You are on page 1of 124

Studies of Water and Solvents at

Liquid/Solid Interfaces by Sum Frequency


Generation Vibrational Spectroscopy
by

Zheng Yang

B.Sc., Peking University, Beijing, 2005

A THESIS SUBMITTED IN PARTIAL FULFILLMENT OF


THE REQUIREMENTS FOR THE DEGREE OF

DOCTOR OF PHILOSOPHY

in

THE FACULTY OF GRADUATE STUDIES

(Chemistry)

THE UNIVERSITY OF BRITISH COLUMBIA

(Vancouver)

May 2011

Zheng Yang, 2011


Abstract

This dissertation studies the surface chemistry of water and organic solvents at

liquid/mineral interfaces using IR-visible sum frequency generation (SFG) vibrational

spectroscopy. The solvents studied include pentane, heptane, tetradecane and toluene, and

the minerals include silica and mica. These liquid/mineral interfaces are relevant for

environmental and industrial processes, such as ice nucleation and oilsands extraction.

Structures of water at water/silica interfaces were studied in the presence of alkali

chloride in solution. Perturbations of the interfacial water structures were observed with

NaCl concentrations as low as 1x10-4 M. Different alkali cations produce different

magnitudes of perturbation, with K+ > Li+ > Na+. This order was explained by the

different effective ionic radii and electrostatic interactions between the cations and silica

surfaces.

The adsorption of water at solvent/silica interfaces was studied at room

temperature. A water layer without detectable free OHs was discovered at toluene/silica

interface. This water layer showed resistance against further adsorption of water

molecules and was very stable at room temperature. However, similar structure of water

was not observed at heptane/silica interfaces.

Water structures on mica were studied with atmospherically relevant sulphuric

acid concentrations. Experimental data showed that ordered water structures on mica

completely disappeared when the concentration of sulfuric acid reached 5 mol/L. The

results partially explain why sulfuric acid coatings influence the ice nucleation properties

of mineral dust particles in the atmosphere.

ii
The competitive adsorption of toluene and n-alkane at solvent/silica interfaces

was studied. The surface coverage of toluene for toluene-pentane, toluene-heptane, and

toluene-tetradecane mixtures were measured over the complete mole fraction range from

0 to 1. Overall, toluene competes favorably on silica, but the molar adsorption free

energy of alkanes increases as the chain length increases.

Finally, experiments were conducted to study the effect of interfacial water on

bitumen liberation from mineral surfaces in water. The bitumen liberation rate increases

when the water content between bitumen and the mineral increases. The liberation also

highly depends on the surface properties of minerals. At the same water content, the rate

of bitumen displacement on different mineral surfaces is: freshly cleaved mica > rinsed

mica > silica.

iii
Preface

A version of Chapter 3 has been published as: Yang, Z.; Li, Q. F.; Chou, K. C.

Structures of Water Molecules at the Interfaces of Aqueous Salt Solutions and Silica:

Cation Effects. Journal of Physical Chemistry C 2009, 113, 8201-8205. The project was

initiated by Dr. Chou. Dr. Li gave valuable suggestions for optimizing experimental setup

and data analysis. My major contributions are: literature survey, design of the research

project, data acquisition, interpretation of the observations, and preparing the manuscripts.

A version of Chapter 4 has been published as: Yang, Z.; Li, Q. F.; Gray, M. R.;

Chou, K. C. Structures of Water Molecules at Solvent/Silica Interfaces. Langmuir 2010,

26, 16397-16400. The project was initiated by Dr. Chou and Dr. Gray. Dr. Li helped in

optimizing the experimental setup. My major contributions are: literature survey, design

of the research project, data acquisition, explanation of the observations, preparing the

manuscripts.

A version of Chapter 5 will be submitted for publication: Yang, Z.; Bertram, A.

K.; Chou, K. C. Why do Sulfuric Acid Coatings Influence the Ice Nucleation Properties

of Mineral Dust Particles in the Atmosphere? The project was initiated by Dr. Bertram

and Dr. Chou. Also Dr. Bertram and Dr. Chou contributed to the design of the project,

and editing of the manuscript. My major contributions are: literature survey, design of the

research project, data acquisition, explanation of the observations, preparing the

manuscripts.

A version of Chapter 6 has been published as: Yang, Z.; Li, Q. F.; Hua, R.; Gray,

M. R.; Chou, K. C. Competitive Adsorption of Toluene and n-Alkanes at Binary

iv
Solution/Silica Interfaces. Journal of Physical Chemistry C 2009, 113, 20355-20359. The

project was initiated by Dr. Chou and Dr. Gray. Dr. Chou developed the computer

program used to fit the SFG spectra. Dr. Li calculated the orientation distribution of the

CH3 groups. Dr. Hua obtained the SFG vibrational spectra of toluene in ppp polarization

configuration. My major contributions are: literature survey, design of the research

project, data acquisition.

A version of Chapter 7 will be submitted for publication. Yang, Z.; Bailey, G.;

Gray, M. R.; Chou, K. C. Effect of Interfacial Water Content on Bitumen Liberation from

Silica and Mica Surfaces. The project was initiated by Dr. Chou and Dr. Gray. Gwen

Bailey developed the computer program used to analyze the contact angle of bitumen

droplet in water. My major contributions are: literature survey, design of the research

project, data acquisition, explanation of the observations, preparing the manuscripts.

v
Table of Contents

Abstract..............................................................................................................................ii

Preface................................................................................................................................iv

Table of Contents..............................................................................................................vi

List of Figures....................................................................................................................ix

Abbreviations..................................................................................................................xiv

Acknowledgments............................................................................................................xv

Dedication...........................................................................................................xvii

Chapter 1 Introduction.....................................................................................................1

1.1 Overview..................................................................................................................2

1.2 Water Structures on Water/Mineral Interfaces........................................................4

1.3 Solvent Adsorption on Minerals..............................................................................5

1.4 Waters Role in Oil Sands Processing.....................................................................7

Chapter 2 IR-visible Sum Frequency Generation Spectroscopy...................................9

2.1 Introduction............................................................................................................10

2.2 Theory of Sum Frequency Generation...................................................................11

2.3 Fresnel Factors.......................................................................................................15

2.4 Molecular Orientation at the Interface...................................................................17

2.5 Experimental Instrumentation................................................................................19

vi
Chapter 3 Structures of Water Molecules at the Interfaces of Aqueous Salt Solutions

and Silica: Cation Effects................................................................................................21

3.1 Introduction............................................................................................................22

3.2 Experimental Section.............................................................................................25

3.3 Results and Analysis..............................................................................................28

3.4 Conclusions............................................................................................................36

Chapter 4 Structures of Water Molecules at Solvent/Silica Interfaces......................37

4.1 Introduction............................................................................................................38

4.2 Experimental Section.............................................................................................40

4.3 Results and Analysis..............................................................................................41

4.4 Conclusions............................................................................................................49

Chapter 5 Why do Sulfuric Acid Coatings Influence the Ice Nucleation Properties of

Mineral Dust Particles in the Atmosphere....................................................................50

5.1 Introduction............................................................................................................51

5.2 Experimental Section.............................................................................................54

5.3 Results and Analysis..............................................................................................56

5.4 Conclusions............................................................................................................64

Chapter 6 Competitive Adsorption of Toluene and n-Alkanes at Binary

Solution/Silica Interfaces.................................................................................................65

6.1 Introduction............................................................................................................66

vii
6.2 Experimental Section.............................................................................................68

6.3 Results and Analysis..............................................................................................70

6.4 Conclusions............................................................................................................83

Chapter 7 Effect of Interfacial Water Content on Bitumen Liberation from Silica

and Mica Surfaces............................................................................................................84

7.1 Introduction............................................................................................................85

7.2 Experimental Section.............................................................................................86

7.2.1. Sum frequency generation spectroscopy measurements..........................................86

7.2.2. Dynamic and equilibrium contact angle measurements..........................................87

7.3 Results and Analysis..............................................................................................89

7.3.1. Sum frequency generation spectroscopy experiments.............................................89

7.3.2. Dynamic and equilibrium contact angle study.........................................................90

7.4 Conclusions............................................................................................................96

Chapter 8 Conclusion......................................................................................................97

References.......................................................................................................................102

viii
List of Figures

Figure 2.1 SFG setup in the reflection geometry. ............................................................11

Figure 2.2 The definition of orientation angle for methyl group. Orientation angle ()

between surface normal and methyl group, axis c is the principal axis of the methyl group.

The ab plane is perpendicular to axis c. The ac plane contains one C-H bond. ...............18

Figure 2.3 Layout of OPG/OPA. DM, RM and BS represent the dichromatic mirror,

reflective mirror, and beam splitter, respectively. ............................................................20

Figure 3.1 Schematic layout of the SFG vibrational spectroscopy. The frequency of the

visible beam was fixed at 532 nm and the frequency of the IR beam was scanned from

2800 3900 cm-1. The 532 nm and IR beams were overlapped both spatially and

temporally on the bottom surface of the solution. ............................................................26

Figure 3.2 SFG spectra from the water/silica interfaces with different concentrations of

three alkali chloride solutions: (a) NaCl, (b) LiCl, and (c) KCl. The solid lines are fitting

curves derived using two Lorentzian peaks at 3200 cm-1 and 3400 cm-1. ........................30

Figure 3.3 Amplitudes (Aq) as functions of the concentration of alkali chloride solutions

for water/silica interfaces obtained by fitting the spectra in Figure 3.2 with Equation

(3.2). ..................................................................................................................................32

Figure 4.1 SFG spectra of heptane/silica interfaces with dry (open circles) and water-

saturated (solid circles) heptane. The peaks between 2800 and 3000 cm-1 are associated

with C-H vibrations of heptane, the peaks at 3200 and 3400 cm-1 are vibrational peaks

associated with hydrogen-bonded water, and the 3660/3680 cm-1 peak is assigned to the

free O-H stretching mode. .................................................................................................42

ix
Figure 4.2 SFG spectra of toluene/silica interfaces with dry (open circles) and water-

saturated (solid circles) toluene. The peaks between 2800 and 3100 cm-1 are associated

with C-H vibrations of toluene, the broad peak centered at 3150 cm-1 is associated with

hydrogen-bonded water, and the 3600 cm-1 peak is assigned to the free O-H stretching

mode. .................................................................................................................................44

Figure 4.3 Adsorption of water on a regular water layer with free OHs. The open circles

represent the SFG spectrum of a toluene/silica interface with dry toluene on an initially

moistened silica surface. In this regular water layer, the 3200 cm-1, 3400 cm-1, and free

OH peaks are all present. Water can condense on such a regular water surface, as

indicated by the solid-circle spectrum, which was taken 30 min. after the toluene was

saturated with water. .........................................................................................................45

Figure 4.4 Schematic drawing of the water-resistant water structure. In the process of

water adsorption, the water molecules minimize their energy by forming the maximal

number of hydrogen bonds with either the neighbouring water molecules or the silanol

groups (SiOH) on silica. At a certain stage, very few free OHs are available. The absence

of free OH groups significantly reduces further water adsorption. ..................................47

Figure 5.1 Panel (a) schematic showing the range of RH values over which uncoated and

coated mineral dust particles are good ice nuclei. The figure is specifically created for

illite particles based on the work of Chernoff et al. A similar behavior was also observed

for kaolinite. The temperature applicable for the ice nucleation studies was approximately

237 K. Panel (b) and (c) show the pH and the molarity of the sulfuric acid coating,

assuming the coating is in equilibrium with the relative humidity. ..................................52

Figure 5.2 SFG spectra of D2O/mica interfaces with D2SO4 concentrations of (a) 0 M, (b)

x
5x10-6 M, (c) 5x10-5 M, (d) 5x10-4 M, (e) 5x10-3 M, (f) 5x10-2 M, (g) 5x10-1 M and (h) 5

M. The inset shows the schematic layout of the spectroscopic setup. The polarizations of

the beams were s-, s-, and p-polarized for SFG, visible, and IR, respectively. ................57

