<<

KTH Chemical Science and

Nanotribology, Surface Interactions and Characterization: An AFM Study

Rubén Álvarez-Asencio

Doctoral Thesis at the KTH Royal Institute of Technology

Stockholm 2014 Akademisk avhandling som med tillstånd av Kungliga Tekniska Högskolan framlägges till offentling granskning för avläggande av teknologie doctorsexamen den 13 juni 2014 kl 10 i hörstal F3, KTH Lindstedtsvägen 26, Stockholm. Avhandling presenteras på engelska.

Nanotribology, Surface Interactions and Characterization: An AFM Study

Rubén Álvarez-Asencio ( [email protected] )

Doctoral Thesis

TRITA-CHE Report 2014:13

ISSN 1654-1081

ISBN 978-91-7595-102-7

KTH Royal Institute of Technology

School of Chemical Science and Engineering

Surface and Corrosion Science

Drottning Kristinas väg 51

SE-100 44 Stockholm

Sweden

ii

Denna avhandling är skyddad enligt upphovsrättslagen. Alla rättigheter förbehålles.

Copyright © 2014 Rubén Álvarez-Asencio. All rights reserved. No part of this thesis may be reproduced by any means without permission from the author.

The following items are reprinted with permission:

Paper I: Copyright © American Chemical Society

Paper II: Copyright © AIP Publishing LLC

Paper III: Copyright © Springer

Printed at US-AB, Stockholm 2014

iii

“God grant me the serenity to accept the things I cannot change; courage to change the things I can; and wisdom to know the difference”.

Reinhold Niebuhr

iv

To Antonio, Ángeles, Antonio-José and Alejandro

v

Abstract

When two surfaces achieve contact, then contact phenomena such as , and can occur, which are of great interest in many disciplines, including , physical , material chemistry, and life and health sciences. These phenomena are largely determined by the nature and magnitude of the surface such as van der Waals, capillary and hydration forces. Moreover these forces are length-dependent, and therefore when the system scales down, their contribution scales up, dominating the interaction between the surfaces.

A goal of my PhD work was to investigate fundamental contact phenomena in terms of the surface forces that regulate their properties. The primary tool applied in this PhD thesis work has been the atomic microscopy (AFM), which (with all of its sub-techniques) offers the possibility to study such forces with high resolution virtually between all types of materials and intervening media. Therefore, in this work it was possible to study the long ranged interactions presented in air between different industrially relevant materials and how these interactions are shielded when the systems are immersed in an ionic .

Also investigated was the influence of microstructure on the tribological properties of metal alloys, where their good tribological properties were related with the vanadium and nitrogen contents for a FeCrVN tool alloy and with the chromium content for a biomedical CoCrMo alloy. Moreover, the effect of the intervening media can significantly affect the surface properties, and when the biomedical CoCrMo alloy was immersed in phosphate buffer saline solution (PBS), repulsive hydration forces decreased the friction coefficient and contact adhesion. On the other hand, with the immersion of the FeCrVN tool alloy in the NaCl solution, small particles displaying low adhesion were generated in specific regions on the

vi surface with low chromium content. These particles are assumed to be related to a prepitting corrosion event in the tool alloy.

The mechanical properties of stratum corneum (SC), which is the outermost layer of the skin, were also studied in this work. The SC presents a highly elastic, but stiff surface where the mechanical properties depend on the nanoscale. A novel probe has been designed with a single hair fibre in order to understand how the skin deforms locally in response to the interaction with such a fibre probe. This study revealed that is mostly the lateral scale of the deformation which determines the mechanical properties of the SC.

Finally, important achievements in this work are the developments of two new techniques - tribological property mapping and the Hybrid method for torsional spring constant evaluation. Tribological property mapping is an AFM technique that provides friction coefficient and contact adhesion maps with information attributed to the surface microstructure. The Hybrid method is an approach that was originally required to obtain the torsional spring constants for rigid beam shaped , which could not be previously determined from their power torsional thermal spectra (conventional method). However, the applicability is shown to be general and this simple method can be used to obtain torsional spring constants for any type of beam shape .

vii

Sammanfattning

När två ytor kommer i kontakt, sker det en interaktion dem emellan varvid grundläggande kontaktfenomen som adhesion, friktion och förslitning uppkommer. Detta är av stort intresse inom flera discipliner till exempel fysik, fysiskalisk kemi, materialkemi och medicinsk forskning. Dessa fenomen bestäms till stor del av naturen och omfattningen hos de krafter som verkar på ytorna såsom van der Waals-, kapillär- och hydrationskrafter. Dessutom är dessa krafter beroende av avståndsberoende så för system i liten skala blir deras vikt mer betydande och kommer att dominera interaktionen mellan ytorna.

Målet med denna avhandlig har varit att undersöka grundläggande kontaktfenomen i termer av de ytkrafter som styr deras unika egenskaper. Det huvudsakliga verktyget som har tillämpats är atomkraftsmikroskopet (AFM) som (med all dess sub-tekniker) erbjuder möjligheten att kunna studera dessa krafter med hög upplösning mellan närapå alla typer av material och mellanliggande media. Därför har det i detta arbete varit möjligt att studera de brett varierande interaktionerna som förekommer i luft mellan olika industriellt relevanta material och hur dessa interaktioner är skärmade när systemen befinner sig i en jonvätska.

En annan studie var inverkan av mikrostruktur på de tribologiska egenskaperna hos metallegeringar. De goda tribologiska egenskaperna fanns vara relaterade till vanadin- och krominnehållet för FeCrVN verktygslegeringen och med kromhalten för den biomedicinska CoCrMo legeringen . Vidare kan effekten av mellanliggande media märkbart påverka ytans egenskaper och när den biomedicinska CoCrMo legeringen var dränkt i fosfat buffertsaltlösning (FBS), reducerades friktionskoefficienterna och kontaktvidhäftningen av repulsiva hydrationskrafter. När FeCrVN

viii verktygslegeringen var dränkt i en NaC1 lösning, alstrades små partiklar inom specifika områden på ytan. Dessa partiklar uppvisade en låg vidhäftning och antas vara relaterade till en tidigare gropfrätning i verktygslegeringen.

Vidare studerades även de mekaniska egenskaperna av stratum corneum (SC), vilket är det yttersta lagret av huden. SC uppvisar en hög elasticitet men med stela ytor där de mekaniska egenskaperna är beroende av nanonivån. En prob tillverkades med ett enstaka hårfiber bundet i änden av kantilevern. Fibern slipades efteråt med hjälp av en fokuserad jonstråle. Denna nya typ av prob användes för att kunna förstå hur huden deformeras lokalt som gensvar på interaktionen med en sådan fiberprob. Studien avslöjade att det till största del är den laterala skalan av deformationen som avgör de mekaniska egenskaperna hos SC.

Slutligen, viktiga delar av detta arbete är utvecklingen av ”tribological property mapping” och hybridmetoden.”Tribological property mapping” är en AFM teknik som ger bilder av friktionskoefficienter och kontaktvidhäftning vilket ger information om mikrostrukturen på ytan. Hybridmetoden däremot är en uppskattning som från början var designad för att erhålla vridmomentsfjäderkonstanterna för styva, rektangulära kantilevrar, vilka tidigare inte kunde bestämmas utifrån deras termiska kraftspektra. Emellertid visar det sig att tillämpningen är allomfattande, och denna metod kan användas för att erhålla vridmomentsfjäderkonstanter för alla typer av rektangulära kantilevrar.

ix

List of Papers

Included Papers

I. Nanotribology: Stiction Suppression and Surface Induced Shear Thinning Álvarez-Asencio, R., Cranston, E. D., Atkin, R., Rutland, M. W. Langmuir, 2012 , 28, 9967-9976

II. Determination of Torsional Spring Constant of Cantilevers: Combining Normal Spring Constant and Classical Beam Theory Álvarez-Asencio, R., Thormann, E., Rutland, M. W. Review of Scientific Instruments, 2013 , 84, 096102

III. Tribological Property Mapping: Local Variation in Friction Coefficient and Adhesion Álvarez-Asencio, R., Pan, J., Thormann, E., Rutland, M. W. Letters, 2013 , 50, 387-395

IV. Nanotribology and Microstructure of a CoCrMo Alloy: A Tribological Properties Mapping Study Álvarez-Asencio, R., Bettini, E., Pan, J., Leygraf, C., Rutland, M. W. Manuscript

V. Role of Microstructure on Pitting Corrosion Initiation of an Experimental Tool Alloy: A PeakForce QNM Atomic Force Microscopy Study Álvarez-Asencio, R., Sababi, M., Pan, J., Ejnermark, S., Rutland, M. W., L. Ekman Corrosion Science, Submitted, 2014

x

VI. Nanomechanical Properties of Human Skin Studied by AFM and a Novel Hair Indenter Álvarez-Asencio, R., Wallqvist, V., Kjellin, M., Luengo, G., Rutland, M. W., Nordgren, N. Manuscript

The author contribution to the included papers:

I. Major part of the experimental work and manuscript preparation II. All the experimental work and major part of manuscript preparation III. All the experimental work and part of manuscript preparation IV. Part of the experimental work and major part of manuscript preparation V. Part of the experimental work and major part of manuscript preparation VI. Part of experimental work and part of manuscript preparation

Other Papers not Included in this Thesis

I. Monolayer Study by VSFS: In Situ Response to Compression and Shear in a Contact Ghalgaoui, A., Shimizu, R., Hosseinpour, S., Álvarez-Asencio, R., McKee, C., Johnson, M., Rutland, M., W. Langmuir, 2014 , 30, 3075-3085

xi

Summary of Papers

The aim of Paper I was to investigate the friction and adhesion between relevant pairs of materials (silica, alumina, polytetrafluoroethylene) and interpret them with regard to the longer ranged interactions between the surfaces. In ambient air the interactions are controlled by attractive van der Waals and strong adhesion, leading to significant frictional forces. In ethylammonium nitrate (EAN) the van der Waals attraction is shielded, and the adhesive/attractive interactions which lead to stiction are almost eliminated, reducing frictional forces 10-fold at high applied load. The friction coefficients in EAN were also significantly reduced and the variation between systems was correlated with . The hydrodynamic forces between the materials surfaces have been also investigated in EAN. A linear increase of these forces with velocity is observed, reducing the probability of stiction. An unexpected result was found when the extracted from the data was almost 3 times lower than the EAN bulk viscosity, indicating a surface ordering effect.

In Paper II , a calibration technique was developed for the calculation of torsional spring constants for AFM cantilevers by combining the normal spring constant measurement with plate/beam theory. Originally, the aim of this method was the determination of torsional spring constants for stiff cantilevers where it is not possible to find the necessary torsional frequency peak because of its very low signal/noise ratio. However, its applicability is more general and it can in fact be used to obtain the torsional spring constant in a simple manner for any beam shaped cantilever.

Tribological property mapping (TPM) is a new imaging technique developed in Paper III that generates friction coefficient and adhesion

xii maps. This technique is based on the combination of a series of lateral atomic force microscopy (LAFM) images obtained as a function of load, which are tiled and pixelwise fitted to a modified Amonton’s law, obtaining both friction coefficient and an adhesion value for each pixel. This imaging technique eliminates the uncertainty of the friction contrast in conventional LAFM imaging since the friction coefficient and adhesion are independent of the load applied during the mapping. As an example of the application of the technique, a heterogeneous commercial powder metallurgical tool alloy has been scanned using a silicon tip, immersed both in air and in tetradecane, in order to obtain friction coefficient and adhesion maps. These data provide unique information related to the heterogeneous microstructure of the alloy as well as an enhanced understanding of the tribological properties of the material.

TPM was used in Paper IV to study the local tribological properties of a biomedical CoCrMo alloy in a phosphate buffer solution (PBS) that mimics the saline conditions of the body. The biomedical alloy turned out to be stable during the measurement in PBS and displayed low friction coefficient and contact adhesion values. These low values were attributed to the chromium oxide surface layer and the hydration forces which originated in PBS, even led to positive values (repulsive) in the contact adhesion map.

In Paper V , the adhesion properties of a FeCrVN tool alloy in and sodium chloride have been studied by PeakForce ® QNM in order to understand the influence of the microstructure on adhesion and corrosion initiation of the tool alloy. It turned out that the adhesion of the alloy is strongly influenced by the vanadium and nitrogen contents in water. However when the alloy is immersed in sodium chloride, other factors affect the system and small particles are formed on the surface with mostly

xiii very low adhesion. These particles are assumed to be related to prepitting events that may lead to passivity breakdown of the alloy.

