diva portal13308/fulltext01.pdf · 2008. 3. 7. · the friction properties in aqueous solution have...

72

Upload: others

Post on 24-Feb-2021

2 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: DiVA portal13308/FULLTEXT01.pdf · 2008. 3. 7. · The friction properties in aqueous solution have been studied for adsorbed layers of synthetic polyelectrolytes, with varying density

������������ �������������������� ���� �������

������������������ �������

�������������������

������� ��������! �"������# �"$��%� ���&''(

Page 2: DiVA portal13308/FULLTEXT01.pdf · 2008. 3. 7. · The friction properties in aqueous solution have been studied for adsorbed layers of synthetic polyelectrolytes, with varying density
Page 3: DiVA portal13308/FULLTEXT01.pdf · 2008. 3. 7. · The friction properties in aqueous solution have been studied for adsorbed layers of synthetic polyelectrolytes, with varying density

Lubrication and Surface Properties of Adsorbed Layers of Polyelectrolytes

and Proteins

TORBJÖRN PETTERSSON

Doctoral Thesis Stockholm, Sweden 2008

Page 4: DiVA portal13308/FULLTEXT01.pdf · 2008. 3. 7. · The friction properties in aqueous solution have been studied for adsorbed layers of synthetic polyelectrolytes, with varying density

ii

Akademisk avhandling som med tillstånd av Kungliga Tekniska högskolan framlägges till offentlig granskning för avläggande av filosofie doktorsexamen torsdagen den 27:e mars 2008 kl. 10.00 i hörsal F3, Lindstedtsvägen 26, Stockholm. Torbjörn Pettersson. Lubrication and surface properties of adsorbed layers of polyelectrolytes and proteins. TRITA CHE-Report 2008:16 ISSN 1654-1081 ISBN 978-91-7178-891-7 Kungliga Tekniska högskolan Skolan för kemi och kemiteknik, Ytkemi Drottning Kristinas väg 51 SE-100 44 Stockholm Denna avhandling är skyddad enligt upphovslagen. Alla rättigheter förbehålles. All rights reserved. No part of this thesis may be reproduced by any means without permission from the author. Copyright © 2008 by Torbjörn Pettersson The following items are reproduced with permission: Paper I: © 2007, American Institute of Physics Paper II: © 2008, American Chemical Society. Paper III: © 2008, American Chemical Society Paper IV: © 2007, American Chemical Society. Paper V: © 2008, American Chemical Society. Paper VI: © 2008, Springer-Verlag Berlin Heidelberg. Cover illustration: Collage of SEM images, a silica particle attached as a colloidal probe on an AFM cantilever with mica substrate as background (SEM images courtesy of Rodrigo Robinson) Printed in 200 copies at E-PRINT, Stockholm, February 2008

Page 5: DiVA portal13308/FULLTEXT01.pdf · 2008. 3. 7. · The friction properties in aqueous solution have been studied for adsorbed layers of synthetic polyelectrolytes, with varying density

iii

Abstract Friction forces between protein / polyelectrolyte layers, adsorption properties of proteins, and conformational changes due to variation in electrolyte concentration have been investigated. The aim was to obtain better understanding of adsorbed layer properties, with focus on the relation between layer structure and lubrication capabilities. The major techniques used were AFM (atomic force microscope) with colloidal probe for normal force and friction measurements together with QCM-D (quartz crystal microbalance with dissipation) for measurement of adsorption and conformational changes of adsorbed layers. A comparison between some techniques for calibration of the AFM instrument for friction measurements was made to find the most suitably one for colloidal probe friction measurements in aqueous solutions. It is suggested that the normal and torsional Sader methods are preferred in combination with torsional detector sensitivity measurement, for which one new methodology has been proposed. Adsorption was studied for bovine serum albumin, cytochrome c, myoglobin and mucin, whereas conformational changes of the adsorbed layer were monitored only for mucin. It was found that it was essential to take into account bulk density and viscosity changes for correct interpretations of QCM data when studying the effect of changes in electrolyte type and concentration on preadsorbed layers of mucin, and also when having different (high) concentrations of proteins in the measuring solution. The adsorbed amount of proteins appears to depend on the strength of the surface attachment, in such a manner that a too high affinity reduces the adsorbed amount. Friction properties in aqueous solution have been studied for adsorbed layers of PEO45MEMA:METAC co-polyelectrolytes, with varying density of grafted PEO45 side chains and varying charge density, as well as for a naturally occurring polyelectrolyte (chitosan) and the glycoprotein mucin. These polymers were used to cover a wide range of different types of adsorbed layers and interactions to gain a better understanding of friction mechanisms and demands on layer properties for achieving favourable lubrication. It was found that the common features of low friction layers are that no attractive forces are present, and that excluded volume and / or electrostatic forces counteract chain interpenetration under load.

Page 6: DiVA portal13308/FULLTEXT01.pdf · 2008. 3. 7. · The friction properties in aqueous solution have been studied for adsorbed layers of synthetic polyelectrolytes, with varying density

iv

List of papers

Included publications Paper I

Comparison of Different Methods to Calibrate Torsional Spring Constant and Photodetector for AFM Friction Measurements in Air and Liquid T. Pettersson, N. Nordgren, M. W. Rutland and A. Feiler Review of Scientific Instruments 78 (2007) p093702.

Paper II Lubrication Properties of Bottle-Brush Polyelectrolytes: An AFM Study on the Effect of Side Chain and Charge Density T. Pettersson, A. Naderi, R. Makuška, and P. M. Claesson Langmuir in press, DOI: 10.1021/la703229n. (Online 16-Feb-2008)

Paper III Normal and Friction Forces between Mucin and Mucin-Chitosan Layers in Absence and Presence of SDS T. Pettersson, A. Dėdinaitė Submitted.

Paper IV Probing Protein Adsorption onto Mercaptoundecanoic Acid Stabilized Gold Nanoparticles and Surfaces by Quartz Crystal Microbalance and ζ-Potential Measurements E. D. Kaufman, J. Belyea, M. C. Johnson, Z. M. Nicholson, J. L. Ricks, P. K. Shah, M. Bayless, T. Pettersson, Z. Feldötö, E. Blomberg, P. Claesson, and S. Franzen Langmuir 23 (2007) p 6053.

Paper V Mucin Layer Response to Electrolytes Studied with QCM-D Z. Feldötö, T. Pettersson, and A. Dėdinaitė Langmuir in press, DOI: 10.1021/la703366k. (Online 12-Feb-2008)

Paper VI

The Effect of Salt Concentration and Cation Valency on Interactions between Mucin-Coated Hydrophobic Surfaces T. Pettersson, Z. Feldötö, P. M. Claesson, and A. Dedinaite Progress in Colloid and Polymer Science in press.

Page 7: DiVA portal13308/FULLTEXT01.pdf · 2008. 3. 7. · The friction properties in aqueous solution have been studied for adsorbed layers of synthetic polyelectrolytes, with varying density

v

The author’s contribution to the papers is as follows: I. Major part of planning, major part of experimental work, major part of

writing II. Part of planning, part of experimental work, part of writing III. All planning, all experimental work, major part of writing IV. Part of planning, part of experimental work (QCM, viscosity and

density), part of writing V. Major part of planning, part of experimental work, major part of writing VI. Major part of planning, major part of experimental work, part of writing

Not included publications Spreading of Individual Toner Particles Studied Using in-situ Optical Microscopy T. Pettersson and A. Fogden Journal of Colloid and Interface Science 287 (2005) p 249 Levelling During Toner Fusing: Effects on Surface Roughness and Gloss of Printed Paper T. Pettersson and A. Fogden, Journal of Imaging Science and Technology 50 (2006) p 202. Friction Measurement Between Polyester Fibres Using the Fibre Probe SPM H. Mizuno, M. Kjellin, N. Nordgren, T. Pettersson, V. Wallqvist, M. Fielden, and M. W. Rutland Australian Journal of Chemistry 59 (2006) p 390–393 Thermal Calibration of Photodiode Sensitivity for Atomic Force Microscopy P. Attard, T. Pettersson and M. W. Rutland Review of Scientific Instruments 77 (2006) p 116110 Effect of Polymer Architecture on the Adsorption Properties of a Nonionic Polymer A. Naderi, J. Iruthayaraj, T. Pettersson, R. Makuška and P. M. Claesson Submitted Aggregate Structure of Highly Charged Cationic Polyelectrolyte and Anionic Surfactant Confined Between Mica-Silica Interfaces: A Systematic Study of Structural Forces and Frictional Forces B. Mohanty, T. Pettersson, A. Dėdinaitė, P. M. Claesson) Manuscript

Page 8: DiVA portal13308/FULLTEXT01.pdf · 2008. 3. 7. · The friction properties in aqueous solution have been studied for adsorbed layers of synthetic polyelectrolytes, with varying density

vi

Table of content Abstract ........................................................................................................ iii List of papers ................................................................................................ iv

Included publications ............................................................................... iv Not included publications ......................................................................... v

Table of content............................................................................................ vi 1 Introduction .......................................................................................... 1

1.1 Surface forces............................................................................... 2 1.1.1 Measuring techniques and geometries..................................... 2 1.1.2 DLVO forces ........................................................................... 2 1.1.3 Polymer induced forces ........................................................... 4

1.2 Friction......................................................................................... 5 1.2.1 Measuring techniques for studying friction............................. 5 1.2.2 Biolubrication .......................................................................... 6 1.2.3 Friction models ........................................................................ 6

2 Instruments and methods ...................................................................... 9 2.1 AFM............................................................................................. 9

2.1.1 Principles ................................................................................. 9 2.1.2 Imaging.................................................................................. 11 2.1.3 Force measurements .............................................................. 12 2.1.4 Friction .................................................................................. 13 2.1.5 Calibration ............................................................................. 15

2.1.5.1 Normal spring constant ................................................. 15 2.1.5.1.1 Cleveland method .................................................... 15 2.1.5.1.2 Sader method ........................................................... 16

2.1.5.2 Torsional spring constant.............................................. 16 2.1.5.2.1 Cleveland method .................................................... 16 2.1.5.2.2 Sader method ........................................................... 16 2.1.5.2.3 Direct methods......................................................... 17

2.1.5.3 Torsional detector sensitivity ........................................ 19 2.2 QCM-D ...................................................................................... 20 2.3 Viscosity .................................................................................... 22 2.4 Density ....................................................................................... 23

3 Materials ............................................................................................. 24 3.1 Surfaces...................................................................................... 24

3.1.1 Mica....................................................................................... 24 3.1.2 Silica ...................................................................................... 24 3.1.3 Thiolated gold........................................................................ 25

3.2 Surfactants ................................................................................. 25 3.2.1 SDS........................................................................................ 25

Page 9: DiVA portal13308/FULLTEXT01.pdf · 2008. 3. 7. · The friction properties in aqueous solution have been studied for adsorbed layers of synthetic polyelectrolytes, with varying density

vii

3.2.2 DTAB .................................................................................... 26 3.3 Proteins and polyelectrolytes ..................................................... 26

3.3.1 Bovine serum albumin, BSA ................................................. 26 3.3.2 Cytochrome c, CytC .............................................................. 26 3.3.3 Myoglobin, Mb...................................................................... 27 3.3.4 Mucin..................................................................................... 27 3.3.5 Chitosan................................................................................. 28 3.3.6 PEO45MEMA:METAC polymers.......................................... 28

4 Summary of papers ............................................................................. 30 4.1 Lubrication measured with AFM............................................... 30 4.2 Adsorption and layer properties measured with QCM-D .......... 31

5 Results and discussion ........................................................................ 33 5.1 AFM friction calibration ............................................................ 33 5.2 Friction and lubrication between polyelectrolye layers ............. 35

5.2.1 PEO45MEMA:METAC-X ..................................................... 36 5.2.2 Mucin and chitosan................................................................ 38 5.2.3 Effect of SDS on layers containing mucin and chitosan ....... 39

5.3 Globular protein adsorption studied with QCM-D .................... 42 5.4 Mucin interaction with metal ions ............................................. 44

6 Concluding remarks............................................................................ 51 6.1 Technical and experimental conclusions ................................... 51 6.2 Physical chemical conclusions................................................... 51

7 Future work......................................................................................... 53 8 Acknowledgment................................................................................ 54 9 References .......................................................................................... 55

Page 10: DiVA portal13308/FULLTEXT01.pdf · 2008. 3. 7. · The friction properties in aqueous solution have been studied for adsorbed layers of synthetic polyelectrolytes, with varying density
Page 11: DiVA portal13308/FULLTEXT01.pdf · 2008. 3. 7. · The friction properties in aqueous solution have been studied for adsorbed layers of synthetic polyelectrolytes, with varying density

1

1 Introduction The aim of this thesis work was to investigate the properties of adsorbed layers of polyelectrolytes and proteins with focus on the lubrication capabilities between such layers. The outcome was a better understanding of the friction mechanisms and on the demands on the layer properties that are needed to obtain low friction. Such understanding is of high interest for medical applications, for example as coatings on artificial implant in joints or coatings on contact lenses, artificial tear fluid / saliva. To this end I have focused on friction measurements in some protein / polyelectrolytes systems, on adsorption properties of proteins, and on their conformational changes due to varying the surrounding electrolyte concentration. The major techniques used in this work were AFM (atomic force microscope) for friction measurements together with QCM-D (quartz crystal microbalance with dissipation) for measurement of adsorption and conformational changes. Calibration of the AFM instrument for friction measurements is very important and time consuming, and various calibrations techniques are availably and are used frequently by different users. Therefore a comparison between some of these techniques was done to find the most suitably one for friction measurements in aqueous solutions. The friction properties in aqueous solution have been studied for adsorbed layers of synthetic polyelectrolytes, with varying density of grafted side chains and varying charge density (some of them acting as a polymer bottle-brush), a naturally occurring polyelectrolyte (chitosan) and the glycoprotein mucin. These polyelectrolytes were used to cover a wide range of different types of surfaces and interactions. The rationale for this is the open random coil structure of mucin that contrasts to the compact globular structure of the other proteins used in my investigation. Adsorption properties were investigated for different proteins (bovine serum albumin, cytochrome c, myoglobin, and mucin) whereas the conformational changes of the adsorbed layer due to change in electrolyte type and concentration were monitored for mucin only. In the following part of this chapter I will briefly discuss surface interactions and also the friction that arises when surfaces are sliding against each other. In chapter 2, the instruments used within this work are described with main focus on AFM and how to use it for friction measurements. This is followed by some comments on the material used in the different studies that are reported in the publications appending this thesis (chapter 3). The included papers are summarised in chapter 4 and this is followed by results and discussion in chapter 5. The main conclusions from the papers are

Page 12: DiVA portal13308/FULLTEXT01.pdf · 2008. 3. 7. · The friction properties in aqueous solution have been studied for adsorbed layers of synthetic polyelectrolytes, with varying density

2

summarised in chapter 6. In chapter 7 some perspectives for future research are briefly outlined.

1.1 Surface forces

1.1.1 Measuring techniques and geometries Several different techniques are available for measuring interactions between surfaces. They use different setups and different measuring geometries. I will briefly mention some of these techniques together with the geometry used.

• SFA (Surface Force Apparatus)1, 2 – Crossed cylinders • AFM (Atomic Force Microscope)3, 4 – can be varied (normally

sphere / flat) • MASIF5 – sphere / sphere or sphere / flat • TFPB (Thin Film Pressure Balance)6 – flat / flat • JKR apparatus7 – sphere / flat

To be able to compare results obtained with different geometries one needs to adopt the Derjaguin approximation,8 which relates the force (F) measured for a specific geometry to the interaction free energy per unit area (W) between two flat surfaces separated the distance D:

normRDFDW

π2)()( = (1.1)

Where Rnorm is the normalized radius, which depends on the geometry used. For instance, sphere/flat: Rnorm = Rsphere, crossed cylinders 21RRRnorn = ; sphere/sphere Rnorm = R1R2/(R1+R2). Of the above mentioned techniques, the most commonly used one is AFM followed by SFA.

1.1.2 DLVO forces In liquid media the interaction between two non-polymer bearing surfaces is commonly dominated by van der Waals attraction and electrostatic double layer forces, and these forces are commonly assumed to be additive. This is the theoretical prediction that forms the basis of the so-called Derjaguin-Landau-Verwey-Overbeek (DLVO) theory.9, 10 Because of different distance dependence of the van der Waals and electrostatic interactions, the total force law as described by DLVO theory, can show several minima and maxima. The van der Waals force is due to interactions of electromagnetic nature (London or dispersion interactions) between the different surface materials.