Figure 5.3 The fitted amplitudes (a) and frequencies (b) of water peaks in the SFG

spectra with various D2SO4 concentrations. Triangles represent ice-like peaks and dots

represent liquid-like peaks. ............................................................................................59

Figure 5.4 Schematics of water molecules and hydrated sulfate (bisulfate) ions at

solution/mica interfaces with different concentrations of H2SO4 solutions. (a) Pure

H2O/mica interface with pH~6. The surface is highly negatively charged. Water

molecules are more ordered near the charged surface. (b)With a low concentration of

H2SO4, the surface charge decreases as the surface is protonated, and the anions interact

with the mineral surface. (c) With a high concentration of H2SO4 (for example 5M),

water molecules are well captured by the sulfate/anions in the solution, and few water

molecules are freely available for the mica surface. In (a) and (b) the dashed lines

separate the ordered water molecules from the water molecules without order. Above the

dashed lines, water molecules have good order because of negative surface potential

induced by the mica surface. In (b) and (c) H+ adsorb on the mica surfaces. The dashed

circles represent hydrated sulfate ions (in low H2SO4 concentration) or hydrated bisulfate

anion (in high H2SO4 concentration). Water molecules in the dashed circle are parts of

the hydrated anions and move together with the core anions. ..........................................63

Figure 6.1 Schematic layout of the spectroscopic setup. The frequency of the visible

beam was fixed at 532 nm, and the frequency of the IR beam was tunable. The 532 nm

and IR beams were overlapped both spatially and temporally on the top surface of the

xi
solution. The thickness of the solvent layer is 3 mm. .......................................................68

Figure 6.2 (A) SFG vibrational spectra of toluene in ssp and ppp polarization

configurations. The ssp and ppp spectra are offset from each other by 0.5 arbitrary units

( 2)
ssp
for clarity. (B) Calculated for CH3 symmetric (solid line) and asymmetric (dashed
(ppp
2)

line) modes as a function of the tilting angle. The orientational distribution of the CH3

( 2)
ssp
groups was assumed to be a delta function. The solid circle indicates the measured
(ppp
2)

value of 4.4 which corresponds to a tilting angle of 25. .................................................71

Figure 6.3 SFG vibrational spectra of toluene-pentane mixtures on silica with toluene

volume fraction toluene


B
= (a) 1, (b) 0.8, (c) 0.6, (d) 0.4, (e) 0.2, and (f) 0. ........................73

Figure 6.4 SFG vibrational spectra of toluene-heptane mixtures on silica with toluene

volume fraction toluene


B
= (a) 1, (b) 0.8, (c) 0.6, (d) 0.4, (e) 0.2, and (f) 0. ........................74

Figure 6.5 SFG vibrational spectra of toluene-tetradecane mixtures on silica with toluene

volume fraction toluene


B
= (a) 1, (b) 0.8, (c) 0.6, (d) 0.4, (e) 0.2, and (f) 0. ......................75

Figure 6.6 Adsorption isotherms of toluene on silica for binary mixtures of pentane

toluene (), heptanetoluene (), and tetradecanetoluene (). The solid curves are

fitting curves using Equation (6.8). ...................................................................................78

Figure 7.1 Schematic layout of the spectroscopic setup. The frequency of the visible

beam was fixed at 532 nm, and the frequency of the IR beam was tunable. The 532 nm

and IR beams were overlapped both spatially and temporally at the bottom surface of

silica plate. ........................................................................................................................87

xii
Figure 7.2 SFG vibrational spectra of water adsorbed on silica surface as a function of

relative humidity (RH). .....................................................................................................90

Figure 7.3 Sequential images extracted from a video of bitumen liberation from silica at

time equal to (a) 0, (b) 20, (c) 80 sec. ...............................................................................91

Figure 7.4 Measured dynamic contact angles for bitumen on (a) silica, (b) rinsed mica, and

(c) cleaved mica. ...............................................................................................................92

xiii
Abbreviations

AFM Atomic force microscopy

BS Beam splitter

CHWP Clark hot water process

DM Dichromatic mirror

KTP KTiOPO4

OPG/OPA Optical parametric generator/amplifier

PMT Photomultiplier tube

PZC Point of zero charge

RH Relative humidity

RM Reflective mirror

SFG Sum frequency generation

SHG Second harmonic generation

STM Scanning tunneling microscopy

YAG Yttrium aluminum garnet

xiv
Acknowledgments

Foremost I would like to express my deep and cordial gratitude to my advisor,

Professor Keng Chang Chou for his continuous support and encouragement throughout

my PhD study in UBC. Professor Chou has always been an inspiring, kind and helpful

mentor. His profound knowledge in laser and spectroscopy has been of great value to me

and inspires me to be a better researcher. I enjoy every discussion with him, for advice

from him always helps me overcome difficulties not only in research but also in daily life.

His great insights in science, logical way in thinking and communication, optimism in

life, kindness and patience as an advisor and courteous attitude to people enlighten me on

characteristics that one should have to be successful.

I hold sincere gratitude to Dr. Qi Feng Li, who has been in our group as a post-

doctoral fellow since I just joined the research group. He offered me a lot of help in both

science and life. His knowledge and skills on optics help me immensely during the

experiments. I will never forget the accomplishments and failures we experienced

together.

I am also truly grateful to Rui Hua, who joined our group one year earlier than me

as a graduate student. I really enjoyed the interesting discussions with him. And I will

never forget his encouragement and support.

I would like to thank former and current group members, Sherry Wu, Gwen

Bailey, Bonnie Leung, Henry Tang, Zhen Wei Wang and Zhe Wang, for sharing their

knowledge, for giving support and help.

xv
In the end, I would to thank Jingyi and my parents who experienced both my

excitements and frustrations, for their warm support during the past years. I pray deeply

for a healthy and enjoyable life for them.

xvi
Dedication

Dedicate this thesis to Jingyi and my parents.

xvii
Chapter 1

Introduction

1
1.1 Overview

Surface phenomena have attracted the interest of numerous researchers in a wide

range of disciplines. Much effort has been devoted to studying the properties of interfaces,

from simple systems such as the adsorption of gases at inert surfaces, to more

complicated systems such as liquid/liquid or solid/liquid interfaces. Solid/liquid

interfaces, in particular, are of great importance in surface science. They are not only

ubiquitous in both the environment and industry, but also highly relevant to numerous

important processes, including cleaning, wetting, adhesion, lubrication, etching, corrosion,

electrochemical reactions, and oil recovery.1 It is expected that the properties of liquid

molecules on a solid surface are very different from those in the bulk.2,3 This is because

the structure of the interfacial liquid can be more readily disturbed by the

attractive/repulsive interactions between a solid surface and liquid molecules.

Detailed investigations of solid/liquid interfaces have been hindered by a lack of

effective experimental probes, which are needed to provide molecular-level information

on the chemical composition and geometric structure of the functional species at buried

liquid interfaces. High surface specificity is necessary to separate the signal of interfacial

molecules from a huge amount of molecules in the bulk. Optical techniques such as

attenuated total reflection spectroscopy, infrared spectroscopy, and ellipsometry have

been used to probe buried liquid interfaces, but they are not intrinsically very surface-

specific and cannot distinguish signals from the bulk. Scanning tunneling microscopy

(STM) and atomic force microscopy (AFM) may be applied to thin films of liquids on

solid substrates but have difficulty in producing clear microscopic images of liquid

surfaces because of molecular movement in most cases.4-6 Recently, nonlinear optical

2
spectroscopic techniques such as sum-frequency generation (SFG) have become the

preferred techniques to study liquid interfaces.7-10 Being highly surface-specific and

applicable to all interfaces accessible by light, SFG has been repeatedly demonstrated to

be the most versatile and powerful analytical tool for investigating liquid interfaces.

SFG, as a second-order nonlinear optical process, is forbidden under the electric-

dipole approximation in the bulk media with inversion symmetry but allowed at surfaces

or interfaces where the inversion symmetry is broken.11 Therefore, this technique is

highly interface-specific with submonolayer sensitivity. Currently, SFG is the only

technique that can provide vibrational spectra of buried liquid interfaces. In this

dissertation SFG is used to study the structures and properties of water and hydrocarbons

at various liquid/mineral interfaces.

3
1.2 Water Structures on Water/Mineral Interfaces

The water/mineral interface is of great importance in surface science since it is

ubiquitous in both the environment and industry and arises in many processes such as

contaminant migration, ice nucleation, soil formation, microorganism growth, lubrication,

and oil recovery.1,12-19 Generally, when water molecules encounter a solid surface, they

will try to reorient to form as many hydrogen bonds as possible to minimize their energy,

which leads to an ordered structure adjacent to the surface. For example, water structures

at a hydrophilic solid surface, such as silica, are understood to be a mixture of a more

ordered and a less ordered hydrogen-bond network, which are associated with vibrational

peaks at 3200 and 3400 cm-1, respectively.20 The 3200 cm-1 peak is due to water

molecules with a full complement of hydrogen bonds in a tetrahedral arrangement and

has often been referred to as ice-like since it exhibits a vibrational peak position similar

to the solid phase of water, ice. The peak at 3400 cm-1 is broad and assigned to water

molecules with either asymmetric hydrogen bonds or bifurcated hydrogen bonds and is

called water-like because the resonance is found at approximately the same frequency

as the infrared absorption of water molecules in the bulk liquid.10,20

The detailed interfacial water structures depend on the property of the surface and

the aqueous solution, such as the hydrophilicity/hydrophobicity of the surface, surface

charge, the species and concentration of ions in solution and so on.10,21 Chapters 3, 4 and

5 of this dissertation present studies of liquid/mineral interfaces relevant to petrolic,

edaphic and atmospheric processes, which include the water structures at aqueous salt

solutions/silica interfaces, solvent/silica interfaces, and high concentration sulphuric acid

solution/mica interfaces.

4
1.3 Solvent Adsorption on Minerals

In contrast to the vast amount of studies focusing on the structures of water at the

interfaces,10,20 there are fewer works studying solvents at solid interfaces using SFG.8,22,23

An important conclusion from previous works is that the polarity of the interfacial region

plays a key role in controlling the structure and orientation of adsorbed organic

molecules.24-27 If there is more than one species of organic solvent existing in the liquid

phase, they will compete for adsorption at the solid interface.

The competitive adsorption of organic solvents (usually hydrocarbons) at

liquid/solid interfaces is important for many industrial and scientific processes, such as

oilsands processing, petroleum recovery, contamination removal, heterogeneously

catalyzed reactions, thin film materials production and many extraction techniques.28-38

Generally the competitive adsorption depends on the molecular properties of solvents,

their respective mole fractions, and the strength of attractions to the solid surface, so at a

binary solution/solid interface, it is expected that the surface chemical composition is

different from the bulk composition because of the different interaction strength with the

surface.39 In many cases, the competitive adsorption of solvents at liquid/solid interfaces

is a critical factor in determining the effectiveness of a technological process. In the

1950s and 1960s, the adsorption isotherms for binary mixtures at liquid/solid interfaces

were studied by various immersion methods, and a number of theories were developed to

elucidate the thermodynamic properties of pure liquids and binary mixtures over solid

surfaces.40,41 Despite this effort, the problem was not completely resolved because the

macroscopic measurements were indirect, and the theories required molecular-level

information about adsorbates as input parameters.41,42 Even with modern technologies, it

5
remains challenging to selectively probe the interfacial region between a liquid and a

solid and to directly measure the surface coverage of a particular component at a

liquid/solid interface. SFG provides a new opportunity to directly measure the adsorption

isotherms at solvent/mineral interfaces.

6
1.4 Waters Role in Oil Sands Processing

Oil sands (also known as bituminous sands or tar sands) are found in large

amounts in many countries throughout the world. It is estimated that there are at least

170 billion barrels (27109 m3) of bitumen that can be recovered from the Athabasca oil

sand deposits in Canada.43 The oil sands contain naturally occurring mixtures of sand,

clay, water, and a dense and extremely viscous form of petroleum referred to as bitumen.

