibration is expected to be

proportional to the drive voltage and the quality factor also in Kanazawa’s model39:

A C Q VC V

Dav av dav d= =* *

*(7)

Page 39: Adsorption of Surfactants at the Solid-Liquid Interface: A ...

31

where Aav is the average amplitude of vibration (average = 1/2 maximum), [m], Cav is found to

be 1.3 pm/V, Q is the quality factor and Vd is the drive voltage, [V].

This value for Cav is rather close to the piezoelectric strain coefficient for an AT-cut quartz

crystal, 3.1 pm/V, which Martin and Hager40 concluded was the same as the Cav. Borovsky,

Mason and Krim41 on the other hand are suggesting an Cav = 0.7 pm/V, so there are still open

questions about the amplitude of vibrations in quartz crystals. Both theoretical and

experimental results put the amplitude of vibration (Cav) for a 5 MHz quartz crystal, with a Q

factor of 100 000, and a drive voltage of 1.0 V, in air in the interval 40-200 nm. Since the Q-

value for the same system in water are around 3 000, this would shrink the amplitude of

vibration to the interval 1-6 nm for the aqueous system. In our case with a drive voltage of 0.7

V, and a dissipation factor equal to 20*10-6 in air and 310*10-6 in water, the resulting average

amplitude is 45 nm in air and 3 nm in water, using Cav.

The Dissipation factor.

There is, beside the resonant frequency, another important parameter that characterises the

oscillatory system, namely the dissipation factor, D or the inverse of Q, the quality factor26, Q

= 1/D. In 1966, Spencer and Smith27 studied the amplitude of an oscillating quartz crystal and

its decay. The amplitude was decaying as an exponential sinusoid.

A t A e ft ct( ) sin( )/= + +−0 2τ π ϕ (8)

where τ is the decay time that depends on f, ϕ is the phase angle and the constant, c, is the dc

offset. The total dissipation factor for the fundamental frequency Dtotal is related to the decay

time, τ , according to:

Dftotal =1

0π τ(9)

where f0 is the fundamental frequency. The dissipation factor is proportional to the power

dissipation in the oscillatory system:

DE

Edissipated

stored

=2π

(10)

Page 40: Adsorption of Surfactants at the Solid-Liquid Interface: A ...

32

Where Edissipated is the energy dissipated during one period of oscillation, and Estored is the

energy stored in the oscillating system. Consequently, D accounts for all mechanisms, that

dissipate energy in the oscillatory system. Which can be summarized as Dtotal:

D Dtotal ii

= ∑ (11)

where Di is the dissipation factor of the i:th mechanism. For instance, an adsorbed film

dissipates energy through its coupling with the solvent10-11. This is probably the largest effect

in thin films, and it will be discussed later. Internal friction in the quartz crystal is another

large factor and so are the losses due to mounting. Further, if the film is not very thin and

viscous, energy is dissipated due to the oscillatory motion induced in the film28. Hence, if the

dissipation factor can be measured correctly it may be possible to obtain additional

information about the visco-elastic properties of adsorbed layers. Other studies indicate that

phase changes within the film can be related to changes in the dissipation factor7-8, which

probably can be related back to the adsorbed films coupling to the solvent liquid. Stockbridge

not only related the frequency shift but also the damping of the oscillations for the quartz

crystal in liquids to liquid properties24:

∆Df

nl l

q q

= 2 02 2

ν ρπυ ρ

(12)

where, ∆D is the total dissipation shift, fn is the frequency, n is the shear wave number

(1,3,5,…), νq is the wave velocity in quartz, ν f is the wave velocity in the film, ρq is the

density of quartz, and finally, ρf is the density of the film. νq = 3340 m/s, ρq = 2650kg/m3, and

for water at 20° C, vl = 1465 m/s.

Martin, Frye and Wessendorf25 used smooth quartz crystals that viscously entrain a layer of

contacting liquid. They observed a response that depends on the square root of the product of

liquid density and viscosity, as predicted by the Stockbridge equation (equation 5). Since the

added liquid is fully coupled to the surface texture, there is no damping of the quartz crystal

oscillations due to the presence of trapped liquid.

Page 41: Adsorption of Surfactants at the Solid-Liquid Interface: A ...

33

The change in dissipation reported in our experiments is the difference between the

dissipation rate after and before addition of solute. Again, under the dilute solution conditions

studied the effects due to changes in bulk viscosity and density can be ignored. Hence, the

values reported are due to changes in dissipation occurring close to the surface. This includes

energy dissipation within the adsorbed layer and differences in coupling between the surface

with the adsorbed layer and the solution compared to between the bare surface and the

solution.

Page 42: Adsorption of Surfactants at the Solid-Liquid Interface: A ...

34

4. Surfactants

4.1 Lipids.

In living systems large molecules, such as proteins, polysaccharides and complex lipids, build

up the structure of the cells and tissues. The word lipid covers an astounding variety of

compounds. One of the most widely used classifications is that of Bloor and later by Deuel1:

i) Simple lipids, which consists of neutral fats (glycerol esters of largely long-chain fatty

acids) and waxes (solid esters of long-chain monohydric alcohols) and ii) Compound lipids

which consists of Phospholipids (lipids containing a phosphate residue), Cerebrosides,

Gangliosides (lipids containing a carbohydrate residue) and Sulphatides (lipids containing a

sulphate residue). The compound lipids are esters of fatty acids with alcohols, which contain

also an additional group, see above.

In this thesis work we will limit ourselves to study the phospholipids, phosphatidylcholine

(PC) and phosphatidylethanolamine (PE) and an emulsion stabilized mainly by such lipids.

Phospholipids are zwitter-ionic. The cationic group in PCs consists of a quaternary

ammonium group whereas in PEs it consists of a primary amine. For this reason PEs become

anionic at high pH, whereas they are cationic at very low pH. In the ionic state, these lipids

with monovalent counterions behave in a similar way to ionic phospholipids. In general, the

acyl chains of the phospholipids are comparably long and the monomer solubility is very low,

approximately 10-11 M, in water for a typical hydrocarbon chain length ∼ C16. As a

consequence of this the phospholipids will self-assemble at low concentrations. Vesicles, or

as they also are called in the phospholipid case, liposomes, are composed of lipid bilayers

enclosing a water core. Liposomes are in general not thermodynamically stable, i.e. they are

not an equilibrium structure. These structures are used in so diverse areas ranging from

mimicking cell membranes to delivering of drugs in the human body.

Phosphatidylcholine (PC).

The last recommendation of a nomenclature for PC is that one from IUPAC-IUB Commission

on Biochemical Nomenclature and that is 1,2-diacyl-sn-glycero-3-phosphocholine; we will

Page 43: Adsorption of Surfactants at the Solid-Liquid Interface: A ...

35

hereafter only call it PC. The molecular structure consists of a zwitter-ionic head-group and

two fatty acyl chains see below.

RICOOCH2 - RIICOOCH - CH2O –POO- –OCH2CH2N

+(CH3)3

Where, RI- and RII- are the fatty acyl substituents. An extensive literature exists about their

aqueous interaction and phase properties 3. For a typical PC the lamellar Lα phase dominates

the phase diagram and it is this phase that is in equilibrium with excess water. At low water

content, and as the temperature is increased it is possible to obtain the transition Lα ⇒ cubic

⇒ H11, where H11 is the reversed hexagonal phase. The driving force for this transition is the

tendency for increased chain divergence when the thermal mobility increases. When the Lα-

phase is cooled, it forms a gel-phase. The transition temperature from the Lα-phase to the gel-

phase on cooling decreases from 41° C to 23° C when the saturated chain length is reduces

from C16 to C142.

Phosphatidylethanolamines (PE).