Page 13: DiVA portal13308/FULLTEXT01.pdf · 2008. 3. 7. · The friction properties in aqueous solution have been studied for adsorbed layers of synthetic polyelectrolytes, with varying density

3

To obtain an expression for the van der Waals interaction between particles one may use the method due to Hamaker,11 in which the summation of all pair wise contributions between the two bodies is carried out as an integration. This approach leads to a very simply analytical expression that for sphere/sphere, sphere/flat, and crossed cylinders takes the form:

26)(

DAR

DF normvdW = (1.2)

Where D is the separation distance, Rnorm as above and A is the material dependent Hamaker constant. The van der Waals force is always attractive between identical surfaces of the same material, but can be repulsive between surfaces of dissimilar materials. The more advanced Lifshitz theory8 results in the same expression (Eq 1.2), but the Hamaker constant will have a different value. The electrostatic interaction originates from the fact that most surfaces in contact with a highly polar liquid (e.g. water) acquire a surface charge. Either by dissociation of ions from the surface into the solution or by preferential adsorption of certain ions from the solution. This surface charge is balanced by a layer of oppositely charged ions (counterions) in the solution within some distance from the surface. In dilute solutions, the characteristic distance of the diffuse ionic cloud around the charged surface is the Debye length, κ-1, which is purely a property of the electrolyte solution. The Debye length falls with increasing ionic strength of the solution. In totally pure water at pH 7, κ-1 is 960 nm, and in 1 mM NaCl solution κ-1 is 9.6 nm.12 Since the Debye length is a measure of the thickness of the diffuse surrounding of counterions near a charged surface, it also determines the range of the electrostatic double-layer interaction between two charged surfaces. The electrostatic double-layer interaction is an entropic force that arises upon decreasing the thickness (volume) of the liquid containing the dissolved ions and opposite charges on the surfaces. The ions stay between the surfaces, generating an osmotic repulsion as their concentration is increased in the gap between the surfaces. The electrostatic interaction free energy at large separation (D > κ-1) between two similarly charged surfaces is repulsive with an exponential decay:

DCeDW κ−=)( (1.3) Where D is the distance between the two surfaces and C is a constant that depends on the geometry of the interacting surfaces, their surface charge density, and the solution conditions. C can be determined by solving the Poisson-Boltzmann equation. At small separation one must use numerical solutions of the Poisson-Boltzmann equation to obtain the exact interaction

Page 14: DiVA portal13308/FULLTEXT01.pdf · 2008. 3. 7. · The friction properties in aqueous solution have been studied for adsorbed layers of synthetic polyelectrolytes, with varying density

4

potential, for which there are no simple and accurate expressions. In the limit of small D, it can be shown that the interaction energy depends on whether the surfaces remain at constant potential, ψ0, or at constant surface charge density, σ, or somewhere between these limits. In the “constant charge limit” the total number of counterions in the compressed film does not change as D is decreased, whereas at “constant potential limit”, the concentration of counterions is constant.8

1.1.3 Polymer induced forces The interactions between adsorbed polymer layers in a liquid medium can be attractive or repulsive, and depend on a large number of parameters such as polymer architecture, molecular weight, surface coverage, polymer-surface affinity and solvency.8 Polymer-induced forces arise when two polymer layers start to overlap, whereby increasing the segment density between the surfaces. This results in a repulsive force contribution from the decreased entropy of mixing. The polymers will also lose some conformational freedom, which will decrease the conformational entropy and contribute to the repulsive force. Further, some segment-solvent contacts will be replaced by segment-segment contacts, and this force contribution can be either repulsive or attractive depending on the solvent quality. The range of the steric force reflects the length of the tails, and the outermost part of the force profile between adsorbed homopolymer layers is predicted to have an exponential decay according to the weak overlap approximation.8 Theories of steric interactions are rather complex, but Alexander – de Gennes have, using scaling arguments, arrived at a simple analytical expression for the steric interaction acting between brush layers.13 The force acting between two polymer brush layers in a good solvent has, according to their, theory the following form:14

⎥⎥⎦

⎢⎢⎣

⎡−⎟

⎠⎞

⎜⎝⎛+⎟

⎠⎞

⎜⎝⎛≈ 12

2527

3516)( 4745

3 LD

DL

skTLC

RDF π

(1.4)

where L and s are the thickness of each brush layer and the average lateral distance between grafting sites of the brushes on each surface, respectively. Eq 1.4 is valid for the sphere on flat geometry used in my AFM study. It should be noted that the numerical prefactor, C, is not predicted by the theory but it is expected to be in the order of unity. Bridging forces can occur when a polymer molecule is attracted to, two interfaces at the same time. As two surface approach sufficiently close to each other, a polymer which is adsorbed onto the first surface may be able to adsorb also onto the other surface,15 forming a bridge between them.

Page 15: DiVA portal13308/FULLTEXT01.pdf · 2008. 3. 7. · The friction properties in aqueous solution have been studied for adsorbed layers of synthetic polyelectrolytes, with varying density

5

Since this results in an increased number of possible conformations for the polymer, it has been argued that the attraction is of entropic origin.16

1.2 Friction Friction is a fundamental physical phenomenon occurring wherever two contacting surfaces are in relative motion. Mankind has struggled against, and used, friction since ancient times. The ability to control and manipulate friction forces is extremely important for many applications. One may wish to reduce or enhance friction. Such control can be technologically important for micromechanical devices and computer disk drives, where the early stages of motion and the stopping processes often exhibit unwanted stick–slip or damage. In contrast, chaotic stick–slip may be desirable, for example, in string instruments. The control of friction forces has been traditionally approached by chemical means, usually by supplementing base lubricants with friction modifier additives. But there are also situations when one wants high friction, as in clutches and brakes, or stick–slip, to enrich the sound of a violin and improve the feel or ‘texture’ of processed food as sensed during biting and chewing. To be able to achieve the desirable high or low friction, it is important to be able to study the friction force for different systems. Classically, friction is often discussed in terms of Amontons’ law, where μ is the coefficient of friction.

loadfric FF μ= (1.5) Where Ffric is the friction force and Fload is the applied load.

1.2.1 Measuring techniques for studying friction There exist various techniques for measuring friction on different length scales and with different sizes of contact areas. Testing instruments have more or less specific design for the different systems of interest. For macro-scale measurements, one often measures the friction force when a test body slides over a surface. These types of measurements are widely used in the industry for process control, for example during paper making.17-22 On micro- and nano-scale sizes the dominating instruments for measuring friction are AFM and SFA with friction attachment (T-SFA). The T-SFA instrument measures friction in micro scale contacts using the geometry of two crossed cylinders covered with atomically flat mica sheets. The mica sheets can be chemically modified with different surface treatment techniques (plasma activation followed by silanation, LB-depositions, gold

Page 16: DiVA portal13308/FULLTEXT01.pdf · 2008. 3. 7. · The friction properties in aqueous solution have been studied for adsorbed layers of synthetic polyelectrolytes, with varying density

6

evaporation followed by thiol adsorption) which gives a wide range of different surface chemistry possibilities.23-28 AFM is used to measure friction at micro or nano-scale contacts. Different geometries are availably with this technique; such as sphere / sphere, fibre / fibre, sphere / flat etc. (This technique will be described in detail in the next chapter since AFM has been employed for the friction measurements presented in this thesis).

1.2.2 Biolubrication In biological systems water-based lubricants function well;29 where the charged hydrophilic surfaces provide an electrostatic “double-layer” repulsion, in addition to the “steric” repulsion of any protruding polymer or polyelectrolytes and the hydration layer of tightly bound water molecules,28 all leading to low friction. Recent experiments30 indicate that brushes of charged polymers (polyelectrolytes) attached to sliding surfaces across an aqueous medium result in superior lubrication (friction coefficient μ < 0.001) at low sliding velocities and pressures up to a few atmospheres. This suggests that some biolubricating systems, such as in the eyes, may be mediated by brush like polyelectrolyte layers.23 However, other experiments18, 25 show that brushes may not function well at high pressures where wear (damage) and/or higher friction coefficients are usually measured, indicating that high-pressure, high-impact systems, such as in the hip or knee joints, may be mediated by more complex mechanisms. Currently, no model system exhibits both low friction and good wear resistance at high pressure (and over a large range of velocities); however, exploring polyelectrolytes seem to be a way to make progress. Recent experiments performed on chitosan attached to surfaces sliding past each other across an aqueous medium resulted in superior lubrication at low sliding velocities and at pressures of up to 6 atm (μ = 0.003).26 However, other experiments involving poly(l-lysine)-g-poly(ethylene glycol),18 cross-linked hyaluronic acid (HA)25 and lubricin layers24 attached chemically to sliding surfaces in aqueous media show that such systems do not generally function well at higher pressures (>3-10 atm), where damage18 and/or high friction24, 25 are measured, although cross-linked HA and lubricin appear to be good wear protectors (but with high friction coefficients).

1.2.3 Friction models The friction force needed to slide a body over a surface is related to the energy dissipation per unit time and unit surface area, W:31

vWA

F efff = (1.6)

Page 17: DiVA portal13308/FULLTEXT01.pdf · 2008. 3. 7. · The friction properties in aqueous solution have been studied for adsorbed layers of synthetic polyelectrolytes, with varying density

7

where Aeff is the effective contact area and v the sliding velocity. It has been suggested32 that the main energy dissipation mechanism between sliding polymer-coated surfaces is the dragging of polymer chains through the interpenetration zone. Other dissipation mechanisms are the flow of solvent through the polymer layer, and the formation and breakage of attractive interactions between the opposing layers. The very low coefficient of friction observed between brush polymer layers compared to layers of adsorbed linear homopolymers has been assigned to lower chain interpenetration (see illustration in Figure 1.1) in the former systems.32 Since the size of the chain interpenetration zone for brushes growths only little with compression,32 the interfacial region is able to maintain a high fluidity (low effective viscosity in the interface region due to the low amount of interpenetrating chains) even under significant load. Lubrication by thin films is a complex phenomena which can conceptually be illustrated by the so-called Stribeck curve,31, 33 see Figure 1.2. In this curve the friction coefficient, μ, is plotted versus a film dependent parameter (ηωP-1; η is the viscosity of the film separating the sliding surfaces, ω is the shear rate and P is the pressure). This treatment suggest that film lubrication can be divided into three different regions.31, 33

• Hydrodynamic lubrication (thick films), where the surfaces are separated by a relatively thick lubrication film whose properties dominate the friction force acting between the surfaces (dependent on viscosity) and leads to an increase in friction with increasing film parameter (ηωP-1).

• Boundary layer friction, where the friction is dominated by the surface properties (e.g. thin adsorbed films or surface asperities).

• Mixed lubrication, a transition between hydrodynamic lubrication and boundary layer friction where both mechanisms are contributing to the friction.

Figure 1.1: Illustration of interpenetration (shaded area) when polymer layers are approaching each other. The case to the right has a smaller interpenetration zone than that to the left. Thus lower friction is expected in the former case.

Page 18: DiVA portal13308/FULLTEXT01.pdf · 2008. 3. 7. · The friction properties in aqueous solution have been studied for adsorbed layers of synthetic polyelectrolytes, with varying density

8

����������� ��������

��

��� �

���

���

���

����

�� �

���

���� Figure 1.2: The Stribeck curve demonstrating the variation of the friction coefficient, μ, as a function of the film parameter ηω/P.

When friction is measured in the boundary layer regime, as was the case in my studies, or in the mixed regime the classical Amontons’ law does not fully work, and one needs to modify this equation to the following form:

loadloadfric FCFF μ+=)( (1.7) where C is a constant able to shift the friction force at zero load. The constant C can be positive or negative, depending on the system studied. For example in Paper II this constant is negative since some of the applied load is carried by electrostatic repulsion between the two surfaces. On the other hand, in Paper III this constant is positive when we have adhesive interaction between the surfaces, and it is expected that the value of C will increase with the strength of the attraction. Different theoretical models attempt to explain the origin of the friction force on a molecular level. The Cobblestone model31 is one such, and here C = ScA; where Sc is the critical shear stress and A is the contact area. In this model Sc can in simple terms be explained as the force needed to push the molecules over each other in analogy to pushing a cart over a road of cobblestones. The cartwheel (representing the molecules of the upper surface) must be made to roll over the cobblestones (molecules of the lower surface) before the cart can be moved.

Page 19: DiVA portal13308/FULLTEXT01.pdf · 2008. 3. 7. · The friction properties in aqueous solution have been studied for adsorbed layers of synthetic polyelectrolytes, with varying density

9

2 Instruments and methods

2.1 AFM The Atomic Force Microscope (AFM) was invented by Binning et al.,3 and it can be used for a wide range of different measurements. It is mainly used for its great ability to image different types of samples at different resolution down to atomic resolution.34-36 But there are a lot more that can be done with an AFM. For example one can measure interactions forces between different objects37, 38 or within proteins.39 It is also possible to use the AFM to manipulate the substrate via etching,40 scratching,41 moving42 different objects or locally modify self organized layers.43, 44

2.1.1 Principles An AFM can look different dependent on the producer of the instrument. But common features are:

• Laser together with a detector • Scanners to move the sample (or the probe) in three perpendicular

directions • Cantilever with a probe (normally a very sharp tip)

����������

���

��� ����������

��

�����

���������

Figure 2.1: Schematics of the different important parts of the AFM: laser, cantilever, probe, detector (photodiode) and scanner. The top right illustration shows the imaging possibility where a tip is used as a probe to sense the morphological structure of a sample surface. The bottom right schematic illustrates the option to use a closed environment for liquid measurement or to use controlled atmosphere during the measurements.

Page 20: DiVA portal13308/FULLTEXT01.pdf · 2008. 3. 7. · The friction properties in aqueous solution have been studied for adsorbed layers of synthetic polyelectrolytes, with varying density

10

Figure 2.2: In a split photodiode one can easily follow the movement of the laser spot in both horizontal and vertical direction. The movement of the laser spot on the photo detector is proportional to the position change (bending and twisting) of the cantilever. The output from the detector is normally in voltage; Vnorm respectively Vlat.

The main principle behind AFM is to reflect a laser beam on the free end of a cantilever. The reflected laser is transferred into a detector system. The detector is sensitive to position changes of the cantilever as it is bending or twisting. The most commonly used detector is a split (four sector) photodiode. The detector system is very sensitive and can for example be used to detect thermal motion of the cantilever. The scanner is used to move the cantilever position relative to the sample. Either the sample is moved or the cantilever and hence the whole laser and detector system is moved. In the following text I will refer to the scanner as it moves the sample. The position of the scanner can be determined with high resolution in all directions (lateral, x/y, respectively in height z). It is normally made of piezoelectric materials. Piezoelectric materials have the drawback that they have hysteresis that needs to be taken into consideration by a suitable calibration procedure. To overcome the hysteresis problem some systems use closed-loops, e.g. positioning sensors with feedback to the driving of the scanner enabling more precise positioning. The cantilever is the sensing part of the AFM, and the cantilever probe is normally mounted as a sharp tip at the free end. The shape of the cantilever can be varied; the most common geometries are rectangular shape and triangular double beam. The cantilever is normally based on silicon based materials; SiO2, SiN3 etc. The cantilever sensitivity can be varied by choosing between different commercially available spring constants in the range of 0.002 N/m to 400 N/m. An important improvement for the surface chemistry society was the invention of the colloidal probe technique by Ducker et al.45, 46 They glued a

Page 21: DiVA portal13308/FULLTEXT01.pdf · 2008. 3. 7. · The friction properties in aqueous solution have been studied for adsorbed layers of synthetic polyelectrolytes, with varying density

11

spherical particle to the cantilever to enable measurement of interactions between a surface and larger particles. The use of the name colloidal is little misleading but there should be a possibility to use small objects. The limitation is only in the skills required to attach/glue the object of interest to the cantilever and or at the substrate, and also to get the objects of interest in contact during the measurements. In general one can attach anything to the cantilever, it is just to use your imagination. Recently we reported on the success of using fibres as probe to measure interaction (force and friction) between two crossed polyester fibres.37 AFM can be used in many different modes, depending of the information one wants to obtain from the measurement. The common techniques of imaging can be of varying kinds. In imaging the probe is rastered over or in contact with the surface using feedback on the z-axis (height), keeping one detector signal constant by raising or lowering the sample. An alternative method compared to imaging is normal force measurement, in which the sample is ramped up and down bringing the probe in contact with the surface to detect forces acting between the sample and the probe. These techniques will be addressed in more detail in the following subsections.

2.1.2 Imaging The two major imaging methods will be briefly described in this section. First I will describe an imaging method called contact mode followed by a description of the tapping mode. In contact mode the probe and sample are in contact with each other during imaging. The operator selects the level of the force that is applied between the probe and sample, via a setpoint (normally in the unit volt) for the vertical detector response. The level of this detector response is then aimed to be constant (at the setpoint value) during the “scanning” when the sample is moved back and forward under the cantilever. If a peak on the sample pass by the probe then the cantilever will react by pushing the cantilever upwards (this should make the detector signal to increase), but the feedback in the system will lower the sample so that the detector signal remains constant. By recording the sample height for each position on the sample one creates a three dimensional image of the surface topography, and by recording the detector response one creates an image of the error within the height image. Tapping mode® (trademark by Veeco Instruments) or AC mode is a very common technique for imaging in air or liquid. The cantilever is oscillated close to the fundamental resonant frequency while the sample is scanned under the oscillating cantilever. The system tries to keep the amplitude of the oscillation constant by using a feedback on the height that raises or

Page 22: DiVA portal13308/FULLTEXT01.pdf · 2008. 3. 7. · The friction properties in aqueous solution have been studied for adsorbed layers of synthetic polyelectrolytes, with varying density

12

lowers the sample if there are any irregularities on the sample, i.e. surface topography.

2.1.3 Force measurements Forces between the sample and the probe can be measured by letting the probe approach the sample in a controlled way. This is done by ramping the sample vertically, from below up towards the probe and down again.

�������� ������� ���������

�����

�������������������������

���

����

������������������������������������������������������������������������������

��������

�������� ��������������������������������������������������������������������������������������������������������������������

�����������������

�� ���������

���������������

����

�����

�������������� �

����

�������������������� !�������� ���"������

����

��������

�!������ !���

����� ���!�������

������

���� � ��

��

Figure 2.3: The insert in the top right corner is a schematic illustration of the force measurement, with the important quantities marked with the respective symbol used in the algorithm to transform primary data into force vs separation data.

Page 23: DiVA portal13308/FULLTEXT01.pdf · 2008. 3. 7. · The friction properties in aqueous solution have been studied for adsorbed layers of synthetic polyelectrolytes, with varying density

13

During the approach the probe will at some point come in contact with the sample. If the sample is continuing to move upwards, then the probe and the sample will move together. This part of the force curve is called the constant compliance region; the slope of this deflection versus displacement curve is referred to as the deflection sensitivity, α, and this quantity is needed to convert the deflection signal into force. The algorithm used to transform the deflection versus displacement data into force versus apparent separation is presented in Figure 2.3. When measuring forces between soft adsorbed layers, it is best to use the deflection sensitivity measured between the hard substrates before adsorbing the layers. The reason for this is that once the layers have formed it is difficult to reach “hard wall” contact, and thus difficult to obtain the true deflection sensitivity. Also, when pressing extra hard with a cantilever one will reach the region in the detector system that is non linear, and hence the deflection sensitivity will be inaccurate. The approach of using the deflection sensitivity determined between the substrate surfaces prior to adsorbing the soft layers has been utilised in Papers II, III, and VI. A good short review of the topic of force measurement is presented in ref 47, a more comprehensive review are presented by Butt et al.48 The measurement of force curves are straightforward, the tricky part is the interpretations of what the force curves actually tell us about the investigated systems.

2.1.4 Friction Friction is measured by scanning the sample in contact mode perpendicular to the cantilever.49 If there is any friction due to the contact, then the cantilever will twist. This is illustrated in Figure 2.4.

Figure 2.4: Schematic illustration of the cantilever as the surface is approaching (left) and makes the cantilever to bend when in contact (middle) when the surface is moving in lateral direction the friction force makes the cantilever to twist (right).