At present, the technology used to extract bitumen from the oil sands deposits is a

variation on the Clark hot water process (CHWP) where a lower processing temperature

(35-55 C) is used.44 The oil sands conditioning stage and bitumen recovery step include:

(1) bitumen displacement along a sand grain, (2) bitumen detachment, (3) bitumen

droplet attachment to an air bubble and (4) bitumen flotation in separation vessels. A

prerequisite to the bitumen extraction process is liberation of bitumen from the mineral

surface, which requires a huge amount of water. Therefore, understanding the

mechanisms behind the displacement of oil by water on minerals is important for bitumen

recovery. Many investigations have been done using dynamic contact angle

measurements to study bitumen displacement and detachment from solid surfaces in the

presence of water containing salt, surfactants and clays at different temperatures and pH

values.45-48 Previous studies have also suggested that the interfacial water content

between mineral surfaces and bitumen is one of the most important factors affecting

bitumen recovery.49 However, the detailed mechanism remains poorly understood. To

study the mechanism, both macroscopic (liberation rate, dynamic and static contact

angle) and microscopic (water amount at mineral surfaces under different relative

humidity, structures of water molecules adsorbed on mineral surfaces) information is

7
needed. The study in chapter 7 provides a molecular-level understanding for the effect of

interfacial water on the liberation process.

8
Chapter 2

IR-visible Sum Frequency Generation

9
2.1 Introduction

Since the first experimental demonstration of surface vibrational spectroscopy via

infrared-visible sum-frequency generation (SFG) published in 1987,50 SFG spectroscopy

has become a versatile spectroscopic technique to study the properties of various surfaces

and interfaces. SFG is extremely surface-specific and sensitive, because as a second-

order nonlinear optical process, SFG is forbidden in a medium with inversion symmetry

under the electric-dipole approximation, but is allowed at a surface or interface where the

inversion symmetry is broken.

In the IR-visible SFG process, two pulsed laser beams (one visible with

frequency Vis and one infrared with frequency IR ) are overlapped spatially and

temporally at an inferface, and an SFG with a frequency of SF = vis + IR is generated.

The intensity of SFG is enhanced when the frequency of the IR beam is resonant with the

vibrational modes of molecules at the interface. Detailed theoretical descriptions of SFG

process are available in the literature.51-53 In this chapter, a brief introduction of SFG

technique is presented.

10
2.2 Theory of Sum Frequency Generation

A typical SFG setup in a reflection geometry is shown in Figure 2.1. The visible

and IR beams, with frequencies vis and IR respectively, are overlapped in time and

space at an interface. The SFG beam generated at the interface has a frequency

of SF = vis + IR , dictated by the conservation of energy between the incoming and

outgoing photons.

Visible SFG

IR
Medium 1

Medium 2

Figure 2.1. SFG setup in the reflection geometry.

Conservation of momentum (phase-matching condition) is also required for the

photons involved in the SFG process:

k SF = k vis + k IR 2.1

where k vis and k IR are the wave vectors of the incident beams.

By combining the conservations of energy and momentum, we can determine the

direction of the SFG beam by the relation:

n SF k SF sin SF = nvis k vis sin vis + n IR k IR sin IR 2.2

11
where i is the angle of the indicated light to the surface normal and k i is the absolute

value of the wave vector of each light.

As a nonlinear optical process, the second-order nonlinear polarization P ( 2 ) (SF ) ,

which generates the SFG output, and the two incident electric fields E (vis ) and

E ( IR ) are related as follows:

P ( 2 ) (SF ) = 0 ( 2 ) : E (vis ) E ( IR ) 2.3

where ( 2 ) is known as the second-order susceptibility and 0 is the electric permittivity

of free space. The SFG intensity is proportional to the square of P ( 2 ) (SF ) , therefore the

measurable SFG intensity I SF is proportional to the intensities of the two incident laser

beams and the square of the effective surface second-order susceptibility eff
(2)
.

The sum frequency intensity in the reflected direction is given by:54

SF
2
2
I SF = eff( 2 ) I vis I IR 2.4
8 0 c cos SF
3 2

with

eff( 2 ) = [L(SF ) e SF ] : ( 2 ) : [L(vis ) evis ][L(IR ) e IR ] 2.5

Here, i is the frequency of the visible, IR and SFG beams, ei is the unit polarization

vector of the optical field at i , and L(i ) is the tensorial Fresnel factor at frequency i .

The macroscopic second-order susceptibility ( 2 ) is a third rank tensor which

consists of 27 elements ijk( 2 ) . The values of the components are the property of the

medium and are invariant under symmetry operations. The number of non-zero ( 2 )

12
elements is often reduced because of symmetry constraints. For example, in a bulk

solution with a centrosymmetric environment, ( 2 ) is invariant under inversion symmetry,

but the electric field and polarization must change signs as they are vectors. Based on

equation (2.3), the inversion operation gives: ( 2 ) = ( 2 ) . So ( 2 ) must equal to 0 in a

centrosymmetric medium. This is the reason why SFG is forbidden in a medium with

inversion symmetry. However, the inversion symmetry is naturally broken at the surface.

There are only four nonequivalent and non-vanishing ( 2 ) values for an isotropic

surface. With the lab coordinates chosen such that z is along the interface normal and x is

in the incident plane, the non-vanishing ( 2 ) are: xxz


( 2)
= yyz
( 2)
, xzx
( 2)
= yzy
( 2)
, zxx
( 2)
= zyy
( 2)
,

and zzz
( 2 ) 52
. These four components can be deduced by measuring SFG with four different

input and output polarization combinations: ssp, (the sum frequency, visible, and infrared

beams are s-polarized, s-polarized, and p-polarized, respectively), sps, pss, and ppp.

ijk( 2 ) is the macroscopic orientational average of the microscopic molecular

hyperpolarizability lmn
(2)
of all the interfacial molecules:

ijk( 2 ) = N s lmn
(2)
(i l )( j m )( k n ) 2.6
l ,m ,n

N s denotes the number density of the interfacial molecules. The subscripts i, j, and k

refer to the lab coordinates, and the subscripts l, m, and n, refer to the axes for the

molecular coordinate system. (i l)( j m )( k n ) is the coordination transformation from

molecular fixed coordinates to laboratory fixed coordinates, and indicates an average

over the molecular orientation distribution.

(2) (2)
Both ijk and lmn consist of a sum of resonant terms and a nonresonant term:

13
A ,ijk
ijk( 2 ) = NR ,ijk +
(2)
2.7
IR i

,lmn
lmn = NR ,lmn +
(2) (2)
2.8
IR i

In the above equations A ,ijk , and are the amplitude of the transition moment,

frequency of the transition and line width of the vibrational mode respectively. The

transition moment amplitude ,lmn is related to the Raman and IR properties of the

vibrational mode, ,lmn is nonzero only when both the Raman polarizability tensor and

the IR transition moment are nonzero. Therefore, a SFG active vibrational mode must be

both Raman and IR active. When the frequency of a vibrational is resonant with the

IR frequency IR , the denominator in Equation (2.7) becomes small and ( 2,ijk) will be

enhanced.

14
2.3 Fresnel Factors

For isotropic surfaces, SFG experiments can be conducted in different input and

output polarization combinations such as ssp, sps, pss, and ppp. Equation (2.5) can be

rewritten for an isotropic interface:

eff( 2 ),ssp = Lyy (SF ) Lyy (vis ) Lzz (IR ) sin IR yyz

eff( 2 ),sps = Lyy (SF )Lzz (vis )Lyy (IR ) sin vis yzy

eff( 2 ), pss = Lzz (SF ) L yy (vis ) L yy ( IR ) sin SF zyy 2.9

eff( 2 ), ppp = Lxx (SF ) Lxx (vis ) Lzz ( IR ) cos SF cos vis sin IR xxz
Lxx (SF ) Lzz (vis ) Lxx ( IR ) cos SF sin vis cos IR xzx
+ Lzz (SF ) Lxx (vis ) Lxx ( IR ) sin SF cos vis sin IR zxx
+ Lzz (SF ) Lzz (vis ) Lzz ( IR ) sin SF sin vis sin IR zzz

The Fresnel factors depend on the refractive indices of the media and the
t
experimental geometry. For isotropic surfaces, only the diagonal elements of L ( i ) need

to be considered:

2n1 (i ) cos i
Lxx (i ) =
n1 (i ) cos i + n2 (i ) cos i

2n1 (i ) cos i
L yy (i ) = (2.10)
n1 (i ) cos i + n2 (i ) cos i

2
2n2 (i ) cos i n1 (i )
Lzz (i ) =
n1 (i ) cos i + n2 (i ) cos i n ' ( )
i

15
In the above equations, ni is the refractive index of medium i as shown in Figure 2.1, i

is the incident angle, and i is the refracted angle. n' ( i ) is the refractive index of the

interfacial layer. Since the interfacial layer is only one or a few monolayers thick, n' ( i )

can be difficult to measure.55 It is therefore an usual practice that n' ( i ) is chosen to be

the bulk refractive index of the material at the interface.

16
2.4 Molecular Orientation at the Interface

As discussed above, SFG measures the macroscopic second-order susceptibility

( 2 ) which is related to the microscopic molecular hyperpolarizability ( 2 ) through

average molecular orientation as shown in Equation (2.6). ( 2 ) is an intrinsic property of

a molecule and the components of ( 2 ) can be determined by the symmetry of the

molecule. If ( 2 ) is known, then the average orientation of the functional group can be

deduced from the measurement of ( 2 ) using Equation (2.6).

For example, the methyl group has C3v symmetry and the symmetric stretch mode

is cylindrically symmetric with symmetry axis along c , so there are only two
nonvanishing independent components in ( 2 ) : aac = bbc and ccc . Therefore, from

Equation (2.6) we find for an azimuthally isotropic surface:56

xxz ,s = yyz ,s = N s ccc [cos (1 + r ) cos3 (1 r )]


1
2

xzx ,s = yzy ,s = zxx ,s = zyy ,s = N s ccc [cos cos3 ](1 r )


1
2.12
2

zzz ,s = N s ccc [r cos + cos 3 (1 r )]

Here aac = r ccc and r is the hyperpolarizability ratio which is equal to 2.3 for the methyl

group.52,57 is the tilt angle which is the angle of the molecular symmetry axis c with

respect to z in the laboratory coordinate system as shown in Figure 2.2:

17
Figure 2.2. The definition of orientation angle for methyl group. Orientation angle () is

between surface normal and methyl group, axis c is the principal axis of the methyl group.

The ab plane is perpendicular to axis c. The ac plane contains one C-H bond.

The absolute value of N s is difficult to determine, so we can cancel it by

measuring the ratios of independent nonvanishing ( 2 ) components. Then we can find the

orientation by assuming a -function distribution for .

18
2.5 Experimental Instrumentation

As shown in Figure 2.3, the SFG experiments were carried out with a visible 532

nm beam and an IR beam tunable from 1800 to 4000 cm-1, which were generated using a

Nd:YAG (yttrium aluminum garnet) laser system (1064 nm, 10 Hz, and 25 ps). The 532

nm beam was produced in a KTiOPO4 (KTP) crystal by second harmonics generation

(SHG) from the fundamental 1064 nm beam. A portion of the 532 nm beam was used to

pump a home-made KTP optical parametric generator/amplifier (OPG/OPA) and a

KTiOAsO4 (KTA) difference frequency generation (DFG) system which mixes the idler

output of the OPG/OPA system and the fundamental 1064 nm beam. Both the visible and

IR beams had a pulse duration of 25 ps and a repetition rate of 10 Hz. The input angles

were set to 45 and 56 with respect to the surface normal for the visible and the IR

beams, respectively. The input energy was ~250J/pulse for the visible beam with a spot

size of 2mm on the sample. For the IR beam the input energy was ~200J/pulse with a

spot size of 0.5mm on the sample. The IR and visible input beams were overlapped both

spatially and temporally at the sample. The SFG output passed through a series of

bandpass filters to eliminate the background light, and then was detected by a

HAMAMATSU R3896 photomultiplier tube (PMT). The signal intensity was recorded

by a gated integrator (SR 280, Stanford Research Systems Inc.) and digitized by a

computer.