The last recommendation of a nomenclature for PE is that one from IUPAC-IUB Commission

on Biochemical Nomenclature and that is 3-sn-phosphatidylethanolamine; we will hereafter

only call it PE. The molecular structure consists of a zwitter-ionic head-group and two fatty

acyl chains see below.

RICOOCH2 - RIICOOCH - CH2O –POO- –OCH2CH2N

+H3

Where, RI- and RII- are the fatty acyl substituents. An extensive literature exists about their

aqueous interaction and phase properties 3. Crystals of didodecyl –PE heated in excess of

water are transformed into the Lα-phase (+H2O) at about 40° C. At about 100° C, a cubic

phase (+H2O) is formed and above 120° C, the reversed hexagonal phase H11 (+H2O) is

obtained3. There are also reports of different PEs with unsaturated chains that exhibit cubic

phases and H11-phases4.

Emulsions.

An emulsion, i.e. a liquid dispersed in another liquid, is not a thermodynamically stable

system. Two processes that can destabilize oil-in-water emulsions are of interest to us. The

Page 44: Adsorption of Surfactants at the Solid-Liquid Interface: A ...

36

first being flocculation of emulsion droplets, which is an aggregation phenomena with the

droplet interface remaining intact and a thin liquid film remaining in between the aggregating

droplets. This normally leads to an increased viscosity and sometimes a gelation. The second

process is the coalescence of flocculated droplets. After coalescence the emulsion droplets

lose their identity. With time the density difference between the coalesced drops and the

aqueous phase leads to accumulation of an oil rich layer at the top of the container. This

phenomenon is called creaming5.

Page 45: Adsorption of Surfactants at the Solid-Liquid Interface: A ...

37

4.2 Cationic surfactants .

The polar moiety in a cationic surfactant has a positive charge, thus they adsorb strongly onto

most solid surfaces (which are usually negatively charged), and can impart special

characteristics to the substrate. Some examples are, softeners, anti-static agents, corrosion

inhibitors, hair conditioners, lubricants and flotation agents1-4. The adsorption of surfactant

ions from aqueous solutions onto hydrophilic negatively charged surfaces is based on two

main interactions: electrostatic and hydrophobic5-7. The electrostatic interaction is involved in

the first step of a two-step adsorption mechanism and its contribution to adsorption depends

largely on the charge of the surfactant ions, surface charge density of the adsorbent,

electrolyte concentration and pH. Hydrophobic interactions are involved in the second step of

the two-step adsorption mechanism where additional surfactants are associated with

electrostatically anchored surfactants. This interaction is mainly influenced by the surfactant

structure and particularly the size of the hydrophobic part8-11. The hydrophobic interaction acts

between the non-polar parts of the surfactant ions (hydrocarbon chain) and it is due to release

of water molecules forming a dynamic cage around the non-polar moiety10. van der Waals

forces also contribute to the association, but this contribution is small compared to the

hydrophobic interaction.

For a deeper discussion on hydrophobic interactions see for instance the book by Tanford and

refrences4, 10, 11. The corresponding attractive forces cause aggregation into micelles in the bulk

phase. At the solid-liquid interface they are responsible for the surfactant aggregation into

surface aggregates. For a further detailed discussion of the structure of the surfactant

adsorption layer, two different mechanisms must be considered: i) Surfactant ions interacting

with hydrophobic surface sites on non-polar or polar/non-polar surfaces11. ii) Surfactant ions

interacting with each other and with those primarily adsorbed by hydrophobic and

electrostatic forces12-13. Obviously on non-polar surfaces with hydrophobic groups (e.g.

methylated silica14), polar/non-polar surfaces containing both hydrophilic (charged) groups

and hydrophobic entities (e.g. low density of chemisorbed alkylsilanes15), the direct

attachment of hydrophobic entities to the surface can be expected. However, on fully hydrated

adsorbents, no significant direct interaction between the surface and the hydrophobic part of

the surfactant is expected.

Page 46: Adsorption of Surfactants at the Solid-Liquid Interface: A ...

38

4.3 Non-ionic surfactants .

The hydrophilic (polar) part of a non-ionic surfactant is usually a poly(ethylene oxide) chain

even though sugar-based surfactants also are becoming common. The hydrophobic (non-

polar) part is, just as for cationics, anionics and zwitter-ionics, most often a hydrocarbon

chain. The interfacial behaviour of ethylene oxide based non-ionic surfactants CnEOm, is

strongly affected by the size of the polar and non-polar parts of the molecules1-6. The adsorbed

amount is found to increase with an increasing number of methylene groups in the

hydrocarbon chain (m), whereas a decrease in the adsorbed amount with increasing number of

ethylene oxide groups in the hydrophilic chain (n) is commonly observed. Polyethylene oxide

based surfactants display a variety of different phases in aqueous solution, depending on the

surfactant structure, temperature, and concentration7-10. Observed phases include micellar

solutions, lamellar, hexagonal, and cubic phases. Depending on whether the polar or the non-

polar part of the surfactant interacts most favourably with the surface, different structures of

the adsorbed surfactant layer is formed. Generally, a monolayer structure is found on

hydrophobic surfaces, and at hydrophilic surfaces different surface aggregate forms such as

micells or bilayers11-14.

Page 47: Adsorption of Surfactants at the Solid-Liquid Interface: A ...

39

5. Results and Discussion

5.1 Adsorption of Emulsions studied with the QCM-DTM.

Adsorption and spreading of emulsions on solid surfaces are of importance in a range of

applications, e.g. drug delivery, lubrication, and during coating of the road surface with

asphalt emulsions, just to mention a few. Hence, understanding interfacial properties of

emulsions and solutions are relevant to a large number of technologies. There are, however,

not very many suitable experimental methods that readily can be applied to this area of

research, particularly for concentrated emulsions at solid surfaces. Hence, optical methods

such as ellipsometry and interferometric surface force measurements are difficult to apply due

to the large scattering of light that characterise these concentrated systems. However, despite

this some rather recent progress has been made. Ellipsometric1 studies of adsorption of dilute

emulsions have been carried out and surface force techniques have successfully been applied2.

In paper I we showed that the QCM-DTM technique, allowing the simultaneous measurement

of changes in resonance frequency and energy dissipation rate, is very suitable for studying

emulsion adsorption and spreading. For the study we used a model oil-in-water emulsion

consisting of purified soybean oil (20 wt.%) dispersed in water. The emulsifier (1.2 wt.%)

used was fractionated egg phosphatides with the major components being

phosphatidylcholines (≈ 70%) and phosphatidylethanolamines. In figure 1 the adsorption

behaviour of the 20 wt.% oil-in-water emulsion is illustrated. The initial adsorption is very

rapid and results in a significant lowering of the resonance frequency and an increase in the

dissipation factor. We note that the initial change in resonance frequency is smaller than for

the liposomes that also were studied (see paper I) whereas the change in dissipation is larger.

This tells us that the adsorbed amount is smaller but that the size of the adsorbed emulsion

droplets (Dav in solution is 300 nm) is larger than that of the adsorbed vesicle (Dav in solution

70 nm). After the initial rapid adsorption a slow increase in resonance frequency and a

decrease in dissipation are observed until equilibrium values are obtained. We suggest that

this is due to a spreading of the emulsion droplets on the surface accompanied by some

material desorption. After 15 minutes we injected another portion of the same emulsion

solution in the measuring chamber again. This exchange of the solution does not result in any

change in the bulk solution composition but nevertheless an increased adsorption, resulting in

Page 48: Adsorption of Surfactants at the Solid-Liquid Interface: A ...

40

a declining frequency shift and a rising change in dissipation, is observed. This indicates

adsorption controlled by hydrodynamic factors. Hence, the injection of the solution provides

the necessary kinetic energy to the emulsion droplets to let them overcome the energy barrier

for further adsorption.

