how water meets a hydrophobic surface:...

152
HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS BY ADELE POYNOR TORIGOE B.S., University of Maryland Baltimore Co, 2001 M.S., University of Illinois at Urbana-Champaign, 2003 DISSERTATION Submitted in partial fulfillment of the requirements for the degree of Doctor of Philosophy in Physics in the Graduate College of the University of Illinois at Urbana-Champaign, 2006 Urbana, Illinois

Upload: phamthuy

Post on 28-Jul-2018

216 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

HOW WATER MEETS A HYDROPHOBIC SURFACE:

RELUCTANTLY AND WITH FLUCTUATIONS

BY

ADELE POYNOR TORIGOE

B.S., University of Maryland Baltimore Co, 2001 M.S., University of Illinois at Urbana-Champaign, 2003

DISSERTATION

Submitted in partial fulfillment of the requirements for the degree of Doctor of Philosophy in Physics

in the Graduate College of the University of Illinois at Urbana-Champaign, 2006

Urbana, Illinois

Page 2: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

ii

Page 3: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

iii

Abstract By definition hydrophobic substances hate water. Water placed on a

hydrophobic surface will form a drop in order to minimize its contact area. What

happens when water is forced into contact with a hydrophobic surface? One theory is

that an ultra-thin low-density region forms near the surface. This depleted region

would have implications in such diverse areas as colloidal self-assembly, and the

boundary conditions of fluid flow. However, the literature still remains divided as to

whether or not such a depleted region exists.

To investigate the existence of this layer, we have employed three surface-

sensitive techniques, time-resolved phase-modulated ellipsometry, surface plasmon

resonance, and X-ray reflectivity. Both ellipsometry and X-ray reflectivity provide

strong evidence for the low-density layer and illuminate unexpected temporal

behavior. Using all three techniques, we found surprising fluctuations at the interface

with a non-Gaussian distribution and a single characteristic time on the order of

tenths of seconds. This information supports the idea that the boundary fluctuates

with something akin to capillary waves.

We have also investigated the dependence of the static and dynamic properties

of the hydrophobic/water interface on variables such as temperature, contact angle,

pH, dissolved gasses, and sample quality, among others, in a hope to discover the root

of the controversy in the literature. We found that the depletion layer is highly

dependent on temperature, contact angle and sample quality. This dependence might

explain some of the discrepancies in the literature as different groups often use

hydrophobic surfaces with different properties.

Page 4: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

iv

Acknowledgement

This Dissertation would not have been possible without the contributions of

many others. I owe a great deal to my advisor Steve Granick. He not only guided me

through the trials and tribulations of this project but he also helped me realize my

goals and become a fully formed physicist. I would like to thank my committee

members, Dr. Yoshi Oono, Dr. Taekjip Ha, and Dr. Ian Robinson, for providing me

with new perspectives on my project. I would also like to thank Dr. Ian Robinson,

along with Dr. Paul Fenter, Dr. Zhan Zhang and Ms. Meng Liang for our great X-ray

reflectivity collaboration. I would never have been able to do it by myself.

In addition, I would like to thank the past and current postdocs in our group. I

am especially indebted to Dr. Sung Chul Bae, without whom I would never have been

able to perform the dynamic SPR measurements; Dr. Ashis Mukhopdhyay, who

taught me everything about ellipsometry; and Dr. Jiang Zhao, who got me started

with SPR in my first year and taught me how to make thiol monolayers. My group

mates have been a great help to me over the years. I am very grateful to Mr. Liang

Hong, who was learning to teach a physicist chemistry; Ms. Yan Yu who helped to

prepare innumerable hydrophobic monolayers; and Dr. Jeff Turner, who did most of

the work for the SFG experiments. I would also like to thank Liangfang Zhang and

Janet Wong for all the useful discussions and friendship. I have had the privilege to

work with several excellent undergraduate collaborators over the years. I would like

to thank Ms. Jolita Šečkutė for her help with the temperature measurements; Mr.

Brendon Hahn for his work on the OTE monolayers; and Ms. Wina Tjen who with

Page 5: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

v

her questions and excitement over three years of working together taught me more

than I thought possible.

I would like to thank my family for all they have done for me; to my parents

Margaret and Edward Poynor, who fostered my curiosity and instilled in me a love of

science; and my brother, Victor Poynor, who kept me from taking myself too

seriously. Finally, I would like to thank my husband, Eugene Torigoe. He was there

for me from the first day of graduate school and together we have come through this

great adventure.

My funding was provided by U.S. Department of Energy, Division of

Materials Science under contract no. DEFG02-91ER45439 with the Frederick Seitz

Materials Research Laboratory at the University of Illinois at Urbana-Champaign

until this year. My research this summer was supported by Dr. Steve Granick’s

Founder Professorship.

Page 6: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

vi

Table of Contents Chapter 1: Introduction ..............................................................................................1

1.1 Theories of Water/Hydrophobic Interfaces .......................................................3 1.2 Simulations of Water/Hydrophobic Interfaces ..................................................6 1.3 Previous Experiments on Water/Hydrophobic Interfaces..................................9 1.4 References........................................................................................................11 1.5 Figures..............................................................................................................14

Chapter 2: Experimental Techniques ......................................................................16 2.1 Ellipsometry.....................................................................................................17 2.2 X-Ray Reflectivity ...........................................................................................25 2.3 Surface Plasmon Resonance ............................................................................28 2.4 Hydrophobic Substrates ...................................................................................34 2.5 References........................................................................................................37 2.6 Figures..............................................................................................................39

Chapter 3: Static Properties of the Water/Hydrophobic Interface.......................44 3.1 Ellipsometric Evidence for the Depletion Layer .............................................44 3.2 X-Ray Reflectivity Evidence for the Depletion Layer ....................................46 3.3 Temperature Dependence ................................................................................50 3.4 Dependence on Dissolved Gases .....................................................................51 3.5 Lateral Properties of the Depletion Layer........................................................53 3.6 Dependence on Sample Quality.......................................................................54 3.7 Conclusions for the Chapter.............................................................................55 3.8 References........................................................................................................56 3.9 Figures..............................................................................................................59

Chapter 4: Dynamic Properties of the Water/Hydrophobic Interface .................74 4.1 Histograms .......................................................................................................74

4.1.1 Dependence of Dynamics on Temperature.............................................77 4.1.2 Dependence of Dynamics on Roughness................................................78 4.1.3 Dependence of Dynamics on Contact Angle ..........................................78 4.1.4 Dependence of Dynamics on Dissolved Gases.......................................80 4.1.5 Dependence of Dynamics on pH ............................................................80

4.2 Power Spectra ..................................................................................................81 4.3 Autocorrelations...............................................................................................82

4.3.1 X-Ray Photon Correlation Spectroscopy................................................83 4.3.2 Surface Plasmon Resonance Correlation Spectroscopy .........................85 4.3.3 Area Dependence ....................................................................................86 4.3.4 Contact Angle Dependence.....................................................................87 4.3.5 Dependence on Dissolved Gases ............................................................87

4.4 Conclusions for the Chapter.............................................................................88 4.5 References........................................................................................................89 4.6 Figures..............................................................................................................91

Chapter 5: Concluding Remarks............................................................................112 5.1 References......................................................................................................115

Appendix A: Use and Maintenance of the Ellipsometer ......................................116 A.1 Alignment and Calibration of the Ellipsometer ............................................116

A.1.1 Aligning the Laser................................................................................116

Page 7: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

vii

A.1.2 Finding the p-Polarization....................................................................117 A.1.3 Finding the Analyzer Angles ...............................................................118 A.1.4 Finding the p-Axis of the Phase Modulator .........................................118 A.1.5 Setting Jo(δo) to Zero............................................................................119 A.1.6 Calibration............................................................................................119

A.2 Taking Measurements with the Ellipsometer................................................120 A.2.1 Cleaning and Assembling the Sample Cell..........................................120 A.2.2 Finding the Brewster Angle .................................................................121 A.2.3 Collection and Analysis of Data ..........................................................121

A.3 Figures...........................................................................................................123 Appendix B: Sample Preparation...........................................................................133

B.1 Hydrophilic Silicon .......................................................................................133 B.2 OTE Monolayers ...........................................................................................134 B.3 Thiol Monolayers ..........................................................................................135

Appendix C: Use of the SPR ...................................................................................136 C.1 Cleaning the Sample Cell ..............................................................................136 C.2 Mounting the Sample ....................................................................................136 C.3 Taking Measurements ...................................................................................137

C.3.1 Static Measurements ............................................................................138 C.3.2 Dynamic Measurements.......................................................................138

C.4 Figures...........................................................................................................140 Author’s Biography .................................................................................................145

Page 8: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

1

Chapter 1: Introduction

Water is one of the most important and ubiquitous chemicals on earth. It

surrounds and permeates our lives; making up over 60% by volume of our bodies,

and covering over 70% of the earth’s surface. The presence of liquid water is thought

to be essential to the development of life. So much so that on the Mars Exploration

Rover mission, the search for evidence of liquid water was considered equivalent to

the search for previous life [1.1].

Still for all its importance, water is not well understood, and exhibits many

anomalous behaviors when compared to other fluids. For example, hydrogen sulfide

(H2S), which has twice the molecular weight of water, would be expected to have a

higher boiling temperature than water but in fact, its boiling point is 160 oC lower

than water [1.2]. Also, water has a much higher surface tension and heat capacity

than comparable liquids. The origins of many of the unusual aspects of water are due

to its unique liquid structure. In its liquid form, water consists of an ever-changing

three-dimensional network of hydrogen bonds.

Water presents some even more puzzling behaviors near hydrophobic

surfaces. Hydrophobic comes from the Greek roots “hydros” meaning water and

“phobos” meaning fear. So literally hydrophobic means fearing water, but in

chemistry it is the tendency to of a substance repel to water. On a more microscopic

level, hydrophobic surfaces cannot form hydrogen bonds. When water is placed on a

hydrophobic surface, its hydrogen-bonding network must be disrupted. This causes

the water to want to minimize its contact with the hydrophobic substance. For this

reason, a drop of water placed on a hydrophobic surface will ball up instead of

Page 9: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

2

spreading out. The angle made between the solid and the liquid surface at the point

of contact is a good measure of the hydrophobicity of the solid. This angle is called

the contact angle (θ) and is illustrated in Figure 1.1. In this work, surfaces with a

contact angle over 90o will be considered hydrophobic and those with a contact angle

under 90o, hydrophilic.

Water/hydrophobic interfaces surround us. You can see many examples in

day to day life. Examples include a dew drop on a leaf, the folding of proteins in our

bodies and even liquids on stain resistant fabrics. On both leaves and stain resistant

fabrics, water can bead up to minimize its interaction; however for a protein the

hydrophobic amino acids are submerged in water. In this case the disruption of

hydrogen bonds seems more problematic, and exactly what happens here is not clear.

Energetically, we would expect the system to form as many hydrogen bonds as

possible resulting in a preferential ordering of the water. Entropically, we would

expect the system to orient randomly and thus sample the maximum number of states.

Which of these two competing interactions dominates? What effect does the

competition have on the dynamic and equilibrium properties of the system? The

answers to these questions are still hotly debated. To help resolve this debate, I have

preformed a series of experiments looking at the density of water near a hydrophobic

surface.

This Dissertation can be roughly divided into five parts. The first part, this

introduction, deals with the background and motivation for this work. It provides an

overview of past theories, simulations, and experiments done in this area. Chapter 2

introduces the experimental considerations and techniques needed to study the

Page 10: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

3

water/hydrophobic interface. In Chapter 3, evidence for a depletion layer is

presented. The dependence of its static properties on such variables as temperature

and dissolved gasses are also explored. Chapter 4 goes on to look at the dynamic

properties of the depletion layer. A short summary is provided in Chapter 5.

1.1 Theories of Water/Hydrophobic Interfaces

Classically, water molecules were thought to form a rigid ice-like cage

enclosing the hydrophobic particle. This shell preserves hydrogen bonding at the cost

of entropy. The effective attraction between hydrophobic particles was then seen as

an entropic force resulting from the release of some water molecules when two shells

overlapped. However Blokzijl and Engberts [1.3] found no increase in ordering of

water around hydrophobic particles, indicating that a different explanation was

required.

In 1973, Stillinger presented a new theory to qualitatively explain what

happens to water near hydrophobic objects [1.4]. Based on calculations from scaled-

particle theory that explicitly includes “the strong and directional hydrogen-bonding

forces in water”, he suggested that large hydrophobic objects will be surrounded by a

“thin film of water vapor”. He predicted that the water in direct contact with the

object would have a density that is only 50,000 times less than that of bulk water.

Furthermore, if the external pressure is close to the saturated vapor pressure for the

given temperature, the film can be thick on the molecular scale because there is

“essentially no driving force to eliminate the vapor film”.

Page 11: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

4

Stillinger envisioned quite a different picture of water around small

hydrophobic particles [1.4]. He believed the water would reorganize where possible

to restore a hydrogen-bonding network. So instead of the hydrogen bonds near a

small hydrophobic particle breaking, they would simply go around the particle. This

reorganization costs entropy because the number of accessible states has been

decreased, but it is much less than the cost of breaking hydrogen bonds.

Lum, Chandler and Weeks [1.5] put forth another theory of hydrophobic

interactions. Like Stillinger, they predicted different interactions for hydrophobic

objects at large and small length scales. However their theory does not include

hydrogen bonds; in fact, it is generic to any confined fluid. For small hydrophobic

particles they believe water can reorganize around the voids created in the solvent to

maintain its density. At larger length scales the reorganization is more difficult and

maintaining the density is impossible. As a result, a layer of water with a density less

than bulk is formed at the interface. They predict that the crossover between small

and large behaviors occurs on the nanometer length scale.

If this low density depleted region exists it would be very important. It would

have many implications not only in scientific areas such as protein folding, colloidal

self-assembly, and the boundary conditions of fluid flow; but also in technological

areas such as microfluidics, hard-disk design, and drug delivery.

Another phenomena associated with hydrophobic water interfaces, is the so-

called hydrophobic interaction. This is where two or more hydrophobic bodies in

water experience a net attraction over long length scales. These long range

interactions are observed to occur in surface forces experiments at distances up to 100

Page 12: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

5

nm [1.6]. Stillinger’s theory does not explain these long-range hydrophobic

interactions [1.4]. However, Lum, Chandler and Weeks [1.5] do propose a

mechanism for these long-range interactions. As two hydrophobic objects approach

one another, fluctuations in the surrounding low-density layers can expel the

remaining water, creating a low density pocket connecting the two objects. The

resulting pressure imbalance on the objects pushes then together, creating an effective

attraction. If the combined vapor layers are large, as Lum, Chandler and Weeks

believe it is for water at ambient conditions, this effective attraction explains long-

range hydrophobic interactions.

Attard and Tyrrell [1.7] suggest long-range hydrophobic attractions are caused

by the presence of nanobubbles. In their view, as two hydrophobic surfaces approach

one another the nanobubbles can bridge the distance and the bridging meniscus then

pulls the surfaces together. Attard and Tyrrell have imaged nanobubbles using

tapping mode Atomic Force Microscopy (AFM). They report the bubbles to be 20-30

nm high and have such a high surface coverage that it is hard to see the bare substrate.

They note that the height of the nanobubbles is comparable to the height at which the

hydrophobic surfaces jump into contact.

Two major difficulties exist with the nanobubble explanation. First,

thermodynamics indicates that the bubbles cannot be in equilibrium because their

internal pressure is too high. According to the Laplace equation the internal gas

pressure of a 10 nm bubble would be 140 atm. Such a high pressure should cause the

nanobubbles to quickly dissolve. The second objection is that the AFM techniques

Page 13: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

6

used to image these bubbles may in fact be nucleating the bubbles or breaking up a

continuous vapor layer.

The dynamics of the depletion layer are of great interest. Lum, Chandler and

Weeks postulate that it is the fluctuations more than the depletion layer itself that

cause the attraction of large hydrophobic objects [1.5]. Also, these dynamics are of

great importance in understanding the kinetics of protein folding and self-assembly.

Chandler and ten Wolde show that it is solvent fluctuations that cause hydrophobic

polymers to go from an extended coil to a collapsed globule, since the nucleus is

stabilized by the formation of a “vapor bubble” [1.8]. More fundamentally, Chandler

proposes that fluctuations are the key distinguishing factor between wet and dry

interfaces [1.9].

1.2 Simulations of Water/Hydrophobic Interfaces

Many computer simulations have been done to try to uncover the causes of the

different hydrophobic effects. The simulations do not always present a consistent

picture. Many groups have looked at water confined between two hydrophobic

surfaces. In 1995, Wallqvist and Berne [1.10] reported that their constant-pressure

molecular dynamics simulation showed a dewetting transition between large

hydrophobic particles (volume ~ 40 water molecules) when they where brought

closer than two water layers. Bratko et al. [1.11] also studied the spontaneous

evaporation between large hydrophobic particles. Using an ambient-pressure, Grand

Canonical ensemble simulation, they report the onset of evaporation at three water

layers. Barrat and Bocquet saw a region of depleted density in their molecular

Page 14: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

7

dynamics simulation, changing the properties of flow at the interface [1.12]. On the

other hand, Choudhury and Pettitt saw a stabilized water layer, and not dewetting,

when water was confined between graphite surfaces [1.13]. Simulations done for

water inside hydrophobic carbon nanotubes also showed stabilized water structures

[1.14, 1.15].

Others have done simulations looking for a vapor layer without confinement.

In their 2003 paper, Huang et al. observed a vapor layer approximately 0.3-nm thick

forming around a nano-sized hydrophobic object [1.16]. They used a constant

temperature and pressure molecular dynamics simulation. Wallqvist, Gallicchio and

Levy [1.17] also saw a vapor layer form when water was enclosed inside a

hydrophobic cavity. The vapor layer was 0.4 nm thick for purely repulsive water-

wall interactions. However when they added a dispersion attraction to the water-wall

interaction, the vapor layer thickness decreased to less than 0.1 nm and the water

seemed to contact the wall.

Research looking at the dynamics of the depletion layer is more rare, although

understanding the dynamic behavior has implications for everything from micelle

formation to protein folding. The simulation done by Huang, Margulis and Berne

showed that a large-scale drying fluctuation was the rate-determining step in

hydrophobic collapse [1.16]. The Janus interface, where water is confined between

hydrophobic and hydrophilic surfaces, was simulated by Thomas McCormick [1.18].

The simulation tests whether the large fluctuations observed by Zhang et al. [1.19]

were caused by fluctuations in the liquid/vapor interface. Using a lattice gas model,

McCormick found that the interface also fluctuates up to twenty-five percent of its

Page 15: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

8

mean and displayed similar power spectrum to that found experimentally, although

the power law is less rapidly decaying function of frequency than observed by Zhang

et al. In the simulation, the fluctuations in the mean were caused by fluctuations in

the interface over length scales from 1.5 to 20 nm.