Nanomechanical properties of the outer most layer of the skin (stratum corneum, SC) have been studied by atomic force microscopy in Paper VI . Nanomechanical mapping reveals that the Young’s modulus of the SC varies somewhat over the surface with a mean value of 0.39 GPa, and the force indentation measurements show that the SC is deformed permanently at high applied loads (above 4 µN). In Paper VI a novel probe has also been designed using a single hair fiber and sharpened with a focused ion beam. The force indentation measurements performed with this probe on SC reveals that the lateral scale of the deformation determines the Young’s modulus of the elastic outermost layer of the skin.

xiv

Table of Contents

Abstract ...... vi Sammanfattning ...... viii List of Papers ...... x Included Papers ...... x Other Papers not Included in this Thesis ...... xi Summary of Papers ...... xii Table of Contents ...... xv Symbols ...... xvii 1 Introduction ...... 1 1.1 Surface Forces ...... 1 1.1.1 van der Waals Forces ...... 1 1.1.2 Capillary Forces ...... 2 1.1.3 Solvation Forces ...... 3 1.1.4 Hydration Forces ...... 4 1.1.5 Hydrodynamic Forces ...... 5 1.2 Adhesion ...... 5 1.3 Tribology ...... 6 1.3.1 Nanotribology ...... 9 2 Experimental ...... 11 2.1 Atomic Force Microscope ...... 11 2.1.1 Imaging ...... 12 2.1.2 Force ...... 12 2.1.3 Friction ...... 14 2.1.4 PeakForce ® QNM ...... 16 2.1.5 Colloidal Probe ...... 17 2.1.6 Determination of Spring Constants ...... 18

xv

2.1.7 Focused Ion Beam (FIB)...... 20 3 Materials and Fluids ...... 22 3.1 Cantilevers and Probes ...... 22 3.2 Substrates ...... 23 3.3 Liquid Media ...... 25 4. Summary of Key Results ...... 27 4.1 Suppression of Surface Interactions by an Ionic Liquid ...... 27 4.2 A Hybrid Route for the Determination of Torsional Spring Constant for AFM Cantilevers ...... 32 4.3 A New Technique to Characterize Heterogeneous Surfaces: Tribological Property Mapping ...... 36 4.4 Biomedical CoCrMo alloy: A Tribological Properties Mapping Study ...... 43 4.5 Surface Study and Corrosion Initiation of an Experimental FeCrVN Tool Alloy ...... 48 4.6 Nanomechanical Properties of Human Skin ...... 56 4.7 A Novel AFM Probe: The Single Hair Fibre Probe ...... 60 5. Conclusions ...... 66 6. Acknowledgment ...... 69

7. References ...... 70

xvi

Symbols

FLoad Applied Load

Vd Average Lateral Output Voltage k Boltzmann Constant

Sc Critical Shear Contact Stress rk Curvature of the Meniscus λ Decay Length ρ Density D Distance, Surface Separation heff Effective Height

Reff Effective Radius W Energy per Unit Area µ Friction Coefficient

FFriction Frictional Force A Hamaker Constant

CH Hydration Constant

FHydration Hydration Force

FHydrodynamic Hydrodynamic Force Imaginary Component of the Hydrodynamic Function for

Normal Vibrations ( ) Imaginary Component of the Hydrodynamic Function for

Torsional Vibrations ( ) δ Lateral Deflection Sensitivity L Length

Vm Molar Volume F Normal Force

Qz Normal Quality Factor

xvii fz Normal Resonance Frequency ν Poisson’s Ratio R Radius of the Sphere A Real Contact Area E* Reduced Young’s Modulus p/p 0 Relative Vapour Pressure vR Relative Velocity G Shear Modulus γ Surface Tension T Temperature t Thickness

Qϕ Torsional Quality Factor fϕ Torsional Resonance Frequency kϕ Torsional Spring Constant

FvdW * Vn Vertical Deflection in Newton

Vn Vertical Deflection in Volts η Viscosity of the Fluid w Width E Young’s Modulus z Z – Piezo Position

xviii

1 Introduction

1.1 Surface Forces

1.1.1 van der Waals Forces van der Waals forces arise from the interaction between electromagnetic fields generated from the surface of any material. They are formed by the sum of three contributions: the orientation force, the induction force and the dispersion or London force.[1, 2]

Dispersion forces usually dominate over orientation and induction forces. Besides, they are always present, playing an important role in a multitude of fundamental phenomena, such as, adhesion and surface tension. Therefore, they are considered the most important contribution to the van der Waals interactions.

The determination of the van der Waals force between molecular pairs is straight forward by adding the orientation, induction and dispersion contributions,[1] but for bulk materials the situation gets more complicated. Hamaker[3] in 1937 was able to estimate the overall van der Waals interaction between two macroscopic by summation of all the molecular pair interactions between the two bodies. His approximation was based on the interaction of a sphere against a flat surface, according to Eq. 1:

(1) = where FvdW corresponds to van der Waals force, D is the distance between the surfaces, R is the radius of the sphere and A the Hamaker constant, where typical values are around 10 -19 J. The addition method suggested by

1

Hamaker for the calculation of A does not consider the influence of neighbouring in the interaction between the molecular pair. Moreover, the Hamaker approach cannot be easily applied to two bodies with an intervening liquid. Therefore, Lifshitz in 1956[1] proposed for the calculation of the Hamaker constant a new theory that ignores the atomic structure and treats the force as a continuous medium. This approximation is able to calculate the Hamaker constant based on only bulk properties, such as, dielectric responses. Hamaker constants are mostly positive, generating an attractive van der Waals. However, there are cases for certain combination of media[4-7] (two different surfaces and an intermediate fluid) where the Hamaker constant is negative, generating a repulsive van der Waals force.

1.1.2 Capillary Forces

Capillary forces tend to arise when a capillary bridge spontaneously forms between two hydrophilic neighbouring asperities in a humid environment.[1, 8] These forces mostly form with the water absorbed on lyophilic surfaces. However, they can also appear in other cases, such as, with hydrophobic surfaces immersed in water connected by air cavities.

The Kelvin equation (Eq. 2) is a useful approximation that describes the thermodynamic equilibrium of a drop. It determines the dimensions of the capillary bridge:

(2) = (/) where rk is the Kelvin radius (mean radius of curvature of the meniscus), p/p 0 the relative vapour pressure, γ is the surface tension, Vm is the molecular volume, k is the Boltzmann constant and T the temperature.

2

Capillary forces depend mostly on the surface roughness and the relative humidity. When the two asperities are in contact, a capillary bridge is formed if the Kelvin radius is larger than the height of the smaller asperity (which is characteristic of surface roughness of both surfaces).[9-11] Only in the case of the interaction between macroscopic smooth spheres, do they become independent of humidity.

Capillary forces have a long range thus they can mask other short range interactions, such as van der Waals. Therefore, in order to measure other forces, the capillary bridge must be eliminated by either working at low relative humidity or immersing the system in liquid.[2]

1.1.3 Solvation Forces

When liquid molecules are confined between two surfaces (Figure 1), they tend to achieve a higher ordering. The increase of this order with the decrease in surface separation generates the solvation force.[1] This interaction depends on the properties of the liquid media, as well as, on the physical and chemical properties of the confining surfaces.

Solvation forces can have an attractive, repulsive or oscillatory nature, becoming dominant at short range. Moreover, as they can be stronger than other forces at small separations (e. g. van der Waals), they can become more important than other interactions, contributing more to the overall adhesion between two particles or surfaces.

3

Figure 1 . Schematic showing the ordering of liquid molecules between two surfaces that generates the solvation forces.

1.1.4 Hydration Forces

The hydration force can be thought of as a type of solvation force where the liquid compressed between the surfaces is water, and is originated by the overlapping of structured water layers between hydrophilic surfaces.[1, 12, 13]

When two hydrophilic surfaces are immersed in a dilute ion solution, the interaction between the surfaces obeys the DLVO theory.[1] However, when the ion concentration increases, the hydrated cations adsorb on the negatively charged hydrophilic surfaces leads to a repulsive hydration force that dominates the interaction. This force depends on the hydration number of the cations adsorbed on the surface, increasing in range and strength with it (Mg 2+ >Li +~ Na +>K +).

The hydration forces have an exponential decay thus can be successfully fit by Eq. 3:

4

(3) () = − where FHydration is the hydration force, λ is the decay length and CH is the hydration constant.

This interaction is strong, repulsive and short ranged, and occurs typically below 2 nm.

1.1.5 Hydrodynamic Forces

The hydrodynamic force ( FHydrodynamic ) is the additional force between two bodies corresponding to displacing the liquid between them.[14] This force, which causes dissipation of energy in hydrodynamic flow, was first experimentally determined by Chan and Horn[15] for the interaction of a sphere against a flat surface as a function of the viscosity of the fluid ( η) and relative velocity ( vR) according to Eq. 4:

(4) = This force becomes so dominant at high relative velocities and fluid that it shields the contribution of shorter ranged forces, such as solvation or van der Waals forces.

1.2 Adhesion

When two surfaces are brought close enough, an attraction induced by intermolecular interactions causes them to stick at contact spots or asperities. The force that is required to overcome this interaction is defined as adhesion or adhesion force. The forces that control the measured adhesion depend mostly on the material pair and their interfacial properties,

5 such as roughness, , cleanliness, crystalline structure, solubility of material in contact with each other, separation rate, time in contact, etc. With the immersion of the system in liquid, these surface forces may be modified or shielded, and new surface interactions may arise (e.g. solvation forces, capillary forces, hydrodynamics, ), contributing to the overall adhesion between the surfaces.

Therefore, the interpretation of adhesion can be challenging due to the number of factors involved, however it can provide useful information about the surface properties, as well as, assist to understand, predict and control other contact phenomena like and friction.

1.3 Tribology

Tribology literally means “the science of rubbing” and was defined in 1966 as “the science and technology of interacting surfaces in relative and of associated subjects and practices”.[16] This science that studies friction, wear and lubrication is of great interest and huge practical significance.

Friction is commonly defined as the force resisting when two bodies are brought in contact, and it has been part of our history since ancient times where, for example, fire was produced by generating high friction between two sliding sticks and large stone building blocks were transported using water lubricated sleds.[16] Friction prevents slipping or sliding and thus allows the possibility to walk, write, drive, grab, push, pull, etc. However, friction can be an inconvenience also because it resists motion. For example friction increases the energy required to drive a car, it can even lead to failure in mechanical devices with such as microelectromechanical motors and gears.

The first systematic study of friction registered was performed by Leonardo D. Vinci (1452-1519). Amontons later in 1699 received the credit and

6 postulated[17] that the friction between two sliding materials ( FFriction ) is proportional to the normal applied load ( FLoad ) according to Eq. 5 (Amontons’ first law):

(5) () = where µ is the friction coefficient defined as the proportionality constant between the friction force and the applied load

This friction coefficient, which is a suitable parameter to compare the lubrication properties between systems, is independent of both the apparent contact area[17] (according to Amontons’ second law) and the sliding velocity[18] (Coulombs’ law of friction). However both these laws are phenomenological and applied at the macroscopic scale. When the adhesion forces in the system are in the range of the applied load, as is the case at the nanoscale, the adhesion behaves as an additional loading force, and Eq. 5 is no longer valid. As a consequence of this additional force, friction forces extend to negative applied loads according to the following equation proposed by Derjaguin[19]:

(6) ( ) = + (0) where FFriction (0) corresponds to the friction force at zero applied load. Eq. 6 is a useful simplification and allows the friction coefficient and adhesion to be obtained independently of each other, where the adhesion or “contact adhesion”[20] is obtained from the intercept of the friction-load relationship with the load axis (Figure 2).

7

Figure 2. Effect of the applied load on the friction force according Derjaguins’, Amontons’ and JKR approximations.

Amontons’ and modified Amontons’ laws are based on experimental observations, but for a single asperity contact, the friction-force relationship is not linear at low loads and thus another approximation is required. One of the first attempts to deal with the adhesive contact of spherical asperities was performed by Johnson, Kendall and Robert (JKR theory, 1971).[1, 21] This theoretical treatment describes quite well the adhesive contact by , even during sliding. This model is simple in principle and postulates that the friction force is not proportional to the load but to the real contact area (Figure 2)[22] according to:

(7) = where Sc is the critical shear contact stress at the contacting interface and A is the real contact area.

However, in measurements when the deformation is small over the measured range, the contact area does not change too much and the modified Amontons’ law describes generally well the adhesive contact

8 interaction. Therefore, this friction-force relationship can be generally used, has been proven in many experimental systems,[11, 23-25] and is one the most common approaches to describe quantitative friction between surfaces.[21, 26, 27]

1.3.1 Nanotribology

Nanotribology is a branch of tribology that studies, friction, adhesion, thin– film lubrication and wear between sliding surfaces, at the molecular and atomic scale.[28] This is an exciting field where scientists of different backgrounds meet in order to understand fundamental contact phenomena that occur when two surfaces are sliding relative to each other.

However, nanotribology faces major challenges; for example, in miniaturized devices with moving parts, such as, micro- and nano- electromechanical systems (MEMS/NEMS), the small length scale and surface-area-to-volume ratio makes interfacial phenomena become dominant. Therefore, a good understanding is required of how surface interactions, such as van der Waals and capillary forces affect friction and adhesion at the molecular and atomic scale.[28-30]

Some light was shed on this topic around 1942 by Bowden and Tabor[31] that demonstrated the real contact between two solids in contact is only a fraction of the apparent contact area due to the surface roughness. At the nanometre/micrometre scale all surfaces are rough and they contact at some microscopic points (asperities). Therefore, the study of the surface interactions between these asperities at the molecular and atomic scale would provide an enhanced understanding of nanotribology.

Since the development of atomic force microscopy (AFM) by Binnig et al . in 1986,[32] much effort has been devoted to the study of the interactions

9 that govern nanotribology. AFM allows the study of the interactions that occur between the tip and the surface, which is used to mimic the interaction between the asperities.[33] Therefore, this technique has provided new insights at length scales not previously accessible, which are essential in order to understand nanotribology. This technique will be discussed in Section 2.1.

10

2 Experimental

2.1 Atomic Force Microscope

Atomic force microscopy (AFM) [32] is a popular, common and extremely versatile technique for analyzing surface properties. AFM has been mostly designed for imaging at the nano- and micro- scales but a more specialized, and no less important area is the study of force and friction with pico- Newton Force resolution.[34]

Typically, the sample is glued on the top of a metal disc that is magnetically attached to the base of the AFM (Figure 3). The scanner can be either located in the head or in the base of the AFM and moves the sample and cantilever relative to each other in x, y and z directions. The cantilever is placed in a holder located over the sample with the tip pointing down. A laser is focused on the cantilever and deflected onto a quartered photodiode. The photodiode voltage reveals the deflection and twist of the cantilever in response to an interaction of the tip where the vertical sections are related with topography and forces and lateral sections with friction.

Figure 3 . Schematic illustration of an AFM where: (a) is the quartered photodiode, (b) is the laser, (c) is the cantilever, (d) is the tip, (e) is the sample and (f) is the scanner (Molworx ®).

11

2.1.1 Imaging

AFM was originally developed to study topography by determining the height changes on the sample surface. The two most common modes for topographical imaging are contact mode and tapping mode. In contact mode, the tip is in permanent contact with the sample, and the scanner is moved up and down to keep the force constant. These height changes in the scanner position provide the topography. This mode is theoretically supposed to give better image resolution and is appropriate for relatively hard samples where the tip cannot damage them. In tapping mode, the cantilever is forced to oscillate with a certain amplitude, making intermittent contact with the sample surface. The topographic image is also generated here by the changes in the scanner height. This mode provides generally the best performance for imaging in general conditions, being recommended for fragile samples. Recently, new modes have been developed such as PeakForce ® Quantitative (QNM) from Bruker ®, and the very promising Intermodulation AFM (ImAFM®) from Intermodulation products ®.

PeakForce ® QNM provides not only topographic information, but other relevant material properties, such as Young’s modulus and adhesion. This mode will be described in Section 2.1.5.

2.1.2 Force

Normal forces are measured in an AFM by moving the surface toward and away (often referred to approach and retraction) from the cantilever tip by the piezoelectric scanner, and the raw data are obtained as cantilever vertical deflection ( Vn) versus Z-piezo position ( z). The raw data are then transformed into force versus separation as described below.

12

Figure 4 . Vertical deflection ( Vn) versus Z-piezo position ( z-top), where A corresponds to the constant compliance region, Z is the point of zero separation and B is the zero deflection region. The inset displays a schematic illustration of a force measurement before contact (A) and in contact (B). Force ( FL) versus separation data (D - bottom) obtained after processing.