Page 24: DiVA portal13308/FULLTEXT01.pdf · 2008. 3. 7. · The friction properties in aqueous solution have been studied for adsorbed layers of synthetic polyelectrolytes, with varying density

14

� � �� �� �� ����

� �������

������ ������

����

�����

����

�����

�����

����

�����

����

�����

�����

� �������

Figure 2.5: Friction vs load curve from one friction ramping measurement. The three dimensional plots represents the raw data for the selected applied loads (red circles).

The friction force is determined by capturing “contact mode images” using a scanning angle perpendicular to the cantilever at different applied loads (deflection set points, VNset). The imaging is done with the slow scan axis disabled. Each “image” consists of 10 lines that record the friction (i.e. the horizontal twisting) when moving the surfaces in both scanning directions. The capturing of image files and the variation of the applied load can be handled by the instrument software using a built-in scripting tool. I adopted the following procedure for a typical friction experiment. Firstly, five force displacement curves were captured. Secondly, a friction ramping experiment (illustrated in Figure 2.5) was carried out. That was executed by firstly increasing the deflection set point stepwise from no applied load up to a desired maximum applied load, and then decreasing it again stepwise until the probe was detached from the surface. The friction was determined at each load, both on loading and unloading. After this friction ramping, five additional force displacement curves were captured. This was done to check the stability of the detector signal at zero load, and to investigate the integrity of the sample after the friction measurements. The resulting friction images were analyzed to extract the friction force and the coefficient of friction, μ, by using the following algorithm: calculate the difference between the friction signals from the two different scanning directions for each scanning lines, remove the first and last x-number of points (x≈20) in each end of the scanning line corresponding to the turning of the probe on the surface. The difference in friction detector signal,

Page 25: DiVA portal13308/FULLTEXT01.pdf · 2008. 3. 7. · The friction properties in aqueous solution have been studied for adsorbed layers of synthetic polyelectrolytes, with varying density

15

ΔVfric (V), is then averaged over the remaining points in each image. The friction force, Ffric, is calculated using the following equation:

eff

fricfric h

kVF 1

2 δφΔ

= (2.1)

Where kφ is the torsional spring constant, δ is the torsional detector sensitivity and heff is the effective height of the probe (i.e. the diameter of the colloidal particle plus half the thickness of the cantilever). The applied load, Fload (N), is calculated from the deflection set point (α is normal deflection sensitivity and kv is normal spring constant):

( ) vloadNsetload kVVF α0−= . (2.2) The effective coefficient of friction, μeff, is calculated as the ratio between Ffric and Fload.

2.1.5 Calibration For all AFM measurements a calibrated scanner is crucial. The scanner is calibrated by imaging well defined grids both in lateral and horizontal directions. For force and friction measurements the spring constants of the cantilever is of high importance since they are used to transfer the cantilever bending, respectively cantilever twisting, to forces. A lot of different methods have been proposed for calibrating the normal respectively torsional spring constants. One also needs to have some conversion factors to convert the detector response into cantilever motions, called detector sensitivity (both for normal and torsional, friction, mode).

2.1.5.1 Normal spring constant

2.1.5.1.1 Cleveland method The Cleveland method is based on measuring the change in resonant frequency due to an added mass.50 The resonant frequency decreases when a mass is attached to the cantilever. The relation between the resonant frequency, fv, and the mass is:

ev

vs m

fk

M −= 2)2( π (2.3)

By attaching several different masses (Ms), e.g. different spheres with known density, and measuring the corresponding normal resonant frequency enables the normal spring constant kv to be determined from the slope of a linear plot of Ms vs (2πfv)2. (Where me is the effective mass with no load on the cantilever.)

Page 26: DiVA portal13308/FULLTEXT01.pdf · 2008. 3. 7. · The friction properties in aqueous solution have been studied for adsorbed layers of synthetic polyelectrolytes, with varying density

16

2.1.5.1.2 Sader method The Sader method is using the principle that a viscous fluid damps the thermal motion of a body, hydrodynamic damping. The cantilever is allowed to vibrate freely due to thermal motion in a fluid such as air. The frequency (fv), and Q-value (Qv) of the fundamental normal resonance is used together with the dimensions (length, l, and width, b) of the cantilever, the density (ρ) and viscosity (η) of the fluid. These parameters are then used to calculate the normal spring constant (kv)51

)(Re)2(1906.0 22v

vivvv flQbk Γ= πρ (2.4)

ηπρ

42

Re2

vv

fb= (2.5)

Γiv(Rev) is the imaginary component of the hydrodynamic function for

normal vibrations as described in detail in ref 52.

2.1.5.2 Torsional spring constant

2.1.5.2.1 Cleveland method The change in resonant frequency due to an added mass is determined similar to the normal spring constant calibration by Cleveland.53

et

s Jf

kr−= 2

5

)2(1528

ππρ φ (2.6)

By attaching several different spheres (with radius r and density ρs) and measuring the corresponding torsional resonant frequency, ft, enables the torsional spring constant, kφ, to be determined from the slope of a linear plot of 28πρsr5/15 vs (2πft)2. (Where Je is the effective mass moment of inertia.)

2.1.5.2.2 Sader method The Sader method for determining torsional spring constants53 is similar to the normal Sader method described above, and based on the hydrodynamic damping of the resonant frequency. The frequency (ft), and Q-value (Qt) of the fundamental torsional resonance is used together with the dimensions (length, l, and width, b) of the cantilever, the density (ρ) and viscosity (η) of the fluid. These parameters are then used to calculate the torsional spring constant (kφ) as:

)(Re2)2(1592.0 4t

titt flQbk Γ= πρφ (2.7)

Page 27: DiVA portal13308/FULLTEXT01.pdf · 2008. 3. 7. · The friction properties in aqueous solution have been studied for adsorbed layers of synthetic polyelectrolytes, with varying density

17

��� ��� ��� ��� ��� ��� ������ ��!���

"#$%��&�������&��

��� ��� �� �� ��� ��' � ��� ��� ��� ���

(������

���

���

���

���

���(���

���

�������

����

� ��� �

��)��

Figure 2:6: Left; torsional frequency response due to added mass. The black spectra corresponds to the unloaded cantilever, while blue and red spectra are obtained with particles with increasing size attached to the cantilever. This makes the resonant frequency to decrease. The right graph illustrates the calibration plot where the linear slope gives the torsional spring constant.

��� ��� �� �� �� �� ��� ��� �'���� ��!���

"#$%��&��

�����&��

��#$%&'(

���%)*&+,

Figure 2:7: Typical thermal torsional frequency spectra measured with a cantilever. A single harmonic function with added white noise is fitted to the data giving the Q-factor and also f0 .

ηπρ

42

Re2

tt

fb= (2.8)

Γit(Ret) is the imaginary component of the hydrodynamic function for

torsional vibrations described in ref 54.

2.1.5.2.3 Direct methods The pivot55 and lever56 methods are both based on the one and the same principle of using the piezoactuator to move a known distance at some distance from the cantilever axis thus providing a turning moment, though

Page 28: DiVA portal13308/FULLTEXT01.pdf · 2008. 3. 7. · The friction properties in aqueous solution have been studied for adsorbed layers of synthetic polyelectrolytes, with varying density

18

the way the resulting data are treated differs slightly between the two methods. They both provide a direct static measurement of the spring constant, measured under similar conditions as when the cantilever is used to measure friction. The deflection sensitivity α0 for a load applied at the centre of the cantilever is compared to the deflection sensitivity αL and the lateral or torsional sensitivity βL when the load is applied at a distance L from the central axis. If δ, the detector lateral sensitivity (V/rad), is known then a torsional constant can be obtained. Alternatively as in the lever method,56 a torsional calibration factor γ (Nm/V) can be used to convert volts to torque directly. (Note that in this case the measurement is geometry and instrument dependent.) The various sensitivities and calibration factors are related by the following equation:55, 56

γδ

αβα

φ

LLkk Lv

L ==0

(2.9)

In the pivot method55, a linearization plot of (αL/(kvβLα0)) vs L gives kφ /δ as the gradient (i.e. 1/γ ). Strictly, δ can be determined in a similar way55 from:

δααβ L

LL =− )/1( 0 (2.10)

however the errors are usually too large to permit this to be done reliably and an independent method is preferred.55, 57-59 In the pivot method a contact mode cantilever tip is glued upside down on a planar substrate and mounted on the piezo scanner. This is henceforth referred to as a pivot. The pivot was then used to twist the tipless cantilever to be calibrated. Force curves were captured (deflection, friction and z-sensor) at different points along the width of the cantilever. This was done automatically with a built-in script handling both the positioning and force curve measurements (Autoramp). The positioning in the script used the Y-offset with an aimed step size of 2 μm. The resulting force curves are then analyzed to calculate α0, αL, βL and thus obtain values for γ and kφ. A plot of 1/β versus L should be linear through the origin, hence the zero position for L was defined such that the line of best fit passed through the origin. This corrected L was also employed for any calculations of γ and kφ. The torsional spring constant was calculated with Eq 2.9 from the slope of the linearization plot with δ given from the external detector calibration in air. While performing force curve measurement on the pivot with the script, the movement of the pivot was imaged with the optical viewing system mounted over the AFM. The resulting images were analyzed to account for nonlinearity of the Y movement and obtain the real position of the pivot in relation to the cantilever (L). The movement was highly nonlinear, but it was extremely well described by a 2nd order polynomial.

Page 29: DiVA portal13308/FULLTEXT01.pdf · 2008. 3. 7. · The friction properties in aqueous solution have been studied for adsorbed layers of synthetic polyelectrolytes, with varying density

19

� ��

��

�����

���

��

-����

�-.�/

��-�

(���0.1�

�� ���� �� ���� � ��� � ��� �

(�����

����

��

����

���

���(���

Figure 2:8: Right schematic illustration of the pivot method, left typical data from a pivot experiment.

��� �

��

Figure 2:9: Schematic illustration of lever method.

In the lever method a single value of L is used corresponding to the length of a glass fibre lever glued to the cantilever and the torsional spring constant is obtained using the following equation (obtained from Eq 2.9 and 2.10).

)1/( 0

2

−=

ααφL

v Lkk (2.11)

When using γ (calculated from a rearranged form of Eq 2.9) instead of the torsional spring constant, the friction force is calculated by: Ff=γΔVlat/2heff.

2.1.5.3 Torsional detector sensitivity Torsional spring constant is only half the issue since it is actually the lateral photodetector signal which is measured and this needs to be calibrated in terms of angle. There is no obvious torsional analogy to the so called sensitivity, or constant compliance measurement against a hard substrate for the normal calibration60 that can be performed in situ during the experiment. Thus, it is crucial to calibrate the torsional detector sensitivity for the medium and geometry to be employed in the experiment. Different methods exist,55, 57-59, 61, 62 and the one used within this work is a modified version of that described in ref 55. First, a cantilever was mounted

Page 30: DiVA portal13308/FULLTEXT01.pdf · 2008. 3. 7. · The friction properties in aqueous solution have been studied for adsorbed layers of synthetic polyelectrolytes, with varying density

20

in the AFM head and the laser and detector were aligned, the cantilever was removed and a mirror was mounted on the scanner to reflect the laser onto the detector. A three-leg scanner was used to tilt the AFM head. Small changes in the torsional tilt angle were achieved with the built-in step motor, the actual tilt was determined optically with image analysis and geometrical calculations to obtain the actual tilt of the AFM head, see Paper I.

2.2 QCM-D QCM stands for Quartz Crystal Microbalance and the name indicates what the technique is about. In this technique one uses a piezoelectric quartz crystal to sense small changes in mass. This is done by oscillating the quartz crystal at different odd-multiples of the resonant frequency. If there is any change in mass, for example due to adsorption, the resonant frequency of the quartz crystal will change. Hence, changes of the resonant frequency reflect the changes in mass that are sensed by the quarts crystal. Normally a QCM instrument uses thin circular quarts crystals (AT-cut) with electrodes on both top and bottom side. An alternating current voltage is applied through the electrodes making the crystal oscillate; the frequency is adjusted to find the resonant frequency. The resonant frequencies of different odd multiples are then monitored with respect to time by the instrument. To enable measurement of adsorption on different materials, one surface of the circular disc is covered by a thin layer of the material of interest. This is normally done by evaporation and the crystals are bought ready made with the surface of interest. QCM measurements can be preformed in different media; liquid, air, gases and vacuum. For instance, protein adsorption at the water / sensor interface can be followed by the frequency changes, but adsorbed protein layers also have some degree of structural flexibility or viscoelasticity, that also affects the frequency. Viscoelasticity can be visualised by measuring the energy loss, or dissipation, of the shear movement of the crystal in water. One way of measuring dissipation is to drive the crystal with A/C current at the resonant frequency followed by disconnection and analysis of the resulting damped sinusoidal curve.63-65 This invention of pulse assisted discrimination of frequency and dissipation makes QCM-D analysis of adsorbed protein layers very simple and gives unique information about the viscoelastic behaviour of the adsorbed protein layers and surrounding water. Very small structural and orientational changes of an adsorbed protein layer, including chemical cross-linking, can be monitored with high accuracy. The amplitude of the oscillating crystal is in the range of a few nm when measuring in water.66, 67 Additionally, the oscillation of the crystal generates an exponentially decaying shear wave into the liquid.68 The decay length of

Page 31: DiVA portal13308/FULLTEXT01.pdf · 2008. 3. 7. · The friction properties in aqueous solution have been studied for adsorbed layers of synthetic polyelectrolytes, with varying density

21

this shear wave is dependent on the overtone number, and on the density and viscosity of the solution. In water the decay length is approximately 240nm, 140nm and 110nm, respectively for the fundamental frequency, for the third overtone, and for the fifth overtone (calculated with Eq 6 in ref 69). The primary data from QCM-D measurements are the change in resonant frequency of the oscillator, ∆f, and the change in dissipation value, ∆D, where the dissipation is defined as: D = Edis/(2πEst), with Edis being the energy dissipated during one period of oscillation and Est the energy stored. When the electrolyte type or the electrolyte concentration is changed, it affects the bulk density, ρ, and viscosity, η, that induce a change in frequency and dissipation that is not due to the adsorbed layer. The frequency shift can, according to Kanazawa and Gordon,68, 70, 71 be calculated as:

2/1

2/30

2/1

002/30 ⎟

⎟⎠

⎞⎜⎜⎝

⎛−⎟

⎟⎠

⎞⎜⎜⎝

⎛=Δ

qq

ss

qq

ssn fnfnf

ρπμηρ

ρπμηρ

(2.12)

where n denotes the overtone number, f0 is the fundamental frequency of the quartz crystal in air which is 4.95 MHz, ρq is the specific density of quartz (2648 kg/m3), μq is the shear wave velocity of quartz (2.95×1010 kg/ms2) and the subscripts s0 and s denote reference electrolyte (buffer or water) and electrolyte, respectively. The change in dissipation due to changes in the bulk has been suggested by Rodahl and Kasemo69 to be given by:

n

ss

qqn

ss

qqn ft

nft

nDπρη

ρπρη

ρ 21

21 00−=Δ (2.13)

where tq is the thickness of the quartz crystal (3.3×10-4 m). The frequency and the dissipation are also affected by adsorption to the crystal surface, and by conformational changes of the adsorbed layer. The Sauerbrey equation72 is commonly used to obtain the sensed mass, Δm, from the change in resonant frequency, Δf:

0nfft

m nqq Δ⋅⋅−=Δ

ρ (2.14)

Where Δfn is the change in resonant frequency for overtone n. The Sauerbrey equation has been suggested to be a good approximation when changes in the dissipation value, ΔD, is less than 10-6 per 5 Hz of Δf.73, 74 Note that the sensed mass includes the mass of the adsorbing species as well as changes in the mass of the solvent that oscillates with the crystal. It is also affected by the viscoelastic properties of the layer.

Page 32: DiVA portal13308/FULLTEXT01.pdf · 2008. 3. 7. · The friction properties in aqueous solution have been studied for adsorbed layers of synthetic polyelectrolytes, with varying density

22

The Johannsmann model75 is more accurate and informative since it uses QCM data from several overtones to create a regression line that derives the true sensed mass. Note that the true sensed mass still contains contributions from both the adsorbing species and solvent oscillating with the crystal. It is an advantage over the Sauerbrey method in that it accounts for the viscoelasticity of the adsorbed layer. Johannsmann derived an equation from acoustic impedance and shear modulus of quartz resonators that related the value of Δf to the various properties of the resonator and the adsorbed layer:

( ) ⎟⎟⎠

⎞⎜⎜⎝

⎛+−≈

38ˆ1ˆ

3233

0

dffJdfff f

qq

ρπρ

μρπδ (2.15)

where fδ is the shift in the complex frequency, d is the thickness of the film, f is the resonant frequency of the crystal in contact with solution and

( )fJ is the complex shear compliance. Eq. 2.15 can be rewritten more conveniently by using the equivalent mass, *m , defined by:

ff

fm

qq ˆ2

*ˆ0

δμρ−= (2.16)

and one obtains:

( ) ⎟⎟⎠

⎞⎜⎜⎝

⎛+=

34ˆ1*ˆ

2220 df

fJmm f πρ (2.17)

The true sensed mass, m0, is calculated under the assumption that ( )fJ is independent of the frequency in the accessible frequency range. A plot of the equivalent mass against the square of the resonant frequency for different overtones gives the true sensed mass as the y-intercept.75 If the measured data show good linearity it confirms that ( )fJ is independent of the frequency. It is important to remember that the resonant frequency is related to the total oscillating mass, hence, monitoring resonant frequencies leads to a detection of “apparent” or “sensed” adsorbed mass, meaning that also the mass of water trapped in the adsorption layer is included. This constitutes an important difference from other techniques such as ellipsometry or reflectometry where the adsorbed amount of the adsorbate alone can be detected. This difference is important to keep in mind when comparing the results obtained with different techniques.

2.3 Viscosity Viscosity is a measure of the resistance of a fluid to deform under shear stress. It is commonly perceived as "thickness", or resistance to flow.

Page 33: DiVA portal13308/FULLTEXT01.pdf · 2008. 3. 7. · The friction properties in aqueous solution have been studied for adsorbed layers of synthetic polyelectrolytes, with varying density

23

Viscosity describes a fluid's internal resistance to flow and may be thought of as a measure of fluid friction. Thus, water is "thin", having a low viscosity (∼1 mPa s), while motor oil is "thick" having a high viscosity (∼250 mPa s). A viscosimeter is an instrument used to measure the viscosity and flow parameters of a fluid. The classical method for determining viscosity is to measure the time it takes for a fluid to flow through a capillary tube, this method has been refined by Ubbelohde and others.76, 77 The glass tube viscometer is still the most common method for the standard determination of the kinematic viscosity of water solutions, and this method was utilized in Paper IV and V.