Grunze and Pertsin have recently simulated the Janus interface using the

Grand Canonical Monte Carlo technique (GCMC) [1.20]. They also observed a low

density layer at the hydrophobic wall. More interestingly, they found giant

fluctuations in the number of particles near this wall. They referred to this

phenomenon as a “wandering interface” because the liquid/vapor interface changes

position between GCMC runs causing the large fluctuations.

Some of the controversies surrounding simulations of hydrophobic

interactions can be resolved by realizing the strength of the water-solute interaction

drastically changes the properties of the depletion layer. Huang et al. found that the

critical distance (Dc) at which the drying transition occurs changed with contact angle

(θc) [1.21]. See Figure 1.2. Simulations with purely repulsive interactions would

have θc=180o. Those simulations which included idealized attractions in the water-

solute interaction would have θc=148o. The paraffin films as modeled by Huang et

al. have θc=115o, while the graphite films used by Choudhury and Pettitt have θc=83o.

The critical distance for the graphite system would be expected to be much smaller

than one layer of water and therefore should not be observed.

Still not all the discrepancies can be explained by the effect of the water

surface interactions. Furthermore, it is unclear which model to use when simulating

water. Wernet et al. found that many molecular dynamics simulations deviate

Page 16: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

9

significantly from experiment [1.22]. In fact they all yield much higher fractions of

tetrahedrally coordinated water than observed experimentally. This uncertainty

makes experimental observation all the more important.

1.3 Previous Experiments on Water/Hydrophobic Interfaces

In the years following the theoretical prediction of a low density layer when

water meets a hydrophobic surface, many people have tried to determine whether or

not it exists. Experimental results have been inconsistent. Some have been

interpreted in favor [1.23-1.27], some as indicating intimate solid-water contact in

places and ‘nanobubbles’ in others [1.7, 1.28], and some against the existence of the

vapor layer[1.29-1.31]. Using an elegant approach, Sur and Lakshminarayanan

provided experimental evidence for the existence of an interfacial vapor layer [1.23].

They found that electrodes coated with hydrophobic layers demonstrated unusually

low interfacial capacitance in water, which was not present in similar systems with

hydrophilic coatings. They attribute this behavior to a “hydrophobic gap” that in

essence creates an additional capacitor in series with the one created by the film,

effectively lowering the capacitance. Unfortunately, they have not been able to

analyze their results in terms of thickness or density of the vapor layer. Castro et al.

observed a depletion layer at the hydrophobic/water interface, using ellipsometry

[1.26]. On thin polystyrene layer (~65 nm), they found a 5 – 10 Å fully depleted

layer. However they saw no depletion on thicker (~300 nm) polystyrene layers.

Recently, groups have probed the interface using reflection techniques.

Reflection techniques, namely neutron and X-ray reflectivity, are good in that the

Page 17: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

10

wavelength employed is on the same order as the predicted thickness of the interfacial

layer. Steitz et al., using neutron reflectivity, observed a non-vanishing scattering

contrast at the interface between deuterated water and polystyrene even though the

materials have matching scattering length densities [1.24]. They explain their result

with the addition of a 2-5 nm region consisting of 88-94% bulk water density. In

another neutron reflectivity experiment, Schwendel et al. found a 2-nm layer

consisting of 9% bulk water density between water-deuterated water mixtures and

alkylated surfaces [1.25]. However they believe that this very low-density layer may

be due to air inclusions. Doshi et al. have seen a reduced density layer adjacent to a

hydrophobic surface using neutron reflectivity [1.27]. They found that the thickness

of the depleted region depended on whether the water contained dissolved gases.

Ambient water produced an 11 Å depletion layer, while degassed water bubbled with

argon had a 2 Å depletion layer.

Using X-ray reflectivity, Jensen et al. were unable to quantitatively establish

the existence of a depletion layer as they had low contrast between their hydrophobic

layer and water [1.30]. Even though they doubt the existence of the layer, they say

that if it does exist they think it would be much thinner than observed by others [1.24,

1.25], with a density near 90% of bulk water and only 0.1 nm thick. Others have also

reported not finding a depleted region. Using ellipsometry, Ducker et al. found that if

the vapor layer was pure air then it would be less than 0.1 nm thick [1.29]. Takata et

al. preformed a very careful ellipsometric study of a hydrophobic alkylsilane

submerged in water and also found no evidence for a vapor layer [1.31].

Page 18: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

11

These experimental ambiguities stem partially from different interpretations of

what is means to have a hydrophobic surface. The effect of dissolved gases,

roughness and surface stability could also be clouding the issue. In Chapter 3, we

explore the static effects different variables have on the depleted region.

There has been less work done on the dynamic properties of the

hydrophobic/water interface. One of the few measurements on the dynamics, an

indirect measurement looking at the change in the force response versus time, was

done by Zhang, Zhu and Granick [1.19]. While studying a Janus interface, they

found films whose viscous response typically fluctuated by twenty-five to fifty

percent of the mean value. They also observed that the power spectrum of the noise

displayed power law behavior, and decayed as 1/f2. It is surprising that these

fluctuations did not average out over the experimental area of ten microns. A more

direct measurement of the fluctuation themselves, has yet to be made.

1.4 References:

[1.1] “Mars Exploration Rover Mission” NASA Jet Propulsion Laboratory,

California Institutute of Technology <http://marsrovers.nasa.gov/science/ >, May 23

2006

[1.2] CRC Handbook of Chemistry and Physics: Special Student Edition, 77th Edition

(CRC Press Inc., Boca Raton, FL; 1996)

[1.3] Blokzijl, W., and J.B.F.N. Engberts Angewandte Chemie International Edition

in English 32 1545 (1993)

[1.4] Stillinger, F.H. Journal Solution Chemistry 2 141 (1973)

Page 19: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

12

[1.5] Lum, K., D. Chandler and J.D. Weeks Journal of Physical Chemistry B 103

4570 (1999)

[1.6] Israelchvili, J. Intermolecular and Surface Forces; Second Edition (Academic

Press, San Diego, CA; 1992)

[1.7] Tyrrell, J.W.G., and P. Attard Physical Review Letters 87 (17) 176104-1 (2001)

[1.8] ten Wolde, P.R., and D. Chandler Proceedings of the National Academy of

Science 99 6539 (2002)

[1.9] Chandler, D. Nature 437 640 (2005)

[1.10] Wallqvist, A., and B.J. Berne Journal Physical Chemistry 99 2893 (1995)

[1.11] Bratko, D., R.A. Curtis, H.W. Blanch and J.M. Prausnitz Journal of Chemical

Physics 115 (8) 3873 (2001)

[1.12] Barrat, J.-L., and L. Bocquet Physical Review Letters 82 (23) 4671 (1999)

[1.13] Choudhury, N., and B.M. Pettitt Journal of the American Chemical Society 127

3556 (2005)

[1.14] Hummer, G., J.C. Rasaiah and J.P. Noworyta Nature 414 188 (2001)

[1.15] Sansom, M.S.P., and P.C. Biggin Nature 414 156 (2001)

[1.16] Huang, X., C.J. Margulis and B.J. Berne Proceedings of the National Academy

of Science 100 (21) 11953 (2003)

[1.17] . Wallqvist, A., E. Gallicchio and R.M. Levy Journal of Physical Chemistry B

105 6745 (2001)

[1.18] McCormick, T.A. Physical Review E 68 061601 (2003)

Page 20: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

13

[1.19] Zhang, X., Y. Zhu and S. Granick Science 295 663 (2002)

[1.20] Grunze, M., and A. Pertsin Journal of Physical Chemistry B 108 (42) 16533

(2004)

[1.21] Huang X., R. Zhou and B.J. Berne Journal of Physical Chemistry B 109 3546

(2005)

[1.22] Wernet, Ph., D. Nordlund, U. Bergmann, M. Cavalleri, M. Odelius, H.

Ogasawara, L.Ǎ. Näslund, T.K. Hirsch, L. Ojamäe, P. Glatzel, L.G.M. Pettersson and

A. Nilsson Science Express Reports (1 April 2004)

[1.23] Sur, U.K. and V. Lakshminarayanan Journal of Colloid and Interface Science

254 410 (2002)

[1.24] Steitz, R., T. Gutberlet, T. Hauss, B. Klösgen, R. Krastev, S. Schemmel, A.C.

Simonsen and G.H. Findenegg Langmuir 19 2409 (2003)

[1.25] Schwendel, D., T. Hayashi, R. Dahint, A. Pertsin, M. Grunze, R. Steitz and F.

Schreiber Langmuir 19 2284 (2003)

[1.26] Castro, L.B.R., A.T. Almeida and D.F.S. Petri Langmuir 20 7610 (2004)

[1.27] Doshi, D.A., E.B. Watkins, J.N. Israelachvili and J. Majewski Proceedings of

the National Academy of Science 102 9458 (2005)

[1.28] Ishida N., T. Inoue, M. Miyahara and K. Higashitani Langmuir 16, 6377

(2000)

[1.29] Mao, M., J. Zhang, R. -H. Yoon and W.A. Ducker Langmuir 20 1843 (2004)

[1.30] Jensen, T.R., M. O. Jensen, N. Reitzel, K. Balashev, G.H. Peters, K. Kjaer and

T. Bjørnholm Physical Review Letters 90 (8) 086101-1 (2003)

[1.31] Y. Takata, J.H.J. Cho, B.M. Law and M. Aratoni Langmuir 22 1715 (2006)

Page 21: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

14

1.5 Figures

Figure 1.1 Illustration of Water at a Hydrophobic Surface

Water will bead up when placed on a hydrophobic surface. The degree of

hydrophobicity can be measured by the contact angle, θ. The contact angle is defined

as the angle between the surface and the tangent to the drop.

θ

Page 22: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

15

Figure 1.2 Critical Distance versus Ellipsoid Radius

The critical distance (Dc) is plotted versus the ellipsoid radius (σ) for different

contact angles (θc). The critical distance decrease with contact angle and is not

expected to be observable at lower contact angles. The figure is taken from

Reference 1.21.

180

90

148

115

Page 23: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

16

Chapter 2: Experimental Techniques

Since Stillinger’s first prediction of a low-density depletion layer forming at

hydrophobic interfaces [2.1] and especially after Lum, Chandler and Weeks further

advancement of the idea [2.2], many people have tried to test this theory. However

direct measurement of phenomena at the water/hydrophobic interface is not

straightforward in the least.

In fact, studying the depletion layer requires techniques that must meet several

criteria. The first of which is the ability to perform in-situ measurements in an

aqueous environment. With many experimental techniques this means that the probe

beam and resulting signal must travel through the water. In this case it is important to

ensure the beam has sufficient energy to penetrate the bulk water without being

significantly diminished.

Also, a sample cell that keeps the liquid in contact with the substrate must be

employed. This is more difficult than it first appears as the sample is very

hydrophobic. Hence, one must counteract the tendency of the water to bead up and

roll off the surface. There are two main strategies to overcome this problem. The

first is to completely submerge the sample in water, ensuring that there is too much

water to form a stable drop. The second strategy is to use a hydrophilic surface to pin

a thinner layer of water to the surface.

Secondly, the experimental technique must be able to distinguish the depletion

layer from the bulk. Both the bulk and depletion layer are formed of water and are

thought only to differ in their densities. Accordingly, the technique must be able to

detect differences in density. This can be done by measuring the mass density (as in

Page 24: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

17

neutron reflectivity), the electron density (as in X-ray reflectivity), or the optical

density (as in ellipsometry).

Two more requirements arise from the fact that the depletion layer is posited

to be several Angstroms to nanometers in thickness. Obviously, in order to observe

the depletion layer, the technique employed must have sub-nanometer resolution.

This rules out diffraction limited techniques in the optical wavelengths. Also, as the

depletion layer is so much smaller than the bulk layer; the signal from the bulk can

quickly overwhelm the depletion layer’s signal. This is why studying hydrophobic

interfaces requires surface selectivity.

Finally, in order to investigate dynamics at the interface, the experimental

technique needs to be able to collect information over a wide range of time scales.

Very little information exists on the time scales that will be important for

hydrophobic interactions, therefore it is essential to explore as many as possible.

2.1 Ellipsometry

When light is reflected off a planar interface, the polarization of the light

changes. Exactly how the polarization changes depends on the structure of the

interface and the incident conditions of the light. For example, initially unpolarized

light reflected off a sharp interface will become completely polarized in the direction

perpendicular to the plane of incidence at the Brewster angle.

In ellipsometry, we measure this change in polarization, and use it to

determine the structure of the sample. The quantity ellipsometry measures is called

the ellipticity ( ρ ) and it is equal to the ratio of the parallel to the perpendicular

Page 25: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

18

reflection coefficient. Phase-modulated ellipsometry makes use of the Brewster angle

to simplify experimental measurements. At the Brewster angle, the real part of ρ

vanishes making ρ equal to the imaginary part of the reflection coefficient ratio, as

given in Equation 2.1

⎟⎟⎠

⎞⎜⎜⎝

⎛=

s

p

rr

Imρ (2.1)

where rp and rs are the complex reflection amplitudes for the electric field

vectors pointing parallel and perpendicular to the plane of incidence, respectively. At

the Brewster angle, rp (and therefore ρ ) will be zero for a sharp interface between

two semi-infinite slabs of constant dielectric coefficient. However, if a thin layer

with a different dielectric coefficient or a more gradual interface is added to the

system, ρ will be non-zero.

In this case, and assuming the interfacial layers are thin compared to the

wavelength of light, ρ can be interpreted rather simply with the Drude equation

[2.3].

∫−−

+−

= dzz

zz)(

])()][([ 31

31

31

εεεεε

εεεε

λπρ (2.2)

where ε(z) is the optical dielectric profile at a distance z from the substrate, λ is the

wavelength of light and ε1 and ε3 are the optical dielectric constants of the incident (z

→ ∞) and the substrate (z → - ∞) materials. The Drude equation will only hold if the

imaginary part of the dielectric profile is smaller than the real part. This condition is

satisfied for most materials with the main exceptions being metals, which have a very

Page 26: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

19

large imaginary component. The Drude equation also assumes that the interfaces are

roughly parallel and flat to minimize scattering.

Ellipsometry is well suited to studying the hydrophobic/water interface. First,

it is very sensitive to the conditions at the surface. It can distinguish the density

difference between the bulk and depleted layer; since in a multi-component layer, the

dielectric constant is related to volume fraction of each component. As can be seen

from the Drude equation (Equation 2.2) the thickness sensitivity of the ellipsometer

depends on the dielectric coefficient of the layer in question. The highest sensitivity

would be achieved for a depletion layer consisting of pure vacuum, and would

decrease as the layer approached the density of bulk water. Still, even if the density

of water was 90% that of bulk water, we would still have sub-Angstrom sensitivity (~

0.6 Å). By using a special vertical sample cell, described by Mukhopadhyay and Law

[2.4], we can make in situ measurements in an aqueous environment. Also the

vertical configuration allows us to submerge the hydrophobic sample under water.

The greatest advantage of using a phase-modulated ellipsometer is that it allows

dynamic measurements to be made. The phase modulator and lockin amplifier enable

the signal to be averaged over only 30 ms giving us improved time resolution.

Moreover using a phase modulated ellipsometer means that we do not have to vary

the incident angle or wavelength in order to measure ρ . This permits us to repeat

measurements very quickly, on order of 50 Hz.

The vertical sample cell has additional features which allow us to study other

aspects of the hydrophobic/water interface. The sample cell has apertures that allow

Page 27: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

20

the introduction of different apparatus; such as a temperature probe, a pH probe or the

inlet and outlet tubing required for in situ fluid exchange.

We have also added a special radiative sample heater, designed by Jim Wentz.

The sample heater works by heating the metal sample holder by conduction. The

sample holder then radiates the heat, symmetrically warming the glass sample cell.

We need symmetric heating for two reasons. First, we want to minimize any

temperature gradients in the glass as this would distort the light entering and exiting

the sample cell. Second, we want to avoid setting up convection currents within the

sample. The sample heater is capable of keeping the sample cell at temperatures up

to 55 oC stable within 0.1 oC for over an hour.

Cooling the cell has proved more difficult. We have only managed it for

static measurements, where the temperature only needs to remain constant for a few

minutes. For cooling, we start by refrigerating the assembled sample cell before

placing it in the ellipsometer. Also we cool the sample holder by placing it into

contact with ice. In this way we have been able to measure temperatures as low as 10

oC.

For all its benefits ellipsometry also has its limitations. One restriction is that

not all materials have a real Brewster angle. In order for a material to have a real

Brewster angle, the imaginary component of its dielectric coefficient must be small

compared to the real component. For example, metals have a large imaginary

component in their dielectric coefficient, and therefore they do not have a real

Brewster angle. Without a Brewster angle the Drude equation is invalid, and

Page 28: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

21

Maxwell’s equations must be solved numerically. For our systems this requires

solving a 4 x 4 complex exponential matrix with six unknowns.

Even when the Drude equation is appropriate, it does not offer a unique

solution for the dielectric profile of the sample from one measurement. We can

overcome this problem to some extent by measuring ρ as the sample is constructed

or by using multiple-index medium technique (MIM) as described by Mao et al. [2.5].

Measuring ρ during construction allows the dielectric profile to be built up step by

step with the sample. MIM makes use of the fact that the Drude equation depends

heavily on the index of the surrounding media, which can be changed without

changing the dielectric profile of the interfacial layers. However there is no way to

find both the thickness and dielectric constant of the depletion layer from our

measurements of ρ . The best we can do is to confine the value to a region of

thickness and index of refraction, and even this depends on how we model the

interface.

Our home-built phase-modulated ellipsometer, pictured in Figure 2.1, is

molded after the one described by Mukhopdhayay [2.6]. It consists of a very stable 2

mW Helium-Neon laser (λ=632.8 nm) which is used as the light source. The initial

polarization of the light is set to p + 45o, where p denotes the direction parallel to the

plane of incidence, by the polarizer. Next, the phase modulator adds a sinusoidally

varying phase shift to the p-component of the electric field. At this stage the electric

field vector (E) can be written as,

E=Eo (s + e iδo sin ωot p) (2.3)

Page 29: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

22

where Eo is the amplitude of the electric field, s denotes the direction perpendicular to

the plane of incidence, δo is the amplitude of the phase shift, and ωo is the phase

shift’s angular frequency. The phase modulator is driven with a frequency (ωo/2π) of

50.1 kHz. The light is then reflected off the sample.

Reflection from the sample changes the amplitude and phase of each

component of the electric filed, embedding information on the interface into the light.

The electric field vector can now be written as

E=Eo (rso eiδs s +rpo e i(δp + δo sin ωot) p) (2.4)

where δs/p and rs/po are the phase and amplitude shifts of the s and p components of the

electric field respectively. Next the light is incident on the analyzer. The analyzer is

rotated between two positions, p + 45o and p - 45o. At p + 45o, the electric field is

proportional to the sum of the s and p components of the electric field; while at p -

45o, the electric field is proportional to the difference of the s and p components of the

electric field.

Finally the photomultiplier tube (PMT) measures the intensity of the light.