There are three important regions in the force curve that are identified in Figure 4 (top); the zero deflection, the constant compliance and the point of zero separation. The zero deflection region corresponds to the value of vertical deflection when the tip is not interacting with the surface and is unperturbed. The constant compliance defines the region after contact when the cantilever deflects by the same amount as the piezoelectric scanner moves and thus the slope of the representation Vn versus z becomes constant. This calibrates the cantilever deflection in terms of distance units. The projections of these two areas intersect at the point of zero separation (Z), which is where the tip comes in physical contact with the studied surface.

The processing of the normal force curve starts with the transformation of

Vn from Volts via the inverse of the slope extracted from the constant compliance region. Afterwards, Hooke´s law (Eq. 8) is applied in order to convert deflection in metres ( ) to force in Newtons, where k is the ∗ z normal spring constant of the cantilever. Finally, z is converted to separation by taking into account the movement of the cantilever in Eq. 9:

13

(8) ∗ = ( ) (9) ∗ = − Although the measurement and processing of force curves are usually straightforward particularly for rigid substrates, interpretation of these force curves is not trivial at all. A good review of this topic is presented by Ralston et al .[33]

2.1.3 Friction

Friction is measured in lateral atomic force microscopy (LFM) by sliding the sample perpendicularly to the cantilever axis while in contact mode at a specific applied load, and recording the resulting cantilever twist. While changes in the vertical deflection of the cantilever give topography and normal forces, variation of the lateral deflection generates the friction information between the tip and the sample. During the scanning of the sample with the tip in both directions (often referred to trace and retrace), the cantilever twists in opposite directions, producing two photodetector output signals with opposite signs for each direction scanned (Figure 5). In the absence of anisotropy, the retrace lateral signal will be a mirror image of the trace.

14

Figure 5. Schematic representation of a typical friction loop. The inversion of the voltage sign is related to the change in the scanning direction ( Paper II ).

The conversion of the vertical deflection to applied load is calculated in the same way as for the vertical deflections show in Section 2.1.2. On the other hand, the average lateral output voltage ( Vd), which is obtained from the difference between the trace and retrace output lateral photodetector signals, is transformed to friction forces by Eq. 10:

(10) = where δ is the lateral deflection sensitivity, kϕ is the torsional spring constant, and heff is the effective height, which corresponds to the height of the tip plus half the cantilever thickness.

Afterwards, friction forces can be displayed as a function of applied load (Figure 6) and the modified Amontons’ law (Eq. 6) can be applied to fit the data and extract from the slope and a measure of the adhesion from the negative applied load where the Ff becomes zero (Figure 2).

15

0

Figure 6 . A Typical relationship between the applied load and the friction force.

2.1.4 PeakForce ® QNM

PeakForce ® QNM or PeakForce ® Quantitative NanoMechanics is an atomic force microscopy mode that allows high resolution topography imaging, as well as providing nanomechanical properties such as adhesion, Young’s modulus and deformation.[35, 36] Therefore, this technique is suitable for the study of heterogeneous surfaces like metal alloys.

In PeakForce ® QNM, the piezo scanner oscillates in the normal direction with a standard frequency of 2 kHz and a force curve is thus generated every 0.5 ms. During the scanning, the AFM feedback loop keeps the chosen maximum applied force constant (Peak Force) by adjusting the overall extension of the piezo. The possibility of controlling the applied force provides the opportunity for non-destructive imaging.

A force curve is generated with every cantilever tap and its analysis provides the nanomechanical properties (Figure 7), which afterwards are presented in the images.

16

Figure 7. Schematic representation of a force curve as a function of separation on approach (solid line) and retraction (dotted line), indicating which part of the force curve provides the nanomechanical properties.

Figure 7 shows that adhesion is extracted from the difference between the baseline and the minimum force during retraction, the deformation from the distance between zero separation and a position with a given percentage of the full deformation of the approaching curve and the Young’s modulus by fitting the linear part of the retraction force curve using the Derjaguin- Muller-Toporov (DMT)[35] contact mechanical model, which describes the relationship between applied force and deformation of a material. Afterwards, a set of images can be obtained for every sample scan, where each image displays a different nanomechanical property of the surface.

2.1.5 Colloidal Probe

The colloidal probe technique developed by Ducker et al .[13, 37] and Butt[38] is based on the exchange of the AFM cantilever tip by a colloidal particle (1-20 µm in diameter), as shown in Figure 8. It has the advantage of allowing forces to be measured with a probe made of virtually any material, as long as the probe has a well defined shape and is almost incompressible,

17 where the magnitude of the measured force scales with the size of the probe.

However, the comparison of forces obtained with different probes is not straight forward, and is thus necessary to normalize them, by applying the Derjaguin’s approximation.[39] This approximation relates the normal force to the energy per unit area ( W) between two flat surfaces, according to:

() (11) = () where Reff is the effective radius that depends on the interacting surfaces. In the case of a colloidal probe interacting against a flat surface ( Paper I ), the spherical shape of the probe simplifies the calculation of effective radius, which can be approximated to the radius of the colloidal probe.

Figure 8 . Scanning image of a 5µm silica colloidal probe glued on a tipless cantilever (Paper I ).

2.1.6 Determination of Spring Constants

In force and friction measurements, the determination of the normal and torsional spring constants is crucial because these constants transform the cantilever bending and twisting to forces (see Section 2.1.2 ). During the last

18 two decades, a large number of solutions have been proposed, which are theoretical, experimental or a combination of both, but there is one approach that has been widely used due to its accuracy and simplicity. [40] This technique, developed by Sader,[41-44] is commonly known as the Sader method.

2.1.6.1 Determination of Spring Constants by the Sader Method

The Sader method is based on how the cantilever vibration frequency response is affected by the surrounding fluid. The cantilever is allowed to vibrate due to thermal motion in a fluid, which is generally air. The normal resonance frequency ( fz) (which can be thought of as the vertical vibration frequency of for example a diving board) and the normal quality factor ( Qz) are obtained by fitting a simple harmonic oscillator function to the normal resonance peak obtained from the thermal power spectra of the cantilever, and afterwards they are combined with the measured length ( L) and width (w) of the cantilever, and the density ( ρ) of the fluid, to determine the normal spring constant, according to Eq. 12[43] where is the imaginary component of the hydrodynamic function for normal ( vibrations.) [45]

Γ (12) = 0.1906 (2) ( )

(13) = The determination of the torsional spring constant is very similar to the calculation of kz presented above, but in this case the torsional resonance frequency ( fϕ) and the torsional quality factor ( Qϕ) are obtained from the torsional resonance peak. Therefore, kϕ is calculated using:[42]

19

Γ (14) = 0.1592 2 ( )

(15) = where Γ is the imaginary component of the hydrodynamic function ( ) for torsional vibrations.[46]

There is a limitation in the determination of the torsional resonance frequency from the power torsional thermal spectra because of its lower resolution, and for stiffer cantilevers this resonance is hard to measure. Section 4.2 and Paper II contains an approach to avoid this issue.

2.1.7 Focused Ion Beam (FIB)

Focused ion beam (FIB)[47] is a technique that employs a focussed beam of ions (usually gallium) to irradiate the sample (Figure 9). One of the applications of FIB is the ability to generate images from the charged particles (ions and electrons) that are released from the sample when the beam of ions hits the surface. Another application is the milling that occurs during beam exposure due to the atomic collision process that removes atoms from the sample surface. An FIB can also be used to deposit material via a gas delivery system that provides a chemical compound close to the beam-sample interaction point. The FIB decomposes the gas locally, and the products are deposited on the surface. In this thesis the FIB has been used solely as a cutting tool for microscopic samples. ( Paper VI )

20

Figure 9 . Schematic representation of the focus ion beam operation principle.

21

3 Materials and Fluids

3.1 Cantilevers and Probes

The AFM cantilevers used for calibration, imaging, force and friction throughout this thesis were crystalline silicon cantilevers with rectangular shape from MikroMasch (Tallinn, Estonia).

The silica and alumina colloidal probes used in Paper I , were provided by Bang Laboratories (Fishers, IN) and Sveriges Tekniska Forskningsinstitut (SP, Stockholm), respectively. These particles were glued on tipless cantilevers using a micromanipulator (Micromanipulator 5171, Eppendorf, Hamburg, Germany) under an optical microscope (Nikon Optiphot 100, Tokyo, Japan).

A hair fibre is composed of a layered structured with three regions: cuticle, cortex and medulla. The cuticle is the outer part of the fibre and is responsible for the surface properties of the hair (Figure 10). The cortex is the main structural component and provides the special mechanical properties of the hair. Finally, the medulla is the central part and is not always present, only for thick hairs, which tend to protrude mostly from the scalp, and its function remain unclear.[48] The facial human hair involved in the Paper VI was provided by L’Oréal (Paris, France).

22

Figure 10 . Scanning electron microscopy image of a human hair cuticle surface.

3.2 Substrates

In Paper I, the silica wafers (model Ultrapack Wafershield H9100-0302) were purchased from Entegris (Dresden, Germany) and the smooth polytetrafluoroethylene (PTFE) surface was prepared by compressing pieces of PTFE between fleshly cleaved mica sheets inside of an oven overnight at 500 °C.[4]

The metal alloy used in Paper III is a nitride tool steel (Vancron ® 40) provided by Uddeholm AB (Sweden), chemical composition of which is given in Table 1. This material consists of a homogeneous distribution of fine nitride particles embedded in a metallic matrix. Vancron ® 40 does not need surface in applications because the surface microstructure of the metal alloy leads to low adhesion against soft working materials. Moreover, it is a hard material with enough ductility and toughness to avoid premature failure.[49, 50]

23

Table 1 . Chemical composition (wt. %) of Vancron ® 40, Fe is balanced. [49]

Wt. % C N Si Mn Cr Mo W V

Vancron ® 40 1.1 1.8 0.5 0.4 4.5 3.2 3.7 8.5

In Paper IV , the biomedical CoCrMo alloy was supplied by Sandvik Material Technology (Sweden) with a chemical composition (wt %) described in Table 2. This CoCrMo alloy consists of two different types of nitride particles embedded in a Co rich matrix that forms a native protective mostly composed of Cr 2O3 and, (in minor amount) Co and Mo oxides.[51, 52] These alloys are widely used for joint replacement due to their corrosion and wear resistance, good mechanical properties, and biocompatibility with the human body.[53, 54]

Table 2 . Chemical composition (wt. %) of the CoCrMo alloy, Co is balanced. [52]

Wt. % Cr C Mo Mn Ni Fe

CoCrMo alloy 28 6.3 0.5 0.2 0.2 0.2

FeCrVN is the experimental nitrogen based tool alloy studied in Paper V provided by Uddeholm AB (Sweden), the chemical composition of which is given in Table 3. This metal alloy is formed by two different types of nitrides (with different vanadium and iron contents), which are embedded in the alloy matrix.[55]

Table 3. Chemical composition (wt. %) of the FeCrVN alloy, Fe is balanced.

Wt. % C Mn Si N Ni Cr Mo V

FeCrVN 0.2 0.3 0.3 4.2 0.05 21.2 1.3 9.0

24

The stratum corneum (SC) is the outer part of the skin and is formed by keratin-rich dead cells called coenocytes, which are inserted in a cement providing high mechanical strength and good elastic properties. This layer has important functions, such as skin barrier, preventing water loss, for the appearance (i.e. optical properties) and photoprotection. It also acts as contact surface for tactile perception.[56-59] The human abdominal SC (~15µm thick) studied in Paper VI was kindly provided by L’Oréal (Paris, France).

3.3 Liquid Media

The ultrapure water used throughout this thesis has a pH of ca. 5.7, resistivity of 18.2 M -cm and carbon content below 2 ppb, and was obtained with a Milli-Q unit (Millipore, Molsheim, France). The ethanol (99.5% purity) used in this work was acquired from Kemetyl (Haninge, Sweden).

Ionic (ILs) are molten organic salts with melting temperatures below 100°C, which have a wide range of applications due to their unusual properties, such as, high temperature stability, thermal conductivity, low volatility and electrical conductivity.[60] Ethylammonium nitrate (EAN), discovered in 1912, is a protic ionic liquid formed by an ethyl ammonium cation and a nitrate anion that forms a 3D H-bonded network structure. This IL has a viscosity that is 30 times higher than water.[61, 62] In Paper I , EAN was synthesized according to a process described previously,[63] combining ethylamine and nitric acid, which were acquired from Sigma Aldrich (Munich, Germany) and Merck (Darmstadt, Germany), respectively.

The tetradecane with a purity of 99% used in Paper III was purchased from Sigma (Haninge, Sweden).

25

The phosphate buffered saline (PBS) is a solution generally used in biological research since it mimics the saline conditions of the body. This salt used in Paper IV contains 8.77 g/L NaCl, 0.2 g/L KCl, 1.28 g/L

Na 2HPO 4 and 1.36 g/L KH 2PO 4, and was obtained from Sigma Aldrich (Munich, Germany).

26

4. Summary of Key Results

4.1 Suppression of Surface Interactions by an Ionic Liquid

In Paper I , the friction and adhesion between pairs of industrially relevant materials (silica, alumina and polytetrafluoroethylene-PTFE) have been analyzed in terms of the long ranged surface interactions, such as, van der Waals and capillary forces. The four systems (with the convention “probe- surface”: silica-silica, silica-PTFE, alumina-silica and alumina-PTFE) were studied in both ambient air and in EAN in order to understand the different types of interactions involved between the surfaces and how EAN lubricates the contact. In Figure 11a is presented a normalized force curve for the silica-PTFE system in air, and in Figure 11b the frictional force as a function of applied load for the same system. When the tip is getting close to the surface (around 5 nm), a jump-in is observed that is characteristic of a van der Waals interaction; when the tip is moved away from the surface, adhesive forces keep the surfaces together until this interaction is overcome. The fitting of the approach curve by using Eq. 1 (Figure 11a, inset) confirms the van der Waals nature of this interaction.

27

Figure 11 . Normalized silica–PTFE force curve in air (a) on approach (closed symbols) and retraction (open symbols). The inset in (a) shows five normalized force curves for the same system on approach fitted with van der Waals theory. (b) Friction as a function of applied load in air for silica-PTFE (Paper I ).