-100

-80

-60

-40

-20

0

20

-20

0

20

40

60

80

100

0 50 100 150 200

f5 (

Hz)

D5 (10-6)

Time (min)

Figure 1: Emulsion adsorption at a gold surface, injections of solution at t = 10, 25, 55, 85 and 125 min and

rinsing with water at 160 min. The frequency shift is represented by () , and the change in dissipation factor

by (- -).

Again, after the initial adsorption a slow spreading and desorption follow. It is worth noting

that the spreading occurs slower at this stage compared to during the preceding stage, and the

probable reason for this being the increased packing density of the emulsion droplets that

makes the spreading process more difficult. Repeated exchanges of the emulsion solution (at t

= 55, 85, and 125 min) give similar results as described above. It is, however, worth noting

that the spreading process becomes slower for each successive exchange of the solution. The

final frequency shift is about –90 Hz, which correspond to an adsorbed amount of

approximately 16 mg/m2 as calculated according to the Sauerbrey relation. Considering the

high dissipation factor it seems likely that the layer on the gold surface consists of deformed

emulsion droplets. We ended the experiment by replacing the emulsion with pure water (at t

=160 min). This resulted in a rapid desorption and the dissipation returned to close to zero

whereas the final frequency shift was about –10 Hz. The latter quantity corresponds to an

adsorbed mass of 1.8 mg/m2, or an area per molecule of 70 Å2. This means that a monolayer

remains on the surface.

Page 49: Adsorption of Surfactants at the Solid-Liquid Interface: A ...

41

5.2 Polymer adsorption on phospholipid coated surfaces.

In paper II, the adsorption of the triblock copolymer F127, poly(ethylene oxide)-

poly(propylene oxide)-poly(ethylene oxide), PEO98PPO67EO98, onto immobilized small

unilamellar vesicles (SUVs) of egg phosphatidylcholine (EPC) was studied by the QCM-DTM.

In figure 2, the adsorption of SUVs onto the hydrophilic gold surface is illustrated. The

liposome solution was introduced into the measuring chamber at t = 10 min. Immediately

after the introduction, the resonance frequency dropped and the dissipation increased. This

shows that the liposomes adsorb to the gold surface and the large changes in frequency and

dissipation indicate that they adsorb as intact liposomes. After the initial rapid adsorption we

can see a slower increase in adsorption. The liposome solution was exchanged for a identical

liposome solution twice at t = 70 min and t = 100 min. This resulted in transient peaks in the

frequency curve due to the temperature (the reservoir is not perfect) and pressure effects due

to the increased volume resting on the oscillating quartz crystal. However, there was no

significant change of the frequency after the solution exchanges were finished, and

temperature and pressure were restored. The slow adsorption of the liposomes at longer times

indicates that the surface was almost completely saturated with liposomes already after the

first introduction of liposomes, probably due to hydrodynamic transport to the surface. At t =

130 min, the liposome solution was exchanged for pure HEPES buffer. This was done to

remove eventual loosely adsorbed liposomes from the gold surface. Evidently there was no or

little desorption of SUVs after rinsing and thus the interaction between the liposomes and the

gold surface is sufficiently strong to make the adsorption irreversible over the experimental

timescale. Further rinsing with pure buffer, at t = 160 min and t = 190 min, did not cause any

desorption of liposomes. Following the above qualitative description of the SUV adsorption

process we shall make some quantitative considerations. The final change in resonance

frequency, ∆fSUV, can be converted into the adsorbed liposome (SUV) mass, mSUV, by means

of the Sauerbrey relation. The mean value of ∆fSUV, obtained from 8 runs, was 343 (± 63) Hz.

Using the Sauerbrey relation this corresponds to an adsorbed mass of 20.35 (± 3.7) mg/m2.

Theoretically, we can estimate the adsorbed mass of the liposomes, including the mass of the

“entrapped” buffer; by assuming that the liposomes have a radius of 15 nm (Dav = 15 ± 5 nm)

and that they form a monolayer of close-packed spheres at the surface. With these

assumptions we get approximately 18 mg / m2 of adsorbed liposomes. In this calculation we

have not accounted for the buffer entrapped between the liposomes and the gold surface

Page 50: Adsorption of Surfactants at the Solid-Liquid Interface: A ...

42

which may also oscillate with the crystal. Nevertheless, the estimate is close to the

experimental value and we conclude that a monolayer of SUVs is formed at the gold surface

in accordance with previously published results1.

After the last rinsing of the SUV monolayer, the polymer solution was introduced at t = 220

min. The frequency dropped immediately after the introduction of the polymer solution

indicating a rapid adsorption of the polymer on the SUV monolayer. Besides the frequency

drop due to the polymer adsorption there was also a transient peak due to temperature and

pressure effects, as discussed above. Furthermore, the dissipation increased as the polymers

adsorbed. The polymer solution was exchanged for an identical polymer solution twice at t =

250 min and t = 280 min. The final frequency drop, ∆fF127, was determined after the third

addition of polymer solution. Using the Sauerbrey relation, this value was converted into the

adsorbed polymer mass, mF127. To investigate the desorption process, we exchanged the

polymer solution for a pure buffer solution at t = 310 min. As can be seen in Figure 2, the

frequency increased and the dissipation decreased, indicating desorption of F127.

Interestingly, the kinetics of the adsorption and desorption processes seem to be about the

same. Evidently, some of the adsorbed polymers are rapidly desorbed during rinsing and the

desorption of the F127 polymers causes little or none desorption of SUVs.

-600

-500

-400

-300

-200

-100

0

100

-5

0

5

10

15

20

25

30

0 50 100 150 200 250 300 350

f15

(Hz)

D15 (10-6)

Time (min)

Figure 2: The adsorption of liposomes (gold surface) after injecting the liposome solution at t = 10, 70 and 100

min, followed by rinsing with HEPES at t = 130, 160 and 190 min. The injection of F127 occurred at t = 220,

250 and 280 min, followed by rinsing with HEPES at t = 310 min. The frequency shift is represented by () ,

and the change in dissipation factor by ( - - -).

Page 51: Adsorption of Surfactants at the Solid-Liquid Interface: A ...

43

It is important to emphasize that there was no measurable adsorption of F127 onto the bare

gold surface. Thus, it is clear that the adsorption of F127 occurs onto the preadsorbed SUVs,

and that the adsorbed amount of F127, follows a Freundlich isotherm at low concentrations,

illustrated in figure 3. The determined polymer-binding isotherm was used to correlate the

adsorbed amount of polymer with the polymer-induced leakage of carboxy fluorescein (CF)

from the SUVs in bulk solution. It was found that the membrane permeability was increased

several fold already at low surface coverage and that the maximum magnitude of the CF

release rate was obtained at, or close to, the F127 concentration needed to reach the maximum

adsorbed amount of polymer.

0

0.05

0.1

0.15

0.2

0 0.2 0.4 0.6 0.8 1 1.2 1.4

mF1

27 /

mS

UV

CF127

/ mg mL-1

Figure 3: Adsorption isotherm for the Pluronic F127. The adsorbed amount was determined as the adsorbed

mass of F127, mF127, divided by the adsorbed mass of SUVs, mSUV. The solid line going through the data points is

only drawn to guide the eye. The horizontal solid line indicates the plateau adsorption value, Γp. The dashed line

is the calculated Freundlich isotherm.

In a follow up experiment (not yet published) we determined the adsorption of the triblock

copolymer F127, poly(ethylene oxide)-poly(propylene oxide)-poly(ethylene oxide),

PEO98PPO67PEO98, onto a bilayer of egg phosphatidylcholine (EPC). With the QCM-DTM

technique we first showed that the same vesicles used above break down to bilayers upon

contact with the silica surface, this is illustrated in figure 4. The concentration of F127 used

was the same as in the SUVs experiment (corresponding to the concentration needed to reach

the plateau value of adsorbed amount) and the normalized adsorbed amount was determined

Page 52: Adsorption of Surfactants at the Solid-Liquid Interface: A ...