Intensity (I) is proportional to the amplitude squared of the electric filed vector as

written in Equation 2.5.

I ≈ ( ) ( )⎥⎥⎦

⎢⎢⎣

⎡+Δ

+±+ too ϖδ sincos

r rr r2

1r r 2po

2so

poso2po

2so (2.5)

where the upper and lower signs refer to p + 45o and p - 45o respectively, and Δ is the

relative phase shift between p and s (i.e. Δ=δp-δs). At the Brewster angle, Δ is

defined to be ± π/2. Using Bessel functions, the cosine term can be expanded as

shown in Equation 2.6.

Page 30: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

23

cos (Δ+δo sin ωot)= cos Δ [Jo(δo) + 2 J2(δo) cos 2ωot + . . .]

-sin Δ [2 J1(δo) sin ωot + 2 J3(δo) sin 3ωot + . . .]

(2.6)

We adjust δo with the phase modulator such that Jo(δo) = 0. In this case, Equation 2.6

becomes

I ≈ ( ) ( ) ( )[ ]…∓ +Δ) ±Δ)+ cos2 cos (J a2sinsin (J a21r r 212po

2so tt οοοο ϖδϖδ

(2.7)

where a is given by

a = ( )( )2112

posopo

so

rrrr

+. (2.8)

At the Brewster angle, (rso/rpo) is much less than unity and a can be approximated as

a ≅ 2 ( )poso rr , (2.9)

so

a sin Δ ≅ 2 ⎟⎟⎠

⎞⎜⎜⎝

s

p

rr

Im = 2 ρ . (2.10)

By using circuitry to keep the DC component of the PMT current (IPMT,DC) constant,

the output current of the PMT (IPMT) can be given by Equation 2.11.

IPMT = IPMT,DC ( ) ( )[ ]…∓ +Δ) ±) cos2 cos (J a2sin (J 41 21 tt οοοο ϖδϖρδ

(2.11)

After the PMT, the signal is given to the lockin amplifier. The voltage

delivered to the lockin is proportional to the A.C. components of IPMT. The lockin

Page 31: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

24

amplifier can easily measure the amplitudes of the components oscillating at ωo (Vωo)

and 2ωo (V2ωo). From Equation 2.11, we see that Vωo and V2ωo can be written as

Vωo = ± (-2 ) b J1(δo) ρ (2.12)

V2ωo = ± 2 a b J2(δo) cos Δ, (2.13)

where b is a constant of proportionality, and the upper and lower signs refer to p +

45o and p - 45o respectively.

When taking data with the ellipsometer, V2ωo is used to find the Brewster

angle. In theory, this could be accomplished by making V2ωo zero for both analyzer

polarizations. However in practice there is always some small residual signal that

keeps V2ωo non-zero, due to the non-ideal nature of the phase modulator [2.7]. So

instead, we adjust the incident angle of the light until V2ωo reads the same small value

for p + 45o and p - 45o. At the Brewster angle we can use Vωo to find ρ . From

Equation 2.12 we see,

ocalωρ οο ϖϖ

21

2

VV∗

+=

−+ (2.14)

where ωocal is given by

ωocal =b*J1(δo) (2.15)

and Vωo+ and Vωo- refer to Vωo measured at p + 45o and p - 45o respectively. The

value of ωocal is independent of the experiment and is simply a constant determined

during the calibration (see Appendix A for details).

Page 32: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

25

During static measurements, a computer program, “Ellipsometer”, is used to

collect Vωo and V2ωo, at both polarizations and calculate ρ . The program measures a

user specified number of ρ ’s (usually between six and ten), and computes the mean

and standard deviation of ρ . The dynamic measurements are made slightly

differently. First, ρ is measured. Then with the analyzer polarization fixed, Vωo is

measured and recorded continuously by the computer program “Fastpoll” for a user

specified amount of time. The collection speed of “Fastpoll” is determined by the

speed at which the DAC card can query the lockin and record the data. This usually

results in a data collection frequency around 50 Hz. After “Fastpoll” has finished

collecting data, ρ is re-measured.

The file collected is in the form of voltage versus time. In order to convert

this into ellipticity, we must first find the average voltage (V2) from the other

analyzer position, which can be calculated from the average voltage collected and ρ

measured before and after collection. Usually the first and second measurement of

ρ agree and the collected data is flat. If not a linear fit is used to calculate a time

dependent V2. From V2 and the collected voltage the ellipticity can be calculated for

the entire time series.

2.2 X-Ray Reflectivity

The reflection of X-rays from a surface is governed by the same fundamental

physical equations as the reflection of visible light. The main difference is in

wavelength; visible light has a wavelength on the order of several hundred

Page 33: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

26

nanometers while X-rays have wavelengths around several Angstroms. In both cases,

after reflection information on the structure of the surface is embedded in the photons.

For example all photons reflected from a thin film will experience complete

destructive interference if the path lengths of the photons reflected from the two

surfaces differ by an odd multiple half a wavelength. Accordingly, X-rays are able to

probe smaller structures than visible light.

The experimental technique of specular X-ray reflectivity makes use of these

facts to probe interfaces. X-rays with a wave number ki are incident on the surface

with an angle α. The X-rays reflected off the surface with a wave number kf at the

same angle are measured. See Figure 2.2 a. The specular reflectivity, R, is defined as

the ratio of reflected to incident X-ray flux.

In X-ray reflectivity, the specular reflectivity is measured as a function of the

momentum transfer, Q. Q is defined as the vectorial difference between ki and kf. As

can be seen in Figure 2.2 b, it is perpendicular to the sample surface. The momentum

transfer is related to the incident angle of the X-rays. It is given by

Q = λπ4 sin(α) (2.16)

where λ is the wavelength of the X-rays. The reflectivity is directly related to the

electron density profile in the surface-normal direction, ρ(z).

( )2

∫∝ dzezR zQiρ (2.17)

where z is the distance perpendicular to the surface. For a single, perfectly flat, sharp

interface the reflectivity is proportional to Q-4. However, multiple interfaces give rise

to interference, which can be seen as a series of dips and peaks on top of the Q-4

Page 34: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

27

decay of the Fresnel reflectivity. These dips are called Kiessig oscillations. The

depth and spacing of the oscillations provide information on the thickness and density

of the different layers. The position of the first dip is a good way to estimate the layer

thickness, D.

D =minQπ (2.18)

where Qmin is the momentum transfer at which the first dip in the reflectivity occurs.

To analyze our reflectivity data we used a fitting program called Parrot 32,

distributed by the Hans-Meitner Institute, Berlin. It uses the Parratt formalism [2.7]

to make least-squares comparisons between experimental and calculated reflectivities.

Basically, the fitting works by assuming an initial model, calculating a reflectivity

curve, and comparing it to the experimental data. Then the model is changed and if

the change brings the model closer to the data, it is kept. This process iterates until

the fit converges, and until additional changes only worsen the agreement of the

calculated and experimental curves.

Specular X-ray reflectivity meets the requirements for testing the hypothesis

of a depletion layer; it has sub-nanometer resolution, surface selectivity, and high

sensitivity to changes in interfacial density profiles. Using a special sample cell we

were able to keep the water in place over our very hydrophobic samples. In our ‘thin-

film’ cell, the sample was held in by an 8 μm thick Kapton membrane which also

confines a ≈ 2 μm thick water layer [2.8] against the surface. The surfaces were so

hydrophobic that the sample cell had to be assembled underwater. Otherwise either

the water would completely dewet the surface and roll off or air bubbles would be

trapped under the Kapton film.

Page 35: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

28

The biggest benefit to X-ray reflectivity over ellipsometry is its ability to

decouple thickness and density. It also provides information on the lateral properties

of the surface, which is averaged out in ellipsometry.

However, X-ray reflectivity also has its limitations. First, it is more difficult

to look at the dynamics of the interface with X-ray reflectivity. One reason for this is

the sample is damaged after continued X-ray exposure. Also, unlike ellipsometry it is

not a table top experiment. In fact the experimental apparatus is over a kilometer in

circumference. The X-ray measurements were performed at the Advanced Photon

Source (BESSRC/XOR beam line 12 BM) at Argonne National Laboratory, in

collaboration with Ian Robinson, Zhan Zhang and Paul Fenter.

2.3 Surface Plasmon Resonance

When light is reflected from an interface between materials with different

dielectric coefficients, the reflected intensity depends on the angle of incidence. This

dependence is described by the Fresnel equations [2.9].

2

12

12

coscoscoscos

⎟⎟⎠

⎞⎜⎜⎝

⎛+−

=ti

tip nn

nnRθθθθ

(2.19)

2

21

21

coscoscoscos

⎟⎟⎠

⎞⎜⎜⎝

⎛+−

=ti

tis nn

nnR

θθθθ

(2.20)

where Rs and Rp denote the reflectivity for light incident with s and p polarizations

respectively, n1 is the index of refraction for the incident media, n2 is the index of

refraction for the transmittive media, θi is the angle of incidence with respect to the

interface normal, and θt is the angle of transmittance with respect to the interface

Page 36: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

29

normal, assuming that the permeability of both materials is equal to the vacuum

permeability μo.

As can be seen from the Fresnel equations, if the light is incident from a

region with a higher index of refraction (that is if n1> n2) the reflectivity will be equal

to one for all angles above an angle, θc, given by

sin θc = 2

1

nn . (2.21)

θc is called the critical angle, and the phenomenon where the reflectivity equals one is

termed total internal reflection.

Even though all the incident light is reflected in total internal reflection, the

intensity of the electromagnetic field in the region of lower refractive index is not

zero. In fact, it oscillates parallel to the surface and decays exponentially

perpendicular to the surface; such a wave is named an evanescent wave. Griffiths

[2.9] shows that if the incident wave is p-polarized the electric field (E) will have the

form

E=Eo e-(z/l) e(iKxx-ωt) (2.22)

where Eo is the amplitude of the field, z is the distance perpendicular to the surface

and x is the distance parallel to the surface, t is the time, ω is the angular frequency of

the light, l is the decay length, and Kx is the in-plane wave number. The decay length,

l, determines how far into the medium the evanescent wave penetrates. It depends on

the refractive index of the two media and the incident angle of the light. It is given by

l = 22

221 sin

1nn

c

i −θϖ (2.23)

Page 37: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

30

where c is the speed of light. The in-plane wave number determines how the wave

propagates along the interface. It is given by

ix cnK θϖ sin1= (2.24)

Evanescent waves have many uses. They can be used to excite dyes close to

the surface in Total Internal Reflection Fluorescence Microscopy. They are used in

Fourier Transform Infrared Spectroscopy in the mode of Attenuated Total Reflection

(FTIR-ATR) to take the vibrational spectrum of molecules near the surface. In

Surface Enhanced Raman Spectroscopy (SERS), they are the source of the

concentrated electric fields used to greatly enhance Raman signals. They can also be

used to excite surface plasmons.

Surface plasmons are longitudinal charge density waves that can be excited at

metal/dielectric interfaces. Surface plasmons can also be thought of as density

fluctuations at the surface of a theoretical high-density electron liquid or plasma

[2.10]. The evanescent wave induces oscillations in the thin metal film’s free

electrons which in turn create an additional electromagnetic field [2.10]. This field

then propagates into the sample coated onto the film [2.10]. Figure 2.3, taken from

Knoll, depicts the electromagnetic field produced by the surface plasmon [2.11].

There are several conditions required for the excitation of surface plasmons.

First, surface plasmons can only be excited when a surface charge density is induced,

which requires an electric filed component perpendicular to the surface [2.11]. As s-

polarized light only has an electric field component parallel to the surface, only p-

polarized light, which has electric filed components both parallel and perpendicular to

the surface, can be used to excite surface plasmons. Furthermore, the two materials

Page 38: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

31

that make up the interface must have dielectric constants whose real parts have

opposite signs to excite surface plasmons [2.12]. Being as the dielectric constants of

metals are generally negative in the visible region, metal/dielectric interfaces are

suitable [2.12]. Finally, light reflected from a surface with an incident angle less than

θc never has sufficient momentum to excite surface plasmons. However, light from

an evanescent wave has higher momentum and under certain conditions can have

sufficient for surface plasmon excitation [2.11]. When all these conditions are met,

resonant coupling between the evanescent wave and surface plasmons can be

obtained. The coupling only occurs when the propagation constant of the surface

plasmon (κx) equals the in-plane wave number of the incoming light (Kx).

)1(

)1(

2

2

2

2

2

2

2

22

ϖϖε

ϖϖεϖκ

p

p

x c−+

−= (2.25)

where ε2 is the dielectric coefficient of the sample and surrounding media, ωp is the

plasma frequency, and εg is the dielectric coefficient of the prism. Equation 2.25

assumes the dielectric coefficient of the metal is described by the Drude model. This

is a good approximation in the visible spectrum because the changes due to electronic

loses are small. The experimental technique, Surface Plasmon Resonance (SPR),

uses this condition combined with the Fresnel equations to determine the optical

thickness of the surface layers form the incident angle at which coupling occurs, or

resonance angle. Experimentally, the resonance angle can be determined by looking

at the reflected intensity as a function of incident angle. Since the coupling removes

energy from the evanescent wave, there is a dip in the reflected intensity at the

resonance angle.

Page 39: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

32

SPR is another surface technique that is well suited to studying how water

meets a hydrophobic surface. It is capable of real time data collection, over many

time scales [2.13]. SPR has great surface sensitivity. Because the evanescent wave

decays perpendicular to the surface in a distance on the order of 200 nm, it only

interrogates the environment quite close to the surface. It also has a high resolution to

changes in thickness and refractive index in this region, which we saw in ellipsometry,

can be translated into changes in density. A standard SPR setup with a rotating prism

and photodetector typically has a resolution in the resonance angle of 10-2-10-3

degrees [2.14]. Finally, it is a non-intrusive technique, and requires no dye or sample

labeling [2.15].

Only two main modifications are required to use SPR for aqueous systems.

First, a higher refractive index prism was required in order to access the angles need

for resonance by high index aqueous solutions. Also, it is necessary to design a

special fluid cell. With it we can make measurements in fluid environments and

observe in situ monolayer growth. It also allows us to vary the bulk media without

disturbing the sample under investigation. It holds the sample substrate against the

fluid reservoir, and can be completely filled with water to ensure no macroscopic

dewetting occurs.

We built a SPR setup using the Kretschmann geometry with a rotating prism

for resonance angle detection. The setup is depicted in Figure 2.4. Light from a 2

mW helium-neon laser is split by a Glan laser polarizer. The p-polarized light is

transmitted and the s-polarized is reflected. We use the reflected light to remove long

time scale fluctuations of the laser from our dynamics measurements. The p-

Page 40: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

33

polarized light then passes through an iris which is used to control the area probed by

the surface plasmons. The incident angle of the light when it hits the prism is

controlled by the rotation stage. Here the light is totally internally reflected, creating

an evanescent wave which excites surface plasmons in the sample. The intensity of

the reflected light is measured by detector 1. Neutral density (ND) filters are used

before both detectors to prevent saturation and to improve signal-to-noise ratios in

fluctuation measurements.

Static measurements are made by measuring the intensity of reflected light for

a range of incident angles. In this way, we can determine the resonance angle and the

shape of the resonance curve. The resonance curve can then be compared to curves

theoretically calculated curves. In this way we can get some idea of the absolute

values for thickness and refractive index of our samples. True quantitative analysis

can only be done for the relative changes in our samples.

Dynamic measurements are made somewhat differently. As we cannot hope

to measure multiple angles quickly enough to see fluctuations on the time scale

desired, we instead sat at a fixed angle and measured changes in reflectivity. The

changes in reflectivity correspond to changes in resonance angle as illustrated in

Figure 2.5. In order to remove long time scale fluctuations, we added an additional

detector to measure the laser intensity. We can then divide the incident and reflected

intensities to get the reflectivity more accurately. The collection and division of the

signals was done with a lockin amplifier. The data was then transferred into the

computer and recorded. Sung Chul Bae provided immense help and guidance in

Page 41: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

34

implementing the dynamic measurements as well as creating the programs to collect

and analyze the data.

2.4 Hydrophobic Substrates

Almost as important as choosing the proper experimental technique is

choosing appropriate hydrophobic surfaces. In order to maximize our chances for

observing the elusive depletion layer, we want to ensure that our surface is as close to

those studied by simulation as possible. We also desire very robust samples to ensure

that any changes in the surface do not mask the presence of the depletion layer.

First, the substrates must be very hydrophobic. Wallqvist et al. showed in

simulations that the strength of the interaction between the water and the surface can

drastically change the thickness of the depletion layer [2.16]. They found that simply

adding a dispersion attraction could decrease the thickness of the vapor layer to 1 Å

from 4 Å. Huang et al. also found a strong dependence of the depletion layer on

contact angle [2.17]. In practice, we use samples with contact angles between 97o

and 110o.

Next, the samples must be well ordered, meaning that they must be fully

coated and densely packed. Surfactant-coated surfaces were used in some earlier

studies, but our selection of a chemically-attached monolayer circumvents the

potential complication that surfactant-coated solid surfaces may reconstruct to form

complex new morphologies when placed in water [2.18, 2.19]. The conformational

changes of loosely packed surfaces make it extremely difficult to decouple the small

Page 42: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

35

change due to the depletion layer from the large change due to the substrate

rearrangement.

Also, the sample must be strongly attached to the surface. This helps prevent

rearrangement as noted earlier but it also ensures that the dynamics are stable. If the

surface is not strongly attached to a rigid surface, any fluctuations in the depletion

layer may be coupled into much smaller movements of the layer itself. It would be

extremely difficult to measure these movements.

Finally, the sample must be coated on substrates appropriate for the

experimental technique for which it is used. In the case of ellipsometry and X-ray

reflectivity, silicon wafers are the best choice. They meet the ellipsometric

requirements for smoothness, dielectric coefficient, and reflectivity. Having a well

defined crystal structure and an electron density about half of water makes them well

suited to X-ray reflectivity. SPR requires a metal-coated transparent substrate that

can be index matched to the prism; therefore we use gold coated high-index flint glass

slides.

On silicon wafers we use self-assembled monolayer of n-

Octadecyltriethoxysilane (OTE). OTE is a methyl terminated molecule consisting of

an eighteen carbon chain and a silane head group. It is the head group that allows

OTE to form well packed monolayers that are chemically bound to the silicon

surface. Each head group can form three silicon-oxygen chemical bonds. It can

make theses bonds either with other OTE molecules or with the substrate. So OTE

basically forms a two-dimensional network that is chemically anchored to the silicon

surface. Using the methods described by Hong et al. near perfect, highly hydrophobic

Page 43: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

36

monolayer were produced [2.20]. These surfaces have a root mean squared (RMS)

roughness less than 2 Å. The OTE surfaces were characterized using AFM and

contact angle measurements.

One of the large drawbacks in using OTE is that while very good layers are

possible, the process is difficult, as silane chemistry is not fully understood, and the

success rate is low. Another problem is that it is extremely difficult to systematically

vary the roughness or contact angle.