The friction response of the silica-PTFE system is also affected by the adhesion forces observed in Figure 11a. Figure 11b presents large friction values, where the large friction at zero applied load is produced by the significant adhesion that acts as an extra load (see Section 1.3 ). The friction coefficient extracted from the gradient of the friction-force data in the linear part of the Figure 11b is constant and thus does not depend on the applied load or frictional force according to Eq. 7. However, a careful examination of the figure reveals that this linear regime is broken at low applied loads. This is expected for PTFE because it is much softer than silica and is deformed after contact, changing the real contact area. As a consequence, the friction force becomes dependent on the area of contact which is increasing due to contact mechanics (see Section 1.3 ).

The hydrophobicity of PTFE makes the silica-PTFE and alumina-PTFE systems representative of minimal adhesion because of the absence of capillary forces that greatly contribute to adhesion forces. However, the other systems studied in air without PTFE (silica-silica and alumina-silica in Paper I ) presented strong capillary forces, leading to high adhesive and friction values which depends on the ambient conditions, such as relative

28 humidity. The differences of the systems studied in air were highly influenced by the materials chemistry, the environment (humidity and temperature), and the number/sequence of measurements.

Figure 12. Normalized force curves on approach (closed symbols) and retraction (open symbols) at an approach rate of 100 nm/s (Paper I ).

Figure 12 shows that with the immersion in EAN, van der Waals and capillary forces in every system were shielded because of the screening effect of the IL. The systems studied in EAN showed no adhesion with friction forces and friction coefficients lower than in air (Figure 13). Since these systems show almost no adhesion at zero applied load, they follow a more classical Amontons’ law (Eq. 5). The differences observed between the systems immersed in EAN were mainly influenced by the sample/probe roughness.

29

Figure 13. Friction measurement between a silica colloidal probe and a silica surface in EAN (Paper I ).

Previous works have been applying adsorbed thin films of ILs,[64-67] but in this work the systems were immersed in the IL. This modification eliminates potential problems, such as, film depletion and atmospheric water absorption.[68]

EAN is a highly viscous liquid, and hydrodynamics (fluid dynamics in this case) thus need to be addressed in the surface interactions. Therefore, in Paper I , an investigation has been performed in order to quantify this effect. Figure 14a shows normal force measurements, on approach, between a silica probe and a silica surface with different scan speeds. They present a relatively long-ranged repulsive force that increased with increasing approach speed. This effect was not observed in the silica-silica force curve presented above in Figure 12a because of the low scanning speed (100 nm/s). At high velocities this hydrodynamic resistance controls the interaction and provides an extra barrier against contact, which from a tribological point of view, adds an extra barrier to avoid adhesion and decrease friction. On the other hand, the opposite effect occurs on retraction where an attractive force is generated because of the viscous resistance to flow into the contact area. These forces during approach and retraction increase monotonically with the ramping speed.

30

This hydrodynamic resistance can be minimized by decreasing the ramping velocity of the AFM scanner which controls the speed of approach. For very slow approaches, where the hydrodynamic force is small (Eq. 4) it is possible to detect steps in the normalized force curve in the inset of Figure 14a that correspond to structural layering of the EAN ion pairs, as shown previously.[62, 69-72] These steps are separated by 0.5 nm which corresponds to the IL ion pair diameter.[69] Thus this can be seen as a type of solvation force where the ion pairs act as molecules (see Section 1.2.3 ). However, these steps produced by EAN are not visible at high rates because they become masked by the much larger and more long ranged hydrodynamic force.

To verify the nature of the hydrodynamic forces in the Figure 14b, a retraction force curve was fitted by Eq. 4, allowing the viscosity to be extracted (all the other parameters in Eq. 4 are fixed. Details of the fitting procedure are given in Paper I ). The inset of Figure 14b shows that the model describes the experimental data remarkably well but the viscosity obtained by the fit is 12 mPa·s, which is much lower than the bulk viscosity of EAN at room temperature (32 mPa·s).[61]

The interpretation of this astonishing result is provided by the consideration of a lamellar ordering of the EAN at the surface which may offer a lower resistance to sliding due to well-defined shear planes.

31

Figure 14 . Normalized forces curves on (a) approach and (b) retraction between a silica probe and a silica surface in EAN for different ramping velocities. The inset in (a) presents steps with the EAN ion pair dimensions obtained at 12 nm/s. Inset in (b) shows the fit using Eq. 6 in the retraction experimental data at 4360 nm/s (Paper I ).

4.2 A Hybrid Route for the Determination of Torsional Spring Constant for AFM Cantilevers

For the determination of the normal and frictional forces in Paper I , III and IV , it was necessary to transform the photodiode voltages (which are related with the deflection and twist of the cantilever) into friction force through the normal and torsional spring constants of the cantilever, ( kz and kθ, respectively, see Sections 2.1.2 and 2.1.3 ).[33, 40, 73] However, due to the high stiffness of the cantilevers used, it was not possible to apply the Sader method for the determination of the torsional constant. As the rigidity of the cantilever increases, the signal/noise ratio of the torsional resonance

32 peak decreases, and at certain rigidity, the peak cannot be found. Therefore, for very rigid cantilevers the Qθ and fθ values cannot be extracted and the torsional resonance frequencies cannot be calculated. A new approach is thus needed for the determination of kθ for rectangular rigid cantilevers that does not rely on the determination of Qθ and fθ.

The determination of these spring constants for rectangular cantilevers can be performed by using theoretical approximations, such as standard beam theory,[74, 75] where the spring constant is related to the beam dimensions and the Young’s ( E) or Shear ( G) moduli of the cantilever material,[76] and t is the thickness of the cantilever:

(16) =

(17) = Another approach is to consider experimental approximations, such as the Sader method (see Section 2.1.6.1 ), which are preferred because the cantilevers are very thin and it is difficult to obtain the thickness accurately or to know how the thickness varies. Furthermore, there is no guarantee that the moduli of a micron thickness cantilever should have the same as the bulk. Besides, the standard beam model for the calculation of kθ does not consider the inherent restraint on axial warping. This term, which becomes important for short cantilevers,[44] has been considered in a derivation from plate theory:[44]

(18) = 1 − ( )

33

Eq. 18 takes into consideration the above issue in the theoretical treatment and improves the accuracy of Eq. 17 but the other problems still remain. Therefore, an appropriate experimental calibration technique is needed to determine the spring constant of stiff cantilevers.

In Paper II , a hybrid route has been developed, where the normal Sader method (nSm) to obtain kz is combined with the theoretical approximation based on the plate theory for the determination of the torsional spring constant for stiff cantilevers. This method, which is referred to as the Hybrid model, does not need to extract any parameters from the power torsional thermal spectra and does not depend on a measured cantilever thickness. The hybrid method was obtained by combining Eq. 18 with Eq. 16 leading to Eq. 19:

(19) = 1 − ( )

To validate the above derived equation for the Hybrid method, it was tested it against the torsional Sader method (tSm) for several cantilevers with different dimensions, and with and without .

34

Figure 15. Comparison between the torsional spring constants calculated by the Hybrid ( ) method with respect to the torsional spring constants calculated by the established Sader method , with dashed linear fit. The solid line in the figure corresponds to a perfect agreement ( ) with the torsional Sader method ( Paper II ).

Figure 15 displays a comparison between the torsional spring constant calculated by the tSm and Eq. 19 where the data are obtained from a set of beam shaped monocrystalline silicon cantilevers, both coated and uncoated (see Paper II ). The error bars were obtained from the standard deviations of three consecutive measurements. This error was omitted for clarity when the error bar was smaller than the symbol. The solid diagonal line in the figure corresponds to the case where both methods return the same values, and the almost indistinguishable dashed fit is the linear regression of

with respect to . Standard deviations of kθ for stiffer cantilevers were larger than for softer ones because the amplitude/noise ratio of the torsional resonance peaks becomes smaller for more rigid cantilevers. Besides, there is a larger error in the x direction than in the y direction that is generated by the greater uncertainty in the tSm method, especially for the more rigid cantilevers. The only required factor in Eq. 19 is the ratio of G/E which is used as a fitting parameter such that the dashed fit is as close as possible to the solid line in Figure 15. The 3.0 value

35 provides the best fit, but the fact that the results are linear is an indication that the approach is correct .

It could also be considered that the quality of the fit in Figure 15 should be affected by the surface coating of the cantilever, however a linear fit was obtained, irrespective of whether coated cantilevers were used or not. Therefore, this coating effect appears to be negligible or to be adequately addressed in the measurement of kz.

In this work, it has been established a working value of G/E for a family of cantilevers by comparing with . It can be seen that this value is general and thereafter, can be safely used in conjunction with the kz to obtain the torsional spring constant, irrespective of whether the cantilever is stiff or not. This agreement between the hydrodynamic approach and the mechanics calculations provides further confidence to the applicability of this model, not only for stiff, rectangular, silicon cantilever, but also for any other type of rectangular beam cantilevers.

4.3 A New Technique to Characterize Heterogeneous Surfaces: Tribological Property Mapping

The friction constant of a material depends on the material properties and through them the strength of the adhesive force at contact, for example the magnitude of the van der Waals force.[77] In Paper I the effect of screening the van der Waals force completely through the use of an IL was clearly observed in the magnitude of the friction coefficient. In that paper homogeneous materials were used, but in an application, a heterogeneous material is usually used where the properties of the surface can vary significantly from place to place, reflecting the local composition. Often, as in a tool steel, the various properties can be harnessed in the performance of

36 the material. Thus, it is useful to be able to map this local variation in the property of a system.

Scanning probe techniques are used for the characterization of material properties at the nanoscale, and lateral atomic force microscopy (see Section 2.1.3 ) has long been a part of this armoury. As the scanning tip transverses the sample surface in LFM, the twisting of the cantilever generates a friction force that depends not only on the applied load but, also depends on the local adhesion between the tip and the surface, which varies with the surface composition and microstructure. Therefore, the friction force depends on the tip, surface and ambient conditions. On the other hand, the friction coefficient is a more useful parameter to compare lubrication properties between different systems (see Section 1.3 ) but is not obtained directly from LFM measurements.

LFM can be used to map different surfaces[78-83] providing qualitative friction images (by sliding the tip across the surface). However, quantitative measurements are more difficult but possible to obtain by transforming the lateral photodetector output to friction force (see Section 2.1.3 ). Both of these measurements provide relevant information of the sample-tip interactions, but to map fully the tribological response of the sample, the adhesion and friction coefficient are required for each pixel of the image. Therefore, tribological property mapping (TPM) has been developed in Paper III for a complete tribological analysis of the surface, where a set of images containing information about the twisting of the cantilever at different applied loads are transformed into two new images in which each pixel contains friction coefficient and adhesion values, respectively. In this study, the commercial powder metallurgical tool alloy, Vancron ®40, was employed because its heterogeneous microstructure allows the identification of local tribological properties.

37

Figure 16. Friction images of Vancron ®40 in air obtained from trace (a, d, g), retrace (b, e, h) and average (c, f, i) friction data applying a load of 15.19, 80.29 and 145.39 nN. The insets show the friction values vary through a cross section of 3.7 µm and the symbols x and + correspond to two positions with lower and higher friction contrast respectively ( Paper III ).

Figure 16 presents LFM trace, retrace and average friction force obtained at three different applied loads, where the average friction was created by the combination of the trace and retrace data ( Paper III ). Afterwards, these average friction images obtained at different applied loads are employed to generate the friction coefficient and adhesion maps. The line sections inserted in every image in Figure 16 demonstrate how the friction contrast varies locally with the applied load, indicating that there are different regions with significantly different frictional properties. This structure of particles embedded in a matrix is consistent with previous studies, which

38 show that Vancron ®40 has a heterogeneous microstructure of fine carbide and carbide-nitride particles inserted in the alloy matrix.[49, 50]

The images in Figure 16 only provide frictional information at the applied load that they were obtained. However, it is unknown what would occur at other applied loads and neither could we predict how the friction would vary with load. Therefore, it is more useful to obtain images containing tribological information which are not related to the applied load. With these images, it is possible to describe fully the frictional properties of the studied system and predict their variation with load.

To show how these images are obtained, the calculation was simplified to two positions on the image, one located over a particle (+) and another over the matrix (x), which display higher and lower friction contrast, respectively. The average friction forces were extracted at each of the two pixels for all of the seven images obtained at different applied load and plotted with respect to the applied load (Figure 17).

Figure 17. Effect of applied load on friction force, at individual pixels. The filled symbols correspond to the particle (+) and the empty symbols to the matrix (x) ( Paper III ).

Afterwards, linear regressions were applied according to Eq. 6 (see Section 1.3 ) and the friction coefficients and the adhesion values were obtained. These two parameters provide load-independent tribological information related to the local microstructure of the two positions analyzed on the

39 images. In Paper III, such treatment was automatically applied to every pixel-position on the seven images, and two maps containing friction coefficients and adhesion values were generated.

Figure 18 . Vancron ®40 surface scanned with a silicon cantilever in air, (a) height, (b) friction coefficient and (c) contact adhesion ( Paper III ).

Figure 18a is a topographic image obtained in contact mode performed in air (19% relative humidity), which presents the microstructure of Vancron ®40. It is composed of a matrix containing particulate features with a diameter smaller than 2 m. These tops of the particles are located both above and below the matrix surface. Figure 18b is the friction coefficient image obtained by TPM for the same area as in Figure 18a, where the features show a larger friction coefficient than the matrix. It appears that whether or not the particles protrude has no effect on the friction coefficient values. Therefore these particles, with friction coefficient values ranged between 0.65 and 0.85, may have the same or similar kind of surface composition (note that composition differences for such materials are nominal, and the composition of the both the matrix and the particles can vary within the same sample).[49, 50] The matrix displays a lower friction coefficient ranged between 0.40 and 0.70.

The other contribution of TPM is the adhesion image shown in Figure 18c. This adhesion is obtained without actually separating the surfaces from contact and for convenience this value is referred as “contact adhesion” (Paper III ). In this case, the correlation with Figures 18a-b is possible but

40 not easy because the contact adhesion values for both the particle and matrix against the silicon tip in air are very similar, around -40 nN. This result is not unexpected because the surface is exposed to ambient relative humidity (19% in this case), and when the tip comes in contact with the surface, a capillary bridge forms due to vapour condensation, increasing the adhesion contribution. This contribution is expected to be relatively similar for all the surface components and thus independent of the scanning position. The magnitude is also expected to be comparable to, or larger than, the influence of other types of surface interactions, such as van der Waals forces.[25] To confirm this contention, normal force measurements were performed between a silicon tip and the Vancron ®40 surface in air. A result is displayed in Figure 19 where the jump in and jump out in the force curve are consistent with this type of interaction. Note that the values of the conventional adhesion extracted from force curves such as that in Figure 19 are slightly lower than the contact adhesion obtained from Figure 18c because of the underestimation of the adhesion expected in the traditional pull off measurements. The lifetime of the capillary condensate is also very different in the two cases, and vapour harvesting[25] may have contributed to a larger condensate with greater adhesion in the TPM images.