44

to be the same as that found for adsorption of F127 on SUVs. Furthermore, the desorption of

the triblock copolymers from the bilayer surface was followed after rinsing the SUV

monolayer with pure buffer. It was found that the desorption process displayed rapid kinetics,

again indicating a weak interaction between the polymers and the lipid membrane. The

desorption of F127 was more complete from the phospholipid bilayer coated surfaces, as

compared to from SUV coated surfaces.

-200

-150

-100

-50

0

50

-1

0

1

2

3

4

5

0 20 40 60 80 100 120 140

f15

(Hz)

D15 (10-6)

Time (min)

Figure 4: Time evolution of the frequency ( ) and dissipation shift (- - ) using silica surfaces. In this

experiment SUVs of EPC was introduced in the measuring chamber at t = 10 and 30 min, followed by rinsing

with HEPES at t = 50 and 70 min. The block copolymer F127 was injected at t = 90 and 110 min. Finally the

polymer solution was replaced by HEPES at t = 130 min.

We suggest that this is due to the different geometries of the vesicle and bilayer coated

surface i.e. some polymers are trapped between the vesicles and the surface. The change in

dissipation factor with polymer adsorption (∆D) was larger when the F127 adsorption

occurred on the vesicle coated surface as compared to on the bilayer. However, when

normalizing the change in ∆D with the change in ∆f occurring upon adsorption of F127, we

reach the same value for the bilayer case and the SUV case. This will be discussed further

below.

Page 53: Adsorption of Surfactants at the Solid-Liquid Interface: A ...

45

We continued our experiments with different tri-block copolymers both at surfaces coated

with SUV and surfaces coated with a bilayer. Our data, that are not yet published,

demonstrate that it is the length of the hydrophilic tail that is responsible for the change in the

dissipation factor occurring upon polymer adsorption. For F88 (PEO103PPO39PEO103), F87

(PEO61PPO40PEO61) and P85 (PEO26PPO40PEO26), the dissipation factor is increasing with

increased PEO chain length, whereas the adsorbed amount is increasing with increasing chain

length of the hydrophobic group (PPO). Further, with the same hydrophilic chain length we

obtain almost similar changes in dissipation value, independent of the hydrophobic chain

length (see table 1). F127 (PEO98P P O50PEO98) ⇒ ∆D = 1.6 ± 0.3*10-6 and F88

(PEO103PO39PEO103) ⇒ ∆D = 1.75 ± 0.3*10-6. Though, the adsorbed amount was considerably

different the observed ∆f (F127) corresponds to 0.95 mg/m2 and ∆f (F88) corresponds to 0.65

mg/m2.

Triblock copolymer ∆∆∆∆f [[[[Hz]]]] ∆∆∆∆D [[[[10-6]]]] ∆∆∆∆D/∆∆∆∆f [[[[10-6/Hz]]]]

F127 (PEO98PPO67EO98,) 16 1.6 0.1

F88 (PEO103PPO39PEO103) 11 1.75 0.16

F87 (PEO61PPO40PEO61) 17 0.8 0.05

P85 (PEO26PPO40PEO26) 23 0.5 0.02

Table 1: A comparison between different triblock copolymers adsorbed onto a phospholipid bilayer.

Interpretation of the dissipation factor.

The energy dissipation factor describe the rate with which energy is dissipated in the system

and the instrument can only measure the total energy dissipation rate. As discussed in the

methods section, many different molecular mechanisms contribute to the measured effect. The

much larger energy dissipation rate observed for adsorbed vesicles compared to adsorbed

bilayers highlights that the coupling to the liquid outside the surface is increased when the

thickness, and roughness, of the surface coating increases. More interesting is that adsorption

of the same polymer on the vesicle coated and the bilayer coated surface gives rise to the

same change in energy dissipation per unit adsorbed polymer mass. This is a strong indication

that the contributions to the energy dissipation from the underlying adsorbed layer

(phospholipid vesicles or bilayers) and from the polymer layer adsorbed on top of it are

additive. A consequence of this result is that adsorption of the polymers to the vesicles does

Page 54: Adsorption of Surfactants at the Solid-Liquid Interface: A ...

46

not affect the vesicles visco-elastic properties as determined at a frequency of 15 MHz. This is

remarkable since the polymer adsorption clearly induces structural changes in the vesicles as

probed by fluorescein leakage.

We further note that it is the PPO block of the polymer that adsorbs to the phospholipid

coated surface, but the length of this block has only a limited effect of the energy dissipation.

On the other, hand the hydrophilic block (PEO) that extends into the solution provides an

efficient pathway to energy dissipation due to its strong interaction with the surrounding

water.

The effect of the packing of surfactant layers, above cmc, for a series of gemini surfactants

were investigated in paper IV. We noted that even though the packing was significantly

different for the different surfactants (see table 3) the change in energy dissipation was very

similar ∆D = 0.5 ± 0.1*10-6. This lack of correlation between layer structure and dissipation is

interesting and indicates that the resolution of the dissipation factor from QCM-DT M

measurements is not sufficient to describe the viscoelastic character of the thin surfactant

film. The reason being that for thin surfactant films other mechanisms dominate the energy

dissipation.

Page 55: Adsorption of Surfactants at the Solid-Liquid Interface: A ...

47

5.3 Adsorption of surfactants studied with the QCM-DTM.

The study of a range of gemini surfactants as described in paper IV, also gave interesting and

unexpected results. In this study, a series of alkanediyl α ω, -bis(alkyldimethylammonium

bromide) or (CnH2n+1)[N+(CH3)2](CH2)s[N

+(CH3)2] (CmH2m+1) 2Br-, (m=n), was used, which will

be referred to as m-s-m in the following. A series consisting of four different spacers, namely

12-3-12, 12-6-12, 12-8-12 and 12-12-12 was investigated.

The study was concerned with adsorption and frictional properties of gemini surfactants at

hydrophilic gold surfaces, using the QCM-DTM, and the Atomic Force Microscopy (AFM)

technique. The molecular packing of a series of gemini surfactants was determined from

QCM-DTM measurements (see Table 1) and the frictional behaviour of the surfactant films

was characterized by employing the AFM.

The results show that by changing the length of the spacer group from 3 to 12 a systematic

change in the molecular packing at the surface is obtained (see table 2). It was found that an

increase in the molecular packing resulted in a lower frictional force between the surfactant

coated surfaces. The frictional results can be viewed upon either as controlled by the rigidity

of the surfactant layer or as a result of defects, holes, in the lubricating film.

gemini Area [[[[Å2]]]]

12-3-12 106 ± 2

12-6-12 131 ± 8

12-8-12 140 ± 4

12-12-12 171 ± 6

Table 2: Area per surfactant versus spacer length for a series of gemini surfactants at a gold surface, assuming a

bilayer model.

In figure 5, the friction versus load measurements between a tungsten probe/sphere and a gold

surface in the presence of gemini surfactant solutions (2 mM) of varying spacer length is

shown. For applied forces over 20 nN all the gemini surfactant coated surfaces show an

almost linear dependence on load. The difference between the surfactants is visible and a

trend is observed between the friction force and spacer length. The observation is that the

Page 56: Adsorption of Surfactants at the Solid-Liquid Interface: A ...

48

friction force at any given load increases with spacer length. Converting the frictional data (in

the 10-70 nN regime) to friction coefficient gives that the gemini with the shortest spacer, 12-

3-12, displays the lowest friction coefficient in this study. Adsorption of the surfactants with

spacer, s = 6 and s = 8 results in similar friction coefficients, and the highest friction

coefficient is observed between layers of the surfactant with the longest spacer 12-12-12.