2.4 Density Anton Paar DSA 5000, is a densitometer combined with measurement of sound velocity. The density is measured with the oscillating U tube principle, i.e. the shift in resonant frequency is related to the mass of the sample. The sample cell is a fixed volume U shaped glass capillary that is temperature regulated and for which the resonant frequency can be measured accurately. The resonant frequency of the U cell is measured by the instrument and report the density of the medium injected in the measuring cell, and this method was utilized in Paper IV and V.

Page 34: DiVA portal13308/FULLTEXT01.pdf · 2008. 3. 7. · The friction properties in aqueous solution have been studied for adsorbed layers of synthetic polyelectrolytes, with varying density

24

3 Materials Many different materials have been used for the work presented in this thesis. I want to give the reader some information of the material used. In all experiment water has been used as the solvent normally with addition of some salt. I do not want to go deep into the properties of water and solutes in this description but I want to mention that clean water and clean glass ware are crucial when one is interested in surface phenomena.

3.1 Surfaces

3.1.1 Mica Naturally occurring muscovite mica is easily cleaved into atomically smooth, macroscopic areas and has hence become an important substrate in many fundamental studies of surface interactions, as in experiments with the surface force apparatus (SFA). In addition, it is widely used as a model substrate for the adsorption of macromolecules, e.g. for atomic force microscopy studies. Mica is a layered dioctahedral aluminosilicate, and, after cleavage, the (001) basal plane is composed of SiO4 tetrahedra arranged in ditrigonal rings. The potassium ions neutralizing the excess lattice charge arising from the substitution of aluminum for silicon in the mica lattice are situated between a pair of these trigonal rings. After cleavage, one-half of these become randomly associated with each of the two opposing surfaces. In aqueous solution, these ions can be dissolved and exchanged for other cations. There are 2.1×1018 m-2 negative sites on the mica surface, but most of these are neutralized by adsorbed cations. The mica surface acquires a high negative surface charge when in contact with water. The apparent surface potential for mica in pure water obtained from fitting DLVO-theory to measured forces is found to be -130 to -150 mV.78 It decreases slightly as the electrolyte concentration increases, being between -80 and -100 mV in 1 mM 1:1 electrolyte.78

3.1.2 Silica Silica colloids (particles) or silica glass is an amorphous material (except crystalline quartz), that often have been used as “model” system when surface chemistry phenomena in a wide variety of areas have been studied. It has been suggested in the literature that a gel-like layer forms when silica is placed in contact with aqueous solutions, composed of polyelectrolyte-like “hairs” protruding out into the solution.79, 80 This will affect the interaction at short separation (<4 nm) between such layers, and at small separation an addition non-DLVO repulsion is most often observed.79, 80

Page 35: DiVA portal13308/FULLTEXT01.pdf · 2008. 3. 7. · The friction properties in aqueous solution have been studied for adsorbed layers of synthetic polyelectrolytes, with varying density

25

Silica is negatively charged due to dissociation of surface silanol groups. The surface charge of silica varies with pH, electrolyte concentration and cleaning process.79-82

3.1.3 Thiolated gold Long-chain alkanethiols adsorb spontaneously from solution onto clean gold surfaces and form densely packed monolayer films with high order.83, 84 This allows us to prepare oriented organic monolayer films containing many functional groups of chemical and biological interest and to exercise a high degree of control over the structure of these films. In particular, by using adsorbates having the structure HS(CH2)-X, one can prepare monolayers presenting an ordered array of the group X at the monolayer-air/or liquid interface. If the polymethylene chain of HS(CH2)-X is at least ten carbon atoms long, the surface properties depend primarily on the tail group X, are independent of the chain length, and are influenced only indirectly by the sulphur-gold interface.85 This thiol modification technique was used on gold coated QCM-D crystals in Paper IV, V and VI, and also on freshly made gold particles in Paper VI. The gold particles were produced from gold wires by using an electronic discharge86 under a flow of filtered nitrogen gas.

3.2 Surfactants A surfactant is an amphiphile compound (normally organic) having separated hydrophilic and hydrophobic parts. Therefore they are soluble to some degree in both organic media and water. At a certain concentration the surfactant will form thermodynamically stable aggregates, i.e. micelles, and this concentration is often referred to as cmc (critical micelle concentration). The driving force for micelle formation is the hydrophobic interaction that results from the stronger affinity between water molecules compared to between water and hydrocarbon.

3.2.1 SDS Sodium dodecyl sulphate (SDS), C12H25SO4Na, is an anionic surfactant that is used in household products such as toothpastes, shampoos, shaving foams and bubble baths for its thickening effect and its ability to create a foam. The molecule has a tail of 12 carbon atoms, attached to a sulphate group, giving the molecule the amphiphilic properties required of a surfactant.12 Since SDS is an ionic surfactant, the cmc will depend on the electrolyte concentration,87, 88 i.e. cmc will decrease with increasing salt concentration (see Table 3.1). This is due to screening of the electrostatic repulsion between the head groups.

Page 36: DiVA portal13308/FULLTEXT01.pdf · 2008. 3. 7. · The friction properties in aqueous solution have been studied for adsorbed layers of synthetic polyelectrolytes, with varying density

26

Table 3.1: CMC for SDS at different NaCl concentrations

NaCl (mM)

CMC for SDSa (mM)

1 8.2 30 3.3

100 1.5 a Values taken from ref 89.

3.2.2 DTAB Dodecyltrimethylammonium bromide (DTAB) is a cationic surfactant, belonging to the same family as the more known CTAB (cetyltrimethyl-ammonium bromide). This type of cationic surfactant finds wide applications as disinfectants or antiseptic agents. They are also used as antistatic agents (hair conditioners), textile softeners, corrosion inhibitors, foam depressants, biocides, emulsifiers and in deinking applications. Similar as for SDS, the cmc for DTAB is varying with salt concentration, e.g. in pure water cmc for DTAB is 14.4 mM, in 10 mM KCl cmc is 10.8 mM, and in 100 mM KCl cmc is 4 mM.90, 91

3.3 Proteins and polyelectrolytes

3.3.1 Bovine serum albumin, BSA BSA is a globular serum albumin protein that can be used as a diluent or a blocking agent in numerous applications. It is also used as a nutrient in cell and microbial culture. In restriction digests, BSA is used to stabilize some enzymes during digestion of DNA and to prevent adhesion of the enzyme to reaction tubes and other vessels. BSA is used because of its lack of effect in many biochemical reactions, and due to its low cost since it is readily available in large quantities from bovine blood. Albumin is essential for maintaining the osmotic pressure needed for proper distribution of body fluids between intravascular compartments and body tissues. It also acts as a plasma carrier by non-specifically binding several hydrophobic steroid hormones, and as a transport protein for hemin and fatty acids. BSA has a molar weight of 66 kDa and the dimension 5.5×5.5×9.0 nm as determine with x-ray diffraction.92

3.3.2 Cytochrome c, CytC Cytochrome c, or CytC is a small globular heme protein found loosely associated with the inner membrane of the mitochondrion. It is a soluble protein, unlike other cytochromes, and is an essential component of the

Page 37: DiVA portal13308/FULLTEXT01.pdf · 2008. 3. 7. · The friction properties in aqueous solution have been studied for adsorbed layers of synthetic polyelectrolytes, with varying density

27

electron transfer chain.93 It is capable of undergoing oxidation and reduction, but does not bind oxygen. It transfers electrons between Complexes III and IV.94 The molar weight of CytC is 12.4 kDa and it consists of 104 amino acids in one continuous polypeptide chain. The dimension determined with x-ray diffraction is 3.0×3.4×3.4 nm.93

3.3.3 Myoglobin, Mb Myoglobin is a single-chain globular protein of 153 amino acids (17.0 kDa), containing a heme (iron-containing porphyrin) prosthetic group in the centre around which the remaining apoprotein folds. It is the primary oxygen-carrying pigment of muscle tissues. Unlike the blood-borne hemoglobin, to which it is structurally related, this protein does not exhibit cooperative binding of oxygen, since positive cooperativity is a property reserved for multimeric proteins. The dimension is 3.3×3.6×6.8 nm.95

3.3.4 Mucin Mucins are a family of large, heavily glycosylated proteins. Although some mucins are membrane-bound due to the presence of a hydrophobic membrane-spanning domain that favours retention in the plasma membrane, the focus in my work is on those mucins that are secreted on mucosal surfaces and in saliva. A schematic structure of mucin is presented in Figure 3.1, two distinctly different regions are found in mucin molecules:

• The amino- and carboxy-terminal regions that are very lightly glycosylated, but rich in cysteines, which are involved in establishing disulfide linkages within and among mucin sub chains.

• Large glycosylated regions in which up to half of the amino acids are serine or threonine. These areas are saturated with hundreds of O-linked oligosaccharides side chains (typical sugars present in mucin are N-acetyl-glucosamine, N-acetyl-galactosamine, galactos, fructose and sialic acid).96 The carbohydrates typically make up 70 – 80% of the mucin mass.96, 97 N-linked oligosaccharides are also found on mucins, but much less abundantly.

Mucins have an anionic polyelectrolyte character due to the presence of sialic acid groups, reported to have a pKa-value98 of 2.6, and a small fraction of sulphate groups (pKa < 1).

Page 38: DiVA portal13308/FULLTEXT01.pdf · 2008. 3. 7. · The friction properties in aqueous solution have been studied for adsorbed layers of synthetic polyelectrolytes, with varying density

28

Figure 3.1: Ttypical structure of a mucin molecule consisting of several subunits connected by disulfide bridges. Note the presence of dense regions of oligosaccharide side chains (of varying lengths) separated by naked regions exposing the protein backbone.

3.3.5 Chitosan Chitosan is produced commercially by deacetylation of chitin, which is the structural element in the exoskeleton of crustaceans (crabs, shrimp, etc.). The degree of deacetylation can be determined by NMR spectroscopy, and the deacetylation in commercial chitosans is typically in the range 60-100%. Chitosan is a linear polysaccharide composed of randomly distributed β-(1-4)-linked D-glucosamine (deacetylated unit) and N-acetyl-D-glucosamine (acetylated unit). The sugar units are linked together with β-(1-4) bonds (as in cellulose), giving the backbone a relatively stiff nature. The persistence length has been determined to be 6.2 nm in 30 mM NaCl.99 Chitosan has a number of commercial and possible biomedical uses. The amino group in chitosan has a pKa value of ~6.5. Thus, chitosan is positively charged and soluble in acidic to neutral solution with a charge density that depends on pH and the degree of deacetylation. Chitosan is bioadhesive and readily binds to negatively charged surfaces such as mucosal membranes. Chitosan enhances the transport of polar drugs across epithelial surfaces, and is biocompatible and biodegradable.

3.3.6 PEO45MEMA:METAC polymers A series of bottle-brush polymers of varying charge densities was used in Paper II. The polymers were synthesized by free-radical co-polymerization of poly(ethylene oxide)-methyl ether methacrylate (PEO45MEMA) and methacryloxyethyl trimethylammonium chloride (METAC) monomers in polymerization tubes under nitrogen atmosphere using AIBN as the thermal initiator.100 Figure 3.2 displays the structures of the two types of segment present in the polymer. Table 3.2 is a list of all the different co-polymers used. The side chains of these co-polymers are well defined, whereas the backbones are polydisperse, which is typical for polymers prepared by free-radical polymerization. For each of these polymers the adsorbed amount is

Page 39: DiVA portal13308/FULLTEXT01.pdf · 2008. 3. 7. · The friction properties in aqueous solution have been studied for adsorbed layers of synthetic polyelectrolytes, with varying density

29

known both on silica and mica101, 102 and from these data the PEO45 chain density is known on each surface, see Figure 5.3.

OO

O

O

n

44

Figure 3.2: Molecular structures of PEO45MEMA (left) and METAC (right) segments in the investigated bottle-brush polyelectrolytes, PEO45-MEMA:METAC-X, where X denotes the amount of the charged METAC component (mol %) in the copolymer.

Table 3.2: List of the PEO45MEMA:METAC-X polyelectrolytes used.

Polyelectrolyte PEO45MEMA (mol%)

METAC (mol%)

Mw (kg mol-1)a

PEO45MEMA:METAC-2 98 2 490 PEO45MEMA:METAC-10 90 10 760 PEO45MEMA:METAC-25 75 25 660 PEO45MEMA:METAC-50 50 50 680 PEO45MEMA:METAC-75 25 75 520 PEO45MEMA:METAC-90 10 90 235 Poly(METAC) 0 100 145

a The weight average molecular weight of the investigated polymers as obtained by static light scattering.101

OO

N

n

Cl+

Page 40: DiVA portal13308/FULLTEXT01.pdf · 2008. 3. 7. · The friction properties in aqueous solution have been studied for adsorbed layers of synthetic polyelectrolytes, with varying density

30

4 Summary of papers The first three papers deal with AFM friction measurements. The results shed some light on lubrication properties of polyelectrolytes, polymer brushes and large glycoproteins. The last three papers have utilized QCM-D to following adsorption and / or following changes in adsorbed layer conformation. Viscosity and density effects of the bulk solution have been extensively discussed. The results of these reports are useful for the correct interpretation of QCM-D data, and they provide new understanding of the effect of ions on mucin layers.

4.1 Lubrication measured with AFM Paper I. In Paper I different methods for calibrating normal and torsional spring constants are compared. The issue of calibrating the torsional detector sensitivity when using some of the torsional methods is addressed. The torsional detector sensitivity is calibrated in air and in liquid, and a ratio to calculate the torsional detector sensitivity factor in a liquid from the air calibration value is derived based on refractive index and optical path length analysis. A new approach to calculate the torsional detector sensitivity by combining the Sader method with the Pivot method is proposed and evaluated. In Paper I it is concluded that the most convenient and fast method for spring constant calibration is to use the normal respectively torsional Sader methods, in combination with lateral detector calibration if friction forces are of interest. This knowledge has been utilized for the friction measurements reported in Paper II and Paper III, and for normal force measurements in Paper II, Paper III and Paper VI. Paper II. The effect of side chain to charge ratio on the friction properties of adsorbed layers formed by bottle-brush polyelectrolytes with poly(ethylene oxide) side chains has been investigated in Paper II. The brush polyelectrolytes were preadsorbed from 0.1 mM NaNO3 solutions onto a mica surface and a silica colloidal probe. Interfacial friction was measured in polyelectrolyte-free solutions with AFM. It was concluded that the decisive factor for achieving favourable lubrication properties is the concentration of non-adsorbing poly(ethylene oxide) side chains in the interfacial region. However, contrary to what may be expected, the results showed that an ideal brush layer structure with the adsorbed polymers adopting comb-like conformation is not necessary for achieving a low coefficient of friction in the asymmetric mica-silica system. In fact, the lowest coefficient of friction, μ, (<0.01) under applied pressures as high as 30 MPa was observed for a system that is incapable of forming brush-like layers.

Page 41: DiVA portal13308/FULLTEXT01.pdf · 2008. 3. 7. · The friction properties in aqueous solution have been studied for adsorbed layers of synthetic polyelectrolytes, with varying density

31

Paper III. Friction properties of adsorbed layers of mucin and chitosan covered mucin have been investigated with the AFM colloidal probe technique as reported in Paper III. Further, the effects of SDS on mucin-chitosan respectively mucin-chitosan-mucin layers haves been studied. Mucin layers were adsorbed on the mica and the silica probe from 30 mM NaCl solution containing 25 ppm bovine submaxillary gland mucin. Chitosan were adsorbed onto the mucin from mucin free 30 mM NaCl solution containing 20 ppm chitosan. Friction was measured in polymer free 30 mM NaCl for all layers and in 30 mM NaCl containing 1 cmc SDS for mucin-chitosan and mucin-chitosan-mucin layers. It will be shown that the favourably low friction obtained for mucin is not obtained for the mucin-chitosan-mucin layer. The layers that showed low friction in this study displayed no attractive forces on separation. We suggest that the main energy dissipative mechanism for the cases with higher friction force is due to formation and breakage of attractive contacts between the cationic (from chitosan) and anionic (from mucin and SDS) residues.

4.2 Adsorption and layer properties measured with QCM-D

Paper IV. Adsorption of three different globular proteins BSA, CytC and Mb onto self assembly monolayers of mercapto unodecanoic acid (MUA) modified gold surface (Au) and MUA modified gold nanoparticles (MUA/AuNP) have been studied with QCM-D respectively with pH titration monitored with ζ-potential measurements. It was concluded that all three proteins formed adsorption layers consisting of an irreversibly adsorbed fraction and a reversibly adsorbed fraction. It was also concluded that BSA is the only of the three proteins studied that was able to sterically stabilize the MUA/AuNP at the isoelectric point. MUA/AuNP coated with the other two proteins were reversibly flocculated at the isoelectric point of the protein MUA/AuNP complexes, whereas the pure MUA/AuNPs were irreversibly flocculated at the isoelectric point. The differences in colloidal stability of the protein MUA/AuNP complex may be correlated with weaker binding in the reversibly bound overlayer in the case of Mb and CytC as compared to BSA. The QCM-D data indicate that the irreversibly adsorbed fraction of Mb is present in form of a bilayer, whereas the irreversibly adsorbed fraction of CytC is a monolayer structure. BSA showed the highest affinity for the MUA/Au surface, forming an irreversibly adsorbed rigid monolayer with a side-down orientation and packing close to that expected by the jamming limit. In addition, BSA showed a large change in the adsorbed mass due to reversibly bound protein. Part of the QCM-D response measured at higher bulk protein concentration (>25μM for BSA) was observed to be due to bulk viscosity and density changes.