Figure 19 . Normal force curves on approach and retraction between a silicon tip and a Vancron ®40 surface in air.

41

Thus it is very important to measure the humidity when the measurements are performed in to order to compare tribological measurements, - this is true for any system where adhesion is important, irrespective of the technique.

On the other hand when the system is immersed in liquid, capillary condensation may not occur and the sole contributor to adhesion is most likely to be the van der Waals interactions between the tip and the surface. Therefore, TPM was performed between Vancron ®40 and a silicon cantilever in tetradecane. Figure 20a shows again a topographic image, and the surface microstructure is very similar to that in Figure 18a. Figure 20b displays the friction coefficient image obtained once again by TPM. As for Figure 18b the particulate features show a larger friction coefficient than the matrix, but in this case, the values are lower than those in air due to the lubricating contribution of tetradecane. Thus, the particles display friction coefficients ranging between 0.25 and 0.55, and the matrix displays values between 0.25 and 0.45. The friction coefficient map also shows less noise than the previous measurement in air probably because of the absence of a capillary condensate during the scanning.

Figure 20 . Vancron ®40 surface scanned with silicon cantilever in tetradecane, (a) height, (b) friction coefficient and (c) contact adhesion ( Paper III ).

In contrast to Figure 18c, the contact adhesion image in Figure 20c shows heterogeneity. The absence of the capillary condensate leads to contrast between the matrix and the particulates, which can be clearly distinguished

42 in the image. The contact adhesion values observed on the asperities ranged between -10 and -40 nN and on the matrix between -15 nN and -25 nN. The considerably lower adhesion values in Figure 20c and the jump in and jump out of the force curve in Figure 21 reflect the absence of the condensate. Thus the major contribution to the contact adhesion is from other forces, such as van der Waals which depend upon the local composition of the sample.

Figure 21. Normal force curves on approach and retraction between a silicon tip and a Vancron ®40 surface in tetradecane ( Paper III ).

4.4 Biomedical CoCrMo alloy: A Tribological Properties Mapping Study

TPM[20] was applied once again to another heterogeneous material in Paper IV . In this case a biomedical CoCrMo alloy was selected because of its special properties such as corrosion and wear resistance, good mechanical properties, and biocompatibility with the human body, and the intermediate medium was PBS because it mimics the saline conditions of the body (see Section 3.3 ). Under these conditions, the local tribological properties could be studied and the stability of the alloy in similar biological environment to which, for example, hip-joint replacements are exposed during their working life.

43

The biomedical CoCrMo alloy is formed by two different types of carbides embedded in a Co-based matrix according to the atomic composition shown in Table 4. These two types of carbides are observed in the backscatter

SEM image in Figure 22, where the bright areas correspond to the M 6C carbides with heavier elements (9.70% molybdenum) and the darker areas are the M 23 C6 carbides and matrix with lighter elements, (2.26% and 3.80%

Mo respectively).[52] Both carbides have face centred cubic structure

(FCC), but the M6C carbides with a like crystalline structure Fd- 3m[84] are the hardest part of the biomedical alloy.

Table 4 . Element composition (at. %) of the respective phases obtained from energy dispersive spectroscopy (EDS) and transmission electron microscopy (TEM).[52]

Element (at. %) Cr Co Mo Mn Ni Fe C Si

M6C 18.66 17.91 9.70 - - - 44.40 9.33

M23 C6 37.42 6.45 2.26 - - - 53.87 -

Alloy matrix 31.20 63.17 3.80 0.52 0.21 0.21 0.98 -

Figure 22 . BSE-SEM image of a CoCrMo alloy, displaying the matrix and the two different types of carbides.

Figure 23a shows a topographic image of the studied medical alloy in contact mode performed in PBS after one hour of immersion. It indicates

44 particulate features contained inside the matrix that protrude above the surface. Two sets of particles with different heights are observed in the image. During the polishing of the sample the harder, more resistant part of the metal abrades less than the softer parts, thus the features of harder particles are higher than the softer matrix. Figure 23b corresponds to the same area as Figure 23a and it displays a TPM image of friction coefficient variation. The particles can be clearly distinguished from the matrix. TPM shows clearly that the particles, which might be speculated as being of different hardness due to their different heights, are in fact of very different nature. One type of particle displays the highest friction coefficient values (between 0.22 and 0.33), and the other type of particle presents the lowest friction coefficient values in the image, ranging between 0 and 0.15. Thus the matrix has frictional properties that are intermediate between the two types of particles described above. The contact adhesion information is presented in Figure 23c, where particles with the highest friction coefficients in Figure 23b also show higher contact adhesion values ranging between 10 and -20 nN. On the other hand, particles with the lowest friction coefficient also present the lowest adhesion, with values that range between 25 and 5 nN. The matrix displays intermediate contact adhesion values between 15 and -5 nN. The contact adhesion according to the definition in Section 1.3 is the intercept of the friction force data with the load axis. If the surfaces experience an adhesion then the value is a negative load, but there is no reason why a “negative adhesion” i.e. an intercept at a positive load, cannot occur. This corresponds to the case of a repulsive force which must be overcome before the surfaces can achieve contact and dissipate frictional energy.[85] I clarify that a true adhesion is denoted for TMP by negative numbers, and a more negative value corresponds to a more attractive interaction. Positive values thus correspond to repulsion.

45

Figure 23 . Si - CoCrMo alloy images in PBS, (a) height, (b) friction coefficient and (c) contact adhesion.

TPM provides relevant tribological information of the CoCrMo alloy but in order to understand why differences phases have different tribological responses, it is necessary to study the influence of the surface microstructure of this alloy.

The chromium content is a very important parameter that affects the tribological properties of the studied alloy. When Cr is exposed to oxygen, it readily oxidizes forming a chromium oxide layer, which decreases the friction coefficient and the contact adhesion. The good tribological properties of the chromium oxide were attributed in previous work[86] to the poor wettability and high hardness of the thin oxide formed on a chromium coated steel tool. Moreover, the Cr content at the interface is slightly enhanced with respect to the bulk due to its preference to form a thin oxide.[87] Therefore, the phases that displayed higher friction coefficients and contact in Figures 23b-c can be identified as the

M6C carbides because of their low chromium content (Table 4). On the other hand, the phases with lower friction coefficient and adhesion values correspond to the M 23 C6 carbides with more than 37% chromium content. This identification agrees with the previous speculation about the relative hardness, where the M 6C carbides with a diamond like hardness will protrude more than the M 23 C6 carbides which abrade more during the polishing. As a conclusion, it is more likely that the thin chromium oxide

46 formed on the biomedical alloy surface is responsible for the good tribological properties of the material. Moreover, the M 23 C6 carbides with the highest chromium content generates the more effective oxide layer displaying the lowest friction coefficient and contact adhesion values in Figures 23 b-c.

Another parameter that affects the tribological response is the intervening medium. After one hour of immersion in PBS, TPM was performed and no indication of corrosion or alteration of the biomedical alloy surface was observed. This supports the inertness of the CoCrMo alloy within this type of biological environment.[54]

Figure 24 . Normal force curves on approach (closed symbols) and on retraction (open symbols) between a silicon tip and a CoCrMo alloy surface in PBS. The continuous line in the approach curve represents the fit according to Eq. 20. The inset shows an enlargement of the data.

Figure 24 shows the surface force at an arbitrary point in PBS. As the tip approaches the surface, it experiences first, from around 7 nm, an attraction attributted to a van der Waals force (see Section 1.2.1 ). Afterwards, an attraction is overcome by a hydration force (see Section 1.2.4 ), generating a repulsion around 2 nm of separation distance. A threshold-force of about 15 nN is needed to reach the hard contact region (constant compliance), which corresponds to the range of positive values (repulsive) observed in Figure 23c and explains why the contact adhesion values are repulsive in this case. The fitting in Figure 24 is not perfect, but verifies that the forces between

47 the tip and the biomedical alloy are likely explicable by the superposition of van der Waals (Eq 1) and hydration forces (Eq 3) according to Eq. 20:

(20) () = + -20 where the fitting parameters are A = 3x10 J, λ = 0.2 nm and CH = 3.7 N. On retraction, the superposition between hydration and van der Waals forces generated a light repulsion below 2 nm and an attractive interaction between 2 and 8 nm. Therefore, the system formed by the silicon tip and the CoCrMo alloy in PBS experiences repulsive interactions mostly because of hydration forces, which generate a decrease in the friction coefficient and contact adhesion values in Figures 23b-c. The effects of these forces are so important than even positive contact adhesion values (repulsion) are observed in Figure 23c, especially in the case of the M 23 C6 carbides.

4.5 Surface Study and Corrosion Initiation of an Experimental FeCrVN Tool Alloy

Mechanical properties describe how materials deform when they are exposed to external forces. These mechanical properties can be combined with tribological information in order to understand the microstructure of a material as well as shed light onto surface phenomena such as corrosion. The aim in Paper V was to study a FeCrVN stainless steel tool alloy by PeakForce QNM ® in water and in NaCl (0.1 M) in order to understand the influence of microstructure on corrosion initiation of the tool alloy.

48

a) b)

5 μm 30 μm

Figure 24. (a) Back-scatter SEM image of FeCrVN experimental tool alloy and (b) Volta potential image of the alloy.

Figure 24a is a back-scatter SEM image of the FeCrVN alloy showing its heterogeneous structure. The alloy is formed by particles of 0.5-3.0 µm enriched in chromium and molybdenum, which are harder than the iron- based martensitic matrix that contain them, and are referred to in this work as nitride particles. There are two types of particles embedded in the matrix and their compositions are displayed in Table 5. The darker ones in Figure 24a have more vanadium and nitrogen, and the lighter ones have more chromium and molybdenum. For simplicity, I will further refer to the two types of hard phase particles as Cr-V rich and Cr-Fe rich nitrides.

Table 5 . Average of at least five analysis points showing the EDS chemical analysis of the matrix and the two different particles that form the alloy (lighter and darker nitrides).

Element (wt. %) Cr V N Mo Fe

Darker nitride 34.1 ± 0.3 33.9 ± 0.7 18.9 ± 0.5 0.8 ± 0.1 13.3 ± 0.5

Lighter nitride 38.2 ± 0.7 5.4 ± 0.3 8.4 ± 0.2 2.2 ± 0.3 43.3 ± 0.8

Alloy matrix 18.7 ± 0.5 1.2 ± 0.3 2.7 ± 0.4 1.7 ± 0.6 73.1 ± 0.5

Figure 24b is a Volta potential image obtained from a Kelvin Force Microscopy (KFM)[88] where the nitride particles have higher Volta

49 potential than the alloy matrix. Differences are also observed between the particles in Figure 24b, where Cr-V rich nitrides present higher Volta potential than the Cr-Fe rich nitrides. Therefore under corrosion conditions the latter ones would likely show more tendency to corrode[55] because corrosion typically starts where the Volta potential is lower, i.e., lower relative nobility.[55, 89-91]

Figure 25a presents a topographic image taken in PeakForce ® mode in MQ- water after one hour of immersion, showing the microstructure of the metal alloy; a matrix containing particles with diameters below 3 µm located above and below the matrix surface.

The adhesion image in Figure 25b displays the adhesion force generated by interactions between the metal alloy and the silicon tip, where the matrix displays high adhesion with values between 25 and 28 nN. The particles contained in the matrix present two different adhesion values, one very similar to the matrix, and other with values between 19 and 25 nN. Furthermore, there are regions around the particles displaying lower adhesion. All the differences in adhesion observed in Figure 25b suggest that there are different compositions in the alloy which agree with the microstructure described below in Figure 24 and Table 5.

50

Figure 25. Peak Force ® QNM images of FeCrVN tool alloy using a silicon tip, in water (a-b) and NaCl solution, 0.1 M (c-h). This series of images in NaCl display how the small particles are gradually generated in the salt solution, until it is impossible to continue imaging due the streakiness presented in (g and h), which is typical of particles adhering to the tip. The closed and open white circles correspond to Cr-V rich and Cr-Fe rich nitride particles, respectively.

51

Vanadium and nitrogen content in the tool alloy are important parameters to consider because they form vanadium nitride (VN), which is characteristic of low adhesion.[92] The origin of this low adhesion is still not fully understood, but it has been reported that VN is prone to lower adhesion than TiN[93] and the surface VN oxide formed at high temperature leads to crystallographic planes that are easy to shear.[94] Therefore, it is most likely that the particles with lower adhesion in Figure 25b correspond to Cr- V rich nitrides and the particles with higher adhesion to Cr-Fe rich nitrides.

In relation to the lower adhesion observed around the rich nitrides in Figure 25b, a possible explanation relates the low adhesion in these corona regions with a different microstructure or a higher surface charge than in the rest of the alloy. This effect is not fully understood, but is probably related to the transition in composition between the particles and the matrix.

The tool alloy surface was immersed in a NaCl solution (0.1 M), and after a waiting time of 60 minutes the sample was scanned three times with PeakForce ® QNM mode, where every image took 8 minutes. After this first scan, Figure 25c shows a surface topography almost identical to the earlier case in water (Figure 25a), but small particles appear on the surface mostly on the matrix and around the nitride particles (especially, around the Cr-V rich nitride particles ). In contrast, the adhesion data (Figure 25d) present important changes where the adhesion decreased dramatically in the three regions. This reduction was larger for the matrix where the adhesion decreased dramatically from 19.98 to 0.95 nN, becoming almost indistinguishable from the Cr-V rich nitride particles.

Figures 25e-f show that the topography and adhesion maps generated after the second scan are rather similar to the previous case (Figures 25c-d). However, in Figure 25f the contrast difference between the Cr-V rich nitride particles and matrix decreased even more, and specific areas located

52 in the matrix and around nitride particles display very low adhesion - around 1 nN.

After a further waiting time of 22 minutes (total immersion 82 minutes), the surface was scanned a third time, and Figures 25g-h present important differences with respect to Figures 25e-f. First, the horizontal artifacts displayed in Figures 25e-f leads to loss of quality on the images. Second, the overall adhesion in Figure 25f decreases, likely as a consequence of the artifacts observed on the images. In all the images obtained in NaCl solution, small spot features are observed on the surface (Figures 25c, e and g), which are mostly on the matrix and around the nitrides (especially around the Cr-V rich nitride particles). After 90 minutes of immersion in NaCl, the probability of attachment of the spots to the cantilever tip increases which prevents imaging. The generation of horizontal artifacts in imaging due to loose particles attaching to the tip is well described in the literature.[95]

With the NaCl solution in the system, new forces may arise contributing to the adhesion, which generally decrease in Figures 25d and f. This is likely a consequence of hydration forces originated by cations absorbed on the surfaces. To understand these changes, the surface chemistry of the silicon tip and the alloy has to be considered. Both surfaces oxidize in contact with oxygen, and the oxide evolves after water exposure.