0

5

10

15

20

0 20 40 60 80

Applied Force nN

Fri

ctio

n nN

Figure 5: Friction-load measurements between a tungsten particle and a gold surface in a 2 mM dimeric

surfactant aqueous solution. Mean values of friction force as a function of spacer length for a series of gemini

surfactants 12-s-12, where results for s = 3 (o), 6 (∆), 8 (-) and 12 (+) are displayed.

A clear trend between the spacer length and thereby the surfactant structure in relation to

friction force or friction coefficient is observed. This gives us reason to believe that it is

possible to predict the friction properties of a surfactant film from the geometry of the

surfactant. The prediction being that surfactants that pack efficiently on the surface, critical

packing parameter close to 1, should provide the lowest friction coefficient.

Page 57: Adsorption of Surfactants at the Solid-Liquid Interface: A ...

49

5.4 Counterion effects on sensed mass and energy dissipation.

In papers III, IV and paper VI, adsorption of cationic surfactants at solid surfaces (gold) and

chemically modified surfaces (SAMs) has been studied, and compared to adsorption of a non-

ionic surfactant (paper III). One of the purposes of these studies was to determine if

adsorption measurements using the QCM-DTM (Quartz Crystal Microbalance-Dissipation)

could reveal complementary information to that revealed by ellipsometry. Of particular

interest was to learn if there was a measurable visco-elastic effect (as obtained from the

dissipation parameter measured with the QCM-DTM) of the surfactant films. This has been

discussed in a previous section; in addition, we also address the role of the counterion for

ionic surfactant adsorption as quantified by the use of the QCM-DTM.

In paper III this is done indirectly by interpreting dissipative changes upon changing the

surface hydrophobicity as a result of associated counterions within the adsorbed layer

structure. Depending on the hydrophobicity of the surface, the Br- ion will be more or less

incorporated in the adsorbed layer structure and thereby changing the rigidity of the layer as

interpreted from changes of the dissipation values. In this study the use of a combination of

hydrophobic and hydrophillic thiols enabled a systematic change of the surface

hydrophobicity of the substrate. Thiohexadecane, HS(CH2)15CH3 and thiohexadecanol

HS(CH2)16OH were mixed in different ratios to obtain variations in surface hydrophobicity. It

has previously been shown that mixtures of molecules with equal chain length have almost

the same SAM (self-assembled monolayers) composition as the composition in the solution.

For this reason we assume that the mole fraction of the SAM forming molecules at the surface

is the same as in bulk solution. For model surfactants we used the cationic DTAB,

(dodecyltrimethylammonium bromide) and the non-ionic C12EO8, (octa-(ethylene oxide)

mono n-dodecyl ether), both having the same hydrophobic chain length, but very different

hydrophilic headgroup and molecular weight.

In table 3 and 4, the contact angle (θ), the adsorbed mass (Γ), and the dissipation factor (∆D),

for the five different mixtures of thiols and both surfactants are displayed. The surfactant

concentration was just above cmc (1.2 cmc). The mixture is defined as the mole percentage of

thiohexadecane, in a mixture of thiohexadecane and thiohexadecanol.

Page 58: Adsorption of Surfactants at the Solid-Liquid Interface: A ...

50

Mixture θθθθ ΓΓΓΓDTAB [[[[mg/m2]]]] ∆∆∆∆DDTAB [[[[10-6]]]]

0% 23 1.40 ± 0.06 1.17 ± 0.61

25% 40 1.04 ± 0.07 1.70 ± 0.38

50% 60 2.17 ± 0.19 2.23 ± 0.52

75% 83 2.01 ± 0.09 2.26 ± 0.25

100% 105 2.19 ± 0.08 2.85 ± 0.16

Table 3: Adsorbed mass ( ΓDTAB ), and the dissipation factor ( ∆DDTAB ) versus contact angle ( θ ), for a 1.2*cmc

solution of DTAB.

Mixture θθθθ ΓΓΓΓC12EO8 [[[[mg/m2]]]] ∆∆∆∆DC12EO8 [[[[*10-6]]]]

0% 23 1.32 ± 0.10 0.31 ± 0.08

25% 40 1.87 ± 0.06 0.17 ± 0.05

50% 60 2.19 ± 0.05 0.40 ± 0.22

75% 83 2.08 ± 0.07 0.60 ± 0.25

100% 105 2.07 ± 0.08 0.22 ± 0.17

Table 4: Adsorbed mass ( ΓC12EO8 ), and the dissipation factor ( ∆DC12EO8 ) versus contact angle ( θ ), for a

1.2*cmc solution of C12EO8.

In determining the area per molecule for DTAB, the counterion, Br-, was assumed to be

incorporated in the surfactant layer and contributing to the frequency response. With this

assumption an area per molecule of about 25 Å2 was observed for the 100% SH-C16 surface.

This estimate considers a perfect degree of counterion association, β , to the monolayer

structure. From surface force measurements it has been observed that between 80-90 % of the

bromide and chloride counterions appear to be bound to similar cationic surfactants layers on

mica surfaces. This makes us estimate the actual degree of counterion binding of the Br- to a

maximum of 90%. Converting the first overtone frequency, 15 MHz to, oscillating period

gives 7*10-8 s, which is in the same order as the residence time for a counterion at the micellar

surface. In addition, the electrostatic potential outside a flat surface is larger than outside

Page 59: Adsorption of Surfactants at the Solid-Liquid Interface: A ...

51

micelles, which should support an even longer residence time. It is also more difficult for ions

to diffuse from a flat surface than from a spherical micelle, simply for geometrical reasons.

All together, this makes the assumption regarding the Br- ion contributing to the adsorbed

mass sensed by the crystal reasonable. We note that even so the mass registered by the QCM-

device is significantly larger than that obtained by ellipsometry (corresponding to about 40-45

Å2 per molecule). We attribute this to the fact that the QCM-device also registers bound and

trapped water as discussed in chapter 5.5. The large dissipation change occurring as a result of

DTAB adsorption, as compared to C12EO8, adsorption is surprising since the hydrophilic part

of the C12EO8 is large and rather strongly hydrated. We suggest that the effect is due to the

large interaction between the charged adsorbed DTAB layer and the oppositely charged

double layer outside the surface. This coupling increases both the sensed mass and the

dissipation factor.

Page 60: Adsorption of Surfactants at the Solid-Liquid Interface: A ...

52

5.5 Bound / trapped water.

It is shown that the frequency shift as obtained from the QCM-DTM experiments results in an

overestimation of the adsorbed mass (paper III-V) if it, erroneously, is interpreted as being

due to the surfactant only This is illustrated by table 5 that shows the adsorbed mass

determined by ellipsometry and the mass sensed by the QCM-DTM device for two surfactant

systems (slightly above their cmc) on two different surfaces. We suggest that this is due to

two different effects, viz., water that is coupled to the adsorbed layer due to hydration of the

polar region of the surfactant, and secondly water that for other reason are trapped within the

adsorbed layer. Both trapped water and water of hydration will contribute to the frequency

shift provided that it moves with the oscillating crystal over the time scale of the experiment,

where one oscillating period for the fundamental mode is of the order of 0.2 µs. T o

understand the hydration of hydrophilic surfactant headgroups it is worthwhile to recapitulate

some of the properties of water.