Page 42: DiVA portal13308/FULLTEXT01.pdf · 2008. 3. 7. · The friction properties in aqueous solution have been studied for adsorbed layers of synthetic polyelectrolytes, with varying density

32

Paper V. Paper V describe the effects of different electrolytes and electrolyte concentrations on a mucin layer preadsorbed from 30 mM NaNO3 onto a hydrophobic thiol-modified Au surface. To this end we investigated three electrolytes with increasing cation valency, NaCl, CaCl2 and LaCl3. At low NaCl concentrations the preadsorbed layer expands whereas at higher concentrations of NaCl the layer becomes more compact. This swelling/compacting of the mucin layer was fully reversible for NaCl. When the mucin layer instead was exposed to CaCl2 or LaCl3 a compaction was observed already at a concentration of 1 mM. For CaCl2 this process is only partly reversible, and for LaCl3 the changes are irreversible. We note that a correct interpretation of the QCM data required that changes in bulk density and viscosity was taken into account. Further, mucin interactions with the cationic surfactant DTAB in aqueous solution of the different electrolytes were evaluated with turbidimetry, demonstrating that the electrolytes used in this work counteracts the association between mucin and DTAB, and that the effect increases with increasing cation valency. The study was furthermore partly aimed to understand the adsorption of mucin on thiolized gold surfaces with different surface chemistry, for this purpose surfaces exposing COOH, OH and CH3 groups were prepared. Mucin adsorbed to all these functionalized gold surfaces, with high affinity to the hydrophobic surface, and to the charged hydrophilic surface (COOH) despite that in the latter case both mucin and the surface is negatively charged. On the other hand, on the uncharged hydrophilic surface exposing OH groups the adsorption of mucin is low. Paper VI. Paper VI reports results from AFM colloidal probe measurements of the forces acting between preadsorbed mucin layers on uncharged, hydrophobic mercaptohexadecane coated gold surfaces. The effects of salt concentration and cation valency, as in Paper V, on these interactions were explored using NaCl, CaCl2 and LaCl3 in the concentration range 1-100 mM. It was shown that for NaCl, the tail length decreases as the salt concentration is increased, as can be rationalized by considering the polyelectrolyte nature of mucin and the screening of intralayer electrostatic interactions between charged groups, mainly anionic sialic acid. When multivalent cations are present in solution a significant compaction of the mucin layer occurs already at low concentrations (1 mM), suggesting binding of these ions to the anionic sites of mucin. The results are discussed in relation to previous data from quartz crystal microbalance measurements on the same systems reported in Paper V.

Page 43: DiVA portal13308/FULLTEXT01.pdf · 2008. 3. 7. · The friction properties in aqueous solution have been studied for adsorbed layers of synthetic polyelectrolytes, with varying density

33

5 Results and discussion This section is intended to give a survey of the results and discussion treated in more detail within the articles appending to the thesis summary, and to especially address some of the results which in my view are the most interesting. Each section is started by a brief summary of the main conclusions reported in the related papers covered by the respectively section. This is followed by a combined result and discussion part that leads to the main conclusions.

5.1 AFM friction calibration In Paper I it is concluded that the most convenient and fastest method for spring constant calibration is to use the normal respectively torsional Sader technique. If friction is of interest the lateral photo detector sensitivity also has to be calibrated. These conclusions are based on the comparison of different methods for calibrating the spring constants. The results from the normal spring constant calibration, comparing the Cleveland method with the Sader method, showed that they have a reasonable agreement, with a linear relation between spring constant and resonant frequency. However, there is an offset between the two different methods, see Figure 1 in Paper I, where the Cleveland method generates slightly higher values. The source of the offset may depend on many factors, and are difficult to exactly determine. Thus, we cannot conclude which of these two methods that will return the most correct value of the spring constant. Although, since the Sader method is the most convenient to use we adopted this method for our future normal spring constant calibrations. The torsional spring constant comparison includes four different techniques, two dynamic, Sader and Cleveland, respectively two direct static, Pivot and Lever, see section 2.1.5.2 for a description of the different methods. The result from this comparison, Figure 3 in Paper I, are that these techniques return torsional spring constants in the same range, and the scatter within a given data set is of the same order between the methodologies. This is, in fact, a confirmation that all the techniques examined have approximately the same merit in terms of accuracy and that none (or all!) of them suffer from any systematic error. The fact that the dynamic methods agree well with the direct methods is highly reassuring and this was the first time that these methods have been directly compared. The Sader method is the least prone to scatter within the data set. Since it is also the simplest, it seems reasonable to suggest that this should be the preferred method henceforth. We highlighted the importance of performing calibration of the torsional detector sensitivity, δ, by modifying a present method where the whole AFM head (that includes the cantilever holder, laser, and detector) is tilted,

Page 44: DiVA portal13308/FULLTEXT01.pdf · 2008. 3. 7. · The friction properties in aqueous solution have been studied for adsorbed layers of synthetic polyelectrolytes, with varying density

34

see section 2.1.5.3 and Paper I for a description of this method. Figure 5.1 presents some of the data measured with this method. However, since this approach is not possible for all types of AFM we also presented an alternative route to obtain the torsional detector sensitivity, by combining a static method (Pivot data was used in Paper I) with the Sader method. In a similar way as for the normal detector sensitivity,103, 104 it is possible to back out the torsional detector sensitivity. This reverse calculation generated data with some scatter, but the average over ten measurements returns the value δair=(3.58±0.3)×103 V/rad which is very close to the average value obtained by the tilting method, δair=(3.56±0.04)×103 V/rad. Figure 5.1 demonstrates how the torsional detector sensitivity varies when measured in air respectively in a liquid cell filled with water. This shows that it is important to take into consideration the media used when measuring friction forces, especially when measuring in a liquid. We addressed this issue by analyzing the change in length of the optical laser path due to different refractive indices and suggested an algorithm to calculate a ratio between the detector sensitivities as presented in the supplemental material. Interestingly, we came to similar conclusions as Tocha et al.105 (published online a couple of month after we submitted our paper).

-0.003 -0.002 -0.001 0.000 0.001 0.002 0.003-10

-8

-6

-4

-2

0

2

4

6

8

10

-0.004 -0.002 0.000 0.002 0.004

-8

-4

0

4

8

6V 0V -3V -6V

Tors

iona

l res

pons

e (V

)

Tilt angle (rad)

Air Water

Figure 5.1: Detector calibration; torsional detector response as a function of tilt angle for 0 V in normal deflection in air and in a liquid cell measurement. (The lines are linear best fits to the measured data which return δair=3.52×103 V/ rad and δwater=4.67×103 V/ rad.) The insert shows the detector response in air when the laser spot is positioned at different locations at the detector; in the case of no tilt the laser spot was varied from 6 to −6 V in the torsional plan, keeping the normal deflection at 0 V.

Page 45: DiVA portal13308/FULLTEXT01.pdf · 2008. 3. 7. · The friction properties in aqueous solution have been studied for adsorbed layers of synthetic polyelectrolytes, with varying density

35

It is also important to know the detector response function, and a linear response is preferable. The data in Figure 5.1 clearly shows that we have a linear response within the detector. In contrast to our findings of a linear response, Cannara et al.61 reported a highly non-linear torsional photodetector response. The reasons for their non-linearity can be numerous but I want to highlight three of them: First, the detector response in their case may really be non-linear. Second, they may possibly have made an error during their evaluation of the data. Third, the method they have used for measuring the detector sensitivity imposes the non-linearity to the torsional detector sensitivity. Actually in the paper from Cannara et al. the calculation algorithm is not well described, but my own calculations using their published data indicate that they can have made a mistake in the calculation algorithm employed. On the other hand, the method employed to determine the detector sensitivity was to press the cantilever with a probe against a hard vertical wall by scanning the sample laterally, generating a “torsional force curve”. This is an elegant idea that have been employed in different works.61, 62 However, in neither of these studies the normal motion of the cantilever during the measurement is reported. In order to get an accurate sensitivity value the probe should not be allowed to slide in the normal plane. If it does, this will likely result in an incorrect sensitivity value. Finally, some word about the time saving obtained by using the Sader calibration techniques compared to other methods. For example the Cleveland techniques are time consuming and requires technical skills to mount particles on cantilevers and replace the particles with new ones. An estimate on the time consumption by using the Cleveland method is that one can calibrate 5 to 10 cantilevers in 8 hours. This can be compared with the suggested Sader methods where the procedure is trivial and one can calibrate 10 cantilevers in only 1.5 hours. The torsional Lever and Pivot techniques are comparably with the Cleveland method with respect to time consumption.

5.2 Friction and lubrication between poly-electrolye layers

In Paper II it is concluded that low friction forces can be achieved by using adsorbed bottle-brush polyelectrolytes with PEO45 side chains. The lubrication properties between such layers on a mica surface and a silica probe are dependent on the total concentration of PEO45 side chains on the surfaces, and also whether mica or silica has the highest coverage of PEO45. The lowest friction was obtained with PEO45MEMA:METAC-10, that adsorbs to a higher extent on silica compared to on mica.

Page 46: DiVA portal13308/FULLTEXT01.pdf · 2008. 3. 7. · The friction properties in aqueous solution have been studied for adsorbed layers of synthetic polyelectrolytes, with varying density

36

For the mucin / chitosan friction study in Paper III the main conclusions are that, the high density of glycosylated and anionic side chains makes the interpenetration between mucin layers unfavourable both due to steric and electrostatic reasons, and this provide good lubrication between mucin layers. Adsorption of chitosan on the mucin layers leads to compaction of the adsorbed layers, which in combination with attractive mucin-chitosan interactions result in high friction. Adsorption of additional mucin to the mucin-chitosan layers does not change the friction since the complex mucin-chitosan and mucin-chitosan-mucin layers do not posses a strictly layered structure. They are rather considerably entangled, thus enabling attractive interactions between oppositely charged sialic acid residues from mucin and amino groups from chitosan residing on opposing surfaces. SDS binds to a great extent to mucin-chitosan and mucin-chitosan-mucin layers. The friction behaviour in presence of SDS is rather complex since both the layer structure and composition can be altered by the combined action of load and shear. In Figure 5.2 and Table 5.1, the friction measurements from Paper II and III are summarized. Figure 5.2 displays the effective coefficient of friction, μeff, measured between adsorbed layers PEO45MEMA:META-X polyelectrolytes in part (a) at the applied load of 2 mN/m, whereas in part (b) data over the whole measurement range are shown. Table 5.1 presents μeff of mucin and chitosan layers (both single component and two component layers, for the two component layer also in presence of SDS and after exposure to SDS solution) at the applied load of 4 mN/m together with friction forces measured during unloading at zero applied loads.

5.2.1 PEO45MEMA:METAC-X When discussing the friction properties of the PEO45MEMA:METAC-X polymers it is valuably to account for the amount of PEG45 segments between the mica and silica surface during the friction measurement. This information is given in Figure 5.3, and is the sum of the number of PEO45 segments adsorbed on the respective surface. The importance of the PEO45 side chains for the friction properties becomes evident when μeff between layers of PEO45MEMA:METAC-2 and layers of poly(METAC) are compared to μeff for the bare mica-silica system in Figure 5.2. The friction force increases with an order of magnitude when poly(METAC) is adsorbed; this is due to electrostatic bridging, which brings an additional energy dissipative process to the system as polymer-surface anchoring points are broken and reformed during shear. When only one of the interacting surfaces is coated with PEO45MEMA:METAC-2, (this polymer does not adsorb on mica, see Figure 5.3), this is enough to decrease the friction compared to the bare mica-silica case.

Page 47: DiVA portal13308/FULLTEXT01.pdf · 2008. 3. 7. · The friction properties in aqueous solution have been studied for adsorbed layers of synthetic polyelectrolytes, with varying density

37

0 20 40 60 80 1000.001

0.01

0.1

1 (a)

μ eff

X (mol%)

0 2 4 6 8 10 12 14 160.001

0.01

0.1

1

10 (b)

μ eff

FLoad Figure 5.2: (a) Effective coefficient of friction, μeff, (filled squares) of the investigated PEO45MEMA:METAC-X systems measured at FLoad/R ≈ 2 mN/m (or at the vicinity of the highest applied load on the system). The unfilled symbol denotes μeff of bare mica-silica. (b) μeff as a function of applied load FLoad: (crosses) bare mica-silica, (filled squares) PEO45MEMA:METAC-2, (filled circles) PEO45MEMA:METAC-10, (filled triangles) PEO45MEMA: METAC-25, (stars) PEO45MEMA:METAC-50, (unfilled squares) PEO45-MEMA:METAC-75, (unfilled circles) PEO45MEMA:METAC-90, (unfilled triangles) poly(METAC)

Additionally, a closer inspection of the results (Figure 5.2 and Figure 5.3) reveals that the lubrication properties depend to a significant degree on which of the surfaces that has the highest concentration of adsorbed side chains. This conclusion is reached from the fact that while the total number

Page 48: DiVA portal13308/FULLTEXT01.pdf · 2008. 3. 7. · The friction properties in aqueous solution have been studied for adsorbed layers of synthetic polyelectrolytes, with varying density

38

0 20 40 60 80 100

0.0

0.2

0.4

0.6

0.8

1.0

1.2

Nr o

f PEO

45 s

egm

ents

/nm

2

X (mol%) Figure 5.3: Number of PEO45 segment per nm2 on silica (squares) and mica (triangles), the dashed line shows the total concentration of PEO45 chains adsorbed on the silica probe and the mica surface.

of PEO45 segments in the interfacial region is approximately the same for brush polyelectrolytes with X=10 to 50 (see Figure 5.3), PEO45MEMA:- METAC-10 which adsorbs the most on silica exhibits better lubrication properties than PEO45MEMA:METAC-50 (which adsorbed the highest on mica). This is consistent with the finding that poly(ethylene oxide) has a significantly larger affinity for silica than for mica. Thus, the silica surface must be fully covered in order to obtain the lowest possible friction against another surface exposing poly(ethylene oxide) chains. The generally low friction forces observed in this study (for X<90) suggests that the adsorbed layers in our study from a friction point of view behave as brush layers, even though they, strictly speaking, are not.101 The reason for this is suggested to be the high side-chain density on the (PEO45MEMA:METAC-X) polyelectrolytes that locally provides a high density of PEO45 side chains, which counteracts chain interpenetration.

5.2.2 Mucin and chitosan In Paper III we suggested that mucin, also from a friction point of view, behaves as a brush layer due to the high density of glycosylated and anionic side chains that makes the interpenetration between mucin layers unfavourable both due to steric and electrostatic reasons. The effective coefficient of friction measured between mucin layers 0.03±0.02 (see Table 5.1) is of the same magnitude as for PEO45MEMA:METAC-X when X = 25 to 75. Between adsorbed layers of chitosan μeff is 0.13±0.02. Kampf et al.26 have reported an even lower friction coefficient (0.07) between chitosan layers adsorbed on mica. We assign the higher value found in Paper III to the attractive interaction that is present in our case, see Figure 3 in Paper III, but not observed in the study of Kampf et al. This difference is suggested to

Page 49: DiVA portal13308/FULLTEXT01.pdf · 2008. 3. 7. · The friction properties in aqueous solution have been studied for adsorbed layers of synthetic polyelectrolytes, with varying density

39

arise from the lower amount of chitosan adsorbed on silica (the probe) than on mica, which facilitates bridging. When chitosan is adsorbed on the mucin layer, the friction force is dramatically increased compared to those encountered between the pure mucin and chitosan layers, with μeff ≈0.4. Interestingly, adsorbing additional mucin on the mucin-chitosan layer does not reduce the friction force that remains characterized by μeff ≈0.4. Adsorbed layers including both mucin and chitosan contain a large number of both anionic and cationic groups. Thus, if the sequential adsorption does not result in a strict layering of the components, then formation and breakage of attractive interactions between sialic acid residues and amine groups can be expected to contribute to the friction. Clearly, these attractive electrostatic interactions between the oppositely charged segments contribute significantly to the friction force observed between these complex layers.

Table 5.1: Effective coefficient of friction and friction force at zero applied load for different layers.

Layers μeff Ff(0) (nN) Loadinga Unloadinga Unloading Mucin 0.03±0.02 0.03±0.02 0.03±0.03 Chitosan 0.13±0.02 0.13±0.02 0.17±0.02 Mucin-chitosan 0.37±0.05 0.42±0.03 1.3±0.4 Mucin-chitosan-mucin 0.42±0.03 0.43±0.04 0.3±0.2 Mucin-chitosan in SDS 0.26±0.03 0.10±0.02 0.26±0.1 Mucin-chitosan after SDS 0.56±0.05 0.61±0.05 5.9±0.5 Mucin-chitosan-mucin in SDS 0.03±0.02 0.19±0.03 0.05±0.05 Mucin-chitosan-mucin after SDS 0.22±0.02 0.23±0.02 0.45±0.1 aAt F/R = 4 mN/m.

5.2.3 Effect of SDS on layers containing mucin and chitosan

Addition of SDS, that readily associates with both chitosan and mucin, brings additional complexity to the two component layers not only because another component is added but also because the surfactant may leave the layer during compression and shear. Figure 5.4 and Figure 5.5 report the friction forces obtained for the two component system in absence and presence of 1 cmc SDS in 30 mM NaCl, and after exposure to SDS followed by rinsing, measured in 30 mM NaCl. After rinsing there will be no SDS left in solution, but some SDS may well remain in the adsorbed layers. The difficulty of removing SDS from adsorbed layers of cationic polyelectrolytes has been noted in previous studies.106

Page 50: DiVA portal13308/FULLTEXT01.pdf · 2008. 3. 7. · The friction properties in aqueous solution have been studied for adsorbed layers of synthetic polyelectrolytes, with varying density

40

0 1 2 3 4 5 60

1

2

3

4

5

6

7

8

9

10

Fn/R (mN/m)

F f (nN

)

0 5 10 15 20

Fn (nN)

a)

0 10 20 30 40 50

10-2

10-1

100

101

Separation (nm)

F n/R (m

N/m

)

b)

Figure 5.4: a) Friction forces vs applied load. Filled symbols represent data obtained during loading and open symbols represent data recorded on unloading. (■ mucin-chitosan, ● mucin-chitosan in SDS, ▲ mucin-chitosan after SDS exposure followed by rinsing with 30 mM NaCl.) b) Force normalized by radius on approach between mucin-chitosan coated surfaces after exposure to SDS followed by rinsing, measured in 30 mM NaCl solution. (▲ before shearing, and × after shearing.)