Despite the difficulties in producing OTE monolayers, we believe it is still the

best monolayer available for silicon surfaces. Many groups use

octadecyltrichlorosilane (OTS) monolayers as their hydrophobic monolayer but this

surface is less well packed than OTE [2.21]. OTE has an area per molecule of 19.9

Å2 [2.21], while OTS has an area per molecule of 23.5 Å2 on silicon [2.22]. There is

also some evidence that OTS can be swelled by water. Kuhl et al. found that on

silicon an average of 2.9 water molecules per OTS molecules where needed to

account for the scattering length density the observed with neutron reflectivity [2.22].

Mao et al. also reported a change in domain height in OTS in the presence of water;

in air domains where ~ 0.2-0.3 nm and in water they were ~0.3-0.5 nm [2.5].

For SPR, we use thiols to form a monolayer on our metal surfaces. Thiols are

chemically attached to our gold surface through a sulfur-gold chemical bond. For

hydrophobic monolayers we use n-octadecanethiol (ODT), which has a contact angle

of over 100o. One of the big advantages of using thiol chemistry that one can

purchase thiols terminated with any desired end. This allows the systematic control

of many variables such hydrophobicity and surface rigidity. By mixing ODT and 11-

Page 44: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

37

mercapto-1-undecanol (R-OH T), we were able to vary the contact angle between

110o and 5o. Also, as the thiol layer takes on the topographical features of the

underlying metal surface, the roughness can be systematically controlled. Thiol

monolayer formation is also much more robust and well understood than OTE

formation, and with very little effort we get nearly 100% success rate. This makes

thiol a great substrate for a physicist. The biggest limitation of the thiols is that they

only form monolayers on coinage metals. Hence, they cannot be used with

ellipsometry except to get rough qualitative results.

Details of the sample preparation are given in Appendix B.

2.5 References

[2.1] Stillinger, F.H. Journal of Solution Chemistry 2 141 (1973)

[2.2] Lum, K., D. Chandler and J.D. Weeks Journal of Physical Chemistry B 103

4570 (1999)

[2.3] Drude, P. K. L. The Theory of Optics (Dover, New York, 1959), p. 292

[2.4] Mukhopadhyay, A., and B.M. Law Physical Review E 62 (4) 5201 (2000)

[2.5] Mao, M., J. Zhang, R. -H. Yoon and W.A. Ducker Langmuir 20 1843 (2004)

[2.6] Mukhopadhyay, A. (2000), "Ellipsometric Study of Surface Phenomena" PhD

thesis, Kansas State University

[2.7] Parratt, G. Physical Review 95 359 (1954)

[2.8] Fenter, P. Reviews in Mineralology and Geochemistry 49 149 (2003)

[2.9] Grifffiths, D.J. Introduction to Electrodynamics Third Edition (Prentice Hall

New Jersey, 1999) p.384-415

Page 45: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

38

[2.10] Salamon, Z., H. A. Macleod and G. Tollin Biochimica et Biophysica Acta 1331

117 (1997)

[2.11] Knoll, W. Annual Review of Physical Chemistry 49 569 (1998)

[2.12] Brockman, J.M., B.P. Nelson and R.M. Corn Annual Review of Physical

Chemistry 51 41 (2000)

[2.13] Green, R.J., R.A. Frazier, K.M. Shakesheff, M.C. Davies, C.J. Roberts and

S.J.B. Tendler Biomaterials 21 1823 (2000)

[2.14] Tao, N.J., S. Boussaad, W.L. Huang, R.A. Arechabaleta and J. D’Agnese

Review of Scientific Instruments 70 (12) 4656 (1999)

[2.15] Homola, J., S.S. Yee and G. Gauglitz Sensors and Actuators B 54 3 (1999)

[2.16] Wallqvist, A., E. Gallicchio and R.M. Levy Journal of Physical Chemistry B

105 6745 (2001)

[2.17] Huang X., R. Zhou and B.J. Berne Journal of Physical Chemistry B 109 3546

(2005)

[2.18] Patrick, H.N., G.G. Warr, S. Manne and I.A. Aksay, Langmuir 15 1685 (1999)

[2.19] Perkin, S., N. Kampf and J. Klein Physical Review Letters 96 038301 (2006)

[2.20] Hong, L., A. Poynor, S. Granick and C. Kessel “Preparation of Smooth

Hydrophobic Self-assembled Monolayers” In preparation

[2.21] Ben Ocko personal communication November 23, 2005

[2.22] Kuhl, T.L., J. Majewski, J.Y. Wong, S. Steinberg, D.E. Leckband, J.N.

Israelachvili and G. S. Smith Biophysical Journal 75 2352 (1998)

Page 46: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

39

2.6 Figures

Figure 2.1 Photograph of the Ellipsometer.

Above is a picture of our home-built phase modulated ellipsometer. Light

from the laser has its initial polarization set by the polarizer. Then a time-dependent

phase difference is added by the phase modulator. The light is reflected off the

sample which changes its polarization, embedding information about the surface into

the light. Next the new polarization is measured by the analyzer and recorded by the

PMT.

PMT Laser

Sample Holder

Phase Modulator

Polarizer Analyzer

Page 47: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

40

a b Figure 2.2 Illustration of X-ray Reflectivity

X-rays with an initial wave number (ki) are incident upon the sample surface

at an angle, α, as shown in a. In specular X-ray scattering, the detector is positioned

so that X-rays with a final wave number (kf) are measured. Part b shows the

definition of the momentum transfer vector, Q. Q is perpendicular to the surface.

ki

kf

Q

α α

ki kf

Page 48: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

41

Figure 2.3 Schematic of a Surface Plasmon

Surface plasmons can be excited at a metal/dielectric interface. The interface

lies in the x-y plane and the plasmon propagates in the x direction. Taken from Knoll

[2.11].

Page 49: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

42

Figure 2.4 Schematic of SPR Apparatus.

We built our SPR using the Kretschmann Geometry. Here light from the laser

becomes p-polarized after the Glan laser polarizer. The size of the beam is controlled

by an iris and then incident on the prism. It hits the gold layer with an angle above

the critical angle, producing an evanescent wave and thus exciting surface plasmons.

The signal is collected by detector 1. The signal from detector 2 is used to remove

long-time scale fluctuations from the data.

LASER

Glan Laser Polarizer

Iris

Mirror 1

Mirror 2 ND

ND

Detector 1

Detector 2

Rotation Stage

Sample Cell

Prism

Page 50: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

43

Figure 2.5 Dynamic SPR Measurements.

A shift in the resonance angle of the system, for example from the blue curve

to the pink curve shown above, indicates the reflectivity at a fixed angle (shown by

the dashed red line) also changes. This allows the shift in resonance angle in time to

be mapped to a shift in reflectivity.

0

0.2

0.4

0.6

0.8

1

50 55 60 65 70

Angle (Degrees)

Ref

lect

ivity

Page 51: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

44

Chapter 3: Static Properties of the Water/Hydrophobic Interface

As stated in Chapter 2, what happens at the water/hydrophobic interface is

still hotly debated. This chapter presents evidence both from ellipsometry and X-ray

reflectivity in support of a depletion layer. It also looks at the static dependence of

the depletion layer on such variables as temperature and sample quality.

3.1 Ellipsometric Evidence for the Depletion Layer

In order to look for evidence of the existence of a depletion layer at the water

hydrophobic surface, we first had to build up a model of our system. First, we

measured the contribution to ρ from the silicon substrate and its oxide layer. To do

this the oxide layer was prepared as it would be for OTE functionalization, details can

be found in Appendix B. The silicon wafer was first soaked in Piranha solution for

one hour at 70 oC, and rinsed with copious amounts of deionized water. The wafers

were further cleaned with UV/Ozone for 30 minutes and finally cleaned in an

oxygen-plasma cleaner. This produced bare silicon wafers with a well-defined oxide

layer approximately 2 nm thick. The bare wafers were completely wet by water

(advancing contact angle = 0o). Next the contribution of the OTE layer was measured

in ethanol. We chose ethanol as it has an index of refraction very close to water,

giving it a Brewster angle almost identical to that of water which allows us to

measure ρ for both liquids with the same incident angle. Also, it is not expected to

produce a depletion layer in contact with OTE as the contact angle between ethanol

and OTE is near 15o. Using a simple three-slab model (one slab for the substrate, one

for the OTE and one for the solvent), we found the ellipsometric thickness of the OTE

Page 52: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

45

layer in ethanol was 2.47 ± 0.03 nm, which is in good agreement with the expected

chemical structure.

Finally with the sample well-parameterized we were able to measure ρ in

water. Using the vertical sample cell, we were able to make the measurement at the

exact same position in water as we had in ethanol. However if we used the same

model and OTE thickness to calculate the expected value of ρ in water, we found

that the calculated value differed from the measured one by 28%. To better quantify

this difference, we define Δ ρ as the measured value minus the calculated value. In

this case, Δ ρ = -0.0040±0.0001. The physical significance of Δ ρ is illustrated in

Figure 3.1. Since Δ ρ is not zero we know that the three-slab model is not correct.

At first, Occam’s Razor suggested to us that this might reflect some kind of defects in

the monolayers, but swelling of the OTE layer by the water would make Δ ρ

positive. Similarly organic contaminants, which usually have an index of refraction

greater than water, would also make Δ ρ positive.

There are only two ways to get a negative value for Δ ρ . The first requires

the OTE layer to shrink. In our case this would mean that the OTE layer went from

being approximately 2.4 nm thick to being 1.4 nm thick, a decrease of over 40%.

This is would require the tilt angle to change from 24° to 58° or massive layer

disordering, which is physically unrealistic as the OTE layer is already well packed.

To ensure this was not the case, we have also done control experiments to test for

changes in the OTE. Using Sum Frequency Generation (SFG) we examined the OTE

surface both in contact with air and water. The SFG spectra in air and water show no

Page 53: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

46

significant differences, as can be seen in Figure 3.2, indicating that no large change in

tilt angle or disordering occurred.

The second reason for a negative Δ ρ requires the addition of another layer

with an index of refraction less than the index of refraction of water. We believe that

our measurement signals the presence of a surface layer of depleted density, directly

from the raw data.

The Drude equation does not yield a unique result for both the thickness and

the density of this depleted zone. Using a four-slab model similar to those used in

neutron reflectivity, we can model the depletion layer, but we cannot uniquely solve

for both the thickness and density. Instead we can relate the values of the thickness

and density as shown in Figure 3.3. As can be seen in Figure 3.3, a fully depleted

layer would have a thickness of 0.180 ± 0.005 nm, while a layer with 90% bulk

density will have a thickness of 2.60 ± 0.07 nm, in agreement with Steitz et al. [3.1].

Alternatively, we can model the vapor layer as a continuous dielectric profile

proportional to the hyperbolic tangent, which is often used in ellispometric modeling.

This provides a more physical, smoothly varying dielectric coefficient that starts out

at the pure vapor value and reaches the bulk value at some distance, L, from the

surface. In this case, L = 0.45 ± 0.01 nm. A comparison of the different models is

shown in Figure 3.4.

3.2 X-Ray Reflectivity Evidence for the Depletion Layer

As stated previously, X-ray reflectivity has an advantage over

ellipsometry in that it can decouple the thickness from the density of the depletion

Page 54: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

47

layer. To use this advantage, X-ray measurements were performed at the Advanced

Photon Source (beam line 12 BM) at Argonne National Laboratory, in collaboration

with Ian Robinson, Paul Fenter and Zhan Zhang. The temperature of the sample was

between 23 ºC and 25ºC. The X-ray beam was reflected at incident angles, α,

generally ranging from 0.1° to 2.3°, corresponding to momentum transfers of 0.03 Å-1

to 0.8 Å-1. The incident beam size ranged from 0.04 mm to 0.4 mm so that the

resulting beam footprint remained well within the middle of the surface. A

monochromatic X-ray beam with energy of 19.0 keV was used to maximize the

transmission through the water layer. The full incident beam flux was 2x1010

photons/sec. Specular reflectivity measurements took place within 20 minutes for a

given beam spot with two additional scans to probe the background with the sample

angle displaced by ±0.05°, during which time no changes to the reflected intensity

were observed. A correction was applied to the data for angle-dependent attenuation

by the water and Kapton layers [3.2]. A scale factor was used in the analysis, as

beam attenuation prevented measurements near the critical angle. Since reflectivity

varied by 10 orders of magnitude, the data are displayed here as normalized

reflectivities, RQ4, to compensate for the ~Q-4 decay of the Fresnel reflectivity

To fix parameters for subsequent analysis, these organic monolayers were first

studied in air (see Figure 3.5). The reflectivity curve shows Kiessig oscillations

owing to interference between X-rays reflected at the silicon-organic and organic-air

interfaces and thus were highly sensitive to the monolayer thickness, D. The

thickness estimated from the position of the first intensity dip, using Equation 2.18,

was 24-27 Å. The derived electron density profile corresponding to the best-fit

Page 55: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

48

reflectivity is shown in Figure 3.6 and includes contributions from the monolayer

thickness and density, and interfacial oxide and monolayer head-group layers.

Bearing in mind that SAM-coated surfaces never attain 100% surface coverage, and

that they contain grain boundaries and other defects, fits to the data were found to be

most consistent when we assumed a monolayer containing 1% less electron density

than perfectly packed monolayers with 100% coverage, owing to such packing

imperfections. However, the observation of deep intensity minima in Figure 3.5

demonstrates that the monolayers were homogeneous over the large area probed by

the X-ray beam.

After placing these same monolayers in water, reflectivity interference

oscillations continued to be observed but with significant shifts in the dip positions

(Figure 3.5). The estimated monolayer thickness, using Equation 2.18, was 14 -15 Å,

less than two-thirds of the thickness in air, which is physically unrealistic as we

explained in Section 3.1. Rather, this signifies a phase shift of the interference

oscillations owing to non-monotonic changes in the electron density normal to the

silicon. Bearing in mind that the electron density of water is close to that of an

organic monolayer, the observed phase shift indicates that electron density varied

from water to monolayer in a non-monotonic fashion, passing through a region whose

electron density was less than either the water or the monolayer. We emphasize that

this qualitative conclusion, obvious to the eye when one inspects Figure 3.5, holds

independently of the quantitative analysis that follows. A schematic sketch is

included in Figure 3.6.

Page 56: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

49

In fitting the curves, the thicknesses, scattering length densities, and

roughnesses for the silicon oxide and bulk silicon layers were fixed to tabulated

values. A “head group” layer was included following the example of Tidswell et al.

[3.3]. The fit for the monolayer in air served to fix the needed parameters for

subsequent analysis when water was added. The same parameters were used in fitting

the reflectivity of these monolayers immersed in water.

In their influence on the quality of fits, the most important parameter in the

first place was roughness of the organic monolayer, ~7 Å; but this was fixed by the

measurements made in air. In water, fits were made separately to the experiments

where the water was saturated with atmospheric gas, or deaerated, with results

summarized in Figure 3.7. The best fits are encouragingly consistent for both

ambient and deaerated water, giving for both cases a depletion layer thickness of 2.7

Å, electron density corresponding to 20% water content, and width of the error

function profile (between the depletion layer and bulk water) of 4 Å. However, there

is considerable room to change these parameters and still get an acceptable fit as the

fits in water were very sensitive not only to the thickness of the depletion zone but

also to its electron density relative to that of bulk. The areas of acceptable fits are

shown in Figure 3.7.

How do the values measured for the depletion layer with ellipsometry

compare to those measured with X-ray reflectivity? As stated earlier, ellipsometry

cannot decouple the depletion layers thickness and density, and the best that could be

done was to confine the value to a curve. This curve has been plotted with the

reflectivity results in Figure 3.7. We see remarkably good agreement between the

Page 57: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

50

results. This demonstrates that depletion layer is a real effect and is visible to more

than one experimental technique. However, what variables control its properties and

why do not all experiments see the depletion layer?

3.3 Temperature Dependence

The formation of the depletion layer is the result of balancing entropic and

energetic interactions. To more fully understand the relative strengths and influences

of these two aspects, it is useful to look at the temperature dependence of the

depletion layer. In a recent molecular dynamics simulation, Mamatkulov et al. found

that the depletion layer increased in thickness and became more depleted as

temperature was increased [3.4].

We measured the ellipticity from 10oC to 50oC, for OTE in both water and

ethanol. In ethanol we saw no significant change in the ellipticity ( ρ ) while in

water we found that ρ increases with temperature. See Figure 3.8. The increase in

ellipticity was seen for many samples with different contact angles and displayed no

hysteresis. See Figure 3.9. There are two simple ways to represent this change. First,

we can use the hyperbolic tangent model and ascertain how the thickness changes

with temperature; fixing the density while allowing the thickness of the depleted

region to vary. Figure 3.10 depicts this interpretation, where the change in thickness

is plotted versus temperature for several samples. As can be seen the thickness of the

depleted region decreases drastically with temperature. In fact, it decreases by more

than 60% over a forty-degree temperature range. Second, we can employ the four-

slab model with a fixed thickness and determine temperature’s effect on the density.

Page 58: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

51

This model is displayed in Figure 3.11. In this case, as the temperature is increased

the depleted region becomes more like the bulk. In reality, the thickness and density

of the depleted region probably change simultaneously. Still we can conclude from

these two simple interpretations that the depletion layer dissipates as the temperature

is increased.

This is exactly the opposite result as seen in the simulation [3.4]. What could

be the reason for the depletion layer dissipating with increasing temperature? In bulk

water, as temperature is increased the number of broken hydrogen bonds also

increases. At a hydrophobic surface this means that few bonds are disrupted

compared to bulk water as the temperature increases. In a sense, this means that

water becomes less “oil-phobic” at higher temperatures [3.5]. Viewed in this way it

is clear that the driving force for creating an interface, namely the minimization of

broken bonds, decreases as temperature is increased and not surprisingly the depletion

layer dissipates [3.6].

3.4 Dependence on Dissolved Gases

Considering that water saturated with dissolved atmospheric gases is reported

to display different hydrophobicity than deaerated water [3.7-3.10], we performed X-

ray and ellipsometry experiments for both cases. These control experiments were

designed to test the uninteresting possibility that a near-surface layer with density less

than bulk might trivially result from migration to the hydrophobic surface of

dissolved gas.

Page 59: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

52

To test this possible influence in ellipsometry, the following control

experiments were performed. First, the hydrophobic surface was immersed in

ethanol, which wets it, in order to displace any possible absorbed gas. Then, without

exposing the surface to air, ethanol was flushed out of the sample cell with copious

amounts of degassed, deionized water, and finally ellipsometry measurements were

made.

Two methods were employed to degas the water prepared by purification

using a Barnstead Nanopure II deionizing system: (i) water was boiled for 30 minutes

and subsequently cooled to room temperature in a filled, sealed vessel; (ii) water was

freeze-dried at liquid nitrogen temperature and subsequently thawed to room

temperature in a filled, sealed vessel, this process was repeated 5-7 times. We

conclude that, although the hypothetical possibility cannot be excluded that some

residual gas persisted, the greatest care by the means available to us was taken to

exclude it. In the measured ellipticity, there was no significant effect of degassing.