Water is far from being a simple liquid if not unique1. The complexity of liquid water is due

to a combination of the small size and distinct polar charge distribution of the water

molecule2. The hydrogen bond is caused by a combination of electrostatic, charge transfer,

dispersion forces and exchange forces1. It is nowadays often assumed that the electrostatic

part is the most important one. The charge distribution of the water molecules can be

modelled by four charges being located in the form of a tetrahedron3, which allows each water

molecule to participate in four strong interactions4-5 with a high degree of spatial

directionality. The energy of a hydrogen bond between two water molecules depends on the

oxygen-oxygen distance and the hydrogen-oxygen angle1. This strong (water-water)

interaction results in a large cohesive energy, a high boiling point, a high surface tension, and

a reluctance to dissolve inert (non-polar, hydrophobic) solutes with which water cannot

interact through similarly strong forces. However, water can also bind to and dissolve polar

and hydrophilic substrates. The binding of water to hydrophilic groups can conveniently be

studied on macroscopic surfaces using various surface force techniques.

By adsorbing surfactants onto the substrate surfaces in such a way that the polar part of the

surfactant is directed towards solution, the hydration of the surfactant layer can be

investigated. Such studies have shown that in addition to attractive van der Waals forces and

Page 61: Adsorption of Surfactants at the Solid-Liquid Interface: A ...

53

repulsive electrostatic double – layer forces, an additional repulsive force is present at short

separations. This force is due to a combination of dehydration of polar groups (a hydration

force)7 and restriction of the perpendicular motion of adsorbed molecules in the gap between

the surfaces, a steric/protrusion force8. The measurable range of this force is typically 1-3 nm,

and it decays roughly exponentially with surface separation having a decay length of about

0.1 - 0.3 nm. Let us for the moment assume that the major cause of this force is dehydration.

The water molecules adjacent to the polar surface, or a surfactant headgroup, are strongly

affected and can be regarded as bound, In e.g. NMR-studies of ethylene oxide-based

surfactants, the amount of bound water is estimated to be about 2 - 8 per ethylene oxide unit9.

However, also water molecules further away will be affected by the presence of a surface or a

polar surfactant headgroup, but less strongly so. It is not seen as “bound” in a NMR

experiment, but it is sufficiently affected to give rise to a “hydration force” in a surface force

experiment. How much of the “hydration water” that is sensed in a QCM-measurement

remains an open question.

Surface ΓΓΓΓ [[[[mg/m2]]]] Ra [[[[nm]]]] θθθθ

Silicaellip 2.12 ± 0.10 1.3 ± 0.1 < 20°

Silanellip. 1.58 ± 0.10 1.1 ± 0.2 98° ± 1°

SilicaQCM 4.25 ± 0.24 1.3 ± 0.2 < 20°

SilanQCM 2.73 ± 0.06 1.0 ± 0.1 98° ± 1°

Table 5: Adsorbed amount ( Γ ), roughness average ( Ra ), and the Contact angle ( θ ), for the silica and

silanised silica model surfaces.

Trapped water on the other hand, involves water molecules that are confined within a limited

geometry such as inside a vesicle, beneath a bilayer, or between tightly packed

micellar/vesicle structures and the surface. This trapped water will also be contributing to the

frequency shift of the QCM-DTM. The experimental observation is that quite a lot of water

contributes to the adsorbed mass sensed by the QCM. For instance, when we compare results

for C14EO6 adsorption as elucidated from the QCM measurements with that obtained from

ellipsometry (paper V) we found that approximately 73 % of the mass sensed by the QCM-

DTM was due to water for silanated silica substrates. For the hydrophilic silica surface this

Page 62: Adsorption of Surfactants at the Solid-Liquid Interface: A ...

54

overestimation was approximately 100 %. Hence, this effect is very large and it can certainly

not be ignored. In order to verify the similarity of the substrates employed using the two

different techniques, surface properties such as surface roughness and contact angles were

also determined. This comparison is elucidated in Table 2. Clearly, the surfaces are similar

enough to expect similar adsorption. Hence, the only possible interpretation is that water

contributes significantly to the mass sensed by the QCM-technique.

Page 63: Adsorption of Surfactants at the Solid-Liquid Interface: A ...

55

6. Concluding remarks

This thesis deals with the adsorption of surfactants, vesicles and emulsions at the solid-liquid

interface. A large part of the work has been concerned with the problems and ambiguities

concerning the interpretations of the frequency shift and the dissipation factor as determined

with a QCM-DTM device. Hence, methods to prepare and clean various solid surfaces are

described. We could use the fact that the electrodes on the piezoelectric quartz crystal are

made of gold, and the gold surfaces makes it possible to use thiol chemistry to make chemical

modifications. Thiol chemistry is probably the best surface modification technique available

at the moment; the possibilities are endless with different functionalised end groups being

available. Studies of adsorption at these well documented, very stable substrates is a dream

that came through for the surface scientist. Studies of concentrated emulsions are possible

with the QCM-DTM technique, simply because there are no need for a transparent solution,

which is a prerequisite for many adsorption techniques today. Another strength of the QCM-

technique is that adsorption onto previously adsorbed layers can easily be followed, which

also is a problem for most light-based techniques, since the modelling parameters increase

rapidly. Such adsorption upon adsorption experiments have to be carefully interpret since the

chemistry can be really complex, and the results ambiguous. One of the big “problems” with

the QCM-technique is that it also registers water associated with the adsorbed layer. This

results in an overestimation of the adsorption of the “pure” adsorbate. This “problem” can

easily be turned into an advantage by comparing with results obtained by light based

techniques, which gives the adsorbed amount without any hydration effects or effects of

trapped water. The result of this correlation gives an estimation of the amount of water

associated with the adsorbed layer. Considering all the literature based on hydration, and

hydration effects this is actually not a “problem”, instead it could be turned into something

really valuable for the scientific community. So, like all other techniques the QCM-DTM has

its drawbacks and advantages, the only real problem is when you are not aware of them.

Page 64: Adsorption of Surfactants at the Solid-Liquid Interface: A ...

56

Acknowledgements

It is now 2 years, 8 months and 10 days since I started my Ph.D study in the Department ofChemistry, Surface Chemistry. Hence, I am truly indebted to more people than I thought that Iwould be. So, this is for the assistance through the oscillations of my life and work.

Per Claesson, for giving me the scientific freedom, that I needed, and for being my supervisor during this time.

Katrin Boschkova, one of the best.

Jan-Christer Eriksson, for all words of wisdom.

The happily graduated surface chemists:Eva B, Magnus B, Thomas E and Andra D.

The unhappily ungraduated surface chemists:Marcus P, Mikael K, Atte K, Jonny E, Torbjörn P, Marie E and Brita R.

Personnel at YKI:Martin M, Thomas A, Lennart B and Britt N.

The so close, and so far away people:Markus J, Nill B, Anna N, Mattias Ö, Ulla J and Pontus E.

YKI is acknowledged for the cookies, copies, printouts and stamps; these things havecertainly made my life easier during this time.

At Q-Sense, Michael Rodahl and Ralf Richter.

And finally The SSF-programme, Colloid and Interface Technology, for its financial support.

Page 65: Adsorption of Surfactants at the Solid-Liquid Interface: A ...

57

References.

Gold

(1) Ibach, H., “Solid-State Physics”, Springer-Verlag, Heidelberg, 1996.

(2) Hayward, D. O., “Chemisorption and Reactions on a Metallic Films”, Anderson, J. R.,

Ed., Academic Press, London, 1971

Self-Assembled Monolayers (SAMs)

1) Ulman, A., “An Introduction to ULTRATHIN ORGANIC FILMS From Langmuir-

Blodgettto Self-Assembly”, Academic Press Inc. San Diego, CA, 1991.

(2) Swalen, J. D.; Allara, D. L.; Andrade, J. D.; Chandross, E. A.; Garoff, S.; Israelachvili, J.;

McCarthy, T. J.; Murray, R.; Pease, R. F.; Rabolt, J. F.; Wynne, K. J.; Yu, H., Langmuir,

1987, 3, 932.

(3) Nuzzo, R. G.; Allara, D. L. J., J. Am. Chem. Soc., 1983, 105, 4481.