Exposure of the mucin-chitosan layer (Figure 5.4) to 1 cmc SDS in 30 mM NaCl, does not affect the friction on loading up to a load of 2 mN/m, after which the friction force becomes close to independent of load. This layer also displays a lower friction force during unloading than during loading, see Figure 5.4a and Table 5.1. This is different compared to all other layers studied in Paper III. These results clearly indicate a structural transition under shear and we offer the following hypothesis: SDS is prone to form bilayer aggregates on oppositely charged surfaces and on adsorbed chitosan layers.107 At low loads, some chitosan chains with associated SDS are extended into solution as evidenced by the presence of a steric repulsion (see Figure 6 in Paper III). Under compression a smoother layer is formed and the combined action of load and shear may result in a situation where sliding occurs between opposing layers exposing the sulphate groups of SDS towards each other. Such a layer structure would result in low friction forces since the load would be carried by electrostatic and hydration forces with no or very limited polymer interpenetration. Rinsing away SDS from the bulk solution with 30 mM NaCl results for the mucin-chitosan in a friction curve that increases strongly when the load exceeds about 0.5 mN/m. At higher loads the slope of the friction vs load curve is similar to that obtained for the mucin-chitosan layers prior to exposure to SDS. However, the magnitude of the friction force is higher. This indicates that some SDS remains in the layer after rinsing. The presence of SDS residues gives rise to attractive interactions both on approach and separation (Figure 5.4b). On loading a low friction force is

Page 51: DiVA portal13308/FULLTEXT01.pdf · 2008. 3. 7. · The friction properties in aqueous solution have been studied for adsorbed layers of synthetic polyelectrolytes, with varying density

41

encountered as long as the load is carried by the steric repulsion (F/R < 0.5 mN/m). However, as soon as the force barrier is overcome, and the surfaces come into intimate contact, a significant increase in friction force is observed. At high loads, (≥ 3 mN/m) when the attractive interactions between the layers are fully developed, an Amontons-like behaviour is recovered (Figure 5.4a, triangles). On unloading significantly higher friction forces are observed at low load, and a strong friction is also observed at zero loads. We assign this to attractive interactions between the hydrocarbon tails of SDS and attractive SDS-chitosan interactions. When the mucin-chitosan-mucin layers are exposed to 1 cmc SDS, the friction force on loading is very low (Figure 5.5a), comparable to that between mucin layers adsorbed on mica and silica surfaces (see Table 5.1). However, at the highest load used in this measurement the friction force is increased and on unloading the friction is significantly higher. The low friction force during loading originates from the large swelling induced by incorporation of the negatively charged surfactant in the layer (see Figure 5.5b). A similar large swelling due to SDS incorporation has previously been noted for layers of a cationic polyelectrolyte where 10% of the segments carry a positive charge.108 In that work it was also noted that the surfactant aggregates suddenly were squeezed out when a critical load was reached, and that the surfactant aggregates did not reappear to any large extent in the layer until the load was significantly released again. No

0 1 2 3 4 5

0

1

2

3

4

5

6

7

Fn/R (mN/m)

F f (nN

)

0 5 10 15

Fn (nN)

a)

0 10 20 30 40 50

10-2

10-1

100

101

Separation (nm)

F n/R (m

N/m

)

b)

Figure 5.5: a) Friction forces vs applied load. Filled symbols represent data obtained during loading and open symbols represent data recorded on unloading. (♦ mucin-chitosan-mucin, ● mucin-chitosan-mucin in SDS, ▲ mucin-chitosan-mucin after SDS exposure followed by rinsing with 30 mM NaCl.) b) Force normalized by radius between mucin-chitosan-mucin coated surfaces in solutions containing 1 cmc SDS and 30 mM NaCl. Forces measured on approach and separation before shearing are shown by filled and unfilled symbols, respectively. The forces measured on approach after shearing are shown by crosses.

Page 52: DiVA portal13308/FULLTEXT01.pdf · 2008. 3. 7. · The friction properties in aqueous solution have been studied for adsorbed layers of synthetic polyelectrolytes, with varying density

42

indication of such an event is observed in the force data in Paper III, (see Figure 5.5b). However, the sudden increase in friction observed at high loads (F/R ≈ 4 mN/m, P ≈ 40 MPa), and the sustainability of high friction under decompression (Figure 5.5a) indicates that under the combined effect of load and shear a large fraction of the incorporated SDS leaves the surface region. When this happens, the friction force becomes similar to that observed after rinsing away the SDS. When the SDS is rinsed away with 30 mM NaCl the friction forces on loading and unloading are similar to the friction that was obtained for the unloading measurement when SDS was present (Figure 5.5a). It appears that the fraction of SDS that is removed by rinsing is similar to the fraction that is squeezed out from the layer under the combined action of load and shear. The lower friction observed between mucin-chitosan-mucin after exposure to SDS followed by rinsing compared to that of the original mucin-chitosan-mucin layer suggests that the remaining SDS is associated with the positive charges of chitosan, and this reduces the number of possible mucin-chitosan bonds that are formed and broken during sliding.

5.3 Globular protein adsorption studied with QCM-D

By studying adsorption of BSA, CytC, and Mb on to mercaptoundecanoic acid modified gold surfaces (MUA/Au), we concluded that BSA adsorbs with higher affinity than CytC and Mb, and with a preferential side-down orientation in an amount corresponding to the jamming limit (Paper IV). CytC forms a closely packed monolayer due to significantly higher mobility on the surface, and Mb adsorbs in a bilayer (or at least significantly more than a monolayer) structure. Below I will present some of the results that finally resulted in the above conclusions. Two different types of adsorption measurement were preformed with QCM-D; adsorption concentration isotherms and single concentration additions at 100 μM protein. I will only focus on the concentration isotherm in the following discussion. Between the different concentrations in the adsorption isotherm the adsorption was interrupted by injection of water to rinse the measuring cell. In Figure 5.6 the adsorption isotherm for BSA is presented (the data for the other proteins are presented in Figure 2 and 3 in Paper IV). The filled symbols are measurements made in the protein solution before rinsing with water, whereas unfilled symbols are data from the water rinsing period. If we compare the filled and unfilled symbols for the concentrations above 0.8 μM one can clearly see that there are some loosely physisorbed fraction of BSA that is removed during the rinsing step leaving a small fraction of adsorbed BSA on the rinsed surface.

Page 53: DiVA portal13308/FULLTEXT01.pdf · 2008. 3. 7. · The friction properties in aqueous solution have been studied for adsorbed layers of synthetic polyelectrolytes, with varying density

43

Figure 5.6: QCM-D data from the third overtone obtained by sequential addition of BSA. The solid squares represent the initial measurement when the protein solution was introduced, and the open squares represent the measurement following a rinsing step. Panel A shows the frequency shift for concentrations in the range from 100 nM to 1 μM. Panel B shows the frequency shift for concentrations in the range from 10 to 100 μM. The triangles show the frequency shift corrected for the density and viscosity of the protein solution. Panels C and D show the corresponding dissipation shifts (×106).

Above 25 μM BSA it is also important to consider viscosity and density changes, compare filled triangles with filled squares in Figure 5.6. The corrected frequencies have reached a plateau value at 50 μM whereas the not corrected values are constantly decreasing. Unfortunately, we only measured viscosity and density for BSA solutions, leaving the similar interpretations for CytC and Mb difficult to prove. However, it is most likely that the viscosity and density for these proteins also start to affect the measurements at similar concentrations. The amount of loosely physisorbed protein is low for CytC and Mb at low concentration and the difference between the values before and after rinse are increasing above 25 μM, part of this effect is likely due to density and viscosity effects. The surface coverage (Γ) of the irreversibly adsorbed proteins is obtained by dividing the sensed mass after rinsing by the molecular mass and by the active area of the QCM crystal; these values are reported in Table 1 in Paper IV. I want to note that these values actually are slightly overestimated due to that the QCM mass includes coupled water. The known side-binding value for BSA109 is 2.0×1012 cm-2, which is lower than the end-on value. When comparing these values with our reported BSA value 0.85×1012 cm-2, it is logical to assume that BSA is adsorbed by a side-binding mechanism onto MUA/Au surfaces. The BSA adsorption is close to the jamming limit,110 i.e. the maximum adsorption achievable if the proteins stick to the surface at the point where they initially attach in random order (in this case approximately

Page 54: DiVA portal13308/FULLTEXT01.pdf · 2008. 3. 7. · The friction properties in aqueous solution have been studied for adsorbed layers of synthetic polyelectrolytes, with varying density

44

50% of full coverage). Thus, we conclude that BSA adsorbs to MUA/Au without significant lateral mobility and with high affinity for the surface. Turning from BSA to CytC the calculated side-on (8.6×1012 cm-2) value and end-on (9.8×1012 cm-2) surface coverage for CytC are close to each other, whereas the measured surface coverage after rinsing for CytC is 11.75×1012 cm-2. Hence, the value is close to the maximum monolayer coverage, the discrepancy is most likely due to sensed coupled water. This indicates that CytC adsorbs in a monolayer, and that CytC are allowed to move laterally on the surface to create a denser layer then predicted by the jamming limit. For Mb the calculated coverage after rinsing is larger than a monolayer. According to the adsorbed mass of 446 ng/cm2 the number of layers of Mb is predicted to be between 1.9 and 3.7 for the end-on and side-on binding models, respectively. Finally I will shortly mention the ζ-potential measurements that also were part of this work. The different proteins were adsorbed on MUA modified gold nanoparticles, MUA/AuNP, and then the ζ-potential was measured with protein in the solution during pH titration. One of the interesting results was that BSA could stabilize the MUA/AuNP even at the isoelectric point. This behaviour provide evidence for steric stabilization.111 Thus, BSA adsorption, including the reversibly adsorbed molecules results in a layer with sufficient coverage and thickness to counteract the van der Waals attraction between the nanoparticles. Interestingly, the larger adsorbed amounts of the other two proteins do not result in steric stabilization, but rather attractive forces result in reversible flocculation. A better understanding of this could be obtained by direct force measurements.

5.4 Mucin interaction with metal ions In this section I will discuss the findings from Paper V and VI. The main conclusions from these two papers are the following. The effects of 1:1 electrolyte (NaCl) on the preadsorbed mucin layer are fully reversibly. The layer contracts with increasing ionic strength, which is due to electrostatic screening of the repulsion between the charge groups in the mucin layer. Multivalent cations, such as Ca2+ and La3+, cause a significant contraction of the tails even in dilute solutions (1 mM). The effects of multivalent ion cannot be explained by only considering the variation of the electrostatic screening due to changes in the Debye-length, suggesting binding of these ions to the anionic sites of mucin. We also concluded that the low adsorption of mucin on the hydrophilic and uncharged surface is due to lack of electrostatic and hydrophobic interactions between mucin and this surface. Adsorption of mucin monitored with QCM-D on different alkylthiol modified gold surfaces (terminated with CH3, OH or COOH groups) from

Page 55: DiVA portal13308/FULLTEXT01.pdf · 2008. 3. 7. · The friction properties in aqueous solution have been studied for adsorbed layers of synthetic polyelectrolytes, with varying density

45

30 mM NaNO3 is presented in Figure 1 and 2 in Paper V. Mucin adsorbs with highest sensed mass to the negatively charged hydrophilic surface with COOH-groups exposed. The adsorption to the uncharged hydrophilic surface exposing OH-groups is remarkably small considering that the sensed mass (1.05 mg/m2) includes both the mass of the mucin and that of water associated with the layer. The driving force for the mucin adsorption is different for the three different surfaces. The main driving force for the adsorption on the non-polar (CH3) surface is hydrophobic interaction, i.e. the removal of contacts between water and the hydrophobic surface as well as between water and non-polar parts of the mucin. Mucin also adsorbs strongly to the negatively charged polar surface and despite the fact that mucin and the surface are both negatively charged; we suggest that electrostatic forces are promoting adsorption of mucin to the negative charged polar (COOH) surface. The reason is that mucin also contains positive amino acid residues. Thus, it seems plausible to suggest that these play an important role in the adsorption process. On the uncharged polar (OH) surface the adsorption is low since neither hydrophobic nor electrostatic forces are important. We suggest that the limited adsorption observed at this surface is caused by changes in free energy due to changes in hydrogen bonding between mucin/water, surface/water, water/water, and mucin/surface, respectively. In the rest of this section, where I will discuss the effect of the change of electrolyte, mucin is initially adsorbed on to hydrophobic (CH3) surfaces from 30 mM NaNO3 solution as in the cases above. Three different electrolytes (NaCl, CaCl2, and LaCl3) were used at concentrations of 1, 10, 50, and 100 mM, both in QCM-D and AFM colloidal probe measurements. QCM-D gives some information on what is happening with the mucin layers, and AFM force measurements provide additional information. By combining the information from QCM-D and AFM we are able to draw some conclusions that neither of these techniques would have provide alone. Figure 5.7 displays results from QCM-D measurement of the preadsorbed mucin layer when it was exposed to NaCl (no mucin is present in the bulk solution). In this figure one can see many interesting features. The need for viscosity and density correction which is visible as the deviation between the data points illustrated by crosses and squares; in particular at 100 mM NaCl. This effect is even more pronounced in Figure 6 and 7 in Paper V that illustrate the presence of viscosity and density effects, that need to be accounted for. There are also some deviations for the sensed mass (calculated via the Sauerbrey equation; filled squares) and the true sensed mass (open squares; obtained from Johannsmann’s model), especially when the dissipation is high as for 1 and 10 mM NaCl. Interestingly, at 1 mM NaCl the sensed mass is lower at the same time as the dissipation is higher than at higher NaCl concentrations. This is not logical for an irreversibly

Page 56: DiVA portal13308/FULLTEXT01.pdf · 2008. 3. 7. · The friction properties in aqueous solution have been studied for adsorbed layers of synthetic polyelectrolytes, with varying density

46

adsorbed layer. (The irreversible adsorption was proven by the complete reversibility with respect to changes in NaCl concentration). To explain this we have to consider the force curves measured with the AFM for the same system, Figure 5.8. The force curves show a long range steric repulsion (700 nm) in 1 mM NaCl, this can be compared with the decay length of the acoustic wave in the QCM, about 200 nm, see Section 2.2. This is considerably shorter than the range of the steric force observed with AFM, demonstrating that the tail region of the layer on each surface is at least

0 20 40 60 80 100 1200

1

2

3

4

5

6

7 Layer response Total response

Dis

sipa

tion

chan

ge (x

106 )

NaCl concentration (mM)0 20 40 60 80 100 120

2.5

2.7

2.9

3.1

3.3

3.5

3.7

3.9

4.1

4.3

4.5

Total response Layer response True sensed mass

Sens

ed m

ass

(mg/

m2 )

NaCl concentration (mM) Figure 5.7: Sensed mass and dissipation change at different NaCl concentrations. “Total response” is the sensed mass obtained from Eq 2.14 and the dissipation change is directly measured. “Layer response” is obtained using Eq 2.12 respectively 2.13, and here the effects due to changes in bulk density and viscosity are subtracted. “True sensed mass” is obtained using Johannsmann’s model that accounts for the layer viscoelasticity. The dashed lines correspond to the values for the mucin layer in contact with 30 mM NaNO3.

0 100 200 300 400 500 600 700 800

10-1

100

101

Separation (nm)

F/R

(mN

/m)

1 mM NaCl10 mM NaCl100 mM NaCl

Figure 5.8: Force normalized by radius between mercaptohexadecane modified gold surfaces carrying a preadsorbed mucin layer as a function of separation. Data are shown for interactions across a range of NaCl solutions.

Page 57: DiVA portal13308/FULLTEXT01.pdf · 2008. 3. 7. · The friction properties in aqueous solution have been studied for adsorbed layers of synthetic polyelectrolytes, with varying density

47

350 nm thick. Thus, the QCM-D is not able to sense the complete adsorbed layer in low ionic strength NaCl solutions, and thus the QCM-D data become unreliable. It seems clear that the high extension of the mucin layer in 1 mM NaCl is due to strong intralayer electrostatic repulsion between anionic groups, mainly sialic acid. By increasing the NaCl concentration to 10 mM a reduction in the range and strength of the interaction is achieved, signifying that the tails contracts. A further increase in NaCl concentration to 50 mM results in further tail contraction. In 100 mM NaCl, a slight reswelling appears to occur as evidenced by a slight increase in the repulsion observed below 10 nm. We conclude that the electrostatic repulsion between the anionic groups is determining the mucin layer extension in 1:1 electrolyte. In Figure 5 to 7 within Paper V it is shown that the changes induced by the different electrolytes are fully reversible for NaCl, and not fully reversible for CaCl2, respectively irreversible when LaCl3 have been used. In cases where both the adsorbed mass (Γ), e.g. obtained by reflectometry, and the true sensed mass (m0) obtained with QCM-D are known, then a QCM-thickness (dQCM) can be evaluated using Eq. 5.1.74

⎟⎟⎠

⎞⎜⎜⎝

⎛ Γ−+Γ

==

00

00

1mm

mmd

smeff

QCM

ρρρ

(5.1)

where ρeff is the effective density of the layer, ρm the density of mucin (a typical mucin value112 of 1.4 g/mL was used in the calculation), and ρs the density of the solvent (reported in Paper V). The equation above builds on a shear model that assumes that all material in the distance range 0 to dQCM oscillates with the crystal and thus contributes fully to the true sensed mass, whereas the material located further away from the surface does not oscillate with the crystal and does not contribute to the sensed mass. This is clearly a simplification, just as the optical models used for evaluating the ellipsometric thickness. It has been shown that the QCM thickness and the ellipsometric thickness are similar for relatively compact and homogeneous layers.113 We do not expect this to be the case for more diffuse polymer layers since the ellipsometric thickness is directly influenced by the segment density profile,114 whereas the QCM thickness is influenced by the amount of water that oscillates with the crystal and this quantity is at present an unknown function of the segment density profile. For the extended mucin layers investigated here we will use the QCM thickness (presented in Figure 5.9) only as a qualitative indicator where a decrease in thickness means a reduction in water content of the adsorption layer, which in turn indicates a contraction of the layer.

Page 58: DiVA portal13308/FULLTEXT01.pdf · 2008. 3. 7. · The friction properties in aqueous solution have been studied for adsorbed layers of synthetic polyelectrolytes, with varying density

48

before 1mM 10mM 50mM 100mM after

0.5

1.0

1.5

2.0

2.5

3.0

3.5

4.0

h QC

M (n

m)

Figure 5.9: The QCM thickness for different mucin-electrolyte combinations, (circles – NaCl, squares – CaCl2, and triangles – LaCl3), calculated by using an adsorbed amount of 1.8 mg/m2, taken from ref 115.