In fact when one position on a sample was measured first in degassed water and then

exposed to air for over an hour and re-measured the change in the thickness of the

depleted region modeled with the hyperbolic tangent was 0.015 ± 0.016 nm.

Similar methods where employed in X-ray reflectivity experiments. As seen

in Figure 3.7, the fits to the reflectivity curves were quite consistent for both ambient

and degassed water.

Page 60: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

53

3.5 Lateral Properties of the Depletion Layer

The ellipsometry and X-ray reflectivity data compares favorably with a long

tradition of theory predicting that water forms a depletion layer in the vicinity of

hydrophobic surfaces [3.11-3.15]. However, from recent AFM-based studies the

alternative hypothesis of ‘nanobubbles’ has emerged in which approximately 50% of

a hydrophobic surface is coated with small bubbles around 5-100 nm in height and

100-800 nm in diameter, provided that the water has not been degassed [3.7, 3.16].

The X-ray data reported in Section 3.2 afford a succinct quantitative test of this

hypothesis. There are two possibilities to consider, depending on whether the

putative bubble size is larger than or less than the coherence length of the X-ray

measurement, as illustrated in Figure 3.12 a. Pershan and coworkers have shown that

the coherence length in X-ray reflectivity along the surface normal direction follows

from geometrical considerations, mainly concerning the detector slit; applied to the

present situation, where the vertical detector aperture was 1 mm at a distance of 750

mm from the sample, it follows that the radial coherence length was approximately

850 Å [3.17]. In the following, we denote the case where the nanobubbles are larger

than this length as the ‘incoherent’ model and the case where they are smaller as the

‘coherent’ model.

In the incoherent model, the measured reflectivity results from a linear

combination of reflectivities for the air/hydrophobic (measured directly) and

water/hydrophobic interfaces, as shown in Figure 3.12 b. This model requires that the

first scattering intensity minimum occur at the same wavevector as when the

monolayer is in air, and it is obvious that the data disagree with this ansatz. A direct

Page 61: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

54

comparison (Figure 3.13) for several different hypothetical nanobubble surface

coverages shows explicitly the unfavorable comparison with experimental data.

Alternatively, we suppose that the nanobubble diameter was less than the

coherence length of the X-rays, < 850 Å. Reflectivity in the latter case is calculated

by combining the density profile of the hydrophobic/air interface with an error

function profile for the electron density of water as shown in Figure 3.12 c, giving an

average electron density, intermediate between air and water. The quantitative

comparison (Figure 3.14) plots RQ4 against Q for several hypothetical nanobubble

surface coverages and shows explicitly the unfavorable comparison to the

experimental data in this case also. Further evidence against the nanobubble

hypothesis comes from the lack of any significant off-specular diffuse scattering

which would be expected since our lateral spatial resolution (>20 μm) is substantially

larger than the reported lateral size of nanobubbles (~100 nm). Together these

observations allow us to rule out nanobubbles as playing a significant role in these

measurements.

3.6 Dependence on Sample Quality

One reason why some groups may not observe the depletion layer is that we

saw that it was highly sensitive to sample quality. We found that extreme care is

needed during sample preparation in order to observe a depletion layer. The details

of sample prepatation are given in Appendix B.

Monolayers that have defects or those that are not strongly attached to the

substrate will swell or blister in water. As discussed in Section 3.1, this will make

Page 62: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

55

Δ ρ more positive and mask the depleted region (see Figure 3.5). Take for example

one bad sample, when measured in air, its OTE layer was 2.39 ± .01 nm thick but

when measured in ethanol was 3.66 ± .02 nm. This sample gave a depleted region

with a negative thickness (L = -0.06± .01 nm, using the hyperbolic tangent model).

Figure 3.15 displays the difficulty of creating a high quality monolayer that

did not swell. Out of the eleven samples shown six did not show signs of swelling,

and could be used to observe a depletion layer. The other samples showed either zero

or negative depletion layer thicknesses.

The presence of sample swelling probably explains why other groups did not

see the depletion layer on octadecyltrichlorosilane (OTS) monolayers [3.18, 3.19].

Mao et al. also reported a change in domain height in OTS in the presence of water;

in air domains where ~ 0.2-0.3 nm and in water they were ~0.3-0.5 nm [3.18]. This

small change in thickness could easily mask the presence of the depletion layer.

Takata et al. also used OTS and failed to see a depletion layer but did not mention if

they observed any swelling in water [3.19].

3.7 Conclusions for the Chapter

We have observed evidence for the depletion layer with both ellipsometry and

X-ray reflectivity. The evidence for the existence of the depletion layer is very

strong, however pinning down the exact thickness and density is much more difficult.

Both techniques only confine the depletion layer to fall within a certain area. Our

best guess is that the depletion layer is 2.7 Å thick and has a density corresponding to

20% water content.

Page 63: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

56

Prior related experiments were controversial. Some ellipsometry experiments

concluded that there is no depletion [3.18, 3.19] while others concluded depletion

exists [3.20]. Experiments based on neutron scattering show effects, but the

magnitudes differ considerably [3.1, 3.9, and 3.21]. X-ray reflectivity measurements

of an organic film floating on an air-water interface showed a density gap < 15 Å in

extent, which in combination with MD simulations was interpreted in terms of 1.2 Å

thick “vacuum” layer [3.22]. Some of these inconsistencies could be due to the

difficulty in properly quantifying the depletion layer. Others are probably due to the

highly sensitive dependence of the depletion layer on sample quality.

Also we have seen no evidence for nanobubbles. This contradicts many AFM

experiments which have observed them [3.7, 3.16]. It is interesting that the

nanobubbles have only been observed during perturbative experiments. One

possibility is that the tapping motion of the AFM tip may have transformed the

depletion layer into small bubbles.

Many groups have posited that the depletion layer was just the trivial

segregation of dissolved gasses to the surface [3.7, 3.9 and 3.10]. We have seen no

effect of dissolved gasses on the depletion layer. This null result indicates that the

depletion layer is in fact a fundamental property of what happens when water meets a

hydrophobic surface.

3.8 References:

[3.1] Steitz, R., T. Gutberlet, T. Hauss, B. Klösgen, R. Krastev, S. Schemmel, A.C.

Simonsen and G.H. Findenegg Langmuir 19 2409 (2003)

Page 64: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

57

[3.2] Fenter P., Reviews in Mineralology and Geochemesitry 49 149 (2003)

[3.3] Tidswell, I.M., T.A. Rabedeau, P.S. Pershan and S.D. Kosowsky Journal of

Chemical Physics 95 2854 (1991)

[3.4] Mamatkulov, S. I., P. K. Khabibullaev and R. R. Netz Langmuir 20 4756 (2004)

[3.5] David Chandler personal communication, June 14 2004

[3.6] Chandler, D. Nature 437 640 (2005)

[3.7] Ishida N., T. Inoue, M. Miyahara and K. Higashitani, Langmuir 16, 6377 (2000)

[3.8] Pashley R. M., J. Phys. Chem. B 107, 1714 (2003)

[3.9] Doshi, D.A., E.B. Watkins, J.N. Israelachvili and J. Majewski., Proceedings of

the National Academy of Science 102 9458 (2005)

[3.10] Stevens, H. R. F. Considine, C. J. Drummond, R. A. Hayes and P. Attard,

Langmuir 21 6399 (2005)

[3.11] Stillinger, F.H. Journal Solution Chemistry 2 141 (1973)

[3.12] Lum, K., D. Chandler and J.D. Weeks Journal of Physical Chemistry B 103

4570 (1999)

[3.13] Huang, X., C.J. Margulis and B.J. Berne Proceedings of the National Academy

of Science 100 (21) 11953 (2003)

[3.14] Dill K.A., T.M. Truskett, V. Vlachy and B. Hribar-Lee Annual Review of

Biophysics & Biomolecular Structure 34 173 (2005)

[3.15] Huang, D. M., and D. Chandler Journal of Physical Chemistry B 106, 2047

(2002)

[3.16] Tyrrell, J.W.G., and P. Attard Physical Review Letters 87 (17) 176104-1

(2001)

Page 65: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

58

[3.17] Tamam L., H. Kraack, E. Sloutskin, B.M. Ocko, P. S. Pershan, A. Ulman and

M. Deutsch, Journal of Physical Chemistry B 109, 12534 (2005)

[3.18] Mao, M., J. Zhang, R.-H. Yoon and W.A. Ducker Langmuir 20 1843 (2004)

[3.19] Takata, Y., J.H.J. Cho, B.M. Law and M Aratoni Langmuir 22 1715 (2006)

[3.20] Castro, L.B.R., A.T. Almeida and D.F.S. Petri Langmuir 20 7610 (2004)

[3.21] Schwendel, D., T. Hayashi, R. Dahint, A. Pertsin, M. Grunze, R. Steitz and F.

Schreiber Langmuir 19 2284 (2003)

[3.22] Jensen, T.R., M. O. Jensen, N. Reitzel, K. Balashev, G.H. Peters, K. Kjaer and

T. Bjørnholm Physical Review Letters 90 (8) 086101-1 (2003)

Page 66: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

59

3.9 Figures

Figure 3.1 Schematic Illustration of Δ ρ Δ ρ is the difference between the measured ellipticity and that calculated

using the OTE thickness measured at the same position on the sample in ethanol.

Δ ρ equal to zero indicates that the three-slab model derived from the ethanol

measurement is correct. While a positive value for Δ ρ suggests that the OTE layer

is swelling. On the other hand, a negative value for Δ ρ points either to a significant

reduction in OTE thickness or to an additional layer with an index of refraction less

than water’s.

d < 2.47 nm

Δρ < 0

Δρ = 0 Δρ > 0

d > 2.47 nm

Δρ < 0

Page 67: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

60

a b

Figure 3.2: Sum Frequency Spectra

The SFG spectra, taken in collaboration with Jeff Turner using a home-built

SFG setup, for the OTE monolayer/air interface (a) and the OTE monolayer/water

interface (b). There is very little difference between the two spectra indicating the

OTE monolayer did not reorder.

2850 2900 2950-1.0

-0.5

0.0

0.5

SFG

Sig

nal (

a.u.

)

IR Energy (cm-1)

2850 2900 2950-1.0

-0.5

0.0

0.5

SFG

Sig

nal (

a.u.

)

IR Energy (cm-1)

Page 68: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

61

Figure 3.3 Depletion Layer Thickness vs. Notional Water Fraction

Ellipsometry cannot uniquely determine the layer thickness and bulk water

volume fraction using the four-slab model. It can only confine these quantities to the

curve shown above.

0

0.5

1

1.5

2

2.5

3

0 0.2 0.4 0.6 0.8 1

Notional Water Fraction

Thic

knes

s (n

m)

Page 69: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

62

Figure 3.4 Three Possible Models for the Depleted Region.

The dashed and dot-dashed lines represent the step-like four-slab models for a

region that is completely depleted (Water volume fraction = 0) and 10% depleted,

respectively. The solid line represents the hyperbolic tangent model, which is more

realistic.

0 1 2 3

1.0

1.2

1.4

1.6

1.8

2.0

2.2D

iele

ctric

Pro

file

Z (nm)

Page 70: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

63

Figure 3.5 Reflectivity Curves for OTE Monolayers in Water and Air

X-ray reflectivity is compared in air (black filled symbols) and water (blue

open symbols) for OTE monolayers on <100> silicon wafers for both ambient (upper

set of curves) and degassed (lower set of curves) water. The solid cyan line is the fit

to the water data. Reflectivity RQ4 is plotted against Q on semilogarithmic scales.

1E-10

1E-7

1E-10

1E-7

R*Q

4

0.0 0.2 0.4 0.6 0.81E-13

1E-10

1E-7

0.0 0.2 0.4 0.6 0.81E-13

1E-10

1E-7

R*Q

4

Q(1/A)

Q(1/Å)

Page 71: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

64

-10 0 10 20 30 40 50 60

0.0

5.0x10-6

1.0x10-5

1.5x10-5

2.0x10-5

z (Å)

ρ (1

/Å2 )

-10 0 10 20 30 40 50

8.0x10-6

1.2x10-5

1.6x10-5

2.0x10-5

ρ(1/

Å2 )

SiSiOx

Head Group + roughness

OTE

Depletion Layer

bulk

z (Å)

a

b Figure 3.6 Calculated Electron Density Profiles

The calculated electron density profile for a monolayer of OTE is shown in a.

It was used to fix the parameters for the OTE in water. The electron density of OTE

in water is shown in b. The labeled regions are a guide to distinguishing the structure.

The non-monotonic dip for the depletion layer is required to produce the phase shift

seen in Figure 3.5.

Page 72: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

65

Figure 3.7 Comparison of the Depletion Layer Properties

The best fit from the X-ray reflectivity for the properties of the depletion layer

are marked with a green star. The shaded regions represent acceptable fits found with

X-ray reflectivity for degassed water (red crossed circles, red shading) and ambient

water (black half filled squares, cyan shading). The fit curve obtained from

ellipsometry is shown (blue triangles).

1 2 3 4

0

10

20

30

40

50

Pre

cent

Bul

k W

ater

Den

sity

Thickness (Å)

Page 73: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

66

Figure 3.8 Temperature Dependence of the Ellipticity.

The ellipticity, normalized to room temperature, is plotted for both OTE in

ethanol (red triangles) and water (blue squares). As can be seen there is no

significant change of the ellipticity in ethanol, while in water the ellipticity increases

greatly with increasing temperature.

.

-0.0015

-0.001

-0.0005

0

0.0005

0.001

0.0015

0.002

10 20 30 40 50

Temperature (C)

Nor

mal

ized

Elli

ptic

ity

Page 74: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

67

Figure 3.9 Temperature Dependence of Δ ρ

The temperature dependence of the ellipticity is plotted for samples with

different contact angles. The squares and diamonds represent a sample with an

advancing water contact angle (θ) of 107 ± 2 in degassed and non-degassed water

respectively. The triangles represent a sample with θ = 104 ± 3, in non-degassed

water. The circles represent a sample with θ = 109 ± 3, in non-degassed water. The

inset shows the reversibility of the measurement on this sample. The filled symbols

are the data series taken while heating the sample and the open symbols where taken

while cooling.

-0.006

-0.005

-0.004

-0.003

10 20 30 40 50Temperature

Δρ

lb28a wgtime

w924h w924a

-0.0038

-0.0036

-0.0034

-0.003230 32 34 36

wgtime heating wgtime cooling

Page 75: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

68

Figure 3.10 Temperature Dependence of Depletion Layer Thickness

The thickness of the depleted region, calculated with hyperbolic tangent

model is plotted versus temperature.

0.25

0.3

0.35

0.4

0.45

0.5

0.55

10 20 30 40 50

Temperature (C)

Thic

knes

s (n

m)

Page 76: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

69

Figure 3.11 Temperature Dependence of the Density of the Depletion Layer

The bulk water volume fraction found with the four-slab model and a 2.6-nm

layer thickness is plotted versus temperature.

0.876

0.886

0.896

0.906

0.916

0.926

0.936

10 20 30 40 50

Temperature

Bul

k W

ater

Vol

ume

Frac

tion

Page 77: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

70

a b c Figure 3.12 Schematic of Coherent and Incoherent Bubble Models

Two models can be used for fitting X-ray reflectivity curves in the presence of

bubbles. If the putative bubbles are large than the coherence length, the incoherent

model applies as illustrated in the left section of a. Then, a linear combination of the

reflectivity curves is taken (b). If the bubbles are smaller than the coherence length,

the coherent model, illustrated in the right section of a, is used. In this case, a linear

combination of the electron densities is taken (c).

Incoherent Model Coherent Model

0 .00 0 .25 0 .50 0 .751E -13

1E -10

1E -7

1E -4

0 .1

R

Q (1 /A )

- 2 0 0 2 0 4 0

0

1 x 1 0 - 5

2 x 1 0 - 5

z (Å )

ρ (1

/Å2 )

Page 78: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

71

Figure 3.13 Incoherent Model

The measured X-ray reflectivity in degassed water for self-assembled OTE

monolayers shown in Figure 3.5 is compared to a nanobubble model in which

scattering occurs incoherently, in part from OTE in contact with bulk water, in part

from OTE in contact with air. Reflectivity RQ4 is plotted against Q on

semilogarithmic scales. The open symbols are experimental data. The solid line is

the calculated fit for 0% surface coverage of water. The dotted line, dash-dotted, and

dashed lines are the fits calculated for 30%, 70% and 100% water surface coverage.

The vertical lines are to guide the eye.

0.00 0.25 0.50 0.75

1E-15

1E-12

1E-9

R*Q

4

Q(1/Å)

70%

30%

0%

100%

Page 79: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

72

Figure 3.14 Coherent Model

The measured X-ray reflectivity in degassed water for self-assembled OTE

monolayers shown in Figure 3.5 is compared to a nanobubble model in which

scattering occurs coherently, in part from OTE in contact with bulk water, in part

from OTE in contact with air. Reflectivity RQ4 is plotted against Q on

semilogarithmic scales. The open symbols are experimental data. The dashed line is

the calculated fit for 0% surface coverage of water. The solid line, dotted, and dash-

dotted lines are the fits calculated for 30%, 70% and 100% water surface coverage.

The vertical lines are to guide the eye.

0.00 0.25 0.50 0.75

1E-14

1E-11

1E-8R

*Q4

QQ(1/Å)

30%

0%

100%

70%

Page 80: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

73

Figure 3.15 Dependence on Sample Quality

Not all OTE monolayers were good enough to display the depletion layer.

The number of occurrences for each depletion layer thickness, calculated with the

four-slab model assuming a fully depleted region (0% bulk water fraction), are shown

above. The negative values indicate swelling in the OTE layer.

-0.6 -0.4 -0.2 0.0 0.2 0.4 0.6 0.80

2

4

6

8

10

12

14

16

18N

umbe

r of O

ccur

ence

s

Depletion Layer Thickness (nm)

Page 81: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

74

Chapter 4: Dynamic Properties of the Water/Hydrophobic Interface

The dynamics of water at a hydrophobic surface has many important

implications especially for protein folding. Yet very few experiments have been done

to probe these dynamics. When we first looked at the depletion layer thickness

versus time with ellipsometry, we noticed that there were fluctuations on many

different time scales, as can bee seen in Figure 4.1. The existence of these large

fluctuations are especially surprising given that ellipsometry inherently averages both

spatially, over a beam size of approximately 10 μm, and temporally, with a time

constant of 30 ms.

It is difficult to really understand the dynamics of the interface by just looking

at the time traces. We have employed three forms of analysis to get a better handle

on the fluctuations; histograms, autocorrelations, and power spectra. Each technique

provides a different perspective on the fluctuations.

4.1 Histograms

Histograms are created by a three step process. First the measured data is

divided into equally spaced bins. Then the number of data points in each bin is

counted. Finally the results are plotted in a bar chart.

Histograms provided much information on the data analyzed [4.1]. They

show the mean and the distribution about that mean. Also they highlight the presence

of outliers, the skewness of the data and whether multiple modes exist.