(4) Alves, C. A; Smith, E. L.; Porter, M. D., J. Am. Chem. Soc., 1992, 114, 1222.

(5) Bain, C. D.; Troughton, E. B.; Tao, Y. T.; Evall, J.; Whitesides, G. M.; Nuzzo, R. G., J.

Am. Chem. Soc., 1989, 111, 321.

(6) Sellers, H.; Ulman, A.; Schnidman, Y.; Eilers, J. E., J. Am. Chem. Soc., 1993, 115, 9389.

(7) Samant, M. G.; Brown, C. A.; Gordon, J. G. I., Langmuir, 1991, 7, 437.

(8) Siepman, J. I.; McDonald, I., R., Molecular Physics, 1993, 79, 457.

(9) Allara, D. L.; Nuzzo, R. G., Langmuir, 1985, 1, 45.

(10) Bensebaa, F.; Voicu, R.; Huron, L.; Ellis, T. H., Langmuir, 1987, 13, 5335.

(11) Bertilsson, L.; Liedberg, B., Langmuir, 1993, 9, 141.

Silica

(1) Iler, R. K., “The Chemistry of Silica”, John Wiley and Sons, New York, 1979.

(2) Yaminsky, V. V.; Ninham, B. W.; Pashley, R. M., Langmuir, 1998, 14, 3223.

(3) Eskilsson, K.; Yaminsky, V. V., Langmuir, 1998, 14, 2444.

Page 66: Adsorption of Surfactants at the Solid-Liquid Interface: A ...

58

Silan coated silica

1) Eskilsson, K.; Tiberg, F., Macromolecules, 1997, 30, 6323.

(2) Parker, J. L.; Claesson, P. M.; Attard, P., J. Phys. Chem., 1994, 98, 8468.

(3) Yaminsky, V. V., Colloids and Surfaces A, 1998.

(4) Yaminsky, V. V.; Claesson, P. M.; Eriksson, J. C., Journal of Colloid and Interface

Science, 1993, 161, 91.

(5) Christenson, H. K.; Claesson, P. M., Advances in colloid and Interface Science, 2001, 91,

391.

(6) Rabinovich, Y. I; Derjaugin, B. V., Colloids and Surfaces, 1988, 30, 243.

(7) Parker, J. L., “ Progress in Surface Science”, 1994, 47, 205.

(8) Claesson, P. M.; Kjellin, U. R. M., “ Modern Characterization Methods of Surfactant

Systems”, ed Binks, B. P., Surfactant Science Series, Marcel Dekker, New York, 1999, 83,

255.

Contact angle

(1) Cassie, A. B. D., Faraday Disc. Chem. Soc., 1948, 3, 11.

(2) Israelachvili, J. N.; Gee, M. L., Langmuir, 1989, 5, 288.

XPS

X-ray Photoelectron Spectroscopy

(1) Siegbahn, K., Science, 1982, 217, 111.

(2) Czanderna, A. W. (Ed.), “Methods of Surface Analysis”, Vol. 1 of ”Methods and

Phenomena. Their Applications in Science and Technology”, Elsevier Publ. Co., Amsterdam,

1975.

(3) Briggs, D., (Ed.), “Handbook of X-ray and Ultraviolet Photoelectron Spectroscopy”,

Heyden, London, 1977.

(4) Carlson, T. A., “Photoelectron and Auger Spectroscopy”, Plenum Press, New York, 1975.

(5) Briggs, D.; Seah, M. P., (Ed), “Practical Surface Analysis”, John Wiley & Sons,

Chichester, 1983.

Page 67: Adsorption of Surfactants at the Solid-Liquid Interface: A ...

59

QCM-DTM

(1) Cady, W. G., “Piezoelectricity”, 2nd ed, Dover, New York, 1964.

(2) Lu, C.; Czanderna, A. W. (eds), “Applications of Piezoelectric Quartz Crystal

Microbalances”, Elsevier, Amsterdam, 1984.

(3) Warner, W. W.; Stockbridge, C. D., “Vacuum Microbalance Techniques”, Plenum Press,

New York, 1963, 3, 55.

(4) Rodahl, M.; Höök, F.; Krozer, A.; Brzezinski, P.; Kasemo, B., Rev. Sci. Instr., 1995, 66,

3924.

(5) Kasemo, B.; Törnqquist, E., Surface Science, 1978, 77, 209.

(6) Kasemo, B.; Törnqquist, E., Phys. Rev. Lett., 1980, 44, 1555.

(7) Bishop, D. J.; Reppy. J. D., Phys. Rew. Lett., 1978, 40, 1727.

(8) Mecea, V. M., Sensors and Actuators A, 1993, 40, 1.

(9) Kleinman, R. N.; Kaminsky, G. K.; Reppy, J. D.; Pindak, R.; Bishop, D. J., Rev. Sci.

Instr., 1985, 56, 2088.

(10) Widom, A.; Krim, J., Phys. Rev. B, 1986, 34,1403.

(11) Krim, J.; Chiarello, R., J. Vac. Sci. Tech. B, 1991, 9, 1343.

(12) Caruso, F.; Rinia, H. A.; Furlong, D. N., Langmuir, 1996, 12, 2145.

(13) Caruso, F.; Serizawa, T.; Furlong, D.; Okahata, Y., Langmuir, 1995, 11, 1546.

(14) Furlong, D.; Aston, J. R., Colloids and Surfaces A, 1982, 4, 121.

(15) Salt, D., “Hy-Q Handbook of Quartz Crystal Devices”, van Nostrand Reinhold Co.

Padstow, Cornwall, 1987.

(16) Brice, J. C., Rew. Mod. Phys., 1985, 57, 105.

(17) Rodahl, M., Thesis, “On the frequency and Q factor response of the quartz crystal

microbalance to liquid overlayers”, Chalmers University of Technology, Gothenburg, 1995.

(18) Lack, F. R.; Willard, G. W.; Fair, I. E., Bell System Technical Journal, 1934, 13, 453.

(19) Sauerbrey, G., Z. Phys. (german), 1959, 155, 206.

Page 68: Adsorption of Surfactants at the Solid-Liquid Interface: A ...

60

(20) Pulker, H. K. Z., Angew, Phys, 1966, 20, 537.

(21) Mecea, V.; Bucur, R. V., Thin Solid Films, 1979, 60, 73.

(22) Nomura, T.; Hattori, O., Analytica Chimica Acta, 1980, 115, 323.

(23) Kanazawa, B. K.; Gordon, J. G., Analytica, Chimica, Acta, 1985, 175, 99.

(24) Stockbridge, C. D., “Vacuum Microbalance Techniques”, Plenum Press, New York,

1966, 5, 147.

(25) Martin, S. J.; Frye, G. C.; Wessendorf, K. O., Sensors and Actuators A, 1994, 44, 209.

(26) Smith, K. L., Electronics & Wireless World, 1986, july, 51.

(27) Spencer, W. J.; Smith, K. L., J. Appl. Phys., 1966, 37, 2557.

(28) Barnes, C.; D’Silva, C.; Jones, J. P.; Lewis, T. J., Sensors and Actuators B, 1991, 3, 295.

(29) Martin, S. J.; Frye, G. C.; Ricco, A. J.; Senturia, S. D., Anal. Chem., 1993, 65, 2910.

(30) Lin, Z.; Ward, M. D., Anal. Chem., 1995, 67, 685.

(31) Lucklum, R.; Schranz, S.; Behling, C.; Eichelbaum, F.; Hauptmann, P., Sensors &

Actuators A, 1997, 60, 40.

(32) Schneider, T. W.; Martin, S. J., Anal. Chem., 1995, 67, 3324.

(33) Bund, A.; Schwitzgebel, G., Analytica Chimica Acta, 1998, 364, 189.