The frequency of IR beam was calibrated by the absorption lines of polystyrene

film (Thermo Electron Corp.). All SFG spectra presented in this dissertation were

normalized against the SFG spectrum of a z-cut quartz.

19
RM
532 nm
RM RM

SHG OPG/OPA
DM
Nd: YAG laser KTP KTP KTP RM
25 ps, 10 Hz, 40mJ/pulse RM BS BS DM
532 nm
1300~1800 nm

1064 nm RM KTA
RM Tunable IR
DFG 2000~4000 cm-1

Figure 2.3. Layout of OPG/OPA. DM, RM and BS represent the dichromatic mirror,
reflective mirror, and beam splitter, respectively.

20
Chapter 3

Structures of Water Molecules at the Interfaces

of Aqueous Salt Solutions and Silica: Cation

Effects*

* A version of this chapter has been published. Yang, Z.; Li, Q. F.; Chou, K. C. Structures of Water
Molecules at the Interfaces of Aqueous Salt Solutions and Silica: Cation Effects. Journal of Physical
Chemistry C 2009, 113, 8201-8205.

21
3.1 Introduction

Buried aqueous interfaces play an important role in many natural and industrial

processes. Among the various aqueous interfaces, water/silica interfaces, which affect

contaminant migration, ice nucleation, soil formation, and microorganism growth, have

been of great interest.12-19 Since it is experimentally challenging to probe a buried

interface, current understanding of buried aqueous interfaces remains limited. For this

reason, a large number of theoretical simulations of water interfaces have been carried

out to provide microscopic information, such as the density profiles and orientations of

water molecules at interfaces.4,58-66 Although the calculations provide qualitative

information, the simulation results are often limited by the capacity of computers, which

are insufficient for determining the detailed structure of interfacial water under practical

conditions involving large numbers of molecules. Previously, the structure of water

molecules at buried aqueous interfaces were studied using X-ray spectroscopy,67-69

electron diffraction,70 second harmonic generation (SHG),71 and sum-frequency

generation (SFG) vibrational spectroscopy11,13,72. Generally, water structures at a

hydrophilic solid surface, such as silica, are understood to be a mixture of a more ordered

and a less ordered hydrogen-bond network, which are associated with vibrational peaks at

3200 and 3400 cm-1, respectively.20 In any case, the detailed structure of water molecules

on silica is not yet completely understood.

Experimentally, the structure of water on silica has been shown to depend on the

surface charges. Silica surfaces possess negative charges because of the deprotonation of

surface silanol groups (SiOH). The surface charges, which create a surface electric field,

induce polar ordering of interfacial water molecules. Ong et al. studied water/silica

22
interfaces with various pH values using second harmonic generation (SHG).71 They

concluded that water molecules near the silica interface were polarized by the interfacial

electric field and responsible for the observed SHG. A few years later, Du et al. studied

OH vibrations of water molecules at water/fused quartz interfaces using IR-visible SFG

vibrational spectroscopy.72 Their results showed that the orientation of the OH bonds and

the ordering of interfacial water molecules are strongly affected by their electrostatic

interaction with the deprotonated surface silanol groups. In a recent study on water/quartz

interfaces, Shen and his coworkers indicated that two different surface sites exist with

different deprotonation pK values on crystalline quartz.73 The peak at 3200 cm-1 seemed

to be associated with surface sites that have higher pK values, and the peak at 3400 cm-1

was closely associated with surface sites having lower pK values. The detailed

mechanism for the formation of these two different sites is not yet understood.

Little is known about the structure of interfacial water molecules under

perturbations by cations. Alkali cations are particularly important, as they are the most

abundant cations in natural water. Previous studies at air/water interfaces with NaCl

solutions showed that a reduced ion density was present near the surface, and the ions

produced little effect on the surface water structure, for molar fractions up to 0.036 (~ 2

M).74,75 The environment of an air/water interface is very different from that of a

water/silica interface, because cations can interact with the silica surface via electrostatic

interactions.76 In this paper, we present studies of water structures on silica surfaces using

IR-visible sum frequency generation vibrational spectroscopy. Significant perturbation of

the water structures was observed with a relatively low concentration of NaCl in water.

Further, different alkali cation species (Na+, Li+, and K+) showed different degrees of

23
impact on the interfacial water structures. The electrostatic interactions between the

hydrated cations and the silica surface as well as the effective ionic radii of the cations

need to be considered to explain the observed phenomena.

24
3.2 Experimental Section

The visible and tunable IR laser beams for SFG vibrational spectroscopy were

obtained from a Nd:YAG (yttrium aluminum garnet) laser with output wavelength of

1064 nm (30 ps, 40 mJ/pulse, and 10 Hz). The laser was used to generate a second

harmonic beam at 532 nm in a KTiOPO4 (KTP) crystal. The tunable IR beam was

produced by difference frequency mixing of the 1064 nm beam with the output of a

home-made KTP optical parametric generator/amplifier (OPG/OPA) pumped by the 532

nm beam. The 532 nm and IR beams were overlapped, both spatially and temporally, on

the sample, as shown in Figure 3.1. The laser fluence was approximately 2 mJ/cm2 per

pulse for the visible beam and 5 mJ/cm2 per pulse for the IR beam. The polarizations of

the beams were s-, s-, and p-polarized for SFG, visible, and IR, respectively. The SFG

intensity was detected by a photomultiplier tube after spatial filtering by an aperture, and

spectral filtering by a bandpass filter. The SFG intensity was normalized against that

from z-cut quartz. Each spectrum shown in the current study was an average of 10 scans

in a 10 cm-1 step, and each scan was obtained by averaging the SFG intensity of 40 laser

shots at each step.

Fused silica plates, with a thickness of 3mm, were cleaned with a commercial

cleaning agent (extran AP12) for 3 min. It was immersed in a 50/50 (v/v) HNO3/H2SO4

solution for ~12 hours, followed by rinsing in pure water (resistivity > 18.2 Mcm,

Millipore). Alkali chloride salts (> 99.8%, certified ACS reagents; purchased from Sigma

Aldrich) were used to prepare solutions with different concentrations. No organic

contamination in the salt solutions was observed in the SFG spectra between 2700 cm-1

and 3000 cm-1. All data presented in the current study were collected within a period of

25
two weeks. During the experimental period, the silica substrates were mostly kept either

in air or acidic solutions to avoid surface quality changes due to prolonged exposures to

pure water, as previously reported by Li et al.77 The SFG spectrum of the pure

water/silica interface was also monitored at the beginning and the end of the experiment

for each electrolyte to ensure that the quality of the silica surface stayed consistent during

the experimental period.

Figure 3.1. Schematic layout of the SFG vibrational spectroscopy. The frequency of the

visible beam was fixed at 532 nm and the frequency of the IR beam was scanned from

2800 3900 cm-1. The 532 nm and IR beams were overlapped both spatially and

temporally on the bottom surface of the solution.

After the pure water/silica SFG spectrum was measured, the SFG spectra of a

series of NaCl solutions with concentrations of 1x10-4 M, 5x10-4 M, 1x10-3 M, 1x10-2 M,

and 1x10-1 M were measured in the sequence from the lowest concentration to the highest

concentration. For each solution with a particular concentration, the cell and silica plates

were rinsed thoroughly with the solution before the spectroscopic measurements. The

26
rinsing process ensured that the bulk solution in the cell had the desired electrolyte

concentration. For each concentration, five scans were collected in a time period of ~1

hour during which no change of the SFG spectrum was observed. Then the cell and silica

plates were cleaned with acids as described above for measuring next electrolyte

solutions (LiCl, for example). The measurements of different electrolytes were carried

out using the same silica substrate within a period of one week, and the experiments were

repeated under the same condition in the second week with freshly prepared solutions and

the same silica substrate. There was no observable difference in the SFG spectra in

comparison to those collected in the previous week. It also confirmed that the surface

quality of the silica substrate had stayed consistent during the whole experiments. Finally,

the 10 spectral scans of the same electrolyte with the same concentration (five scans in

the first experiment and five scans in the repeated experiment) were averaged to improve

the signal-to-noise ratio.

As the pH values of solutions can affect the surface water structure, great

attention has been made to monitor the pH values of the solutions to assure that the

observed changes of SFG spectra were not a pH effect. The pH value of water was 7

when freshly obtained from the Milli-Q system. However, it is known that water exposed

to air is mildly acidic because water readily absorbs carbon dioxide from the air. It

ultimately leads to a pH of approximately 5.7.78,79 All salts studied in this paper, such as

LiCl, NaCl, and KCl, are known as neutral salts, which cause little change of the pH

values in aqueous solutions. The pH values of all solutions used in the current study were

around 5.74 with a standard deviation of 0.05. A table presenting the pH value for each

individual solution was given in the supporting document.

27
3.3 Results and Analysis

Figure 3.2A shows the SFG spectra of water/silica interfaces with pure water and

aqueous NaCl solutions of 1x10-4 M, 5x10-4 M, 1x10-3 M, 1x10-2 M, and 1x10-1 M. The

spectra of water exhibit two peaks at 3200 cm-1 and 3400 cm-1.11,72 As the concentration

of NaCl increased, the intensity of both peaks decreased. It has been interpreted that the

peak at 3200 cm-1 represents OH in a more ordered hydrogen-bond network, and the peak

at 3400 cm-1 represents a less ordered hydrogen-bond network.10,20,80 It is known that

silanol groups on silica play a critical role in determining the molecule adsorption on the

surface in an aqueous solution.76 The dissociation of protons from the surface silanol

groups creates negative charges on a silica surface. The reaction can be described as

follows:

-SiOH + mH2O = -SiO -- mH2O + H+ (3.1)

where -SiOH is the surface silanol group and m describes the number of water molecules

associated with the -SiO.81,82 The silanol groups are estimated to have a surface density

of ~5 1014 cm-2 on the silica surface, which is equivalent to one silanol group per 20

2.76,83 When the pH value of the aqueous solution increases, the silanol groups become

deprotonated, and the surface charge increases. Experimentally, both the 3200 cm-1 and

3400 cm-1 peaks were observed to increase with increasing pH value, suggesting that a

larger surface electrical field induces a better ordered hydrogen-bond network on

silica.11,73,84 Previous studies by Ong et al., using second harmonic generation (SHG),

showed that there were two types of silanol groups at a water/silica interface, with pK

values of 4.9 and 8.5, populating 19% and 81% of the surface area, respectively.71

Further studies by Ostroverkhov et al., using a phase-sensitive SFG technique, indicated

28
that the 3200 cm-1 peak is associated with surface sites that have a higher pK value, and

the 3400 cm-1 peak is associated with surface sites with a lower pK value.73

To obtain quantitative information, the spectra in Figure 3.2 were fitted by two

Lorentzian peaks at 3200 cm-1 and 3400 cm-1:

2
Aq ,ijk
I ( SFG ) (2)
NR ,ijk +
q =1, 2 q +i q
(3.2)
IR

where NR
( 2)
is the nonresonant contribution; IR is the frequency of the input infrared

laser beam; q represents the qth vibrational mode; Aq is the amplitude; q is the width;

and q is the resonant frequency. The fitting curves are shown in Figure 3.2 as solid lines,

and the fitted amplitudes Aq are plotted in Figure 3.3.

Parts B and C of Figure 3.2 show the SFG spectra collected with LiCl and KCl

solutions. The magnitudes of the decreases are different when the cation species are

changed. Similar experiments with different anions, such as NaBr and NaI, were also

carried out, but, within measurement error, the spectra of NaBr and NaI were the same as

those of the NaCl solutions. (Spectra of NaBr and NaI solutions are not shown.)

Therefore, the cation, which interacts with the silica surface via electrostatic interactions,

is the key factor in perturbing the hydrogen-bond network at the water/silica interface.