Silicon forms silicon oxide with oxygen exposure. Afterwards it gets hydrolyzed by water and forms silanol groups on the surface. These groups ionize, generating a negative charge on the surface according to:[96, 97]

+ ↔ +

53

Therefore, the oxide layer on the silicon tip is negatively charged, and when it is immersed in the NaCl solution it attracts and accumulates hydrated Na + ions.

The alloy surface has also a native oxide-like passive layer which is formed in contact with water. However, the structure of this native oxide layer is different than the silicon one, and is formed by a gel-like structure with a large amount of bound water with different bridging structures.[87, 98] In NaCl, the Cl - ions start to attack weak sites in the native oxide, creating metal-salt or corrosion products,[99] which can be assumed to be negatively charged and thus also attract and accumulate Na + ions.

During a PeakForce ® measurement, the tip and the alloy achieve contact and the interaction between the hydrated Na + ions accumulated on the surfaces generates a repulsion or a hydration force[2, 12, 100, 101] that could be responsible for the decrease in adhesion observed in Figures 25d and f. This adhesion force appears to be region specific, and thus related to the microstructure of the experimental tool alloy.

The corrosion of this tool alloy generally leads to metal dissolution and pitting formation.[55] However, in this work, small protrusions produced by metal salts or corrosion products appear on the surface, which are mostly correlated with small spots of very low adhesion displayed in the adhesion images. It is hypothesized that these areas are related to pre-pitting events due to the action of the NaCl solution that generates the formation of probably some kind of oxides, hydroxides, and/or chloride complex compounds.

Chromium and nitrogen are two elements that are commonly used to enhance the passivity of stainless steel[87, 102, 103] leading to a high corrosion resistance. The matrix is the region of the tool alloy where the chromium content is lower in the passive layer, thus this area is weaker

54 against Cl - attack,[55, 87, 102, 103] and has more tendency to form metal salts or corrosion products. Moreover, these protrusions appear to concentrate around the nitrides which correspond to regions in the matrix with deficiency of Cr and N (Figure 25f).[55, 87, 104-106] Likely, pre- pitting events preferentially take place in these areas, resulting in higher concentration of negatively charged corrosion products that lead to stronger hydration forces and thus lower adhesion, as described previously. However, a new study is necessary to verify this hypothesis where the challenge is the in-situ chemical analysis of such small features on the tool alloy surface.

Adhesion, topography and chemical composition provide us with relevant information about microstructure and phenomena such as corrosion, but when deformation and Young’s modulus were expected to provide additional information, a difficulty arose. PeakForce ® QNM provides the Young’s modulus of a material by deforming the studied substrate with the cantilever tip applying a threshold force. This threshold will increase with the hardness of the studied material. Therefore, by using an AFM tip softer than the studied surface, the tip tends to deform and even to break (Figure 26) without reaching the necessary threshold force to deform the surface. Consequently, no relevant information based on the Young’s modulus and deformation of the tool alloy was obtained by using an AFM silicon tip in PeakForce ® QNM mode.

55

Figure 26. SEM image of a silicon tip after (a) and before (b) a PeakForce ® QNM measurement applying high force.

4.6 Nanomechanical Properties of Human Skin

The limitation for obtaining the Young’s modulus and deformation faced in Paper V is of course lifted for softer materials. For unambiguous studies the material of the AFM tip should be harder than the surface scanned. The rigidity of the chosen cantilever has to be selected carefully because first, the applied force by the AFM tip has to be high enough to induce a relevant deformation on the surface. Furthermore the cantilever laser-reflection must always be located in the linear regime of the AFM-photodetector, otherwise the load applied would incorporate a high error. Therefore, it is necessary to use a cantilever made of a harder material than the sample, where the cantilever should be of the right rigidity.

The stratum corneum (SC - see Section 3.3 ) has mechanical properties that contribute significantly to its special functions as the outer protective layer such as skin barrier and photoprotection. Of particular interest are biointeractions, for example how much deflection is caused as a human hair interacts with the skin, but there are limited data available in the area.[56- 58] In order to gain a better understanding of the mechanical properties of

56 the SC, surface indentation and PeakForce ® QNM measurements have been performed.

Indentation[107-109] is a commonly used technique to study the mechanical properties of materials (such as hardness and elastic modulus) where the load and the indentation depth are obtained simultaneously during the load and unloading process. Indentation can be designed for use with very small indenters, applying very small loads in the range of nanometres, causing nanometre-indentation depths. In this way was born nanoindentation[109, 110] where the atomic force microscopy became a very important asset for this type of measurement.

The aim of this study was to investigate the nanomechanical properties and the magnitude of the force needed to induce elastic/plastic deformation of the outer layer of skin (SC) by combining AFM imaging and force/nanoindentaion measurements.

Figure 27. AFM PeakForce QNM images with Topography (a-c) and DMT modulus histograms (c-e) at scan sizes of 20µm, 10µm and 5µm respectively . The vertical axes in (a-c) indicate the maximum difference between the darkest and the brightest parts.

57

Figures 27 (a-f) shows the heterogeneous topography of the outer part of the SC layer obtained after a PeakForce ® QNM measurement. The reduced Young’s moduli ( E* in Eqs. 21-22)[111] were obtained by fitting the contact mechanical theory of Derjaguin, Muller and Toropov (DMT) to the force curve obtained at each pixel. The extracted values which were obtained at the nanoscale can be plotted as histograms (see Figures 27 d-f), where the mean reduced Young’s modulus obtained was 0.51 GPa.

Afterwards, the values of the reduced Young’s moduli were transformed into Young’s moduli (E SC ) for comparison purposes according to Eq. 22 (this equation was derived from Eq. 21[112] by ignoring the contribution of the silicon because of its much higher relative stiffness):

ν ν (21) ∗ = +

ν (22) ∗ = where the subscript S corresponds to the silicon, SC to the stratum corneum and ν to Poisson’s ratio ( νSC = 0.48) . [113]

Therefore, the reduced Young´s modulus of 0.51 GPa was transformed into a value of 0.39 GPa, which is consistent with the relative high stiffness of the SC reported in literature.[56, 113, 114]

Afterwards, the deformation of the SC surface was studied by normal force measurements (nanoindentation) using the same silicon cantilever. First, the topography of a 1 m2 region of the surface was obtained. Afterwards, nanoindentation was performed, and finally the surface was imaged again in order to detect any type of permanent deformation. This nanoindentation experiment was performed at different locations by applying 2.5 N once and then twice applying 4.8 N.

58

Figure 28. PeakForce topography images (a-d from left to right) of SC using a sharp tip. (a) 1 µm scan of neat SC sample (b) 1 µm scans after force indentation using a sharp tip at 2.5 µN (c) 1 µm scan after force indentation using a sharp tip at 4.8 µN (d) 1 µm scan after repeat force indentation in a new location using a sharp tip at 4.8 µN. The vertical axes indicate the maximum difference between the darkest and the brightest parts in the topographic images.

Figure 28a presents the topographic image before indentation and Figure 28b after applying 2.5 N, and it can be seen that no permanent deformation occurred. Afterwards, a larger force of 4.8 N was applied in two adjacent spots and Figure 28d demonstrates that this nanoindentation force was enough to produce permanent deformation in both areas with similar shape and depth.

Figure 29 shows the characteristic normal force curves obtained from the nanoindentation measurements from which mechanical properties of the SC can also be extracted.

Figure 29 . (a) Force indentation using a sharp tip on approach at a ramp size of 2 µm (1 Hz). Maximum applied force 2.5 µN (black curve) and 4.8 µN (grey curve). Inset: Indention curve on retraction for 4.8 µN maximum applied force (grey curve). (b) Indentation profile (4.8 µN) obtained from the cross section of the topography image that corresponds to the permanent deformation of the SC (Figure 28c), where the tip image (which corresponds to a SEM image of a cantilever tip) in inset indicates maximum indentation depth.

59

The black curve in Figure 29a is obtained by applying a load of 2.5 N, which causes an indentation depth of approximately 125 nm. This indentation has mostly an elastic deformation since no permanent deformation was observed in Figure 28b. On the other hand, the grey curve generated with an applied load of 4.8 N, which corresponds to an applied pressure of 0.47 GPa ( Paper VI ), presents an even larger indentation depth, of around 200 nm. Figures 28 c-d show that this value is large enough to produce a permanent deformation on the SC. On retraction, the elastic recovery of the SC reduced the indentation depth to the permanent deformation of around 37 nm observed in the cross section of the topography in Figure 29b. When the tip contacts a new spot on the SC surface, the surface may undergo both plastic and elastic deformation. On retraction the material tries to recover its original shape, but is partially prevented because of the permanent or plastic deformation. The only recovery undergone by the material is due to its elastic relaxation, which corresponds to the apparent indentation depth on retraction.[108] Therefore, a way to extract the permanent deformation is by taking the difference of the apparent indentation depths of the force curves measured during approach and retraction (Figure 29a). This treatment generates a value of 45 nm, which is quite consistent with the 37 nm observed in the cross section of Figure 29b and thus confirms the elastic-plastic behavior of the SC.

4.7 A Novel AFM Probe: The Single Hair Fibre Probe

Another advantage of AFM is the possibility to use different probes in order to understand the tip-sample interactions. A material which often interacts with SC is wool, the fibres of which can cause discomfort. Our own hair, or that of other individuals, can also cause discomfort and to be able to treat this issue it is necessary to understand how the SC deforms locally in response to interaction with such a “probe”.

60

Thus as part of Paper VI , a nanoindentation AFM study was performed using a novel probe, which was designed from only a single hair fibre (Figure 30).

Figure 30. SEM images of (a) a single vertical hair fibre attached to the end of a cantilever pre-cut using sharp scissors and (b) the attached fibre cut after using FIB.

The indentation measurements were initiated by imaging the topography of the SC sample using a silicon tip. Afterwards, a single force measurement was performed in the same area using the novel hair probe at an applied load of 4.8 N. Subsequently, the surface area was imaged again with the silicon tip.

Figure 31. PeakForce ® topography images. 100 µm 2 scans of SC (a) before and (b) after force measurement using the single hair fibre indenter.

Figure 31 presents the topographic images obtained before and after the nanoindentation measurement, and a comparison of both images suggests

61 that there is no apparent alteration of the surface. This result might be a bit unexpected considering the permanent deformation observed in Figure 28, however there are more parameters to be considered, such as the tip radius and effective area of contact. Thus, in order to further analyze this nanoindentation experiment, the force distance profile should also be considered.

Figure 32. Force indentation (grey curve) on retraction using single hair fibre probe indenter on SC (ramp size of 2 µm and rate 1 Hz). The black curve represents reference measurement on retraction of the single hair fibre probe against a bare mica substrate.

Figure 32 displays an example of normal force curves on retraction obtained during the nanoindentation AFM measurement using the hair fibre probe. The grey curve corresponds to the measurement against the SC surface (negative separation corresponds to the indentation depth) and the black one against a stiff mica surface. The sharp 90º angle in the force upon contact is characteristic behavior of a stiff non-deformable substrate. The analysis of the force curve is performed on the assumption that constant compliance is achieved, and the linearity of the constant compliance region strongly suggests that the probe does not deform. If it does then the deformation is pseudo linear and thus the deformation is taken into account in the deflection sensitivity which is then employed for the interaction with SC. In contrast, the grey curve shows an indentation depth of around 550 nm and a nonlinearity which is presumably related to the viscoelastic nature

62 of the SC.[115] Comparison of the maximum indentation depths of the hair- SC force profiles during approach and retraction shows that they differ by about 164 nm, which corresponds to the plastic deformation that is observed only during approach (Figure 32). However, in contrast to the apparent plastic deformation observed in Figures 28c-d, this 164 nm of plastic deformation are not visible in Figure 31 because the deformation is spread over a larger area, which reflects the much larger radius of the hair probe (at least in one dimension). At first, the difference in penetration depths observed for the different probe diameters is difficult to justify, but it can be resolved by considering that the Young’s moduli are actually a function of the interaction size.

Three different combinations of Hertzian mechanical models were assumed in Paper VI for the approximation of appropriate Young’s moduli for the hair probe against the SC. The first model (Figure 33a) is based on a sphere against an elastic half space, where the hair probe is assumed to behave as a sphere of radius corresponding to the large radius of curvature of the probe. This model simplifies the situation because it does not account for the large topographic variation of the SC, and generates a Young’s modulus of 0.98 MPa and a maximum applied pressure of 0.09 MPa. The second (Figure 33b) and the third (Figure 33c) models take into consideration the topography of the SC, assuming that the hair probe contacts the SC surface, deforming it at two levels; both individual features and a more general flat surface of the SC, simultaneously. These interactions can be coarsely approximated by a sphere-sphere contact (i.e. the hair probe interacting with a SC asperity) in series with a sphere-flat contact (i.e. the hair probe interacting with the underlying SC surface). The differences between second and third models are that in the second it is assumed the same Young’s modulus for the SC asperity and the SC surface and in the third they are different ( Paper VI ).

63

Figure 33 . Schematic showing the different Hertzian models applied between the hair probe (sphere) and SC (surface) before (left) and after indentation (right). (a) A sphere interacting with an elastic surface, where EH is the Young’s modulus of the hair and ESC2 is a fitting parameter that corresponds to the Young’s modulus of the SC. (b) A combination of a sphere interacting against an elastic sphere (SC-asperity), and a sphere interacting against an elastic surface, where ESC3 is a fitting parameter that corresponds to both the Young’s modulus of the SC-asperity and the underlying SC surface. (c) A combination between a sphere interacting against an elastic sphere

(SC-asperity), and a sphere interacting against an elastic surface, where ESC1 is the Young’s modulus of SC-asperity (obtained from Figure 27) and ESC4 is a fitting parameter that corresponds to the Young’s modulus of the SC.

64

Finally, the second and third models generate Young’s moduli of 2.79 and 1.01 MPa, and maximum applied pressure of 0.70 and 17.7 MPa, respectively.

It is stressed that there is no good contact model for the geometry employed, and the above values are not more than an approximation, but should be of the right order of magnitude. The most important development here is that it has been identified the scale of applied pressures that are appropriate to describe hair induced deformation of the SC.