The first feature to note is that the QCM thickness (Figure 5.9) is orders of magnitude lower than the range of the steric force detected with AFM (see Figure 5.8 and Figure 6 to 8 in Paper V). This is not unexpected. To measure a thickness of an adsorbed polymer layer that is diffuse is by necessity dependent on the measuring technique. Surface force measurements, and measurements of hydrodynamic thicknesses using e.g. dynamic light scattering, are very sensitive to the longest tails and thus provide a large value for the thickness. On the other hand, the models used to evaluate the QCM thickness, and also the ellipsometric thickness, are insensitive to a small fraction of long tails. Thus, the range of the measured forces provides information related to the length of the longest tails whereas the QCM thickness in our case provides complementary information related to the amount of water hydrodynamically coupled to the layer. The QCM thickness is consistently lower in CaCl2 solutions than in NaCl solutions. Thus, the presence of Ca2+ ions results in expulsion of water, most likely from the inner part of the adsorbed layer where trapped water is expected to be found, as well as in a reduction in the range of the steric force monitored by AFM (Figure 7, Paper V) that reflects the length of the tails. The QCM thickness in 1 mM LaCl3 is lower than found for any of the other salts, demonstrating the compacting effect of the trivalent cation. In contrast to what was found for the other electrolytes, the QCM thickness increases with LaCl3 concentration, indicating incorporation of water due to swelling of the inner part. We attribute this to recharging of the layer as also demonstrated for the tail region by the force measurements reported in Figure 7 in Paper V. The strong influence of multivalent cations on the structure of mucin layers show that the presence of these ions is likely to affect the lubricating and

Page 59: DiVA portal13308/FULLTEXT01.pdf · 2008. 3. 7. · The friction properties in aqueous solution have been studied for adsorbed layers of synthetic polyelectrolytes, with varying density

49

hydrating properties of mucin coatings. This is a topic that deserves further attention. To this end I will present some of my unpublished data on this topic; the friction data presented in Figure 5.10 is measured on the same systems as used in Paper VI (hydrophobic thiolated gold). Hence, the surfaces are rougher compared to the surfaces used in Paper III where we have studied the friction between mucin layers adsorbed onto hydrophilic mica and silica. By comparing the friction force measured between mucin layers form Figure 10 in Paper III with Figure 5.10 (crosses) we can see that the friction is considerably higher for the latter case. It has previously been reported that the friction between surfaces of similar chemistry but with different surface roughness is very different, and follows a logarithmic increase of the coefficient of friction with increasing surface roughness.38, 116 Nevertheless, my purpose by showing Figure 5.10 is to illustrate the trends observed when changing electrolyte concentration and type. For 1 mM NaCl the friction force is extremely low up to an applied load of 5 mN/m. From the force curve in Figure 5.8 it can be seen that this is approximately where we go from a situation where the load is carried by the dilute tail region to a situation where the load is carried by compressed layers. The slope found at the upper part of the friction curve is similar to the slope of the other salt concentrations. The friction increases with NaCl concentration, and this is attributed to the contraction of the tail region. For LaCl3 we can see the compaction of the mucin layer as an increase of the friction for all concentrations used compared to the initial mucin layer in 30 mM NaNO3 (crosses). To summarize, we have despite the surface roughness of the probe/sample observed really low friction, μeff=0.02, at loads up to 5 mN/m in 1 mM NaCl. This is possible since part of the applied load is carried by the strong steric repulsion between the two mucin layers that have very extended tail regions (C in Eq 1.7 is negative). It have recently been demonstrated that extremely low friction (μeff<0.0003) is obtained between surfaces separated by repulsive van der Waals forces.117 This indicates that by adsorbing mucin at a high ionic strength and then later use the layer for lubrication at lower ionic strength, the decreasing electrostatic screening will result in expanded layers with good lubricating properties. This effect can likely be used for other polyelectrolytes as well, as long as the surface attachment is irreversible. This interesting result generates some additional questions that I do not yet know the answer to. Does mucin has the same behaviour when it is adsorbed at hydrophilic and charged surfaces compared with hydrophobic surfaces? How rough can the surfaces be and still be effectively lubricated by the extended mucin layer? What other polyelectrolytes (type of structures) will show similar behaviour?

Page 60: DiVA portal13308/FULLTEXT01.pdf · 2008. 3. 7. · The friction properties in aqueous solution have been studied for adsorbed layers of synthetic polyelectrolytes, with varying density

50

0 2 4 6 80

10

20

30

40

50

60

70

80

Fn/R (mN/m)

F f (nN

)

mucin in 30 mM NaNO

3

1 mM NaCl10 mM NaCl100 mM NaCl

0 10 20 30 40

Fn (nN)

a)

0 1 2 3 4 5 6 7 80

10

20

30

40

50

60

70

80

90

Fn/R (mN/m)

F f (nN

)

mucin in 30 mM NaNO

3

1 mM LaCl3

100 mM LaCl3

0 10 20 30 40

Fn (nN)

b)

Figure 5.10: Friction force measured between mercaptohexadecane modified gold surfaces (a flat and a colloidal probe) carrying a preadsorbed mucin layer as a function of applied load. Data are shown for interactions across 30 mM NaNO3, and across a range of a) NaCl and b) LaCl3 solutions.

Page 61: DiVA portal13308/FULLTEXT01.pdf · 2008. 3. 7. · The friction properties in aqueous solution have been studied for adsorbed layers of synthetic polyelectrolytes, with varying density

51

6 Concluding remarks This thesis work has resulted in several conclusions, both concerning on the one hand technical aspects and experimental procedures, and on the other hand physical chemical aspects of the investigated systems.

6.1 Technical and experimental conclusions For AFM friction calibrations any residual doubt as to whether the direct static methods for spring constant determination return the same value as the dynamic ones are now essentially laid to rest. This opens up the possibilities to use the quicker and far more trivial Sader technique for calibrating both the normal and torsional spring constants. By combining one direct static method with the torsional Sader method one can backwards calculate the torsional detector sensitivity; this is important since far from all AFM instruments allow the operator to use the method of tilting the detector system when calibrating torsional the detector sensitivity. When measuring force and friction with AFM on soft adsorbed layers it is best to use the normal deflection sensitivity (constant compliance) measured between the surfaces before adsorbing the layers. QCM can encounter difficulties to monitor changes in polyelectrolyte layers if the layers have tails that are protruding further from the surface than the decay length of the QCM acoustic wave. The combination of QCM-D and AFM colloidal probe measurements can allow understanding of some structural changes that neither of these two techniques is capable to confirm alone. Density and viscosity corrections are essential for correct interpretations of QCM data when studying the effect of changes in electrolyte type and concentration on preadsorbed layers of mucin, and also when having different concentrations of proteins in the measuring solution.

6.2 Physical chemical conclusions The adsorbed amount of proteins appears to depend on the strength of the surface attachment, in such a manner that a too high affinity reduces the adsorbed amount. This can partly be due to surface denaturation, but also because that the proteins become immobilized on the spot of attachment. In such a case a monolayer cannot be more densely packed than allowed by the jamming limit. This was the case for BSA that adsorbs with high affinity (low mobility) on the MUA/Au surface with a preferential side-down orientation. CytC forms a closely packed monolayer due to higher mobility on the surface and Mb adsorbs in a bilayer (or at least greater than monolayer) structure. Of course, a very low affinity between the protein and

Page 62: DiVA portal13308/FULLTEXT01.pdf · 2008. 3. 7. · The friction properties in aqueous solution have been studied for adsorbed layers of synthetic polyelectrolytes, with varying density

52

the surface also results in low adsorption. In my studies this situation was encountered when mucin was allowed to adsorb on OH-terminated thiol surfaces, where the lack of both electrostatic and hydrophobic driving forces for the adsorption resulted in the very low adsorption. When a preadsorbed mucin layer interacts with 1:1 electrolyte (NaCl) in the range of 1 to 100 mM the changes induced in the layer structure are fully reversibly. The layer expands with decreasing ionic strength, which is due to electrostatic repulsion between the charge groups in the mucin layer. Multivalent cations, such as Ca2+ and La3+, cause a significant contraction of the layer even in dilute solutions (1 mM). This cannot be explained by a variation of the electrostatic screening due to changes in the Debye-length, but suggests specific binding of these ions. In my thesis work I have encountered several cases where very low friction coefficients have been achieved. The common features of these systems are that no attractive forces are present, and that excluded volume and / or electrostatic forces counteract chain interpenetration under load. For instance, the friction between PEO45MEMA:METAC-X polymer layers on mica and silica was found to be small when the PEO45 side chain density was sufficient high. Another layer that display a low coefficient of friction is that formed by mucin. In this case the high density of glycosylated and anionic side chains on mucin makes the interpenetration between mucin layers unfavourable both due to steric and electrostatic reasons. Adsorption of chitosan on the mucin layers leads to compaction of the adsorbed layers and thus steeper and more short-range repulsive forces. This, in combination with attractive mucin-chitosan interactions, generates high friction between mucin-chitosan layers. Adsorption of additional mucin to the mucin-chitosan layers does not change the friction since the complex mucin-chitosan and mucin-chitosan-mucin layers do not posses a strictly layered structure. Instead, they are considerably entangled thus enabling attractive interactions between oppositely charged sialic acid residues from mucin and amino groups from chitosan residing on opposing surfaces, which provides for high friction.

Page 63: DiVA portal13308/FULLTEXT01.pdf · 2008. 3. 7. · The friction properties in aqueous solution have been studied for adsorbed layers of synthetic polyelectrolytes, with varying density

53

7 Future work During the work included in this thesis I have encountered some lack in instrumental methods and some appealing results. Some of these are interesting enough to point out within this chapter of the thesis as future work. From the instrumental point of view I want to highlight the need of two things. First, a procedure for measuring the torsional detector sensitivity in an easy way in analogy with the method for normal detector sensitivity103, 104 where two different indirect spring constant calibration methods are combined to calculate the normal detector sensitivity. Second, there is a need for an extensive evaluation of the non-linearity in the cantilever / detector system when measuring normal forces with a cantilever that is deflected to large extent. Now over to some of my suggested measurements that will increase the knowledge of the lubrication behaviour of mucin, and mucin interaction with different electrolytes. Here I mean that a study of the friction and normal forces between mucin adsorbed on hydrophobic or hydrophilic surfaces, while keeping the surface roughness the same would be of interest. I expect that the nature of the surface will affect the mucin conformation, but how will this influence the lubrication? I would also like to see a systematic study of the effect of surface roughness to answer the question: How rough can the surfaces be and still be effectively lubricated by the extended mucin layer? Finally I would like to try to find a molecular design for a polymer / polyelectrolyte that will work as an excellent lubrication layer in, for example, artificial saliva, artificial tear fluid or as a coating in prosthetic joints. One may regard the results obtained for the PEO45MEMA:METAC-X polymers as promising in this respect.

Page 64: DiVA portal13308/FULLTEXT01.pdf · 2008. 3. 7. · The friction properties in aqueous solution have been studied for adsorbed layers of synthetic polyelectrolytes, with varying density

54

8 Acknowledgment I would like to express my sincere gratitude and gratitude to all the people who have helped and supported me while working with this thesis. To mention a few, though not forgetting anyone, you are all in my mind. I would especially like to mention the following: All my love to my wonderful wife Anna and my fantastic children Erik and Kalle. Thank you for inspiring and encouraging me and being the sunshine of my life and making it all worthwhile. My supervisors Dr. Andra Dėdinaitė and Prof. Per Claesson, thank you for being very inspiring, fantastic researchers, and taking me on this project. Thanks to all my co-authors, it has been a pleasure to work with you all. Actually my interest in surface chemistry was born during my employment at Akzo Nobel Surface Chemistry Asphalt Applications and I want to thank Thomas Wallin for giving me the possibility to learn about emulsions, adhesion and wetting. I also want to thank all persons that I worked together with during that time. There have been many persons that have helped me, inspired me, and contributed to a scientific encouraging atmosphere. I would like to specially thank some past and present members in the surface force group at KTH and YKI in a random order: Niklas Nordgren, Adam Feiler, Peter Weissenborn, Tobias Halthur, Rodrigo Robinson, Luis Bastardo, Eric Tyrode, Joseph Iruthayaraj, Mark Rutland, Jessica Elversson, Mikael Sundin, Brita Blomqvist, Joakim Voltaire, Ali Naderi, Andrew Fogden, Zsombor Feldötö, Johan Fröberg, Pär Wedin. My family and friends, who I have been way too bad in keeping in contact with, please forgive me. Thanks to everyone in Boogielovers DF that helped me to clear my mind while dancing, during my thesis writing. I would like to thank the Swedish research council (Vetenskapsrådet) for financial support. I also want to acknowledge Mathworks and National Instruments for producing excellent research tools, Matlab and LabView, which have been useful for everything from data acquisition and data evaluation to re-naming files on the computer. Finally, numbers of other people that have supported my work in one way or the other and even though not mentioned here by name, their contributions have been very valuable, for which I am grateful.

Page 65: DiVA portal13308/FULLTEXT01.pdf · 2008. 3. 7. · The friction properties in aqueous solution have been studied for adsorbed layers of synthetic polyelectrolytes, with varying density

55

9 References 1. Klein, J., "Forces between mica surfaces bearing layers of adsorbed polystyrene in cyclohexane".

Nature 1980, 288, 248. 2. Israelachvili, J.N. and Adams, G.E., "Measurement of forces between 2 mica surfaces in aqueous-

electrolyte solutions in range 0-100 nm". J. Chem. Soc., Faraday Trans. 1978, 74, 975. 3. Binning, G., Rohrer, H., Gerber, C., et al., "Surface studies by scanning tunneling microscopy".

Phys. Rev. Lett. 1982, 49, 57. 4. Hansma, P.K., Elings, V.B., Marti, O., et al., "Scanning tunneling microscopy and atomic force

microscopy - application to biology and technology". Science 1988, 242, 209. 5. Parker, J.L., "Surface force measurements in surfactant systems". Prog. Surf. Sci. 1994, 47, 205. 6. Mysels, K.J. and Jones, M.N., "Direct measurement of the variation of double-layer repulsion with

distance". Discuss. Faraday Soc. 1966, No. 42, 42. 7. Chaudhury, M.K. and Whitesides, G.M., "Direct measurement of interfacial interactions between

semispherical lenses and flat sheets of poly(dimethylsiloxane) and their chemical derivatives". Langmuir 1991, 7, 1013.

8. Israelachvili, J., Intermolecular and surface forces. 1991, Amsterdam: Academic Press. 9. Verwey, E.J.W. and Overbeek, J.T.G., Theory of the stability of lyophobic colloids. 1948. 216 pp. 10. Derjaguin, B. and Landau, L., "Theory of the stability of strongly charged lyophobic sols and of the

adhesion of strongly charged particles in solutions of electrolytes". Acta Physicochimica URSS 1941, 14, 633.

11. Hamaker, H.C., "The london-van der waals attraction between spherical particles". Physica 1937, 4, 1058.

12. Evans, D.F. and Wennerström, H., The colloidal domain. 2:nd ed. 1999, New York: Wiley-VCH. 13. De Gennes, P.G., "Polymers at an interface; a simplified view". Adv. Colloid Interface Sci. 1987,

27, 189. 14. Kamiyama, Y. and Israelachvili, J., "Effect of ph and salt on the adsorption and interactions of an

amphoteric polyelectrolyte". Macromolecules 1992, 25, 5081. 15. Scheutjens, J.M.H.M. and Fleer, G.J., "Interaction between two adsorbed polymer layers".

Macromolecules 1985, 18, 1882. 16. Dahlgren, M.A.G., Waltermo, Å., Blomberg, E., et al., "Salt effects on the interaction between

adsorbed cationic polyelectrolyte layers - theory and experiment". J. Phys. Chem. 1993, 97, 11769. 17. Johansson, A., "Paper friction - influence of measurement conditions ". Tappi J. 1998, 81, 175. 18. Lee, S., Müller, M., Ratoi-Salagean, M., et al., "Boundary lubrication of oxide surfaces by poly(l-

lysine)-g-poly(ethylene glycol) (pll-g-peg) in aqueous media". Tribol. Lett. 2003, 15, 231. 19. Walker, P.S., Dowson, D., Longfield, M.D., et al., "Rheological behaviour of human joints". Rheol.

Acta 1969, 8, 234. 20. Moro, T., Takatori, Y., Ishihara, K., et al., "Surface grafting of artificial joints with a biocompatible

polymer for preventing periprosthetic osteolysis". Nat. Mater. 2004, 3, 829. 21. Krishnan, R., Caligaris, M., Mauck, R.L., et al., "Removal of the superficial zone of bovine

articular cartilage does not increase its frictional coefficient". Osteoarthritis and Cartilage 2004, 12, 947.

22. Jay, G.D., Harris, D.A., and Cha, C.J., "Boundary lubrication by lubricin is mediated by o-linked beta(1-3)gal-galnac oligosaccharides". Glycoconjugate J. 2001, 18, 807.

23. Gourdon, D., Lin, Q., Oroudjev, E., et al., "Adhesion and stable low friction provided by a subnanometer-thick monolayer of a natural polysaccharide". Langmuir 2008, 24, 1534.

24. Zappone, B., Ruths, M., Greene, G.W., et al., "Adsorption, lubrication, and wear of lubricin on model surfaces: Polymer brush-like behavior of a glycoprotein". Biophys. J. 2007, 92, 1693.

25. Benz, M., Chen, N.H., and Israelachvili, J., "Lubrication and wear properties of grafted polyelectrolytes, hyaluronan and hylan, measured in the surface forces apparatus". J. Biomed. Mater. Res. Part A 2004, 71A, 6.

26. Kampf, N., Raviv, U., and Klein, J., "Normal and shear forces between adsorbed and gelled layers of chitosan, a naturally occurring cationic polyelectrolyte". Macromolecules 2004, 37, 1134.

27. Raviv, U., Frey, J., Sak, R., et al., "Properties and interactions of physigrafted end-functionalized poly(ethylene glycol) layers". Langmuir 2002, 18, 7482.