29
Figure 3.2. SFG spectra from the water/silica interfaces with different concentrations of

three alkali chloride solutions: (a) NaCl, (b) LiCl, and (c) KCl. The solid lines are fitting

curves derived using two Lorentzian peaks at 3200 cm-1 and 3400 cm-1.

30
The interaction of a hydrated cation with a silica surface has been shown to

promote negative charge development.76 The reaction can be described as:



-SiOH + nH2OM+ = -SiO -- nH2OM++ H+ (3.3)

where M+ denotes a cation and n describes the number of water molecules solvating M+.

The notation, -SiO -- nH2OM+, indicates that cations are located at a small but finite

distance from the silica surface. Overall, the surface charges created by the cations are

mostly neutralized by the cations themselves. Therefore, the surface charges developed

by the cations are not expected to enhance the ordering of surface water molecules, or the

observed SFG intensity.

The pH of the point of zero charge for amorphous silica is about 2 - 3.85 With a

solution pH value of 5.7, the silica surface is negatively charged. The surface -SiO

groups, described in Equation (3.1), interact with the cations via electrostatic

interaction:76

-SiO -- mH2O + nH2OM+ = -SiO -- nH2OM++ mH2O Kassc (3.4)

where Kassc is the equilibrium constant. As a result of the electrostatic interaction, the

cations reduce the surface electrical field. Overall, the ordering of the original hydrogen-

bond network is perturbed because of the decrease in the surface electrical field and the

replacement of the more ordered water molecules by the hydrated cations. Qualitatively,

this mechanism explains that the SFG intensity of interfacial water peaks decreases as the

salt concentration increases.

31
Figure 3.3. Amplitudes (Aq) as functions of the concentration of alkali chloride solutions

for water/silica interfaces obtained by fitting the spectra in Figure 3.2 with Equation (3.2).

32
As shown in Figures 3.2 and 3.3, for the same concentration, the SFG intensities

of interfacial water in alkali chloride solutions are in the order: Na+ > Li+ > K+. To

explain the different behaviors for Li+, Na+, and K+, both the equilibrium constant Kassc in

Equation (3.4) and the effective ionic radii of the hydrated cations need to be considered.

First, the equilibrium constant Kassc describes the interaction between the surface -SiO

groups and the hydrated cation nH2OM+. Previous studies by Dove et al. showed that the

equilibrium constant Kassc for alkali chloride solutions is in the order: Kassc, KCl > Kassc, NaCl

> Kassc,LiCl.86 A larger equilibrium constant Kassc indicates a higher density of -SiO--

nH2OM+ on the silica surface; thus, a larger perturbation of the water structures, and

consequently, a smaller SFG intensity. Therefore, if only the equilibrium constant Kassc is

considered, the SFG peak intensity would be in the order: Li+ > Na+ > K+. Second, the

sizes of the hydrated cations should be considered. Ions in aqueous solution are hydrated.

Generally, the smaller and higher-charged ions attract more water molecules. The

hydrated ionic radii of Li+, Na+, and K+ are approximately 0.6 nm, 0.4 nm, and 0.3 nm,

respectively.87 In an aqueous solution, the radius of a hydrated Li+ is roughly twice that

of a hydrated K+. Therefore, a hydrated Li+ displaces more ordered H2O at the surface.

Consequently, if only the effective ionic radii are considered, one would expect that the

SFG intensity of water be in the order: K+ > Na+ > Li+. Overall, the balance between

these two effects gives the observed SFG intensity in the order: Na+ > Li+ > K+. The SFG

intensity is lower for K+ and Li+ solutions because K+ has a stronger interaction with the

surface -SiO groups and Li+ has a larger effective ionic radius. Nevertheless, these two

effects, even though they perturb the interfacial water ordering in different ways, are not

totally independent. The size of a hydrated cation is determined by the electrostatic

33
interaction between the cation and water molecules, and the strength of the interaction

between the hydrated cation and the silica surface depends on the size of the hydrated

cation.88-91 The structures of the water of hydration remain as an active research area, and

many theoretical studies of the detailed structures can be found in the references. 92-97

As shown in Figure 3.3, the peak at 3200 cm-1 is more vulnerable to the cation

perturbation. The amplitude of the 3200 cm-1 peak experiences a ~ 20% decrease with a

concentration of 1x10-4 M, while the 3400 cm-1 peak does not have an significant

decrease until 1x10-2 M. Additionally, the perturbation for the 3200 cm-1 peak reaches its

saturation at a concentration of 0.01 M, while the perturbation for the peak at 3400 cm-1 is

not saturated until 0.1 M. Eventually, both peaks have a 50% decrease in their amplitudes

Aq , which is equivalent to a 75% decrease in the measured SFG intensity.

Since the peak at 3200 cm-1 is more vulnerable to the cation perturbation, the

water structure associated with this peak is more likely to exist in the region where the

surface electrical field is higher. As described above, cations interact with the silica

surface via electrostatic interaction, and therefore, the surface number density of cations

is higher in the region where the surface electrostatic field is higher. Previous studies

have suggested that the silanol groups with lower pK values are isolated silanol groups

because they are relatively easier to dissociate.71,76,98 Vincinal silanol groups, which

locate closely and can be coupled to each other through hydrogen bonds, have higher pK

values. The vincinal silanol groups are likely to create a higher local surface charge

density because of their higher local number density. Therefore, vincinal silanol groups

are more likely to create a higher local electrical field on the silica surface, consequently

a better ordered structure and attract more cations to the surface. This model is consistent

34
with previous observations by Ostroverkhov et al., indicating that the peak at 3200 cm-1 is

associated with surface sites that have a higher pK value.73

35
3.4 Conclusions

IR-visible sum frequency vibrational spectroscopy was applied to study the

structure of water molecules at water/silica interfaces with the presence of alkali chloride

in the solutions. Significant perturbations of the interfacial water structures were

observed with a 1x10-4 M NaCl solution. The cations play a key role in perturbing the

hydrogen-bond network at the water/silica interfaces as they interact with the silica

surface via electrostatic interaction. Different alkali cation species produce different

degrees of perturbation in the order: K+ > Li+ > Na+. This order can be explained by

considering the electrostatic interaction between the cations and silica surfaces and the

effective ionic radii of the cations. The peak at 3200 cm-1 experiences lager perturbation

suggesting that the more ordered structure at 3200 cm-1 is associated with the vincinal

silanol groups, which produce a higher surface electrical field and have a higher pK value

compared to isolated silanol groups.

36
Chapter 4

Structures of Water Molecules at Solvent/Silica

Interfaces*

* A version of this chapter has been published. Yang, Z.; Li, Q. F.; Gray, M. R.; Chou, K. C. Structures of
Water Molecules at Solvent/Silica Interfaces. Langmuir 2010, 26, 16397-16400.

37
4.1 Introduction

Surface chemistry at solvent/mineral interfaces plays a critical role in many

industrial and environmental processes, such as oil sand processing28,29,34,38, petroleum

recovery32, contamination removal30,33,35,37, and many other extraction techniques31,36.

Water molecules, either existing naturally or being added in a process, form a thin layer

on the mineral surface and may affect the effectiveness of the process. For example,

previous studies have suggested that the loss of water from the mineral interfaces can

significantly decrease the yield of oil sand extraction.99 Despite its importance, the

behaviour of water molecules at solvent/mineral interfaces is poorly understood, mainly

because it is technically challenging to probe the buried interfaces. Recent developments

in sum frequency generation (SFG) vibrational spectroscopy have proven that SFG is an

effective technique to study buried liquid interfaces.20,72 Previous studies at water/silica

interfaces using SFG have shown that the structure of water is strongly affected by the

surface charges on silica.84 In addition to the liquid-water vibrational peak at 3400 cm-1,

an ice-like peak located at 3200 cm-1 was observed at water/silica interfaces. The ice-

like peak grows significantly with higher surface charge density, suggesting that the

surface charges on silica induce a better ordered water network with increased hydrogen

bonds. On the other hand, the SFG spectra of water at hydrophobic interfaces exhibit a

narrow peak near 3650 cm-1, which is associated with the free OH dangling bonds, and a

broad figure between 3000 and 3600 cm-1, which is associated with hydrogen-bonded OH.

Studies by Scatena et al. at hexane/water interfaces have shown that the hydrogen

bonding between water molecules at the interface is weaker than that at air/water

interfaces.100 Therefore, the structure of water at solvent/silica interfaces will depend on

38
the interactions of water with the hydrophilic mineral surfaces and the hydrophobic

solvents.

In this paper, the adsorption of water at toluene/silica and heptane/silica interfaces

was studied using SFG. Toluene and heptane were chosen because of their popularity in

laboratories and industries. The properties of toluene and heptane are also fundamentally

interesting for a comparison, as toluene is aromatic and heptane is a straight-chain alkane.

Silica was used in the current study because of its abundance in nature. The surface

chemistry of silica is important for oil sand processing28,29,34,38 and liquid

chromatography101-104. We observed that the structure of water adsorbed at the

solvent/silica interface was strongly affected by the organic solvent. We discovered a

highly hydrogen-bonded water structure at the toluene/silica interface with no detectable

free OHs. Surprisingly, this structure of water showed resistance against further

adsorption of water and was extremely stable at room temperature. On the other hand, a

similar structure was not observed at heptane/silica interfaces.

39
4.2 Experimental Section

The visible and tunable IR laser beams for SFG vibrational spectroscopy were

obtained from a Nd:YAG (yttrium aluminum garnet) laser with output wavelength of

1064 nm (30 ps, 40 mJ/pulse, and 10 Hz). The laser was used to generate a second

harmonic beam at 532 nm in a KTiOPO4 (KTP) crystal. The tunable IR beam was

produced by difference frequency mixing of the 1064 nm beam with the output of a KTP

optical parametric generator/amplifier (OPG/OPA) pumped by the 532 nm beam. The

532 nm and IR beams passed through the silica window and were overlapped, both

spatially and temporally, at the solvent/silica interface. The laser fluence was

approximately 2 mJ/cm2 per pulse for the visible beam and 5 mJ/cm2 per pulse for the IR

beam. The polarizations of the beams were s-, s-, and p-polarized for SFG, visible, and IR,

respectively. The SFG intensity was detected by a photomultiplier tube after spatial

filtering by an aperture, and spectral filtering by a bandpass filter. The SFG intensity was

normalized against that from a z-cut quartz.

Optically-smooth fused silica windows (ISP optics, USA) with a thickness of 3mm

and sample cell made of glass (capacity: 1 ml) were cleaned with a commercial cleaning

agent (extran AP12) for 3 min. They were then immersed in a mixture of sulfuric acid

(98%) and nochromix reagent (GODAX Laboratories, Inc.) for 24 hours, followed by

rinsing in pure water (resistivity > 18.2 Mcm, Millipore), and finally dried at 100 for 3

h to remove residual surface water. After these treatments, the silica plates were kept in

heptane or toluene (HPLC grade, Fisher) to prevent further water adsorption on the

surface. All spectra were taken at room temperature.