65

5. Conclusions

The combination of a nanotribology approach with corrosion showed that this synergy can lead to new insights and research avenues. Significant technique development in atomic force microscopy (AFM) was achieved which enabled investigation of metal surfaces in a new light. Such expertise has the ability to be applied over a wide variety of fields-for example in understanding interactions in ionic liquid, as well as how the stratum corneum deforms in response to probes of different sizes.

The long ranged interactions between pairs of materials have been studied by AFM in air and after immersion into ethylammonium nitrate (EAN). In air the tribological properties depends on the long ranged interactions present between them (van der Waals and capillary forces), which are significantly affected by the environmental conditions, material properties and number/sequence of measurements. However, in EAN the ionic liquid suppress the long ranged interactions and reduces the friction and adhesion, which become only affected by the roughness of the surfaces in contact. When the surfaces were approaching each other (in EAN) with a relative fast speed, there was a repulsion that precluded contact because of hydrodynamic forces. The analysis of the hydrodynamic force provided an interfacial viscosity which was almost a 3-fold reduction compared to the bulk viscosity. It is assumed that this low viscosity value is due to a lamellar ordering of the EAN at the interface that reduces the resistance to sliding.

Tribological property mapping (TPM) is a new technique that generates friction coefficient and especially contact adhesion maps, by using lateral atomic force microscopy images. TPM provides tribological information which can be related to the sample microstructure. This technique has been

66 applied to two different metal alloys showing their heterogeneous microstructure which is formed by hard particles embedded in a softer matrix. In case of the CoCrMo alloy, it was shown that the chromium oxide and the phosphate buffer saline solution are responsible for the low friction coefficient and contact adhesion.

Another AFM technique, PeakForce ® QNM was used in water to study the surface microstructure of an experimental FeCrVN tool alloy, and in NaCl solution to understand the influence of microstructure in the corrosion initiation of this tool alloy. The results show that the low adhesion is attributed to the vanadium and nitrogen contents, and the corrosion initiation or prepitting depends on the chromium content. Moreover, this prepitting originates in specific areas leading to small particles with low adhesion.

Force and PeakForce ® QNM measurements with a Si tip were also used to understand the mechanical properties of the human stratum corneum (SC). The SC is heterogeneous at the nanoscale where substancial permanent deformation occurs at an applied pressure of about 0.47 GPa.

A novel probe was developed by attaching a single hair probe to the end of a cantilever to perform nanoindentation measurements on the SC sample. These measurements show that the SC is extremely elastic at the nanoscale, and that the lateral deformation regulates the effective Young’s modulus. Three models were proposed in order to calculate the pressure applied for the novel hair probe in SC.

A new approximation referred as Hybrid method was originally developed to obtain the torsional spring constants of rigid cantilevers where the power torsional thermal spectra are difficult to obtain because of the high resonance frequency and low signal/noise ratio. This method that combines the normal spring constant with the length and the width of the cantilever to

67 obtain the torsional spring constant can be used for the determination of any type of rectangular cantilever with or without coating.

68

6. Acknowledgments

First of all, I would like to express my sincere gratitude to my supervisors Prof. Mark Rutland and Jinshan Pan, for giving me the opportunity to do my PhD in KTH. Especially to Prof. Mark Rutland for his support and guidance, as well as, the attitude to find always the positive side of any situation has really inspired me. I also want to thanks Emily Cranston and Esben Thormann (co-supervisors) for patiently teaching me to use the AFM and help me to become the professional that I am now. Thanks to my friends Deborah Wakeham and Beatrice Johannson for helping with my Thesis. There are so many people who have contributed in their own way to the completion of my PhD, coauthors, friends, and colleagues. They work not only at the Division of Surface and Corrosion Science, but also at SP, Nanologica AB, L´Oreal and Stockholm University. You know who you are and even I don’t name you will be always with me wherever I go. I want to especially thank my dear friend Rodrigo Robinson, for being willing to listen and share complaints and for giving me good advices. SSF (Swedish foundation for Strategic Research) is gratefully acknowledged for sponsoring this project. Alfredo Metere at Molworx (www.molworx.com ) is also acknowledged for the design of Figure 3. My sweet Linnéa Bengstsson, thank you for your love, support, faith and patience. I could never have made it without you. The last and not the least, my family because they have always believed in and supported me. Os quiero y os echo de menos, siempre estaréis conmigo esté donde esté.

69

7. References

1. J.N. Israelachvili, Intermolecular and Surface Forces. 3rd ed. 2008: Academic Press. 2. B. Cappella and G. Dietler, Force-Distance Curves by Atomic Force Microscopy. Reports, 1999. 34 (1-3): p. 1-104. 3. H.C. Hamaker, The London-van der Waals Attraction between Spherical Particles. Physica, 1937. 4(10): p. 1058-1072. 4. A.A. Feiler, L. Bergstrom and M.W. Rutland, Superlubricity Using Repulsive van der Waals Forces. Langmuir, 2008. 24 (6): p. 2274-2276. 5. A. Milling, P. Mulvaney and I. Larson, Direct Measurement of Repulsive van der Waals Interactions Using an Atomic Force Microscope. Journal of Colloid and Interface Science, 1996. 180 (2): p. 460-465. 6. A. Meurk, P.F. Luckham and L. Bergstrom, Direct Measurement of Repulsive and Attractive van der Waals Forces between Inorganic Materials. Langmuir, 1997. 13 (14): p. 3896-3899. 7. S.-W. Lee and W.M. Sigmund, Repulsive van der Waals Forces for Silica and Alumina. Journal of Colloid and Interface Science, 2001. 243 (2): p. 365-369. 8. H.-J. Butt and M. Kappl, Capillary Forces, in Surface and Interfacial Forces. 2010, Wiley-VCH Verlag GmbH & Co. KGaA. p. 127-161. 9. Y.I. Rabinovich, J.J. Adler, A. Ata, R.K. Singh and B.M. Moudgil, Adhesion between Nanoscale Rough Surfaces: II. Measurement and Comparison with Theory. Journal of Colloid and Interface Science, 2000. - 232 (- 1): p. - 24. 10. Y.I. Rabinovich, J.J. Adler, A. Ata, R.K. Singh and B.M. Moudgil, Adhesion between Nanoscale Rough Surfaces: I. Role of Asperity Geometry. Journal of Colloid and Interface Science, 2000. 232 (1): p. 10- 16. 11. A.A. Feiler, P. Jenkins and M.W. Rutland, Effect of Relative Humidity on Adhesion and Frictional Properties of Micro- and Nano-scopic Contacts. Journal of Adhesion Science and Technology, 2005. 19 (3-5): p. 165-179. 12. J.J. Valle-Delgado, J.A. Molina-Bolívar, F. Galisteo-González, M.J. Gálvez-Ruiz, A. Feiler and M.W. Rutland, Hydration Forces between Silica Surfaces: Experimental Data and Predictions from Different Theories. The Journal of Chemical Physics, 2005. 123 (3): p. -. 13. W.A. Ducker, T.J. Senden and R.M. Pashley, Measurement of Forces in Liquids Using a Force Microscope. Langmuir, 1992. 8(7): p. 1831-1836. 14. H.-J. Butt and M. Kappl, Hydrodynamic Forces, in Surface and Interfacial Forces. 2010, Wiley-VCH Verlag GmbH & Co. KGaA. p. 163-187.

70

15. D.Y.C. Chan and R.G. Horn, The Drainage of Thin Liquid Films between Solid Surfaces. J. Chem. Phys., 1985. 83 : p. 5311-5324. 16. B. Bhushan, Introduction to Tribology. 2002: John Wiley and Sons. 17. G. Amontons, Proceedings of the French Royal Academy of Science, 1699: p. 257−282. 18. D. Downson, History of Tribology. 1979, London: Longman. 19. B. Derjaguin, Molekulartheorie der Äußeren Reibung. Zeitschrift für Physik A Hadrons and Nuclei, 1934. 88 (9): p. 661-675. 20. R. Álvarez-Asencio, J. Pan, E. Thormann and M. Rutland, Tribological Properties Mapping: Local Variation in Friction Coefficient and Adhesion. Tribology Letters, 2013. 50 (3): p. 387-395. 21. J. Gao, W.D. Luedtke, D. Gourdon, M. Ruths, J.N. Israelachvili and U. Landman, Frictional Forces and Amontons’ Law: From the Molecular to the Macroscopic Scale. The Journal of Physical Chemistry B, 2004. 108 (11): p. 3410-3425. 22. A.M. Homola, J.N. Israelachvili, M.L. Gee and P.M. McGuiggan, Measurements of and Relation Between the Adhesion and Friction of Two Surfaces Separated by Molecularly Thin Liquid Films. Journal of Tribology, 1989. 111 (4): p. 675-682. 23. M.A. Plunkett, A. Feiler and M.W. Rutland, Atomic Force Microscopy Measurements of Adsorbed Polyelectrolyte Layers. 2. Effect of Composition and Substrate on Structure, Forces, and Friction. Langmuir, 2003. 19 (10): p. 4180-4187. 24. G. Bogdanovic, F. Tiberg and M.W. Rutland, Sliding Friction between Cellulose and Silica Surfaces. Langmuir, 2001. 17 (19): p. 5911-5916. 25. A.A. Feiler, J. Stiernstedt, K. Theander, P. Jenkins and M.W. Rutland, Effect of Capillary Condensation on Friction Force and Adhesion. Langmuir, 2006. 23 (2): p. 517-522. 26. A. Berman, C. Drummond and J. Israelachvili, Amontons’ Law at the Molecular Level. Tribology Letters, 1998. 4(2): p. 95-101. 27. S. Yamada and J. Israelachvili, Friction and Adhesion Hysteresis of Fluorocarbon Surfactant Monolayer-Coated Surfaces Measured with the . The Journal of Physical Chemistry B, 1998. 102 (1): p. 234-244. 28. C.M. Mate, Tribology on the Small Scale : A Bottom Up Approach to Friction, Lubrication, and Wear. 2008, Oxford: Oxford University Press. xiii, 333 p. 29. S. Izabela, C. Michael and W.C. Robert, Recent Advances in Single- Asperity Nanotribology. Journal of Physics D: Applied Physics, 2008. 41 (12): p. 123001. 30. Handbook of Micro/Nano Tribology, ed. B. Bhushan. 1999: CRC Press.

71

31. F.P. Bowden and D. Tabor, Mechanism of Metallic Friction. Nature 1942. 150 : p. 197-199. 32. G. Binnig, C.F. Quate and C. Gerber, Atomic Force Microscope. Physical Review Letters, 1986. 56 (9): p. 930-933. 33. J. Ralston, I. Larson, M.W. Rutland, A.A. Feiler and M. Kleijn, Atomic Force Microscopy and Direct Surface Force Measurements - (IUPAC Technical Report). Pure and Applied Chemistry, 2005. 77 (12): p. 2149- 2170. 34. J.R. Smith, C. Larson and S.A. Campbell, Recent Applications of SEM and AFM for Assessing Topography of Metal and Related Coatings - a Review. Transactions of the Institute of Metal Finishing, 2011. 89 (1): p. 18-27. 35. M. Sababi, J. Kettle, H. Rautkoski, P.M. Claesson and E. Thormann, Structural and Nanomechanical Properties of Paperboard Coatings Studied by Peak Force Tapping Atomic Force Microscopy. ACS Applied Materials & Interfaces, 2012. 4(10): p. 5534-5541. 36. B. Foster, New Atomic Force Microscopy (AFM) Approaches Life Science Gently, Quantitatively, and Correlatively. American Laboratory, 2012. 44 : p. 4. 37. W.A. Ducker, T.J. Senden and R.M. Pashley, Direct Measurement of Colloidal Forces Using an Atomic Force Microscope. Nature, 1991. 353 (6341): p. 239-241. 38. H.-J. Butt, Measuring Electrostatic, van der Waals, and Hydration Forces in Electrolyte Solutions with an Atomic Force Microscope. Biophysical Journal, 1991. 60 (6): p. 1438-1444. 39. B.V. Derjaguin, Kolloid. Zh., 1934. 69 : p. 155-164. 40. T. Pettersson, N. Nordgren, M.W. Rutland and A.A. Feiler, Comparison of Different Methods to Calibrate Torsional Spring Constant and Photodetector for Atomic Force Microscopy Friction Measurements in Air and Liquid. Review of Scientific Instruments, 2007. 78 (9): p. 093702. 41. J.E. Sader, I. Larson, P. Mulvaney and L.R. White, Method for the Calibration of Atomic Force Microscope Cantilevers. Review of Scientific Instruments, 1995. 66 (7): p. 3789-3798. 42. C.P. Green, H. Lioe, J.P. Cleveland, R. Proksch, P. Mulvaney and J.E. Sader, Normal and Torsional Spring Constants of Atomic Force Microscope Cantilevers. Review of Scientific Instruments, 2004. 75 (6): p. 1988-1996. 43. J.E. Sader, J.W.M. Chon and P. Mulvaney, Calibration of Rectangular Atomic Force Microscope Cantilevers. Review of Scientific Instruments, 1999. 70 (10): p. 3967-3969.