According to the Central Limit Theorem, a distribution of N statistically

independent measurements will be Gaussian for large N. As we measured over one

Page 82: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

75

hundred thousand points in each time trace, we would expect our histograms to

Gaussian. Deviations from the expected Gaussian shape can reveal interesting

insights. Two of the main causes for non-Gaussian distributions are underlying

deterministic models, such as sinusoidality in the data, and the mixing of probability

models, due to multiple states [4.1]. With sinusoidal variation, the system spends

more of its time in the extrema of the waveform where the velocity is the slowest than

it does near the mean value, thus producing non-Gaussian distributions. In the case of

a system with multiple states, as the system transitions between the states, their

various distributions will be combined. The resulting distribution will be bimodal for

a two-state system but in general can take on a variety of different non-Gaussian

distributions.

Figure 4.2 shows histograms for three different systems, taken from

ellipsometry measurements. The low amplitude tails seem on all three histograms are

characteristic of all the histograms measured with ellipsometry and probably

represent the distribution of noise convoluted with the signal. The first system, the

hydrophobic/ethanol system, consists of OTE in ethanol (Figure 4.2 a). As stated

earlier, ethanol has a low contact angle with OTE and a depletion layer is not

expected. This system has the expected histogram. It is sharply peaked with a well-

defined mean and Gaussian-like distribution around the mean. In contrast, the

hydrophobic/water system has a very different histogram (Figure 4.2 b). It is not

sharply peaked and has a flattened top. It has a non-Gaussian distribution. This

histogram is especially puzzling when compared to the water/hydrophilic system,

which consists of silicon oxide in water (contact angle = 0o), shown in Figure 4.2 c.

Page 83: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

76

The hydrophilic/water histogram looks like the hydrophobic/ethanol histogram, with

a sharply peaked distribution.

What could be the cause of the non-Gaussian distributions in the

water/hydrophobic system? Since the other two systems are not expected to have and

depletion layer, while the water/hydrophobic system is, we believe the distribution is

caused by some property of the depletion layer. As stated above, non-Gaussian

distributions can arise from sinusoidality in the data. In the water/hydrophobic

system, sinusoidal variation may be arising from capillary waves forming at the

water/depletion layer interface. This kind of fluctuating interface is suggested by

McCormick’s simulation [4.2]. He found that removing spatial Fourier modes with

wavelengths between 1.5 nm and 19 nm significantly changed the temporal Fourier

behavior, indicating the interface fluctuates on many length scales.

Also non-Gaussian distributions can arise from multiple states existing within

the system. In the hydrophobic/water system, multiple states may be evidence for a

wandering interface as described by Grunze [4.3]. His simulations show shifting of

the density distribution between runs. This gives a picture of the depletion layer

jumping between different thicknesses as time goes by.

We have used SPR to confirm the non-Gaussian behavior of the

water/hydrophobic histograms. As shown in Figure 4.3, the histogram of ODT in

ethanol (Figure 4.3 a) has the expected Gaussian shape, while the histogram of ODT

in water is bimodal (Figure 4.3 b). Figure 4.4 also shows the characteristic bimodal

histogram in water and a Gaussian histogram in air.

Page 84: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

77

Now that we know the water/hydrophobic interface demonstrates non-

Gaussian fluctuations we can explore how other factors influence the dynamics of the

depletion layer.

4.1.1 Dependence of Dynamics on Temperature

We saw that temperature had a very large effect on the static properties of the

depletion layer. What effect does it have on the dynamic properties? To test the

temperature dependence, we took time traces for two temperatures separated by 30 oC

at the same spot on the sample. The resulting histograms are plotted in Figure 4.5.

As can be seen in Figure 4.5, the histograms are very similar in appearance. The one

taken at 56 oC is slightly broader; however histograms taken on other samples show

slight broadening with a decrease in temperature.

It is surprising that temperature has such a small effect on the fluctuations

when it had such a large effect on the static properties of the depletion layer. In the

static situation the depletion layer dissipated as temperature was increased, however,

it did not disappear over the temperature range available to us. In 2005, Chandler

proposed that although in some situations the depletion layer may become negligibly

thin its main defining characteristic, the size of its fluctuations, would remain

unchanged [4.4]. It seems this is what is occurring here. The depletion layer is

becoming thinner with increased temperature yet it continues to have the same size

fluctuations.

Page 85: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

78

4.1.2 Dependence of Dynamics on Roughness

In the course of our experiments, we found that surface roughness has a

profound effect on the shape of the histogram. Figure 4.6 shows the histogram from a

rough surface taken with ellipsometry. The histogram is obviously very different

form the histograms of the smooth monolayers shown above. It has two widely

spaced narrow peaks and a fairly U-shaped distribution.

How can the roughness cause such a drastic change in the histogram? We

found that a histogram constructed from a single sine wave produces a very similar

U-shaped distribution (see the inset in Figure 4.6). We think that the protrusions on

the monolayer surface may be pinning the capillary waves and creating standing

waves at the interface.

4.1.3 Dependence of Dynamics on Contact Angle

We expect the fluctuations to disappear when the sample is no longer

hydrophobic enough to form a depletion layer. In fact, Huang et al. found that the

strength of the liquid surface interaction strongly affected the characteristics of the

depletion layer [4.5]. With ellipsometry, we preformed a series of experiments on

mixed monolayers of octadecanethiol (ODT) and 11-mercapto-1-undecanol (R-OH T)

to look at the effect of hydrophobicity on the fluctuations. Figure 4.7 shows the

histograms for two different mixtures in water. The surface used in Figure 4.7 a has

the same broadened flat-topped peak as seen with OTE in water. The second

histogram (Figure 4.7 b) is narrower but surprisingly it still has a flat-topped

distribution even though the water contact angle is only 19o.

Page 86: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

79

To get a clearer picture of what was happening; we have plotted the standard

deviation, which is the distance where the distribution drops to approximately 60% of

its maximum value, against the contact angle, as shown in Figure 4.8. The

distribution becomes narrower as the contact angle decreases. At approximately 50

degrees, the standard deviation drops below that of the OTE/Ethanol system,

indicating that the depletion layer no longer exists. It appears that, as expected, the

distribution becomes more sharply peaked as the contact angle is decreased; however

the fluctuations persist much longer than expected.

Another way to look at the effect of the surface liquid interaction strength on

the fluctuations is to vary the liquid while leaving the surface unaltered. This

reciprocal situation has an advantage for ellipsometry in that we do not need to use

thiols. The results are shown in Figure 4.9. The fluctuations are not well correlated

to the solvent contact angle.

However, if instead we plot the standard deviation against the relative polarity

of the solvent, we find a much better correlation. See Figure 4.10. The relative

solvent polarity (p) is defined relative to water (i.e. for water p=1), and indicates the

magnitude of the dipole moment of the solvent. We observe that the distribution

becomes much narrower as the solvent becomes less polar.

Why would the fluctuations change with solvent polarity instead of contact

angle? One clue can be found in an experiment done by Cho et al. [4.6]. They found

that the slip length, the extrapolated distance into the solid where the fluid velocity

becomes zero, on a hydrophobic surface correlated not with contact angle as expected

but with solvent polarity. They also found that the more polar liquids had unusually

Page 87: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

80

strong repulsions at very slow piezodrive speeds. They believe that this strong

repulsion is due to dipole-dipole image interactions, which in this case are repulsive

and tends to align the dipoles parallel to the surface. Cho et al. posit that it is the

highly structured nature of the dipole liquids that make it more difficult to slip. In our

case if the inherent desire of high polarity liquids to order is frustrated by an

incommensurate surface, there will be a larger incentive to form a second interface

and create a depletion layer. Also the stronger repulsion between the surface and

liquid enhances the formation of a depletion layer.

4.1.4 Dependence of Dynamics on Dissolved Gases

The effect of absorbed gases on hydrophobic surfaces could have implications

on a wide variety of phenomena, such as fluid flow and lubrication. To see what

effect absorbed gas would have on our fluctuations; we first used ellipsometry to

measure the fluctuations in degassed water. Then we allowed the water to equilibrate

with ambient conditions and remeasured the fluctuations. The results are shown in

Figure 4.11. Both histograms are qualitatively the same. This indicates that absorbed

gases have little effect on the fluctuations at a water/hydrophobic interface.

4.1.5 Dependence of Dynamics on pH

We also looked at how changing the pH of the solution would effect the

fluctuations. Varying the pH will change the surface charge density of the

hydrophobic surface, as explained by Grunze et al. [4.7]. Auto-ionization of water

creates hydronium and hydroxide ions that can then preferentially absorb onto the

Page 88: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

81

hydrophobic surface. This creates significant surface charging; for ODT at pH 6 with

10-3 M KCl, there are 1.8x10-2 excess charges per nm2. The literature suggests a pH

dependent surface charging on hydrophobic surfaces with a pKa ≈ 4. We measured

the fluctuations with ellipsometry for a range of pH. Figure 4.12 displays the results.

When we decrease the pH from 8 to 6, that is as the surface becomes less negative,

the histograms take on a more “flat-topped” shape. This change in shape may

indicate that the surface is going from a wet state (Gaussian shape) to a dry state (flat-

topped shape) due to decreased absorption of hydroxide ions with decreasing pH.

Absorbed hydroxide ions could allow hydrogen bonding to the surface in effect

decreasing the hydrophobicity of the surface, and therefore reducing the fluctuations.

4.2 Power Spectra

Power spectra are another way to investigate the dynamics at the

water/hydrophobic interface. The power spectrum is made by squaring the Fourier

transform of the time trace and plotting it versus frequency on a logarithmic scale.

This graph indicates the frequencies and the relative amounts of the sine waves

making up the time trace. The power spectrum of a pure sine wave, for example

sin(2πft), will have one spike positioned at the frequency of the sine wave, in this

case f. A linear combination of a series of sine waves will be made up of many

spikes. However, many power spectra show not spikes but functions that vary

smoothly with frequency. These spectra are a result of time traces made up of non-

discrete frequencies. The power spectrum of white noise is a horizontal line, as it

consists of all frequencies equally.

Page 89: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

82

Using ellipsometric time series, we have calculated the power spectra for

many systems. The power spectrum from a water/hydrophilic interface made up of

water on silicon oxide is shown in Figure 4.13. It has the horizontal shape indicative

of white noise. Figure 4.14 displays the power spectrum of water at a hydrophobic

surface of OTE. The power spectrum displays a linear shape with a slope equal to

negative two. At the OTE/ethanol interface a different shape is observed. As seen in

Figure 4.15, the power spectrum is not completely flat as for the water/hydrophilic

system but does not have the slope characteristic of the water/hydrophobic interface.

This slope of – 2 suggests that the fluctuations are due to discrete entities as

smooth variations would decay more rapidly [4.8]. However it could also be a result

of long time scale drift in the data as this often turns up as a slope of negative two.

We found no satisfactory way to remove this possibility from the data. Instead we

switched to a different technique, autocorrelation. The power spectrum and the

autocorrelation function form a Fourier transform pair. Therefore much of the same

information can be gathered from autocorrelation curves that could be found in power

spectra.

4.3 Autocorrelations

Autocorrelations are good at determining whether a data set is random

[4.9, 4.10]. If it is not random, it is also good at determining what type of time series

model would be appropriate for that data set [4.9, 4.10]. Autocorrelations measure

the correlation between measurements in a data set with a specific time lag (τ). The

autocorrelation function, G(τ), can be defined as

Page 90: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

83

( )( ) ( )[ ] ( ) ( )[ ]

2F

tFtFtFtFG

−−−−=

τττ (4.1)

where F(t) is the fluctuation as a function of time, and <F> is the average value of the

fluctuations [4.11] . This equation can be simplified to

( ) ( ) ( )12 −

−=

F

tFtFG

ττ (4.2)

The autocorrelation function is plotted against the time lag to make the

autocorrelation curve. If the autocorrelation curve is flat and very close to zero, then

the system is random [4.9]. Sinusoidal behavior will produce an autocorrelation

curve with an alternating series of positive and negative spikes [4.9]. For a multiple

state system we expect that the system will remain in a single state for a characteristic

amount of time (t1) and then will switch to another state. In this case we would

expect the autocorrelation curve to be roughly step shaped, with positive

autocorrelation up until t1 and then dropping to near zero.

Looking at the shape of the autocorrelation curve could help us determine

which proposed model, capillary waves or wandering interface, best describes the

dynamics at the water hydrophobic interface.

4.3.1 X-Ray Photon Correlation Spectroscopy

Coherent scattering from a disordered surface produces a speckle pattern that

reflects the spatial arrangement of the surface [4.12]. X-ray photon correlation

spectroscopy (XPCS) measures the time variation of the speckle pattern, and

therefore provides information on the dynamics of the interface [4.12].

Page 91: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

84

We used XPCS to attempt to observe the fluctuations of the

water/hydrophobic interface using a thin layer of water sandwiched between Kapton

and the OTE monolayers. We made two types of measurements with XPCS. In the

first set, the static measurements, the entire measurement was collected from one spot

on the sample; leading to an exposure time of approximately one minute. The static

measurements were made five times before moving to a new location on the sample.

In the second set, the moving measurements, the sample was translated during the

collection time; leading to a one to three second exposure time due to the finite size of

the beam. The moving measurements were repeated once before moving to a fresh

area.

The static measurements showed a very weak correlation around of 1.002 at

approximately 175 ms but only for the first set of scans on a sample. See Figure 4.16

a. Subsequent scans on the same sample over ten microns way failed to show this

correlation as displayed in Figure 4.16 b. The photon intensity also showed different

behaviors for the first scan on the sample. It would increase slightly for 30 seconds

and then decrease even more gradually over the next four scans. This did not occur

again when we moved to a new spot on the sample. This strange behavior could be

due to beam damage. As reported by Tidswell, photoelectrons produced in the silicon

wafer could cause widespread damage of the monolayer [4.13]. In this case the

correlations we observed could be of some X-ray induced reaction.

In an attempt to minimize beam damage we switched to moving

measurements. Autocorrelation curves taken from the moving measurements

Page 92: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

85

displayed no correlations as can be seen from Figure 4.17. The series of spikes

displayed in each set of curves is due to the pulsed nature of the electron beam.

What is the meaning of our negative result? Of course it could mean that

there are no characteristic times in the fluctuations to observe. However the amount

of beam damage produced could also be interfering with the observation of such

dynamics. As stated in Section 3.6, the observation of the depleted region is highly

dependent on the sample quality. The much higher flux required for the XPCS

observably damaged the sample within a few seconds and may have destroyed the

depletion layer.

4.3.2 Surface Plasmon Resonance Correlation Spectroscopy

By switching to SPR, we can look at the autocorrelations without having to

worry about damaging the sample. In SPR, the dynamic measurement recorded the

change in reflected intensity versus time. The change in intensity corresponds to a

change in the resonance angle as explained in Chapter 2. As the thickness or density

at the water/hydrophobic interface change, the resonance frequency is shifted.

Therefore, the time-scale of the change in reflected intensity corresponds to the time-

scales for fluctuation at the interface.

We collected dynamic measurements of ODT in water and in ethanol. The

autocorrelation curve from ODT in ethanol is shown in Figure 4.18 a. At small time

lags, around 1x10-3 s, there is a peak we see in all autocorrelation curves taken with

the SPR. We think it may correspond to some sort of electrical noise. The rest of the

curve is flat and close to zero. This indicates there is no characteristic time for

Page 93: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

86

ethanol on ODT. The autocorrelation curve from ODT in water is shown in Figure

4.18 b. Again we see the noise peak at small time scales. However, there is a very

definite step in the correlation in the tenths of seconds region. This indicates that the

water/ODT system has a characteristic time of 0.5-0.6 s.

We did not expect to see a characteristic time that was so long. We would

expect that any fluctuations to occur on the millisecond or microsecond time scales.

Yet the XCPS also showed a slight correlation on the tenths of seconds time scale, so

it seems to be real.

What could this characteristic time indicate? It could mean that the surface

switches between states on the order of two to three times per second. Capillary

waves will also have a characteristic time [4.12]. In this case, the characteristic time

changes with the area interrogated. In order to try to distinguish between these two

possibilities, we looked at how the characteristic time changed with area.

4.3.3 Area Dependence

We used an iris to change the size of the incident laser beam in order to

change the plasmon excitation area. The results are shown in Figure 4.19. Although

the correlation is not as sharp as on other samples, there is no noticeable shift in the

characteristic time for all the areas we where able to employ. However, we were only

able to change the lateral size by less than a factor of 2.5 from 1.9 mm to 4.5 mm,

which may not have been enough to observe the effect.

Page 94: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

87

4.3.4 Contact Angle Dependence

We saw that the histograms were effected by changing the contact angle of the

hydrophobic substrate in Section 4.1.3. Does the contact angle have any effect on the

characteristic time?

To answer this question we made a series of mixed thiol monolayers with

contact angles between 102 degrees and 70 degrees. We then collected dynamic

measurements in water. The results are displayed in Figure 4.20.

As can be seen in Figure 4.20, the strength of the correlation decreases with

decreasing contact angle. The correlation has disappeared by the time the contact

angle reaches 70 degrees. Interestingly, the characteristic time does not change with

contact angle.

4.3.5 Dependence on Dissolved Gases

As stated previously, the effected of dissolved gases is thought to have an

effect on hydrophobic interfaces. To see if there was any dependence of the

autocorrelation curves on dissolved gases, we preformed an experiment where

degassed water was used to measure the time series with SPR than the experiment

was repeated with aerated water. The results are shown in Figure 4.21. Although the

correlation is not very sharp on this sample, there was very little difference between

the two curves.

Page 95: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

88

4.4 Conclusions for the Chapter

We have seen that the water/hydrophobic interface fluctuates in unusual ways.

First, using histograms to analyze the time series data collected, we observed that

water/hydrophobic systems have non-Gaussian shapes very different from those

observed in wetting systems. These fluctuations took on more Gaussian shapes as the

contact angle or the dipole moment were decreased. Temperature and degassing had

no effect.

Two possibilities have been put forward to explain the histograms. In the

first, capillary waves forming at the interface were proposed as the origin of the

fluctuations. Secondly, the idea of a wandering interface was suggested. In this case,

the depletion layer shifts between different thicknesses. Unfortunately, the histogram

analysis was not able to support one over the other.

Autocorrelation curves provided another chance at trying to pick between

these two possibilities. The autocorrelation curves of water/hydrophobic systems

were found to have step-like decay indicating a characteristic time in the tenths of

seconds time scale. As contact angle was decreased, the strength of the correlation

decreased but there was no change in the characteristic time scale. Changing the

beam size and degassing the water had little effect on the autocorrelation curves.

At first sight, the results from the autocorrelations make both capillary waves

and a wandering interface seem very unlikely. It is hard to imagine an interface

shifting between different thicknesses only a few times per second. Also the lack of

area dependence seems to rule out capillary waves. However, we only changed the

area of the beam size by a factor less than 2.5; according to Sinha et al., capillary

Page 96: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

89

waves logarithmically depend on the interrogated area [4.14]. In this case, the

expected change would have only been a factor of 0.86. Considering this evidence it

seems that capillary waves are the most likely source of fluctuations at the

water/hydrophobic interface.