28. Raviv, U. and Klein, J., "Fluidity of bound hydration layers". Science 2002, 297, 1540.

Page 66: DiVA portal13308/FULLTEXT01.pdf · 2008. 3. 7. · The friction properties in aqueous solution have been studied for adsorbed layers of synthetic polyelectrolytes, with varying density

56

29. Urbakh, M., Klafter, J., Gourdon, D., et al., "The nonlinear nature of friction". Nature 2004, 430, 525.

30. Raviv, U., Giasson, S., Kampf, N., et al., "Lubrication by charged polymers". Nature 2003, 425, 163.

31. Ruths, M., Berman, A.D., and Israelachvili, J.N., Surface forces and nanorheology of molecularly thin films, in Springer handbook of nanotechnology, B. Bhushan, Editor. 2004, Springer-Verlag: Berlin Heidelberg. p. 543.

32. Klein, J., "Shear, friction, and lubrication forces between polymer-bearing surfaces". Annu. Rev. Mater. Sci. 1996, 26, 581.

33. Tsukruk, V.V., "Molecular lubricants and glues for micro- and nanodevices". Adv. Mater. 2001, 13, 95.

34. Carnally, S., Barrow, K., Alexander, M.R., et al., "Ultra-resolution imaging of a self-assembling biomolecular system using robust carbon nanotube afm probes". Langmuir 2007, 23, 3906.

35. Yi, I., Sugimoto, Y., Nishi, R., et al., "Atomic structure of ge clusters on si(111)-(7x7) by non-contact afm". Nanotechnology 2007, 18.

36. Milhiet, P.E., Gubellini, F., Berquand, A., et al., "High-resolution afm of membrane proteins directly incorporated at high density in planar lipid bilayer". Biophys. J. 2006, 91, 3268.

37. Mizuno, H., Kjellin, M., Nordgren, N., et al., "Friction measurement between polyester fibres using the fibre probe spm". Aust. J. Chem. 2006, 59, 390.

38. Stiernstedt, J., Nordgren, N., Wagberg, L., et al., "Friction and forces between cellulose model surfaces: A comparison". J. Colloid Interface Sci. 2006, 303, 117.

39. Lee, C.K., Wang, Y.M., Huang, L.S., et al., "Atomic force microscopy: Determination of unbinding force, off rate and energy barrier for protein-ligand interaction". Micron 2007, 38, 446.

40. Clement, N., Tonneau, D., Dallaporta, H., et al., "Electronic transport properties of single-crystal silicon nanowires fabricated using an atomic force microscope". Physica E 2002, 13, 999.

41. Yan, Y.D., Sun, T., and Dong, S., "Study on effects of tip geometry on afm nanoscratching tests". Wear 2007, 262, 477.

42. Rubio-Sierra, F.J., Heckl, W.M., and Stark, R.W., "Nanomanipulation by atomic force microscopy". Adv. Eng. Mater. 2005, 7, 193.

43. Xu, S. and Liu, G.Y., "Nanometer-scale fabrication by simultaneous nanoshaving and molecular self-assembly". Langmuir 1997, 13, 127.

44. Salaita, K., Wang, Y.H., and Mirkin, C.A., "Applications of dip-pen nanolithography". Nat. Nanotechnol. 2007, 2, 145.

45. Ducker, W.A., Senden, T.J., and Pashley, R.M., "Measurement of forces in liquids using a force microscope". Langmuir 1992, 8, 1831.

46. Ducker, W.A., Senden, T.J., and Pashley, R.M., "Direct measurement of colloidal forces using an atomic force microscope". Nature 1991, 353, 239.

47. Senden, T.J., "Force microscopy and surface interactions". Curr. Opin. Colloid Interface Sci. 2001, 6, 95.

48. Butt, H.-J., Cappella, B., and Kappl, M., "Force measurements with the atomic force microscope: Technique, interpretation and applications". Surf. Sci. Rep. 2005, 59, 1.

49. Mate, C.M., McClelland, G.M., Erlandsson, R., et al., "Atomic-scale friction of a tungsten tip on a graphite surface". Phys. Rev. Lett. 1987, 59, 1942.

50. Cleveland, J.P., Manne, S., Bocek, D., et al., "A nondestructive method for determining the spring constant of cantilevers for scanning force microscopy". Rev. Sci. Instrum. 1993, 64, 403.

51. Sader, J.E., Chon, J.W.M., and Mulvaney, P., "Calibration of rectangular atomic force microscope cantilevers". Rev. Sci. Instrum. 1999, 70, 3967.

52. Sader, J.E., "Frequency response of cantilever beams immersed in viscous fluids with applications to the atomic force microscope". J. Appl. Phys. 1998, 84, 64.

53. Green, C.P., Lioe, H., Cleveland, J.P., et al., "Normal and torsional spring constants of atomic force microscope cantilevers". Rev. Sci. Instrum. 2004, 75, 1988.

54. Green, C.P. and Sader, J.E., "Torsional frequency respons of cantilever beams immersed in viscous fluids with applications to the atomic force microscope". J. Appl. Phys. 2002, 92, 6262.

55. Bogdanovic, G., Meurk, A., and Rutland, M.W., "Tip friction - torsional spring constant determination". Colloid Surf. B-Biointerfaces 2000, 19, 397.

56. Feiler, A., Attard, P., and Larson, I., "Calibration of the torsinal spring constant and the lateral photodiode response of frictional force microscopes". Rev. Sci. Instrum. 2000, 71, 2746.

Page 67: DiVA portal13308/FULLTEXT01.pdf · 2008. 3. 7. · The friction properties in aqueous solution have been studied for adsorbed layers of synthetic polyelectrolytes, with varying density

57

57. Cain, R.G., Reitsma, M.G., Biggs, S., et al., "Quantitative comparison of three calibration techniques for the lateral force microscope". Rev. Sci. Instrum. 2001, 72, 3304.

58. Cain, R.G., Biggs, S., and Page, N.W., "Force calibration in lateral force microscopy". J. Colloid Interface Sci. 2000, 227, 55.

59. Liu, Y.H., Evans, D.F., Song, Q., et al., "Structure and frictional properties of self-assembled surfactant monolayers". Langmuir 1996, 12, 1235.

60. Rutland, M.W., Tyrrell, J.W.G., and Attard, P., "Analysis of atomic force microscopy data for deformable materials". J. Adhes. Sci. Technol. 2004, 18, 1199.

61. Cannara, R.J., Eglin, M., and Carpick, R.W., "Lateral force calibration in atomic force microscopy: A new lateral force calibration method and general guidelines for optimization". Rev. Sci. Instrum. 2006, 77.

62. Ecke, S., Raiteri, R., Bonaccurso, E., et al., "Measuring normal and friction forces acting on individual fine particles". Rev. Sci. Instrum. 2001, 72, 4164.

63. Rodahl, M., Hook, F., Fredriksson, C., et al., "Simultaneous frequency and dissipation factor qcm measurements of biomolecular adsorption and cell adhesion". Faraday Discuss. 1997, 229.

64. Rodahl, M. and Kasemo, B., "A simple setup to simultaneously measure the resonant frequency and the absolute dissipation factor of a quartz crystal microbalance". Rev. Sci. Instrum. 1996, 67, 3238.

65. Rodahl, M., Hook, F., Krozer, A., et al., "Quartz-crystal microbalance setup for frequency and q-factor measurements in gaseous and liquid environments". Rev. Sci. Instrum. 1995, 66, 3924.

66. Borovsky, B., Mason, B.L., and Krim, J., "Scanning tunneling microscope measurements of the amplitude of vibration of a quartz crystal oscillator". J. Appl. Phys. 2000, 88, 4017.

67. Kanazawa, K.K., "Mechanical behaviour of films on the quartz microbalance". Faraday Discuss. 1997, 77.

68. Kanazawa, K.K. and Gordon, J.G., "The oscillation frequency of a quartz resonator in contact with a liquid ". Anal. Chim. Acta 1985, 175, 99.

69. Rodahl, M. and Kasemo, B., "Frequency and dissipation-factor responses to localized liquid deposits on a qcm electrode". Sensors Actuat. B 1996, 37, 111.

70. Kanazawa, K.K. and Gordon, J.G., "Frequency of a quartz microbalance in contact with liquid". Anal. Chem. 1985, 57, 1770.

71. Reed, C.D., Kanazawa, K.K., and Kaufman, J.H., "Physical description of a viscoelastically loaded at-cut quartz resonator". J. Appl. Phys. 1990, 68, 1993.

72. Sauerbrey, G., "Verwendung von schwingquarzen zur wagung dunner schichten und zur mikrowagung". Z. Phys. 1959, 155, 206.

73. Brewer, S.H., Glomm, W.R., Johnson, M.C., et al., "Probing bsa binding to citrate-coated gold nanoparticles and surfaces". Langmuir 2005, 21, 9303.

74. Höök, F., Kasemo, B., Nylander, T., et al., "Variations in coupled water, viscoelastic properties, and film thickness of a mefp-1 protein film during adsorption and cross-linking: A quartz crystal microbalance with dissipation monitoring, ellipsometry, and surface plasmon resonance study". Anal. Chem. 2001, 73, 5796.

75. Johannsmann, D., Mathauer, K., Wegner, G., et al., "Viscoelastic properties of thin-films probed with a quartz-crystal resonator". Phys. Rev. B 1992, 46, 7808.

76. Ubbelohde, L., Viscometer, 1936, US Patent, 2048305. 77. Ubbelohde, L., "Viscosity of illuminating oils, and an apparatus for its determination". Petroleum

1909, 4, 861. 78. Pashley, R.M., "Hydration forces between mica surfaces in aqueous electrolyte solutions". J.

Colloid Interface Sci. 1981, 80, 153. 79. Yaminsky, V.V., Ninham, B.W., and Pashley, R.M., "Interaction between surfaces of fused silica

in water. Evidence of cold fusion and effects of cold plasma treatment". Langmuir 1998, 14, 3223. 80. Vigil, G., Xu, Z.H., Steinberg, S., et al., "Interactions of silica surfaces". J. Colloid Interface Sci.

1994, 165, 367. 81. Taran, E., Donose, B.C., Vakarelski, I.U., et al., "Ph dependence of friction forces between silica

surfaces in solutions". J. Colloid Interface Sci. 2006, 297, 199. 82. Bowen, W.R., Hilal, N., Lovitt, R.W., et al., "An atomic force microscopy study of the adhesion of

a silica sphere to a silica surface - effects of surface cleaning". Colloid Surf. A-Physicochem. Eng. Asp. 1999, 157, 117.

83. Nuzzo, R.G. and Allara, D.L., "Adsorption of bifunctional organic disulfides on gold surfaces". J. Am. Chem. Soc. 1983, 105, 4481.

Page 68: DiVA portal13308/FULLTEXT01.pdf · 2008. 3. 7. · The friction properties in aqueous solution have been studied for adsorbed layers of synthetic polyelectrolytes, with varying density

58

84. Porter, M.D., Bright, T.B., Allara, D.L., et al., "Spontaneously organized molecular assemblies. 4. Structural characterization of n-alkyl thiol monolayers on gold by optical ellipsometry, infrared spectroscopy, and electrochemistry". J. Am. Chem. Soc. 1987, 109, 3559.

85. Bain, C.D. and Whitesides, G.M., "Molecular-level control over surface order in self-assembled monolayer films of thiols on gold". Science 1988, 240, 62.

86. Raiteri, R., Preuss, M., Grattarola, M., et al., "Preliminary results on the electrostatic double-layer force between two surfaces with high surface potentials". Colloid Surf. A-Physicochem. Eng. Asp. 1998, 136, 191.

87. Gunnarsson, G., Joensson, B., and Wennerstroem, H., "Surfactant association into micelles. An electrostatic approach". J. Phys. Chem. 1980, 84, 3114.

88. Newbery, J.E., "Variation of the critical micelle concentration of sodium dodecyl-sulfate with ionic-strength monitored by selective-ion membrane electrodes". Colloid Polym. Sci. 1979, 257, 773.

89. Bastardo, L., Claesson, P., and Brown, W., "Interactions between mucin and alkyl sodium sulfates in solution. A light scattering study". Langmuir 2002, 18, 3848.

90. Abuin, E.B. and Scaiano, J.C., "Exploratory-study of the effect of poly-electrolyte surfactant aggregates on photochemical behavior". J. Am. Chem. Soc. 1984, 106, 6274.

91. Koopal, L.K., Goloub, T., de Keizer, A., et al., "The effect of cationic surfactants on wetting, colloid stability and flotation of silica". Colloid Surf., A 1999, 151, 15.

92. Sugio, S., Kashima, A., Mochizuki, S., et al., "Crystal structure of human serum albumin at 2.5 angstrom resolution". Protein Eng. 1999, 12, 439.

93. Dickerson, R.E., Takano, T., Eisenberg, D., et al., "Ferricytochrome c. I. General features of the horse and bonito proteins at 2.8 ang. Resolution". J. Biol. Chem. 1971, 246, 1511.

94. Lehninger, A.L., Nelson, D.L., and Cox, M.M., Principles of biochemistry. Second ed. 1993, New York: Worth Publishers.

95. Vojtechovsky, J., Chu, K., Berendzen, J., et al., "Crystal structures of myoglobin-ligand complexes at near-atomic resolution". Biophys. J. 1999, 77, 2153.

96. Strous, G.J. and Dekker, J., "Mucin-type glycoproteins". Crit. Rev. Biochem. Mol. Biol. 1992, 27, 57.

97. Lee, S., Muller, M., Rezwan, K., et al., "Porcine gastric mucin (pgm) at the water/poly(dimethylsiloxane) (pdms) interface: Influence of ph and ionic strength on its conformation, adsorption, and aqueous lubrication properties". Langmuir 2005, 21, 8344.

98. Scheinthal, B.M. and Bettelheim, F.A., "Multiple forms of sialic acids". Carbohydr. Res. 1968, 6, 257.

99. Bastardo, L.A., Self assembly of surfactants and polyelectrolytes in solutions and at interfaces. PhD thesis, KTH, Royal Institute of Thechnology, 2005. ISBN: 91-7178-127-7.

100. Iruthayaraj, J., Poptoshev, E., Vareikis, A., et al., "Adsorption of low charge density polyelectrolyte containing poly(ethylene oxide) side chains on silica: Effects of ionic strength and ph". Macromolecules 2005, 38, 6152.

101. Naderi, A., Iruthayaraj, J., Vareikis, A., et al., "Surface properties of bottle-brush polyelectrolytes on mica: Effect of side chain and charge density". Langmuir 2007, 23, 12222.

102. Olanya, G., Iruthayaraj, J., Vareikis, A., et al., "Adsorption properties of bottle-brush polyelectrolytes on silica". Submitted to Langmuir (la-2007-03739v) 2007.

103. Attard, P., Pettersson, T., and Rutland, M.W., "Thermal calibration of photodiode sensitivity for atomic force microscopy". Rev. Sci. Instrum. 2006, 77, 116110.

104. Higgins, M.J., Proksch, R., Sader, J.E., et al., "Noninvasive determination of optical lever sensitivity in atomic force microscopy". Rev. Sci. Instrum. 2006, 77, 013701.

105. Tocha, E., Song, J., Schonherr, H., et al., "Calibration of friction force signals in atomic force microscopy in liquid media". Langmuir 2007, 23, 7078.

106. Dedinaite, A., Claesson, P.M., and Bergstrom, M., "Polyelectrolyte-surfactant layers: Adsorption of preformed aggregates versus adsorption of surfactant to preadsorbed polyelectrolyte". Langmuir 2000, 16, 5257.

107. Dedinaite, A. and Ernstsson, M., "Chitosan-sds interactions at a solid-liquid interface: Effects of surfactant concentration and ionic strength". J. Phys. Chem. B 2003, 107, 8181.

108. Kjellin, U.R.M., Claesson, P.M., and Audebert, R., "Interactions between adsorbed layers of a low charge density cationic polyelectrolyte on mica in the absence and presence of anionic surfactant". J. Colloid Interface Sci. 1997, 190, 476.

Page 69: DiVA portal13308/FULLTEXT01.pdf · 2008. 3. 7. · The friction properties in aqueous solution have been studied for adsorbed layers of synthetic polyelectrolytes, with varying density

59

109. Glomm, W.R., Moses, S.J., Brennaman, M.K., et al., "Detection of adsorption of ru(ii) and os(ii) polypyridyl complexes on gold and silver nanoparticles by single-photon counting emission measurements". J. Phys. Chem. B 2005, 109, 804.

110. Vigil, R.D. and Robert, M.Z., "Random sequential adsorption of unoriented rectangles onto a plane". J. Chem. Phys. 1989, 91, 2599.

111. Bjerneld, E.J., Földes-Papp, Z., Käll, M., et al., "Single-molecule surface-enhanced raman and fluorescence correlation spectroscopy of horseradish peroxidase". J. Phys. Chem. B 2002, 106, 1213.

112. Creeth, J.M., Bhaskar, K.R., and Horton, J.R., "The separation and characterization of bronchical glycoproteins by density-gradient methods". Biophys. J. 1977, 167, 557.

113. Lundin, M., Macakova, L., Dedinaite, A., et al. Langmuir 2008, in Press (la-2007-02653m.R1). 114. Fleer, G.J., Cohen Stuart, M.A., Scheutjens, J.M.H.M., et al., Polymers at interfaces. 1993,

London: Chapman & Hall. 115. Dedinaite, A. and Bastardo, L., "Interactions between mucin and surfactants at solid-liquid

interfaces". Langmuir 2002, 18, 9383. 116. Nordgren, N., Eklöf, J., Zhou, Q., et al., "Top-down grafting of xyloglucan to gold monitored by

qcm-d and afm: Enzymatic activity and interactions with cellulose". Biomacromolecules 2008, in press, DOI: 10.1021/bm701214e (Web Release Date: 09-Feb-2008).

117. Feiler, A.A., Bergstrom, L., and Rutland, M.W., "Superlubricity using repulsive van der waals forces". Langmuir 2008, in press, DOI:/10.1021/la7036907 (Web Release Date: 16-Feb-2008).

Page 70: DiVA portal13308/FULLTEXT01.pdf · 2008. 3. 7. · The friction properties in aqueous solution have been studied for adsorbed layers of synthetic polyelectrolytes, with varying density
Page 71: DiVA portal13308/FULLTEXT01.pdf · 2008. 3. 7. · The friction properties in aqueous solution have been studied for adsorbed layers of synthetic polyelectrolytes, with varying density
Page 72: DiVA portal13308/FULLTEXT01.pdf · 2008. 3. 7. · The friction properties in aqueous solution have been studied for adsorbed layers of synthetic polyelectrolytes, with varying density

����������� �����������������������

������������������

������ �!