72

44. J.E. Sader, Susceptibility of Atomic Force Microscope Cantilevers to Lateral Forces. Review of Scientific Instruments, 2003. 74 (4): p. 2438- 2443. 45. J.E. Sader, Frequency Response of Cantilever Beams Immersed in Viscous Fluids with Applications to the Atomic Force Microscope. Journal of Applied Physics, 1998. 84 (1): p. 64-76. 46. C.P. Green and J.E. Sader, Torsional Frequency Response of Cantilever Beams Immersed in Viscous Fluids with Applications to the Atomic Force Microscope. Journal of Applied Physics, 2002. 92 (10): p. 6262-6274. 47. L.A. Giannuzzi and F.A. Stevie, Introduction to Focused Ion Beams. 2005, New York: Springer. 48. S.G. Luengo, A. Galliano and C. Debief, Aqueous Lubrication in Cosmetic, in Aqueous Lubrication D.N. Spencer, 2014, L´Oréal: Zurich. 49. S. Hatami, A. Nafari, L. Nyborg and U. Jelvestam, Related Surface Properties of Powder Metallurgical Tool Steels Alloyed with and without Nitrogen. Wear, 2010. 269 (3-4): p. 229-240. 50. I. Heikkilä, E. Van der Heíde, E.D. Stam, H. Giraud, G. Lovato, N. Akdut, F. Clarysse and P. Caenen, Tool Material Aspects in Forming of Stainless Steel with Easy-to-Clean , in Innovations in Metal Forming. 2004. 51. C. Valero Vidal and A. Igual Muñoz, Electrochemical Characterisation of Biomedical Alloys for Surgical Implants in Simulated Body Fluids. Corrosion Science, 2008. 50 (7): p. 1954-1961. 52. E. Bettini, T. Eriksson, M. Boström, C. Leygraf and J. Pan, Influence of Metal Carbides on Dissolution Behavior of Biomedical CoCrMo alloy: SEM, TEM and AFM Studies. Electrochimica Acta, 2011. 56 (25): p. 9413-9419. 53. J.A. Disegi, R.L. Kennedy and R. Pilliar, Cobalt-Base Alloys for Biomedical Applications. 1999, Virginia: ASTM STP1365. 54. J.J. Jacobs and L.T. Craig, Alternative Surfaces in Total Joint Replacement. 1998, Fredericksburg, VA: ASTM. 55. M. Sababi, S. Ejnermark, J. Andersson, P.M. Claesson and J. Pan, Microstructure Influence on Corrosion Behavior of a Fe–Cr–V–N Tool Alloy Studied by SEM/EDS, Scanning Kelvin Force Microscopy and Electrochemical Measurement. Corrosion Science, 2013. 66 (0): p. 153- 159. 56. Y. Yuan and R. Verma, Measuring Microelastic Properties of Stratum Corneum. Colloids and Surfaces B: Biointerfaces, 2006. 48 (1): p. 6-12. 57. A. Potter, S.G. Luengo, R. Santoprete and B. Querleux, Stratum Corneum , in Skin Moisturization. 2009. p. 259-278.

73

58. W. Tang and B. Bhushan, Adhesion, Friction and Wear Characterization of Skin and Skin Cream Using Atomic Force Microscope. Colloids and Surfaces B: Biointerfaces, 2010. 76 (1): p. 1-15. 59. L. Norlén, Stratum Corneum Keratin Structure, Function and Formation – a Comprehensive Review. International Journal of Cosmetic Science, 2006. 28 (6): p. 397-425. 60. M. Palacio and B. Bhushan, A Review of Ionic Liquids for Green Molecular Lubrication in . Tribology Letters, 2010. 40 (2): p. 247-268. 61. T.L. Greaves and C.J. Drummond, Protic Ionic Liquids: Properties and Applications. Chemical Reviews, 2008. 108 (1): p. 206-237. 62. O. Werzer, G.G. Warr and R. Atkin, Compact Poly(ethylene oxide) Structures Adsorbed at the Ethylammonium Nitrate-Silica Interface. Langmuir, 2011. 27 (7): p. 3541-3549. 63. P. Niga, D. Wakeham, A. Nelson, G.G. Warr, M. Rutland and R. Atkin, Structure of the Ethylammonium Nitrate Surface: An X-ray Reflectivity and Vibrational Sum Frequency Spectroscopy Study. Langmuir, 2010. 26 (11): p. 8282-8288. 64. K. Ueno, M. Kasuya, M. Watanabe, M. Mizukami and K. Kurihara, Resonance Shear Measurement of Nanoconfined Ionic Liquids. Physical Chemistry Chemical Physics, 2010. 12 (16): p. 4066-4071. 65. Y. Mo, W. Zhao, M. Zhu and M. Bai, Nano/Microtribological Properties of Ultrathin Functionalized Imidazolium Wear-Resistant Ionic Liquid Films on Single Crystal Silicon. Tribology Letters, 2008. 32 (3): p. 143- 151. 66. W. Zhao, L. Wang, M. Bai and Q. Xue, Micro/Nanotribological Behaviors of Ionic Liquid Nanofilms with Different Functional Cations. Surface and Interface Analysis, 2010. 43 (6): p. 945-953. 67. J. Pu, D. Huang, L. Wang and Q. Xue, Tribology Study of Dual-Layer Ultrathin Ionic Liquid Films with Bonded Phase: Influences of the Self- Assembled Underlayer. Colloids and Surfaces A: Physicochemical and Engineering Aspects, 2010. 372 (1-3): p. 155-164. 68. M. Palacio and B. Bhushan, Molecularly Thick Dicationic Ionic Liquid Films for Nanolubrication. Journal of Vacuum Science & Technology A: Vacuum, Surfaces, and Films, 2009. 27 (4): p. 986-995. 69. D. Wakeham, R. Hayes, G.G. Warr and R. Atkin, Influence of Temperature and Molecular Structure on Ionic Liquid Solvation Layers. The Journal of Physical Chemistry B, 2009. 113 (17): p. 5961-5966. 70. R. Atkin and G.G. Warr, Structure in Confined Room-Temperature Ionic Liquids. The Journal of Physical Chemistry C, 2007. 111 (13): p. 5162- 5168.

74

71. J.A. Smith, O. Werzer, G.B. Webber, G.G. Warr and R. Atkin, Surprising Particle Stability and Rapid Sedimentation Rates in an Ionic Liquid. Journal of Physical Chemistry Letters, 2010. 1(1): p. 64-68. 72. R. Hayes, G.G. Warr and R. Atkin, At the Interface: Solvation and Designing Ionic Liquids. Physical Chemistry Chemical Physics, 2010. 12 (8): p. 1709-1723. 73. S.S. Perry, Scanning Probe Microscopy Measurements of Friction. Mrs Bulletin, 2004. 29 (7): p. 478-483. 74. R.J. Roark and W.C. Young, Formulas for Stress and Strain. McGraw- Hill. 1975, Nee York. 75. R.G. Cain, S. Biggs and N.W. Page, Force Calibration in Lateral Force Microscopy. Journal of Colloid and Interface Science, 2000. 227 (1): p. 55- 65. 76. D. Faoite, D. Browne, R.F. Chang-Díaz and K. Stanton, A Review of the Processing, Composition, and Temperature-Dependent Mechanical and Thermal Properties of Dielectric Technical Ceramics. Journal of , 2012. 47 (10): p. 4211-4235. 77. E.W. van der Vegte and G. Hadziioannou, Scanning Force Microscopy with Chemical Specificity: An Extensive Study of Chemically Specific Tip−Surface Interactions and the Chemical Imaging of Surface Functional Groups. Langmuir, 1997. 13 (16): p. 4357-4368. 78. J.R. Cannara, M. Eglin and W.R. Carpick, Lateral Force Calibration in Atomic Force Microscopy: A New Lateral Force Calibration Method and General Guidelines for Optimization. Review of Scientific Instruments, 2006. 77 (5): p. 053701. 79. K. Sasaki, Y. Koike, H. Azehara, H. Hokari and M. Fujihira, Lateral Force Microscope and Phase Imaging of Patterned Thiol Self-Assembled Monolayer Using Chemically Modified Tips. Applied Physics A: Materials Science & Processing, 1998. 66 (0): p. 1275-1277. 80. R.L. McMullen and S.P. Kelty, Investigation of Human Hair Fibers Using Lateral Force Microscopy. Scanning, 2001. 23 (5): p. 337-345. 81. J.R. Smith and J.A. Swift, Lamellar Subcomponents of the Cuticular Cell Membrane Complex of Mammalian Keratin Fibres Show Friction and Hardness Contrast by AFM. Journal of Microscopy, 2002. 206 (3): p. 182- 193. 82. F.Z. Sidouni, N. Nurdin, P. Chabrecek, D. Lohmann, J. Vogt, N. Xanthopoulos, H.J. Mathieu, P. Francois, P. Vaudaux and P. Descouts, Surface Properties of a Specifically Modified High-Grade Medical Polyurethane. Surface Science, 2001. 491 (3): p. 355-369. 83. M.D. Levi, Y. Cohen, Y. Cohen, D. Aurbach, M. Lapkowski, E. Vieil and J. Serose, Atomic Force Microscopy Study of the Morphology of

75

Polythiophene Films Grafted onto the Surface of a Pt Microelectrode Array. Synthetic Metals, 2000. 109 (1-3): p. 55-65. 84. L. Ping, Phase Analysis in Steel Using Analytical Transmission Electron Microscopy. 2004, Sandviken: Sandvik Materials Technology. 85. N. Nordgren, P. Eronen, M. Österberg, J. Laine and M.W. Rutland, Mediation of the Nanotribological Properties of Cellulose by Chitosan Adsorption. Biomacromolecules, 2009. 10 (3): p. 645-650. 86. P. Beer, In Situ Examinations of the Friction Properties of Chromium Coated Tools in Contact with Wet Wood. Tribology Letters, 2005. 18 (3): p. 373-376. 87. P. Marcus, Corrosion Mechanisms in Theory and Practice. 3rd ed. Corrosion technology. 2012, Boca Raton, Fla.: CRC. xii, 929 p. 88. M. Nonnenmacher, M.P. O’Boyle and H.K. Wickramasinghe, Kelvin Probe Force Microscopy. Applied Physics Letters, 1991. 58 (25): p. 2921- 2923. 89. N. Sathirachinda, R. Pettersson, S. Wessman and J. Pan, Study of Nobility of Chromium Nitrides in Isothermally Aged Duplex Stainless Steels by Using SKPFM and SEM/EDS. Corrosion Science, 2010. 52 (1): p. 179- 186. 90. M. Li, L.Q. Guo, L.J. Qiao and Y. Bai, The Mechanism of Hydrogen- Induced Pitting Corrosion in Duplex Stainless Steel Studied by SKPFM. Corrosion Science, 2012. 60 (0): p. 76-81. 91. V. Guillaumin, P. Schmutz and G.S. Frankel, Characterization of Corrosion Interfaces by the Scanning Kelvin Probe Force Microscopy Journal of the Electrochemical Society, 2001. 148 (5): p. B163-B173. 92. I. Heikkilä, The Possitive Effect of Nitrogen Alloying of Tool Steels Used In Sheat Forming, in Department of Engineering Sciences. 2013, Uppsala Universitet: Uppsala. 93. L. Vitos, K. Larsson, B. Johansson, M. Hanson and S. Hogmark, An Atomistic Approach to the Initiation Mechanism of Galling. Computational Materials Science, 2006. 37 (3): p. 193-197. 94. A. Glaser, S. Surnev, F.P. Netzer, N. Fateh, G.A. Fontalvo and C. Mitterer, Oxidation of Vanadium Nitride and Titanium Nitride Coatings. Surface Science, 2007. 601 (4): p. 1153-1159. 95. S. Vijendran, H. Sykulska and W.T. Pike, AFM Investigation of Martian Soil Simulants on Micromachined Si Substrates. Journal of Microscopy, 2007. 227 (3): p. 236-245. 96. L.T. Zhuravlev, The Surface Chemistry of Amorphous Silica. Zhuravlev Model. Colloids and Surfaces A: Physicochemical and Engineering Aspects, 2000. 173 (1–3): p. 1-38. 97. R.K. Iler, The Chemistry of Silica : Solubility, Polymerization, Colloid and Surface Properties, and Biochemistry. 1979, New York: Wiley.

76

98. G. Okamoto, Passive Film of 18-8 Stainless Steel Structure and Its Function. Corrosion Science, 1973. 13 (6): p. 471-489. 99. J. Sedriks, A., Corrosion of Stainless Steel. 1996, Virginia: John Wiley & Sons. 100. K. Holmberg, D.O. Shah and M.J. Schwuger, Handbook of Applied Surface and Colloid Chemistry. 2002, Chichester: Wiley. 101. J.N. Israelachvili and R.M. Pashley, Molecular Layering of Water at Surfaces and Origin of Repulsive Hydration Forces. Nature, 1983. 306 (5940): p. 249-250. 102. C.R. Clayton, G.P. Halada and J.R. Kearns, Passivity of High-Nitrogen Stainless Alloys: the Role of Metal Oxyanions and Salt Films. Materials Science and Engineering: A, 1995. 198 (1–2): p. 135-144. 103. S. Yuan, B. Liang, Y. Zhao and S.O. Pehkonen, Surface Chemistry and Corrosion Behaviour of 304 Stainless Steel in Simulated Seawater Containing Inorganic Sulphide and Sulphate-Reducing Bacteria. Corrosion Science, 2013. 74 (0): p. 353-366. 104. C. García, F. Martín, Y. Blanco and M.L. Aparicio, Effect of Ageing Heat Treatments on the Microstructure and Intergranular Corrosion of Powder Duplex Stainless Steels. Corrosion Science, 2010. 52 (11): p. 3725-3737. 105. S.S.M. Tavares, F.J. da Silva, C. Scandian, G.F. da Silva and H.F.G. de Abreu, Microstructure and Intergranular Corrosion Resistance of UNS S17400 (17-4PH) Stainless Steel. Corrosion Science, 2010. 52 (11): p. 3835-3839. 106. S.X. Li, Y.N. He, S.R. Yu and P.Y. Zhang, Evaluation of the Effect of Grain Size on Chromium Carbide Precipitation and Intergranular Corrosion of 316L Stainless Steel. Corrosion Science, 2013. 66 : p. 211- 216. 107. Y.-T. Cheng and C.-M. Cheng, Scaling, Dimensional Analysis, and Indentation Measurements. Materials Science and Engineering: R: Reports, 2004. 44 (4–5): p. 91-149. 108. A.C. Fischer-Cripps, Nanoindentation. 2nd ed. series. 2004, New York ; London: Springer. xxii, 263 p. 109. W.C. Oliver and G.M. Pharr, Measurement of Hardness and Elastic Modulus by Instrumented Indentation: Advances in Understanding and Refinements to Methodology. Journal of Materials Research, 2004. 19 (01): p. 3-20. 110. J.R. Withers and D.E. Aston, Nanomechanical Measurements with AFM in the Elastic LImit. Advances in Colloid and Interface Science, 2006. 120 (1–3): p. 57-67.

77

111. B.V. Derjaguin, V.M. Muller and Y.P. Toporov, Effect of Contact Deformations on the Adhesion of Particles. Journal of Colloid and Interface Science, 1975. 53 (2): p. 314-326. 112. K.L. Johnson, Contact Mechanics. 1985, Cambridge: Cambridge University Press. 113. F. Xu and T. Lu, Introduction of Skin Biothermomechanics and Thermal Pain. 2011: Science Press Beijing and Springer-Verlag Berlin Heidelberg. 114. P.G. Agache, C. Monneur, J.L. Leveque and J. Rigal, Mechanical Properties and Young's Modulus of Human Skin in Vivo. Archives of Dermatological Research, 1980. 269 (3): p. 221-232. 115. T. Jee and K. Komvopoulos, Skin Viscoelasticity Studied in Vitro by Microprobe-Based Techniques. Journal of Biomechanics, 2014. 47 (2): p. 553-559.

78