4.5 References:

[4.1] Section “1.3.3.14 Histograms” NIST/SEMATECH e-Handbook of Statistical

Methods, <http://www.itl.nist.gov/div898/handbook/>, May 4 2006

[4.2] McCormick, T.A. Physical Review E 68 061601 (2003)

[4.3] Grunze, M., and A. Pertsin Journal of Physical Chemistry B 108 (42) 16533

(2004)

[4.4] Chandler, D. Nature 437 640 (2005)

[4.5] Huang X., R. Zhou and B.J. Berne Journal of Physical Chemistry B 109 3546

(2005)

[4.6] Cho, J.-H.J., B. M. Law and F. Rieutord Physical Review Letters 92, 166102

(2004)

[4.7] Chan, Y.-H.M., R. Schweiss, C. Werner and M. Grunze Langmuir 19 7380

(2003)

[4.8] Zhang, X., Y. Zhu and S. Granick Science 295 663 (2002)

[4.9] Section “1.3.3.1 Autocorrelations” NIST/SEMATECH e-Handbook of Statistical

Methods, <http://www.itl.nist.gov/div898/handbook/>, May 9 2006

[4.10] Section “1.3.5.12 Autocorrelations” NIST/SEMATECH e-Handbook of

Statistical Methods, <http://www.itl.nist.gov/div898/handbook/>, May 9 2006

Page 97: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

90

[4.11] “Fluorescence Correlation Spectroscopy (FCS) Technical Manual” Third

Edition June 2000 ISS Champaign IL

[4.12] Madsen, A., T. Seydel, M. Tolan and G. Grübel Journal of Synchrotron

Radiation 12 786 (2005)

[4.13] Tidswell, I.M., T.A. Rabedeau, P.S. Pershan and S.D. Kosowsky Journal of

Chemical Physics 95 2854 (1991)

[4.14] Sinha, S.K., E.B. Sirota, S. Garoff and H.B. Stanley Physical Review B 38

2297 (1988)

Page 98: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

91

4.6 Figures

Figure 4.1 Time Traces of Ellipsometry Measurement

Times traces calculated using the hyperbolic tangent model are plotted on

many different time scales.

500 1000 15000.52

0.54

0.56

Time (s)

510 530 5500.52

0.54

0.56 502 506 5100.52

0.54

0.56

Dep

letio

n La

yer T

hick

ness

(nm

)

500.2 500.6 501.00.52

0.54

0.56

Page 99: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

92

a b c Figure 4.2 Histograms of Three Systems

The histograms for the change in thickness are shown for three different

systems; ethanol and OTE (a), water and OTE (b), and water and silicon oxide (c).

Δ h (nm)

Water/Hydrophobic

Δh (nm)

Water/Hydrophilic

Δ h (nm)

Ethanol/Hydrophobic

Page 100: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

93

a

b

Figure 4.3 Histograms Taken with SPR

The histograms were taken with SPR of ODT in ethanol (a) and in water (b).

-0.03 -0.02 -0.01 0.00 0.01 0.02 0.030

1000000

2000000

3000000

4000000

Num

ber

Δ R

-0.03 -0.02 -0.01 0.00 0.01 0.02 0.030

300000

600000

900000

1200000

1500000

1800000

Num

ber

Δ R

Page 101: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

94

a b Figure 4.3 Histograms Taken with SPR in Air and Water.

The histograms were taken with SPR of ODT in air (a) and in water (b).

-0.010 -0.005 0.000 0.005 0.0100

200000

400000

600000

800000

Num

ber

Del ReflectivityΔ Reflectivity

-0.010 -0.005 0.000 0.005 0.0100

200000

400000

600000

800000

1000000

1200000

1400000

1600000

Num

ber

Δ Reflectivity

Page 102: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

95

a b Figure 4.5 Effect of Temperature on the Histograms

Histograms taken on OTE in water are shown for two different temperatures;

56 oC (a) and 26 oC (b), on the same spot on the sample.

-0.015 -0.010 -0.005 0.000 0.005 0.0100

2000

4000

6000

8000

10000

12000

Num

ber

Δh(nm)

-0.015 -0.010 -0.005 0.000 0.005 0.010 0.0150

2000

4000

6000

8000

10000

12000

Num

ber

Δh(nm)

Page 103: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

96

Figure 4.6 Effect of Roughness on Histograms

The histogram of fluctuations in height for a rough OTE surface in water was

taken with ellipsometry. The inset shows a 5 x 5 μm height contrast AFM image of

the surface.

Δ h (nm)

Page 104: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

97

a b Figure 4.7 Contact Angle Dependence in Histograms

The histograms were taken with ellipsometry for thiols in water. In a, a 2:1

mixture of ODT and R-OH T was used, while in b a pure R-OH T layer was used.

Δ h (nm) -0.10 -0.05 0.00 0.05 0.100

200400600800

10001200140016001800

Num

ber

θ = 97.3 ± 1.1

Δ h (nm)-0.10 -0.05 0.00 0.05 0.100

500

1000

1500

2000

Num

ber

θ = 19.0 ± 1.4

Page 105: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

98

Figure 4.8 Standard Deviation versus Substrate Contact Angle

The standard deviation, illustrated in the inset, for histograms taken by

ellipsometry on mixed thiol monolayers in water was plotted against substrate contact

angle. The trend is highlighted by the blue line. The pink line represents the standard

deviation of ethanol/OTE system.

0.001

0.002

0.003

0.004

0.005

0 50 100Advancing Contact Angle

Stan

dard

Dev

iatio

n

-0.02 0.00 0.020

10000

20000

30000

40000

Num

ber

Δh (nm)

Page 106: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

99

Figure 4.9 Standard Deviation versus Solvent Contact Angle

The standard deviation of histograms taken on OTE with ellipsometry in

different solvents is plotted versus the solvent contact angle.

0 20 1000.00

0.01

0.02S

tand

ard

Dev

iatio

n

Contact Angle

Page 107: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

100

Figure 4.10 Standard Deviation versus Relative Solvent Polarity

The standard deviations of histograms taken with ellipsometry for OTE in

various solvents are plotted against relative solvent polarity.

0.000.01

0.010.02

0.020.03

-0.1 0.1 0.3 0.5 0.7 0.9 1.1

Relative Solvent Polarity

Stan

dard

Dev

iatio

n

Cyclohexane

Toluene Ethanol

Water

Page 108: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

101

a b Figure 4.11 Histogram Dependence on Dissolved Gasses

Histograms of OTE in freshly degassed water (a) and then remeasured after

the water was allowed to equilibrate with the atmosphere (b) were taken with

ellipsometry at the same position on a single sample. The hyperbolic tangent model

was used.

H (nm)

Page 109: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

102

a

b c Figure 4.12 Histograms Dependence on pH

Histograms taken with ellipsometry on OTE in aqueous solution for pH equal

to 8.3 (a), 7.7 (b) and 6.4 (c) are displayed. The hyperbolic tangent model was used.

Δh (nm)

-0.005 0.000 0.0050

5000

10000

15000

20000

25000

Num

ber

-0.005 0.000 0.0050

5000

10000

15000

Num

ber

Δh (nm)

-0.005 0.000 0.0050

5000

10000

15000

20000

Num

ber

Δh (nm)

Page 110: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

103

Figure 4.13 Power Spectrum from a Water/Hydrophilic System

The power spectrum was calculated from ellipsometric data for a system

consisting of water and silicon oxide, which is hydrophilic. The line is included as a

guide for the eye.

Page 111: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

104

Figure 4.14 Power Spectrum from a Water/Hydrophobic System

The power spectrum was calculated from ellipsometric data for a system

consisting of water and OTE, which is hydrophobic. The line is included as a guide

for the eye.

-2

Page 112: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

105

Figure 4.15 Power Spectra from OTE in Water and Ethanol

The power spectra were calculated from ellipsometric data for OTE in water

(blue) and ethanol (red).

Page 113: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

106

a b Figure 4.16 X-ray Autocorrelations Static Sample

These X-ray autocorrelations were taken on a static sample. The first time a

scan was taken on the sample a slight correlation of 1.002 was seen at a time lag of

approximately 175 ms (a). Subsequent scans showed no correlation (b).

50 100 150 200

1.00

1.02

g(τ)

Time (ms)

50 100 150 200

1.00

1.02

g(τ)

Time (ms)

Page 114: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

107

Figure 4.17 X-Ray Autocorrelations on Moving Sample

This X-ray autocorrelation was taken on a moving sample, in an effort to

prevent beam damage. The scan shows no autocorrelation. Subsequent scans were

similar.

50 100 150 2000.90

0.92

0.94

0.96

0.98

1.00

1.02

1.04

1.06

1.08

1.10

g(τ)

Time (ms)

Page 115: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

108

a b Figure 4.18 Autocorrelation Curves from SPR

Autocorrelation curves for ODT monolayers with ethanol (a) and water (b) are

shown. The red lines are a guide for the eye.

1E-3 0.01 0.1 10.000

0.005

0.010

0.015

0.020

G(τ

)

τ(s)

1E-3 0.01 0.1 1 10

0.000

0.001

0.002

0.003

0.004

G(τ

)

τ(s)

Page 116: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

109

Figure 4.19 Effect of Beam Size on the Autocorrelation Curves

Autocorrelation curves for ODT in water were taken with different incident

beam sizes. Measured laser beam diameters of 1.9 mm (blue circles), 2.8 mm (red

triangles), and 4.5 mm (black squares) were used.

0.1 10.0001

0.0002

0.002

0.0030.1 1

0.0001

0.0002 d = 4.5 mm d = 1.9 mm

G(τ

)

τ(s)

d = 2.8 mm

Page 117: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

110

Figure 4.20 Effect of Contact Angle on Autocorrelation Curves

Autocorrelation curves are shown for monolayers of mixed thiols with

different contact angles in water. The blue squares, green circles and pink triangles

are for thiol layers with contact angles of 102, 97 and 70 degrees respectively. For

clarity, the blue curve was shifted +.0005 and the magenta curve was shifted -.0005

vertically.

1 10-0.001

0.000

0.001

0.002

0.003

G

(τ)

τ(s)

102 degrees 97 degrees 70 degrees

Page 118: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

111

Figure 4.21 Effect of Dissolved Gasses on Autocorrelation Curves

SPR autocorrelation curves taken with on the same ODT sample first with

degassed water (red squares) and then with aerated water (black circles).

0.1 1 100.0000

0.0005

0.00100.1 1 10

0.0000

0.0005

G(τ

)

τ (s)

Page 119: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

112

Chapter 5: Concluding Remarks

Water/Hydrophobic interfaces surrounds us. You can see many

examples in day to day life for example dew drop on a leaf, the folding of proteins in

our bodies and even liquids on stain resistant fabrics. Yet no well accepted picture of

what happens at such interfaces has been developed.

One theory that has been put forth is that a thin depleted region forms around

extended hydrophobic objects [5.1, 5.2]. If it exists this depletion layer would have

implications in such diverse areas as colloidal self-assembly, and the boundary

conditions of fluid flow. Another idea is that nanobubbles form at the interface [5.3].

Many experiments have seen evidence of the depletion layer [5.4-5.8]. Others have

seen evidence for nanobubbles [5.3, 5.9]. Yet others have no evidence for anything

different at water/hydrophobic interfaces [5.10-5.12].

In an effort to help clarify this situation, we have investigated

water/hydrophobic surfaces using three different experimental techniques;

ellipsometry, X-ray reflectivity, and SPR. These three techniques complement each

other, making up for each others weaknesses. For example, ellipsometry, being a

table top technique, is great for systematically changing variables such as pH and

temperature; however it cannot easily decouple thickness and density. On the other

had X-ray reflectivity can decouple thickness and density but it is impractical to run

an entire series of pH experiments.

In Chapter 3, we saw very strong ellipsometric evidence for the qualitative

existence of a depleted region, but could not quantitatively determine its properties.

Using X-ray reflectivity we also saw qualitative evidence for the existence of a

Page 120: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

113

depleted region. The best fit of the reflectivity data gave a thickness of 2.7 Å,

electron density corresponding to 20% water content, and width of the error function

profile (between the depletion layer and bulk water) of 4 Å. However there was still

considerable room to change the parameters and still get an expectable fit. After

combining the ellipsometry data and the reflectivity data, we found they agree quite

well and we can be fairly confident that the depletion layer has a thickness between

1.5 Å and 4 Å, and a density less than 45 % that of bulk water.

We also took advantage of the information on the lateral characteristics of the

interface provided by X-ray reflectivity. Using this information, we were able to rule

out the possibility of nanobubbles forming at the water/hydrophobic interface. This

finding contrast with AFM studies that saw evidence of nanobubbles [5.3, 5.9]. In

truth this result is not too surprising as nanobubbles have only been observed with

tapping mode AFM which may have nucleated the bubbles during the measurement.

We found that the depletion layer is highly dependent on temperature, contact

angle and sample quality. This dependence might explain some of the discrepancies

in the literature as different groups often use hydrophobic surfaces with different

properties.

Additionally, we looked at the dynamics of the hydrophobic/water interface.

Very few experiments have looked at the dynamics in this system despite its

fundamental importance. In fact, Chandler proposes that fluctuations at the

hydrophobic/water interface are the key distinguishing factor between wet and dry

interfaces [5.13]. We used both histograms and autocorrelation functions to get a

handle on how the interface was varying in time.

Page 121: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

114

In Chapter 4, we found that the histograms of time series taken from

water/hydrophobic systems had a strange non-Gaussian shape. While those taken

from wetting interfaces showed the expected Gaussian shape. This was observed

both in time series data collected with ellipsometry and SPR. We proposed two

possible reasons for the non-Gaussian shapes; (1) capillary waves at the boundary

between the water and depleted region, or (2) a wandering interface, where the

depleted region jumps between states with different thicknesses.

To try to distinguish which possibility was more reasonable, we switched to

looking at autocorrelations. We found that the autocorrelations indicated a single

characteristic time of 0.5-0.6 s. This time seems too long to attribute to a wandering

interface. Therefore, we believe the capillary wave model is most likely. However,

we were not able to observe the characteristic time change with area as expected with

capillary waves, most likely because we did not change the area by a large enough

factor.

Finally, we are able to build up a detailed picture of what happens at a

water/hydrophobic interface. We see that the water pulls away from the surface

leaving behind a depletion layer. Also we found that this depletion layer fluctuates

with something akin to capillary waves. All in all, we can say that water meets a

hydrophobic surface reluctantly and with fluctuations.

Page 122: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

115

5.1 References

[5.1] Stillinger, F.H. Journal Solution Chemistry 2 141 (1973)

[5.2] Lum, K., D. Chandler and J.D. Weeks Journal of Physical Chemistry B 103

4570 (1999)

[5.3] Tyrrell, J.W.G., and P. Attard Physical Review Letters 87 (17) 176104-1 (2001)

[5.4] Sur, U.K. and V. Lakshminarayanan Journal of Colloid and Interface Science

254 410 (2002)

[5.5] Steitz, R., T. Gutberlet, T. Hauss, B. Klösgen, R. Krastev, S. Schemmel, A.C.

Simonsen and G.H. Findenegg Langmuir 19 2409 (2003)

[5.6] Schwendel, D., T. Hayashi, R. Dahint, A. Pertsin, M. Grunze, R. Steitz and F.

Schreiber Langmuir 19 2284 (2003)

[5.7] Castro, L.B.R., A.T. Almeida and D.F.S. Petri Langmuir 20 7610 (2004)

[5.8] Doshi, D.A., E.B. Watkins, J.N. Israelachvili and J. Majewski., Proceedings of

the National Academy of Science 102 9458 (2005)

[5.9] Ishida N., T. Inoue, M. Miyahara and K. Higashitani, Langmuir 16, 6377

(2000).

[5.10] Mao, M., J. Zhang, R.-H. Yoon and W.A. Ducker Langmuir 20 1843 (2004)

[5.11] Takata, Y., J.H.J. Cho, B.M. Law and M Aratoni, Langmuir 22 1715 (2006)

[5.12] Jensen, T.R., M. O. Jensen, N. Reitzel, K. Balashev, G.H. Peters, K. Kjaer and

T. Bjørnholm Physical Review Letters 90 (8) 086101-1 (2003)

[5.13] Chandler, D. Nature 437 640 (2005)

Page 123: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

116

Appendix A: Use and Maintenance of the Ellipsometer

The purpose of this appendix is to explain the use and maintenance of the

ellipsometer so that in the future others can continue to make use of the experimental

apparatus.

A.1 Alignment and Calibration of the Ellipsometer

The alignment and calibration must be done on a regular basis, on average

once a year. Also if the absolute values of the voltages measured at the two analyzer

angles disagree significantly, the ellipsometer needs to be realigned. In the following

I lay out step by step instructions for the alignment and calibration.

A.1.1 Aligning the Laser

The normal operating arrangement of the ellipsometer is shown in Figure A.1.

First, the nut at the center joint of the arms must be removed, and then the arms are

carefully lifted off and oriented parallel to each other. Next, remove all elements

from the incident arm except the laser. Be very careful when moving the phase

modulator. It is very delicate and could break if shaken or jarred. Make sure to sit it

flat on the bench so it cannot fall. Cover the analyzer and the photomultiplier tube

(PMT) with post-it notes. This ensures that the laser cannot shine directly on the

PMT which could burn it. This configuration is shown in Figure A.2. Using two

irises of the same height align the laser. This is done by making sure the laser beam

passes through both openings when one is close to the laser and the other is far away

(i.e. on the second arm). The alignment of the laser can be controlled as shown in

Page 124: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

117

Figure A.3. After the laser is aligned put the arms back on the center joint. There is

no need to replace the nut yet.

A.1.2 Finding the p-Polarization

Now align the arms such that the laser hits the silicon sample at 75 degrees

and is reflected at 75 degrees and hits the analyzer. This is shown in Figure A.4. The

sample may need to be rotated to achieve this geometry. This ensures the light is at

the Brewster angle for silicon. In this section it is important to use a fresh and clean

silicon sample. Details of silicon cleaning are given in Appendix B. Replace the

polarizer on the incident arm. Make sure that the light reflected from the polarizer

hits the laser just above or below the incident beam. Adjust the polarization until the

reflected light is extinguished. It is important that the reflected beam is clean (i.e.

very little scattering). If it is not, clean the glass sample holder and make a new

silicon sample. The angle that light is extinguished is the p-polarization angle, p. Do

this step twice and average the values. They may be displaced by 180 degrees; if this

is true, subtract 180 from the larger and then average them. It is important to

remember that the laser itself is also polarized and so this can also cause a minimum

in the reflected light. To avoid this problem, check that the light after the polarizer

but before the sample is still strong when the reflected light is extinguished. If the

only minimum occurs when the light after the polarizer is weak, rotate the laser body

30 degrees. You will have to go back and realign the laser in this case and then find

the p-polarization. Write down the p-polarization angle.