40
4.3 Results and Analysis

Figure 4.1 shows the SFG spectra of heptane/silica interfaces with dry (open

circles) and water-saturated (solid circles) heptane. The peaks between 2800 and 3000

cm-1 are associated with C-H vibrational modes from heptane. The detailed study of

heptane/silica interfaces has been reported previously.105 The current analysis will focus

on the water peaks between 3000 and 3700 cm-1. The dry heptane/silica interface contains

a small amount of water, and a free OH peak is visible at 3680 cm-1. The water-saturated

spectrum was collected at ~30 min after adding a small water droplet in the cell to

saturate the water content in heptane. The solubility of water in heptane is only ~ 0.056%

(mole fraction) at 25C.106-108 However, the adsorption of water on silica is favorable

because the silanol groups (either SiOH or SiO) on the silica surface interact with water

molecules strongly. With the water-saturated heptane, three major water peaks can be

identified at 3200, 3400, and 3660 cm-1 as shown in Figure 4.1 (solid circles). The peaks

at 3200 and 3400 cm-1 have been commonly observed at water/silica interfaces.20 These

two peaks are located at the same positions as the IR absorption peaks of bulk ice and

liquid water and hence are sometimes labeled as ice-like and liquid-like peaks,

respectively.20 It has been proposed that the 3200 cm-1 peak represents more ordered

water molecules with symmetric tetrahedral coordination, while the 3400 cm-1 peak

represents the more disordered asymmetrically bonded molecules.80,100 The free OH peak

at 3660 cm-1 has been widely observed at water surfaces in contact with hydrophobic

environments, such as air/water11 or oil/water100 interfaces. Therefore, the 3660 cm-1 peak

in Figure 4.1 represents free OHs in contact with the hepane phase. The spectra of water

at heptane/silica interfaces, as shown in Figure 4.1, are basically consistent with our

41
current understanding of interfacial water molecules and show a combination of SFG

spectra from water at a typical water/hydrophilic-surface and a typical

water/hydrophobic-surface interfaces.

16
14
SFG Intensity (arb. u.)

12
10
8
6
4
2
0
2800 3000 3200 3400 3600 3800
-1
Wavenumber (cm )

Figure 4.1. SFG spectra of heptane/silica interfaces with dry (open circles) and water-

saturated (solid circles) heptane. The peaks between 2800 and 3000 cm-1 are associated

with C-H vibrations of heptane, the peaks at 3200 and 3400 cm-1 are vibrational peaks

associated with hydrogen-bonded water, and the 3660/3680 cm-1 peak is assigned to the

free O-H stretching mode.

Figure 4.2 shows the SFG spectra of toluene/silica interfaces with dry (open

circles) and water-saturated (solid circles) toluene. The vibrational peaks between 2800

and 3100 cm-1 are associated with toluene. The detailed peak assignments for toluene

have been reported previously and can be found in the reference.105 Again, a small

amount of water was present at the interface with the dry toluene and showed a free OH

42
peak near 3600 cm-1. The position of the free OH peak is red-shifted and broadened

compared to that at heptane/silica interfaces (Figure 4.1). The redshift of the free OH

mode indicates that the solvent does interact with the interfacial water molecules. When

water was added into the dry toluene, water molecules adsorbed on silica and formed a

new structure which had not been previously observed. The solubility of water in toluene

is ~ 0.28% (mole fraction) at 25C.109,110 As expected, the water in toluene adsorbed on

the silica surface, and the vibrational peak associated with water appeared, as shown in

Figure 4.2 (solid circles). Surprisingly, the adsorbed water produced a very different

spectrum compared to that at the heptane/silica interface shown in Figure 4.1 (solid

circles). At the toluene/silica interface, water formed a broad peak centered at 3150 cm-1.

Based on our current understanding of water surfaces, two important features are missing

in the water-saturated spectrum shown in Figure 4.2. First, the free OH peak, which exists

at an oil/water interface, is no longer detectable. Second, the weakly bonded water peak

at 3400 cm-1, which exists at the water/silica interface, is also missing. The spectrum

suggested that water formed a highly-hydrogen bonded structure at the toluene/silica

interface with nearly no free OHs.

An important character of a water surface without free OHs is that the interaction

between the water surface and an incoming water molecule is significantly reduced.111

The water structure shown in Figure 4.2 (solid circles) was found to be very stable in the

water-saturated toluene. For more than 12 hours, the spectrum remained unchanged, and

the water surface did not accumulate more water to become a regular liquid water layer

similar to that at the heptane/silica interface. For comparison, the same experiment was

carried out on an initially moistened silica surface (by exposing to N2 of 60% relative

43
humidity), where the 3200 cm-1, 3400 cm-1, and free OH peaks were all present initially,

as shown in Figure 4.3 with open circles. The solid circles in Figure 4.3 show the SFG

spectrum taken 30 min after water was added in toluene. It was clear that, on this regular

water surface, further adsorption of water continued when the water content in toluene

was saturated.

10
SFG Intensity (arb. u.)

0
2800 3000 3200 3400 3600 3800
Wavenumber (cm-1)

Figure 4.2. SFG spectra of toluene/silica interfaces with dry (open circles) and water-

saturated (solid circles) toluene. The peaks between 2800 and 3100 cm-1 are associated

with C-H vibrations of toluene, the broad peak centered at 3150 cm-1 is associated with

hydrogen-bonded water, and the 3600 cm-1 peak is assigned to the free O-H stretching

mode.

44
6

SFG Intensity (arb. u.)


4

0
2800 3000 3200 3400 3600 3800
-1
Wavenumber (cm )

Figure 4.3. Adsorption of water on a regular water layer with free OHs. The open circles

represent the SFG spectrum of a toluene/silica interface with dry toluene on an initially

moistened silica surface. In this regular water layer, the 3200 cm-1, 3400 cm-1, and free

OH peaks are all present. Water can condense on such a regular water surface, as

indicated by the solid-circle spectrum, which was taken 30 min. after the toluene was

saturated with water.

The water-saturated spectrum in Figure 4.2 (solid circles) suggests that the

formation of a water surface with no free OHs is possible without an atomically smooth

surface. On an atomically smooth mica surface, it has been proposed that the condensed

water molecules may form a two-dimensional (2D) ice-like structure which is stabilized

via the interaction between water molecules and the solid surface.112,113 This structure

was latter confirmed by SFG vibrational spectroscopic measurements for D2O on mica,

with a strong ice-like peak located 2375 cm-1 accompanied by the disappearance of

45
both the liquid-like peak at 2510 cm-1 and the free O-D peak at 2740 cm-1.114 A similar

2D flat ice structure was also observed at -160 C on an atomically smooth crystalline

Pt(111) surface, on which free OHs were not present because all water molecules bound

either directly to the metal surface or to each other through the in-layer hydrogen bonds.

This 2D ice layer was reported by Kimmel et al. as a hydrophobic water monolayer

because it depressed further growth of ice on the surface.111 The structure shown in

Figure 4.2 (solid circles) shares many similar properties with these 2D water structures,

such as the disappearance of both the liquid-like and free OH peaks and the depression

of further water adsorption. However, one major difference was that the silica surface

used in the current study was not atomically smooth. The optically flat silica surface was

mechanically polished and has a roughness of a several nanometers under an atomic force

microscope. The rough surface cannot provide a lattice matching condition to form a 2D

ice structure similar to that on an atomically smooth mica or Pt surface. A new

mechanism is needed to explain the observed phenomena.

A possible explanation is that the surface defects (scratches, cavities, or pits) on

silica may help the adsorbed water molecules to minimize the number of free OHs. When

water molecules adsorb on the silica surface one by one to fill up the defects, the water

molecules will minimize their energy by forming the maximal number of hydrogen bonds.

At a certain stage, the OHs bind either to other water molecules or to the silanol groups

(protonated or deprotonated) on silica, as demonstrated in Figure 4.4, and these

nanometer holes are then sealed for further water adsorption. This model explains the

long-term stability of the highly hydrogen-bonded structure shown in Figure 4.2 (solid

circles). On the other hand, if the defects are initially overfilled with too many available

46
free OHs, as shown in Figure 4.3 (open circles), the free OH peaks will always exist on

the regular water layer, and then further adsorption of water is favorable. Although the

water-resistant water structure centered at 3150 cm-1 has a vibrational frequency near

that of ice, it is unlikely to have an icelike structure with symmetric tetrahedral

coordination because the tetrahedral coordination would have to be highly distorted to fit

in the defect sites, which explains the broad width of the peak. Nevertheless, the large red

shift of the OH stretching frequency, in comparison to free OHs, indicates that the

strength of the hydrogen bonds is likely comparable to that of ice.

Figure 4.4. Schematic drawing of the water-resistant water structure. In the process of

water adsorption, the water molecules minimize their energy by forming the maximal

number of hydrogen bonds with either the neighbouring water molecules or the silanol

groups (SiOH) on silica. At a certain stage, very few free OHs are available. The absence

of free OH groups significantly reduces further water adsorption.

47
Toluene plays an important role in stabilizing the water-resistant water surface.

When experiments were carried out at an air/silica interface with the relative humidity

(RH) increasing from 0%to 100%, the free-OH peak was observed at all RH values,

indicating the water surface without free OHs either could not form or was unstable at an

air/silica interface. Therefore, toluene plays a critical role in minimizing the number of

free OHs. Our data also indicated that the hydrophobicity of toluene was not the only

driving force to minimize the number of free OHs on the water surface, as we did not

observe a similar structure at heptane/silica interfaces. Further experimental and

theoretical studies will be needed to better determine whether the shape of toluene or the

stronger interaction between toluene and water also plays a role in forming the water

layer without free OHs.

48
4.4 Conclusions

The adsorption of water at solvent/silica interfaces was studied using SFG

vibrational spectroscopy. Experimental data showed that the interactions between

solvents and water molecules can significantly change the interfacial water properties.

We discovered a water layer between toluene and silica with no detectable free OHs. The

water layer without free OHs showed resistance against further adsorption of water

molecules. However, this special structure of water was not observed at heptane/silica

interfaces, at which free OHs were always observed.

49
Chapter 5

Why do Sulfuric Acid Coatings Influence the Ice

Nucleation Properties of Mineral Dust Particles

in the Atmosphere*

* A version of this chapter has been published. Yang, Z.; Bertram, A. K.; Chou, K. C. Why do Sulfuric
Acid Coatings Influence the Ice Nucleation Properties of Mineral Dust Particles in the Atmosphere?
Journal of Physical Chemistry Letters 2011, 2, 1232-1236.

50
5.1 Introduction

Mineral dust particles are abundant in the atmosphere. Common minerals found in

aerosolized dust include quartz, illite, muscovite, chlorite, kaolinite, and calcite.115-125

During their lifetime in the atmosphere, mineral dust particles can become coated with

inorganic material such as sulfuric acid. 126,127 Recently several studies have investigated

the effect of sulfuric acid coatings on the ice nucleation properties of mineral dust

particles.118,128-133 Laboratory studies with supermicron particles have shown that

uncoated minerals particles such as illite and kaolinite are good ice nuclei both below and

above water saturation.134,135 But once the supermicron particles are coated with sulfuric

acid they are only good ice nuclei at close to and above 100% relative humidity (RH) for

temperatures around 235 K.129,132 Experiments with submicron mineral particles also

show that particles coated with sulfuric acid are poor ice nuclei below water saturation,

but can act as ice nuclei above water saturation (although the ice nucleation ability may

be reduced above water saturation compared to the uncoated case due to the permanent

loss of some active sites).133 This difference between uncoated and coated mineral dust

particles in terms of ice nucleation ability is illustrated in Figure 5.1a, for the specific

case of supermicron illite particles from the work of Chernoff et al.132 Figures 5.1b and

5.1c show the pH and H2SO4 concentration of the coating as a function of RH, calculated

from the Aerosol Inorganic Model.136 A comparison of panel a, b and c illustrate that

coated mineral dust particles are poor ice nuclei when the coating material has a low pH

(less than 1) and a high acid concentration (H2SO4 concentration > 0.1 M).

51
Figure 5.1. Panel (a) schematic showing the range of RH values over which uncoated and

coated mineral dust particles are good ice nuclei. The figure is specifically created for

illite particles based on the work of Chernoff et al.132 A similar behavior was also

observed for kaolinite. The temperature applicable for the ice nucleation studies was

approximately 237 K. Panel (b) and (c) show the pH and the molarity of the sulfuric acid

coating, assuming the coating is in equilibrium with the relative humidity.