(34) Tessier, L.; Patat, F.; Schmitt, N.; Feuillard, G., Anal. Chem., 1994, 66, 3569.

(35) Mecea, V. M.; Carlson, J. O.; Bucur, R. V., Sensors and & Actuators A, 1996, 53, 371.

(36) Martin, B. A.; Hager, H. E., J. Appl. Phys., 1989, 65, 2627.

(37) Ward, M. D.; Buttry, D. A., Science, 1990, 249, 1000.

(38) Frye, G. C.; Martin, S. J., J. Appl. Spec. Rev., 1991, 26, 73.

(39) Kanazawa, K., Faraday Discussions, 1997, 107, 77.

(40) Martin, B. A.; Hager, H. E., J. Appl. Phys., 1989, 65, 2630.

(41) Borovsky, B.; Mason, B. L.; Krim, J., J. Appl. Phys., 2000, 88, 4017.

Page 69: Adsorption of Surfactants at the Solid-Liquid Interface: A ...

61

Lipids

(1) The whole issue of,“Chemistry and Physics of Lipids”, 1991, 57, 1.

(2) Lewis, R. N. A. H.; Mak, N.; Edhaney, R. N. M., Biochemistry, 1987, 26, 6118.

(3) Seddon, J. M.; Cevc, G.; Kayl, R. D.; Marsh, D., Biochemistry, 1984, 23, 2634.

(4) Lindblom, G.; Rilfors, L., Biochim. Biophys. Acta, 1989, 988, 221.

(5) “Lipids-Molecular Organization, Physical Functions and Technical Applications”,

Larsson, K, The Oily Press Ltd, Dundee, Scotland, 1994.

Cationic surfactants

(1) Rosen, M. J., “ Surfactants and Interfacial Phenomena”, John Wiley, New York, 1989, 33.

(2) Mittal, K. L., “ Solution Chemistry of Surfactants” Pharm Press, New York, 1979, 1.

(3) Florence, A. T., Attwood, D., “ Physiochemical Principles of Pharmacy”, Macmillan,

Houndsmill, 1985, 174.

(4) Yeskie, M. A.; Harwell, J. H., J. Phys. Chem., 1988, 92, 2346.

(5) Koopal, L. K.; Keltjens, L., Colloids and Surfaces, 1986, 17, 371.

(6) Gao, Y.; Du, J.; Gu, T., J. Chem. Soc., Faraday Trans. I, 1987, 83, 2671.

(7) Gu, T.; Gao, Y.; He, L., J. Chem. Soc., Faraday Trans. I, 1988, 84, 4471.

(8) Gu, T.; Huang, Z., Colloids and Surfaces, 1989, 40, 71.

(9) Gu, T.; Zhu, B.Y., Colloids and Surfaces, 1990, 44, 81.

(10) Koopal, L. A.; Ralston, J., J. Colloid Interface Sci., 1986, 112, 362.

(11) Rupprecht, H.; Kindl, G., Progr. Colloid Polymer Sci., 1976, 60, 194.

(12) Hunter, R. J., “ Foundation of Colloid Sciences”, Oxford Sciences, Oxford, 1987, 1, 591.

(13) Koopal, L. K.; Wilkinson, G. T.; Ralstone, J., J. Colloid Interface Sci., 1988, 126, 493.

(14) Zhu, B. Y.; Gu, T., J. Chem. Soc. Faraday Trans. I, 1989, 85, 3813.

(15) Lee, G.; Rupprecht, H., J. Colloid Interface Sci., 1985, 105, 257.

Non-ionic surfactants

(1) Desbene, P. L.; Portet, F.; Treiner, C. J., J. Colloid Interface Sci., 1997, 190, 350.

(2) Böhmer, M. R.; Koopal, L. K.; Janssen, R.; Lee, E. M.; Thomas, R. K.; Rennie, A. R.,

Langmuir, 1992, 8, 2228.

Page 70: Adsorption of Surfactants at the Solid-Liquid Interface: A ...

62

(3) Partyka, S.; Zaini, S.; Lindheimer, M.; Brun, B., Colloids Surf., 1984, 12, 255.

(4) Penfold, J.; Staples, E.; Tucker, I.; Cummins, P., J. Phys. Chem., 1996, 100, 18133.

(5) Portet, F.; Desbene, P. L.; Treiner, C., J. Colloid Interface Sci., 1996, 184, 216.

(6) Tiberg, F.; Jönsson, B.; Tang, J.; Lindman, B., Langmuir, 1994, 10, 2294.

(7) Mandell, L.; Fontell, K.; Ekwall, P., “ Ordered Fluids and Liquid Crystals”, Gould, R. F.

ed., American Chemical Society, Washington D. C., 1967, 63, 89.

(8) Ekwall, P., “ Advances in Liquid Crystals”, Brown, G. H. ed., Academic Press, New York,

1975, 1, 1.

(9) Luzzati, V.; Mustacchi, H.; Skoulios, A.; Husson, F., Acta Cryst., 1960, 13, 660.

(10) Luzzati, V.; Mustacchi, H.; Skoulios, A., Disc. Farad. Soc., 1958, 25, 43.

(11) Tiberg, F., J. Chem. Soc, Faraday Trans., 1996, 92, 531.

(12) Levitz, P., Langmuir, 1992, 7, 1595.

(13) Levitz, P.; Van Damme, H.; Keravis, D., J. Phys. Chem., 1984, 88, 2228.

(14) Levitz, P.; Van Damme H., J. Phys. Chem., 1986, 90, 1302.

Adsorption of emulsions studied with the QCM-DTM

(1) Malmsten, M.; Lindström, T.; Wärnheim, T., J. Colloidal Interface Sci., 1995, 173, 297.

(2) Blomberg, E.; Claesson, P. M.; Wärnheim, T., Colloids Surf, (in press).

Polymer adsorption on phospholipid coated surfaces

(1) Keller, C.; Kasemo, B.; Biophys. J., 1998, 75, 1397.

Counterion effects on sensed mass and energy dissipation

(1) Bertilsson, L.; Liedberg, B., Langmuir, 1993, 9, 141.

Bound /trapped water

(1) Frank, F., “ Water: A Comprehensive Treatise”, Plenum Press, New York, 1972-1982, 1-

7.

(2) Marcelja, S., “ Liquids at Interfaces”, Elsevier Science, Amsterdam, 1990, 2,99.

(3) Coulson, C. A., “ Valence”, Oxford Uni. Press, London, 1961, 12.

Page 71: Adsorption of Surfactants at the Solid-Liquid Interface: A ...

63

(4) Tanford, C., “ The hydrophobic Effect”, Wiley, New York, 1980.

(5) Hunter, R. J., “ Foundations of Colloid Science”, Claredon, Oxford, 1989, 1.

(6) Israelachvili, J. N., “ Intermolecular and Surface Forces with Applications to Colloidal

and Biological Systems”, Academic Press, London, 1992.

(7) Rand, R. P.; Parsegian, V. A., Biochim. Biophys. Acta, 1989, 988, 351.

(8) Israelachvili, J.; Wennerström, H., Nature, 1996, 379, 219.

(9) Derjaugin, B. V.; Churaev, N. V., “ Fluid Interfacial Phenomena”, Wiley, Chichester,

1986, 15, 663.

(10) Nilsson, P. G.; Wennerström, H.; Lindman, B., J. Phys. Chem., 1983, 87, 1377.

(11) Drost Hansen, W., “ Biophysics in Water”, Wiley, Chichester, 1982, 163.

(12) Israelachvili, J. N., Chemica Scripta, 1985, 25, 7.

(13) Israelachvili, J. N.; Pashley, R. M., “ Biophysics of Water”, Wiley, Chichester, 1982,

183.


Recommended