Page 125: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

118

A.1.3 Finding the Analyzer Angles

Set the polarizer at p+45o. Remove the arms from the center joint and make

them horizontal. Remove the post-it notes from the analyzer but not from the PMT.

Adjust the polarizer angle until the transmitted light is extinguished. Do this twice.

The average is a. Set the polarizer at p-45o, and find the minimums, the average is a

+ 90o.

A.1.4 Finding the p-Axis of the Phase Modulator

Replace the phase modulator on the incident rail, aligning it such that the

reflected light hits just above the incident beam on the laser, and the beam passes

through the center of the phase modulator. Turn on the power supplies. On the

power supply (large brown box) turn on the power (green light), wait 30 s and then

turn on the high voltage (red light). Then turn on the current controller. See Figure

A.5. Add an n-d filter between the laser and the polarizer. Remove the post-it notes

from the PMT. Set the polarizer at p, and the analyzer at a. Set the lockin harmonic

to 2 and the phase to 2ω. Unlock the set screw on the phase modulator and adjust the

voltage until R is a minimum (X will go negative and positive). Record the angle of

the minimum. Repeat with the analyzer at a + 90o. The average of these two

measurements is the p-axis of the phase modulator. Carefully tighten the set screw

without changing the polarization of the phase modulator.

Page 126: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

119

A.1.5 Setting Jo(δo) to Zero

Set the polarizer at p+45o. On the power supply, switch the power supply

from amplifier to supply. Turn the voltage with the large dial until the current on the

current controller reads 0.7. See Figure A.6. Measure the DC voltage by connecting

the coaxial cable to the back of the current controller (labeled DC) and the voltmeter.

Rotate the little dial on the Beaglehole controller until it reads 0.0. Make a chart of

the voltage recorded from each setting of the Beaglehole controller for both a and a +

90o. An example is shown in Figure A.7. The goal is to find the controller position

that makes the voltage the same for both a and a + 90o. Once this position is found

lock the controller and switch the power supply back to constant current mode. To do

this dial the voltage on the power supply until the current reads 0.0, and switch back

to amplifier. This is the end of the alignment.

A.1.6 Calibration

With the analyzer set at a and the polarizer set at p+45o, change the phase on

the lockin until the voltage reads zero. Make sure the harmonic was set at 2. Then

push the – 90o button on the lockin and record the phase and voltage. Repeat with the

analyzer set at a + 90o. The average voltage is V2ω,cal and the average phase is 2ω.

Next we need to find ω. The arms should still be parallel. Put a quarter wave-

plate after the phase modulator. Make sure the reflected beam is just above the

incident beam, to do this you will have to turn off the power supply and remove the n-

d filter. Be sure to replace the n-d filter before turning on the PMT. Turn on the

PMT. Find the fast axis of the wave-plate by minimizing the voltage at the two

Page 127: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

120

analyzer angles with the harmonic set at 2. Lock the wave-plate. Set the lockin

harmonic at 1. Minimize the voltage with the analyzer set at a. Then push the + 90o

button on the lockin and record the phase and voltage. Repeat with the analyzer set at

a + 90o. The average voltage is Vω,cal and the average phase is ω. This is the end of

the calibration. Replace the arms on the center joint and replace the nut.

A.2 Taking measurements with the Ellipsometer

Taking measurements with the ellipsometer requires several intermediate

steps. The first is cleaning and assembling the sample cell. Next is finding the

Brewster angle. Finally, you can collect and analyze the data.

A.2.1 Cleaning and Assembling the Sample Cell

The parts of the sample cell are pictured in Figure A.8. The glass and Teflon

parts are cleaned by soaking them in base bath overnight and then rinsing with large

amounts of de-ionized water. The base bath is made by adding 1 part KOH to 5 parts

isopropyl alcohol. The glass should not be stored for long periods in the base bath as

this will cause it to become cloudy. The metal part is cleaned by soaking in nitric

acid overnight and then rinsing with large amounts of de-ionized water. After

cleaning, all parts are dried in the oven for over 30 min and then allowed to cool to

room temperature.

To assemble the sample holder first put the o-ring on the metal part. Next,

mount the sample, as shown in Figure A.9. The sample should be mounted loosely

enough to pull out easily but tightly enough not to shift when held vertically. It is

Page 128: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

121

important not to mount the sample too tightly as this will cause stress induced

birefringence and distort the ellipticity. Next, slip the metal part inside the glass

holder and secure it by tightening the bushing. The bushing should only be hand

tightened. The assembled sample cell is shown in Figure A.10.

A.2.2 Finding the Brewster Angle

Next the sample cell is placed inside the sample holder in the ellipsometer.

The set screws can be used to ensure the light is hitting the sample in the center and

reflecting out the sample holder opening toward the PMT. The vertical position of

the sample can be adjusted with the translation stage attached to the backside of the

sample holder. Once the sample is positioned and light is hitting the PMT, turn on

the power supply. Remember not to turn the PMT on if the n-d filter is out. The

lockin should be set with the harmonic at 2 and the phase at 2ω. Rotate the incident

arm until the voltage for both analyzer positions are the same in sign and magnitude.

While you are changing the incident angle, make sure the reflected light is hitting the

PMT. This position is the Brewster angle.

A.2.3 Collection and Analysis of Data

Once the ellipsometer is set at the Brewster angle, you can start data

collection. It is easiest to do this with the computer program “ellipsometer”, but I

will explain how to do it by hand. Set the lockin harmonic to 1 and the phase to ω.

Record the voltage for both analyzer positions. Repeat this procedure several times in

order to find the deviation in the data.

Page 129: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

122

The ellipticity can be calculated from the above voltages. First take the

absolute values of the voltages. Next take the average of the a and a + 90o voltages

for each measurement. Divide the average by two times Vω,cal, and this quantity

equals the ellipticity.

Page 130: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

123

A.3 Figures

Figure A.1 Photograph of the Ellipsometer.

Above is a picture of our home-built phase modulated ellipsometer in its

operating position.

PMT Laser

Sample Holder

Phase Modulator

Polarizer Analyzer

N-d filter

Page 131: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

124

Figure A.2 Aligning the Laser

This photo shows the geometry for aligning the laser. Post-it notes are used to

protect the PMT.

Page 132: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

125

Figure A.3 Adjusting the Laser

This photo shows the how to adjust the laser. Turning the top knob moves the

laser left and right (red arrows). The middle knob allows you to adjust the rotational

orientation of the laser (blue arrows). Turning the lower knob moves the laser up and

down (yellow arrows).

Page 133: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

126

Figure A.4 Finding the p-Polarization

This photo shows the geometry for finding the p-polarization. Both arms are

set at 75 degrees as shown.

75o

Page 134: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

127

Figure A.5 Turning on the PMT

This photo shows the current controller (a) and the power supply (b) in regular

operating mode.

a

b

Page 135: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

128

Figure A.6 Supply Mode

This photo shows the power supply in supply mode. The arrow indicates the

knob used for setting the current.

Page 136: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

129

Controller a a+90o

0 3.3 mV 1.48 V 100 46 mV 1.45 V 300 166 mV 1.35 V 500 .33 V 1.19 V 600 .52 V 1.01 V 550 .71 V .82 V 525 .87 V .66 V 537 .79 V .74 V 531 .755 V .783 V 534 .761 V .761 V

Figure A.7 Controller Chart

Here is an example chart used to find the proper value for the Beaglehole

controller. We are trying to find the controller value where the voltages are equal. It

helps to start by changing the controller value in large steps and seeing where the

voltages cross each other. Then you can use smaller steps as you close in on the

proper value.

Page 137: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

130

Figure A.8 Components of the Ellipsometer Sample Cell

The components of the sample cell are pictured above. The o-ring is covered

with Teflon.

Glass Part

Metal Part

O-Ring

Bushing

Page 138: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

131

Figure A.9 Metal Part/Sample Assembly

This picture shows the correct way to mount the sample on the metal part.

Remember to put the o-ring on the metal part before mounting the sample.

Page 139: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

132

Figure A.10 Assembled Sample Cell

The fully assembled sample cell is shown. Note when adding liquid do not

fill it above the bottom of the metal piece expect for degassed fluids.

Page 140: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

133

Appendix B: Sample Preparation

The purpose of this appendix is to provide information on sample preparation

detailed enough so that future students can continue preparing high quality samples.

This appendix is broken into three parts; preparing hydrophilic silicon, preparing

OTE monolayers and preparing thiol monolayers.

B.1 Hydrophilic Silicon

Preparing clean hydrophilic silicon substrates is important not only for use as

hydrophilic surfaces but also as a first step in making OTE monolayers. In order to

produce clean silicon, you need to use clean glassware. You will need one

scintillation vial per sample and several pipettes. The glassware should be cleaned by

soaking in base bath overnight, rinsing thoroughly with deionized water and drying in

the oven.

Next, the silicon is soaked in Piranha solution heated to 70 oC for at least one

hour. Piranha solution is made by mixing three parts hydrogen peroxide with seven

parts sulfuric acid. This reaction is exothermic and extreme caution should be taken.

Next, the silicon should be rinsed twenty times with deionized water and dried in the

oven for two to three hours.

Then the silicon should be cleaned in the UV/ozone cleaner. Place the silicon

samples shiny side up on to the Teflon o-rings inside the UV/ozone cleaner. Adjust

the lab jack until the samples are approximately a half inch from the UV lamp. Flow

a small amount of oxygen into the cleaner. Close the cleaner, turn it on and let it run

for half an hour.

Page 141: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

134

The final step is plasma cleaning. Put the sample into the plasma cleaner, and

vacuum it down for 45 minutes. Then add oxygen for 5 minutes. Purge the system

for 1 minute. Add more oxygen for 30 seconds; purge for 15 seconds; add oxygen for

15 seconds and then turn off the oxygen. Immediately turn on the plasma cleaner to

HI. The plasma should be a bright pink or purple color. After 10 minutes, turn off

the plasma. Allow the samples to cool for 30 minutes. Turn off the vacuum then add

oxygen until the samples can be removed. The samples should now be very

hydrophilic, with a contact angle around zero degrees.

B.2 OTE Monolayers

The best OTE monolayers are formed with vacuum distilled OTE. The first

step is to make a prehydrolysis solution by diluting 0.21g distilled OTE and 0.125g of

1.3 M HCl to a volume of 25mL with THF. Next age the prehydrolysis solution for

approximately 44 hours. Next, make the dipping solution by dissolving 0.6g of the

prehydrolysis solution in 10g cyclohexane. This can be done in a disposable plastic

vial rinsed with acetone several times and blown dry in nitrogen, or in a clean glass

beaker. The clean silicon samples are then submerged in the dipping solution for 35

min. The dipping solutions should be overflowed with cyclohexane to remove a

partial film of excess OTE at the surface. Then the samples can be pulled vertically

and slowly from the solution. This is done to minimize the chances of any surface

layer from attaching to the surface.

After removing the silicon wafers, ultrasonicate them for three minutes each

in cyclohexane, isopropanol, and acetone to remove micelles and any other

Page 142: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

135

polymerized clusters adsorbed on the surface. Following ultrasonication, blow dry

the samples with pure nitrogen and bake under vacuum at approximately 120°C for

two hours.

Once you have made the OTE monolayers, it is important to characterize them

to see if they are of high quality. The contact angle of good monolayers is high and

steady with time. A contact angle that changes with time indicates that the OTE is

not well attached to the silicon.

B.3 Thiol Monolayers

Making the thiol monolayers is much easier than making the OTE

monolayers. The first step is to sputter a clean glass slide with a titanium or

chromium adhesion layer, 1-2 nm thick, followed by approximately 60 nm of gold.

The gold layer thickness is not critical; it will work as long as it is between 20 nm and

100 nm thick. Make a 1 mM solution of thiol in ethanol. Place the gold coated slide

in a clean vial and add enough thiol to completely submerge it. Let it sit for 30

minutes. Then rinse it five times with ethanol. Fill the vial with ethanol and

ultrasonicate the sample for five minutes. Rinse the sample with ethanol five more

times and then blow it dry. This simple procedure has almost a 100% success rate. If

you want to make mixed thiol monolayers, be sure to mix the thiols together in

solution before adding the glass slide. Otherwise the finished monolayer will have

much more of the first thiol you added than you expected.

Page 143: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

136

Appendix C: Use of the SPR

The purpose of this appendix is to explain to future graduate students how to

continue experiments with the SPR. The preparation of samples for SPR was

explained in Appendix B. There is little maintenance or calibration required for the

SPR set-up. The batteries in the photodetectors need to be replaced periodically and

everything should be kept free of dust and fingerprints.

C.1 Cleaning the Sample Cell

The sample cell components are pictured in Figure C.1. The Teflon piece and

the tubing should be cleaned in base bath. The metal pieces do not contact the sample

and therefore do not need to be cleaned. The O-ring is not lined with Teflon and

should only be cleaned with water, unless it was exposed to hydrocarbons in which

case soap may be used followed by very thorough rinsing in water.

C.2 Mounting the Sample

Mounting the sample is more of an art than a science. The flint glass slides

are brittle and often break during the mounting process. I hope that the tips outlined

below will minimize its difficulty.

The first step is to couple the gold coated slide to the prism. The high-index

prism requires we use a high-index optical coupling substance. We use Cargille

“Meltmount” with refractive index of 1.704 and a melting point of 60 oC. The first

time the prism is coupled to a sample you first must melt the coupler and spread a few

drops onto the uncoated side of the sample. Then place the preheated prism on to the

Page 144: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

137

slide. The heat of the prism should help spread the coupler between the slide and

prism. If it does not, gently heat (~50 oC) the slide until the coupler forms a thin

layer. If the prism is already coupled to a sample and you simply want to change

samples gently heat (~50 oC) the sample/prism. After a few seconds the coupler will

soften enough that the prism can be removed by gently twisting and sliding it off.

The area between the prism and slide will look golden if properly coupled. Silvery

spots indicate trapped air. To remove air bubbles, while the coupler is still warm,

press the prism down onto the slide. You can also try sliding it to remove bubbles

close to the edges. Figure C.2 pictures the prism properly mounted to the slide.

Next the prism/slide must be put into the sample holder. With the thumb-

screws fully loosened and one removed, position the slide so that it covers the o-ring

and the prism is centered. Make sure the o-ring is properly seated in its hole. It

should appear level when seen from both sides. Next, making sure the washers are

not touching the slide; gently tighten the thumb-screws until the metal bars just touch

the washers. This should be tight enough to retain water. If not, tighten the screws a

quarter turn more. Be sure to tighten the screws equally during this process, or else

the slide will break. The assembly is shown in Figure C.3.

C.3 Taking Measurements

Now the sample cell can be placed in the SPR set-up. To do this move the

translational stage on the rotation stage all the way back away from the prism holder.

Lower the sample cell into place such that the prism sits in the prism holder.

Advance the translational stage just until the sample cell is held tight. At this point

Page 145: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

138

the cell can be filled with liquids if desired. You are now ready to begin making

measurements. The final configuration is pictured in Figure C.4.

C.3.1 Static Measurements

Static measurements are made by measuring the reflected intensity versus

angle. We make these measurements by hand. Simply adjust the angle in small

increments and record the intensity. I usually use 1 degree increments until I notice

the intensity starting to change when I switch to 0.1 degree increments. The

measured curve can be compared to curves calculated on Corn’s website

http://unicorn.ps.uci.edu/calculations/fresnel/fcform.html.

C.3.2 Dynamic Measurements

Dynamic measurements are made by recording the change in reflectivity

versus time at one angle. In this case, it is important to set the angle such that it is

within the linear region of the SPR curve to get maximum contrast between the

different states. Once the angle is chosen the n-d filters should be adjusted such that

the signal from detector 1 is the same magnitude as that from detector 2. See Figure

C.5. Also make sure the photodetectors have not been saturated. A lockin is used to

divide the two signals and also to avoid problems due to impedance mismatching.

The data is then recorded using a program written by Dr. Sung Chul Bae. When

collecting data for long periods of time it is important to ensure that tension from the

coaxial cables does not cause the photodetector to shift. To prevent this pull up slack

Page 146: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

139

in the cables before locking down the detector. Also placing pedestals beside the

detector and along the cable can help.

Page 147: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

140

C.4 Figures

Figure C.1 SPR Sample Cell Components

The sample cell components are shown above. There are four screws, four

thumb screws, and 16 washers. Per screw, three washers go between the metal bar

and the Teflon sample cell and one goes between the Teflon sample cell and the

thumb screw.

Sample Cell

O-Ring

Thumb Screw

Screw

Metal Bar

Washer

Page 148: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

141

a

b

Figure C.2 Prism/Sample Assembly

The proper way for the prism to be mounted on the sample is illustrated

above. The schematic (a) illustrates the relative orientation of the components. The

photograph shows a mounting free of bubbles.

Optical Coupling Prism

Slide

Au Coating

Thiol Layer

Page 149: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

142

Figure C.3 Mounted SPR Assembly

This figure shows the proper way to assembly the SPR sample cell.

Page 150: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

143

Figure C.4 Sample Cell in Place in the SPR

This figure shows the final configuration of the sample cell in the SPR.

Page 151: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

144

Figure C.5 Schematic of SPR Apparatus

We built our SPR using the Kretschmann Geometry. Here light from the laser

becomes p-polarized after the Glan laser polarizer. The size of the beam is controlled

by an iris and then incident on the prism. It hits the gold layer with an angle above

the critical angle, producing an evanescent wave and thus exciting surface plasmons.

The signal is collected by detector 1. The signal from detector 2 is used to remove

long-time scale fluctuations from the data.

LASER

Glan Laser Polarizer

Iris

Mirror 1

Mirror 2 ND

ND

Detector 1

Detector 2

Rotation Stage

Sample Cell

Prism

Page 152: HOW WATER MEETS A HYDROPHOBIC SURFACE: …research.physics.illinois.edu/.../copies/Poynor_Torigoe_Adele.pdf · HOW WATER MEETS A HYDROPHOBIC SURFACE: RELUCTANTLY AND WITH FLUCTUATIONS

145

AUTHOR'S BIOGRAPHY

Adele Poynor Torigoe was born in Columbus, Ohio, on December 27,

1978. She received a B.S. in Physics Summa Cum Laude from the University

of Maryland Baltimore County in 2001. She continued her education at the

University of Illinois at Urbana-Champaign, earning her Master of Science in

Physics in 2003. During her time at the University of Illinois, she was a

teaching assistant for two semesters of Physics 112 and one semester of

Physics 213/214, for which she won two UIUC Department of Physics

Excellence in Teaching Awards; the Mavis Memorial Fund Scholarship; and

was twice listed on “The Incomplete list of Teachers Ranked as Excellent by

their Students”. She also received a travel stipend from the Chair’s travel

fund to present at the ACS Meeting in 2004. Following the completion of her

Ph.D., Adele will begin a tenure-track professorship at Allegheny College in

Pennsylvania.