Despite the evidence that sulfuric acid coatings influence the ice nucleation

properties of mineral dust particles, a molecular-level understanding is still lacking. A

molecular-level understanding is preferred so that laboratory results can be confidently

52
extrapolated to the atmosphere. In the following we investigate why sulfuric acid

coatings influence the ice nucleation properties of mineral dust particles at molarities >

0.1 M. We probed the structure of water at the mineral aqueous acid interface as a

function of the sulfuric acid concentration using IR-visible sum frequency generation

(SFG) vibrational spectroscopy, which is a highly surface-specific optical technique for

studying liquid/solid interfaces.20 We specifically focused on differences in water

structures for concentrations less than and greater than 0.1 M. For these studies mica was

chosen for the mineral interface since mica has a surface structure similar to illite which

is reported to make up more than 50% percentage of dust in some regions.137

Additionally, atomically flat mica surfaces are relatively easy to obtain by cleaving a

mica sheet.

53
5.2 Experimental Section

The visible and IR laser beams for SFG vibrational spectroscopy were obtained

from a Nd:YAG (yttrium aluminum garnet) laser with output wavelength of 1064 nm (30

ps, 40 mJ/pulse, and 10 Hz). The laser was used to generate a second harmonic beam at

532 nm in a KTiOPO4 (KTP) crystal. The tunable IR beam was produced by difference

frequency mixing of the 1064 nm beam with the output of a KTP optical parametric

generator/amplifier (OPG/OPA) pumped by the 532 nm beam. The 532 nm and IR beams

were overlapped, both spatially and temporally, on the sample, as shown in the inset of

Figure 5.2. The energy of the laser beams were ~ 200 uJ/pulse for both the visible and IR

beams. The polarizations of the beams were s-, s-, and p-polarized for SFG, visible, and

IR, respectively. The SFG intensity was detected by a photomultiplier tube after spatial

filtering by an aperture and spectral filtering by a bandpass filter. Each spectrum shown

in the current study is an average of 6 scans in a 10 cm-1 step, and each scan is an average

of 50 laser shots per step.

Deuterium oxide D2O (99.9% deuterated water; Sigma-Aldrich) was used, instead

of H2O, to avoid the IR absorption in mica due to the O-H stretching at 3620 cm-1.138

Deuterated sulfuric acid D2SO4 (Sulfuric acid-d2, 96-98 wt.% deuterated, solution in D2O,

99.5 atom% deuterated, Sigma-Aldrich) was used to prepare solutions for the SFG

measurements. Mica sheets were cleaved right before the SFG measurements. A series of

D2SO4 solutions with concentrations of 0, 5x10-6, 5x10-5, 5x10-4, 5x10-3, 5x10-2, 5x10-1

mol/L were measured in the sequence from the lowest concentration to the highest

concentration. For each solution, the spectroscopic cell and mica sheet were rinsed

thoroughly with the solution before the spectroscopic measurements. The rinsing process

54
ensured that the concentration of D2SO4 in the cell is the same as the solution used. For

each concentration, 6 scans were collected in a period of ~1 hour during which no

significant change of the SFG spectrum was observed. All spectra presented in the paper

were taken using the same mica sheet. All experiments were carried out at room

temperature.

55
5.3 Results and Analysis

Figure 5.2 shows the SFG vibrational spectra of deuterium oxide (D2O)/mica

interfaces with D2SO4 concentrations of 0, 5x10-6, 5x10-5, 5x10-4, 5x10-3, 5x10-2, 5x10-1,

and 5 M. Deuterated water and sulfuric acid was used, instead of H2O, to avoid the IR

absorption peak of mica at 3620 cm-1.138 The spectra of D2O exhibit two peaks located

near 2375 and 2550 cm-1, which represent the peaks at 3200 and 3400 cm-1, respectively,

for interfacial H2O.11 The redshift of the D2O peaks with respect to those of H2O is a

result of the isotopic substitution. These two peaks are sometimes labeled as the ice-

like and the liquid-like peaks, indicating that their peak positions are similar to those

of ice and liquid water, respectively.20,72 It is generally accepted that the former

represents water in a more ordered hydrogen-bonding network, and the latter represents a

less ordered hydrogen-bonding structure.10,20 The detailed water structures associated

with these two peaks are not completely understood, but it was proposed that the 2375

cm-1 peak represents water molecules with symmetric tetrahedral coordination, while the

2550 cm-1 peak represents the asymmetrically bonded molecules.80,100

56
Figure 5.2. SFG spectra of D2O/mica interfaces with D2SO4 concentrations of (a) 0 M, (b)

5x10-6 M, (c) 5x10-5 M, (d) 5x10-4 M, (e) 5x10-3 M, (f) 5x10-2 M, (g) 5x10-1 M and (h) 5

M. The inset shows the schematic layout of the spectroscopic setup. The polarizations of

the beams were s-, s-, and p-polarized for SFG, visible, and IR, respectively.

To obtain a more quantitative analysis, the spectra in Figure 5.2 were fitted with

2
Aq
Lorentzian lineshape I ( SFG ) ( 2)
NR + , where NR
( 2)
is the nonresonant
q IR q +i q

57
contribution, Aq is the amplitude, q is the width, and q is the resonant wavenumber for

the qth vibrational mode. The fitting curves were shown in Figure 5.2 (solid lines), and

the fitting amplitudes and resonant wavenumbers were shown in Figure 5.3. The best fits

were obtained with a negative Aq for the ice-like peak and a positive Aq for the

liquid-like peak. These assignments are consistent with the phase-sensitive SFG

measurements at low pH reported by Ostroverkhov et al.73 As shown in Figure 5.2 and

5.3a, when the concentration of D2SO4 increased, both the liquid-like and the ice-

like peaks intensities decreased. With a D2SO4 concentration of 5 M (pH ~ -0.7), both

peaks completely vanished, indicating that the ordered hydrogen-bonding network of

water no longer exists. Figure 5.3b shows that the frequencies of the water peaks are pH

dependent. In general, both the surface potential of mica and dipole-dipole coupling

between water molecules can shift the frequency139, but studies of the frequency shifts are

beyond the focus of the current study.

The drastic reduction of the ordered water structures (Figure 5.2) with D2SO4

concentrations of 0.5 M is consistent with previous observations by Eastwood et al. and

Chernoff et al. indicating H2SO4-coated mineral surfaces with H2SO4 concentrations >

0.1 M are poor ice nuclei (see Figure 5.1).129,132,135 The results suggest that the ordered

surface water structures may have played an important role in the ice nucleation process.

Although the drastic reduction and eventual disappearance of the ordered water structures

offers a microscopic explanation for why H2SO4-coated mineral surfaces are poor ice

nuclei, the question remains to be answered is why the ordered surface water structures is

drastically reduced with a high acid concentration. Based on our current understanding,

there are three possible effects responsible for the disappearance of ordered water

58
structures: (a) the decrease of mica surface potential, (b) the adsorption of sulfates on the

surface, and (c) solvations of sulfates in water. These effects are discussed below.

Figure 5.3. The fitted amplitudes (a) and frequencies (b) of water peaks in the SFG

spectra with various D2SO4 concentrations. Triangles represent ice-like peaks and dots

represent liquid-like peaks.

It is known that a decrease in the surface potential leads to a less ordered water

structure on material surface.84,140-142 The surface potential on a mineral surface is pH

dependent. The structure of muscovite mica constituted of octahedral hydroxyl-aluminum

59
sheets lying between two silicon tetrahedral layers. One in four silicon atoms is

substituted by an aluminum atom in the silicon tetrahedral layer, and the substitution

results in a negative charge which is neutralized by K+ ions located between the silicon

tetrahedral layers. When mica is cleaved, a cleavage plane happens in the potassium

layer. The K+ ions are equally distributed between the two surfaces, and the surface is

overall neutral.143 When placed in water, hydrated potassium ions dissociate from the

mica surface,144,145 and the surface, which consists of Si, Al, and O connected by Si-O

and Al-O bonds, become negatively charged, then partially neutralized by H+ ions in

water. The surface charge of mica is therefore dependent on the pH of the solution.

Overall, the surface charge decreases when the pH value decreases. The phenomena have

also been observed on other mineral surfaces.71 Measurements for the point of zero

charge (PZC) of mica were not conclusive, but it is generally accepted that the PZC of

muscovite mica is less than 3.146 One of the difficulties in measuring the PZC for mica is

the increased solubility of lattice aluminium ions at low pH,147 which creates negative

surface charges on the surface. However, it was clearly observed that the increased

concentration of hydrogen ion at low pH decreased the surface potential of mica, which

can be attributed to the protonation of the Si-O groups at the surface to form ionizable

surface silanol groups.146 Therefore, the decrease of the surface potential on mica at

lower pH is one of the mechanism responsible for the decrease of the water peaks in

Figure 5.2.

When the mica surface approached neutral or positively charged, the adsorption

of anions on mica surface may also affect the structure of water on mica. It is known that

the adsorption of sulfate on mineral surface is mainly associated with Al and Fe oxy-

60
hydroxides and with allophanic constituents.148,149 Sulfate adsorption generally increases

when the pH decreases.148,150,151 The vibrational peaks of sulfate are too weak to be

measured directly by the current optical setup. To further study whether the

disappearance of the ordered water structure was related to any particular properties of

sulfate, the same experiment with 5 M of hydrochloric acid (HCl) was carried out for

comparison. The experiment showed the SFG peaks of water also disappeared with 5 M

of HCl solution. Therefore, the vanishing of the ordered water structures was likely not

linked to any particular properties of sulfates. However, the anions should interact with

the mineral surface when the surface approached neutral or become positively charged.

With a high concentration of H2SO4, the solvations of sulfates anions (SO42-)

become an important factor affecting the water structure on mica. Previous ab initio

studies by Cannon et al, showed that 13 water molecules are present in the first solvation

shell of SO42-.152 With a H2SO4 concentration of 5 M, the water-sulfate mole ratio is

roughly 8:1. At this concentration, the mica surface must compete for water with the

sulfate ions in the bulk solution. The phenomenon can affect ice nucleation processes in

two ways. First, the number of water molecules available to the mica surface is reduced.

Second, as the mica and sulfates are competing for water molecules, most water

molecules are well captured either by the surface charges on mica or by the anions in the

solution. The situation creates an energy barrier for ice nucleation because formations of

ice nuclei would have to overcome the electrostatic interactions to rearrange the ordering

of water molecules.

When the concentration of sulfuric acid increases, the above three mechanisms all

work against ordering of water molecules on mica surface. When pH decreases, the

61
surface potential decreases, and consequently the ordering of interfacial water decreases

which was indicated by the decrease of the SFG peaks. As the surface potential decreases,

the adsorption of anions on the mica surface becomes significant, and the adsorbed

anions displace ordered water molecules on the surface. Finally, when the concentration

of sulfuric acid reaches a critical concentration, in which the solvation of sulfate ions

consumes large amount of water, the ordered water structure disappears. At this stage,

nearly all water molecules are captured by the anions, and few water molecules are freely

available to the mica surface. The process is illustrated in Figure 5.4 showing water

molecules and hydrated sulfate (bisulfate) anions at solution/mica interfaces with

increasing concentrations of H2SO4.

62
Figure 5.4. Schematics of water molecules and hydrated sulfate (bisulfate) ions at

solution/mica interfaces with different concentrations of H2SO4 solutions. (a) Pure

H2O/mica interface with pH~6. The surface is highly negatively charged. Water

molecules are more ordered near the charged surface. (b)With a low concentration of

H2SO4, the surface charge decreases as the surface is protonated, and the anions interact

with the mineral surface. (c) With a high concentration of H2SO4 (for example 5M),

water molecules are well captured by the sulfate/anions in the solution, and few water

molecules are freely available for the mica surface. In (a) and (b) the dashed lines

separate the ordered water molecules from the water molecules without order. Above the

dashed lines, water molecules have good order because of negative surface potential

induced by the mica surface. In (b) and (c) H+ adsorb on the mica surfaces. The dashed

circles represent hydrated sulfate ions (in low H2SO4 concentration) or hydrated bisulfate

anion (in high H2SO4 concentration). Water molecules in the dashed circle are parts of

the hydrated anions and move together with the core anions.

63
5.4 Conclusions

The structures of water on mica surfaces in the presence of H2SO4 with

atmospheric relevant concentrations were studied using SFG vibrational spectroscopy.

We found that ordered water structures significantly decreased with 0.5 M of H2SO4 and

disappeared with 5 M of H2SO4. The study provided a molecular-level understanding for

previous laboratory studies showing minerals particles coated with sulfuric acid are

relatively poor ice nuclei. The observed phenomenon was explained by a combined effect

of the decreased mica surface potential at low pH, the adsorption of sulfates on mica, and

the lack of free water molecules in high concentration of acidic solution.

64
Chapter 6

Competitive Adsorption of Toluene and n-

Alkanes at Binary Solution/Silica Interfaces*

* A version of this chapter has been published. Yang, Z.; Li, Q. F.; Hua, R.; Gray, M. R.; Chou, K. C.
Competitive Adsorption of Toluene and n-Alkanes at Binary Solution/Silica Interfaces. Journal of Physical
Chemistry C 2009, 113, 20355-20359.

65
6.1 Introduction

The competitive adsorption of hydrocarbons at liquid/mineral interfaces plays a

critical role in many industrial and environmental processes, such as oilsands

processing28,29,34,38, petroleum recovery32, contamination removal30,33,35,37, and many

extraction techniques31,36. Toluene and alkanes are particularly important because they

are the most commonly used solvents for both industrial and scientific applications. At a

binary solution/mineral interface, it is expected that the surface chemical composition is

different from the bulk composition because of the different interaction strength with the

surface.39 In many cases, the competitive adsorption of solvents at liquid/solid interfaces

is a critical factor determining the effectiveness of a technological process. In the 1950s

and 1960s, the adsorption isotherms for binary mixtures at liquid/solid interfaces were

studied by various immersion methods, and a number of theories were developed.40,41

Despite this effort, the problem was not completely resolved42, because the macroscopic

measurements were indirect, and the theories require molecular-level information about

adsorbates as input parameters.41 Even with modern technologies, it remains challenging

to directly measure the surface coverage of a particular component at a liquid/solid

interface. Recent developments in IR-visible sum frequency generation (SFG) vibrational

spectroscopy have shown that SFG is an effective technique to obtain molecular-level

information at buried liquid interfaces, and many studies have been done at water

interfaces.8 Compared with water interfaces, little is known about solvent/solid

interfaces.153

SFG vibrational spectroscopy has been used for a broad range of studies by a

rapidly growing number of research groups.20,51,154-156 As a second-order nonlinear

66
optical process, SFG is forbidden in centrosymmetric media, such as liquids, but the

symmetry is broken at an interface. For this reason, SFG is highly surface-specific and

capable of measuring surface vibrational spectra under ambient conditions. For liquid

surfaces, it has been shown that SFG from a water surface is dominated by the top

monolayer.157 With the monolayer sensitivity and short probing depth, SFG provides a

new opportunity to directly measure the adsorption isotherms at solvent/mineral

interfaces. In this chapter, SFG vibrational spectroscopy was used to study the

competitive adsorption of toluene and alkanes on silica, which is one of the most

abundant minerals and has also been widely used in liquid chromatography101-104.

The adsorption process at a liquid/solid interface is significantly different from

that at a gas/solid interface because there is no empty site at a liquid/solid interface. A

change in the bulk composition results in the replacement of one component by the other

component. The Langmuir isotherm has often been used to describe the adsorption

process for dilute solutions at liquid/solid interfaces153,158-163, but it is a good

approximation only for strongly adsorbed molecules. In the current study, the adsorption

free energies for the solvents are of the same order of magnitude. Therefore, the

Langmuir isotherm is not suitable for the current study. In this paper, we obtained the

surface coverage of toluene using SFG and used the Everett isotherm to fit the measured

adsorption isotherm over the complete toluene fraction range.

67
6.2 Experimental Section

The visible and IR laser beams for SFG vibrational spectroscopy were obtained

from a Nd:YAG (yttrium aluminum garnet) laser with an output wavelength of 1064 nm

(30 ps, 40 mJ/pulse, and 10 Hz). The laser was used to generate a second harmonic beam

at 532 nm in a KTiOPO4 (KTP) crystal. The tunable IR beam was produced by difference

frequency mixing of the 1064 nm beam with the output of a KTP optical parametric

generator/amplifier pumped by the 532 nm beam. The 532 nm and IR beams were

overlapped, both spatially and temporally, on the sample, as shown in Figure 6.1. The

energy of the laser beams were ~ 200 J/pulse for both the visible and IR beams. The

SFG intensity was detected by a photomultiplier tube and normalized against that from a

z-cut quartz. Each spectrum shown in the current study was an average of four scans in a

5 cm-1 step, and each scan was obtained by averaging the SFG intensity of 40 laser shots

at each step.

Figure 6.1. Schematic layout of the spectroscopic setup. The frequency of the visible

beam was fixed at 532 nm, and the frequency of the IR beam was tunable. The 532 nm

and IR beams were overlapped both spatially and temporally on the top surface of the

solution. The thickness of the solvent layer is 3 mm.

68
Fused silica plates, with a thickness of 3 mm, were cleaned with a commercial

cleaning agent (extran AP12) for 3 min. Then, they were immersed in a 50/50 (v/v)

HNO3/H2SO4 solution for ~12 hours, followed by rinsing in pure water (resistivity > 18.2

Mcm) and finally dried at 100C for 2 hours to remove residual surface water. After

these treatments the silica plates were kept in toluene to prevent further water adsorption

on the surface. Toluene, pentane, heptane and tetradecane (Fisher; HPLC grade) were

used as received to prepare mixtures with different volume fractions. The SFG spectrum

of the pure toluene/silica interface was monitored at the beginning and the end of the

experiment for each toluene-alkane mixture to ensure that the sample had stayed

consistent during the experimental period.

After the pure toluene/silica SFG spectrum was measured, the SFG spectra of a

series of toluene-heptane mixtures with toluene volume fraction toluene


B
= 0.8, 0.6, 0.4, 0.2,

and 0 were measured in the sequence from the highest toluene fraction to the lowest. For

each binary mixture with a particular toluene fraction, the cell and silica plates were

rinsed thoroughly with the mixture before the spectroscopic measurement. The rinsing

process ensured that the bulk mixture in the cell had the intended toluene fraction. For

each toluene fraction, four scans were collected in a period of 30 minutes during which

no change of the SFG spectrum was observed. Then the cell and silica plates were

cleaned with acids as described above for experiments with a different alkane. All spectra

were taken at room temperature.

69
6.3 Results and Analysis

The SFG vibrational spectra of toluene/silica interfaces in ssp (SFG, visible, and

IR polarizations are s-, s-, and p-polarized, respectively) and ppp are shown in Figure

6.2A. The peaks are assigned as follows: 2860 cm-1 and 2875 cm-1 to the

combination/overtone modes, 2920 cm-1 to the symmetric stretch of the CH3, 3022 cm-1

to the 20a CH stretching mode of the phenyl group, and 3075 cm-1 to the 2 CH stretching

mode of the phenyl group.164 Previously, Hommel et al. have observed a peak at 2945 cm-
1
at air/toluene surface using SFG and assigned the peak to the CH3 asymmetric mode.

However, the ssp and ppp intensity ratio of the 2945 cm-1 peak in Figure 6.2A is not

consistent with the CH3 asymmetric mode. Assuming the orientation distribution of the

CH3 groups is a delta function, the ratio of the second order nonlinear susceptibility in ssp

( 2)
ssp
and ppp configurations can be calculated as a function of the CH3 tilting angle52,
(ppp
2)

and the results are shown in Figure 6.2B. (The detailed calculation is available in the

( 2)
ssp
supporting materials.) For the CH3 asymmetric mode, the calculated ratio is always
(ppp
2)

less than 1, even with a finite distribution width, but the spectra in Figure 6.2A indicates

( 2)
ssp
a ratio of ( 2)
~ 3. Therefore, the 2945 cm-1 peak could be a Fermi resonance associated
ppp

with the symmetric mode, instead of an asymmetric mode. For the CH3 symmetric mode,

( 2)
ssp
the measured ratio is ~ 4.4, which corresponds to a tilting angle of 25 with respect
(ppp
2)

to the surface normal, as indicated by the solid circle and the dotted lines in Figure 6.2B.

This tilting angle is consistent with the molecular dynamic simulation for adsorption of

70
toluene on silica, showing the plane of the phenyl ring mostly adopts an upright geometry

with a tilting angle of about 30 with respect to the surface normal because of the

interaction of its electrons with the silica surface.165

Figure 6.2. (A) SFG vibrational spectra of toluene in ssp and ppp polarization

configurations. The ssp and ppp spectra are offset from each other by 0.5 arbitrary units

( 2)
ssp
for clarity. (B) Calculated for CH3 symmetric (solid line) and asymmetric (dashed
(ppp
2)

line) modes as a function of the tilting angle. The orientational distribution of the CH3

( 2)
ssp
groups was assumed to be a delta function. The solid circle indicates the measured
(ppp
2)

value of 4.4 which corresponds to a tilting angle of 25.

71
Figure 6.3a-f show the ssp SFG spectra of toluene-pentane mixtures with the bulk

toluene volume fraction toluene


B
= 1, 0.8, 0.6, 0.4, 0.2, and 0, respectively. Previously,

Selfer et al. studied the adsorption of alkanes on silica and showed that hexadecane lies

flat on the silica surface.22 When the axis of the CH3 group is along the surface, the

asymmetric peak will dominate, and the symmetric peak will be missing. If the axis of

the CH3 group is along the surface normal, the situation will be reversed. The spectrum of

pure pentane on silica in Figure 6.2f shows two peaks at 2857 cm-1 and 2951 cm-1, which

are consistent with the CH2 symmetric and CH3 asymmetric modes, respectively.166

Figure 6.3a-f show that the peak intensities of the toluene 20a and 2 modes decrease as

toluene
B
decreases. The spectra shown in Figure 6.3 were collected in the sequence from

the highest toluene volume fraction ( toluene


B
= 1) to the lowest ( toluene
B
= 0). Within the

measurement error, similar spectra were obtained for experiments carried out in a

reversed order. Therefore, the competitive adsorption process is reversible, and the

surface composition depends only on the bulk mole fraction of toluene, not on the history

of the system.

Experiments were repeated for toluene-heptane and toluene-tetradecane binary

mixtures on silica, as shown in Figures 6.4 and 6.5, respectively. While the spectrum of

pure heptane on silica (Figure 6.4f) is similar to that of pure pentane on silica in Figure

6.3f, significant CH3 symmetric peak at 2875 cm-1 and CH2 asymmetric peak at 2920 cm-
1
were observed at tetradecane/silica interfaces.166 Therefore, there is an increasing

conformational disorder for longer alkane chains.

72
Figure 6.3. SFG vibrational spectra of toluene-pentane mixtures on silica with toluene

volume fraction toluene


B
= (a) 1, (b) 0.8, (c) 0.6, (d) 0.4, (e) 0.2, and (f) 0.

73
Figure 6.4. SFG vibrational spectra of toluene-heptane mixtures on silica with toluene

volume fraction toluene


B
= (a) 1, (b) 0.8, (c) 0.6, (d) 0.4, (e) 0.2, and (f) 0.

74
Figure 6.5. SFG vibrational spectra of toluene-tetradecane mixtures on silica with toluene

volume fraction toluene


B
= (a) 1, (b) 0.8, (c) 0.6, (d) 0.4, (e) 0.2, and (f) 0.

75
Since the spectra of both toluene and alkanes have CH peaks, the CH peaks are

not good indicators for the adsorbed chemical species on silica. On the other hand, the

20a and 2 peaks from the phenyl group are the unique signature of toluene and allow us

to quantitatively study toluene absorbed on the silica surface. Because the intensity of the

20a peak is higher than that of the 2 peak, the following analysis will focused on the 20a

peak to obtain the absorption isotherm of toluene on silica.

To obtain quantitative information, further theoretical analysis of the SFG spectra

is required. The detailed theoretical background of SFG can be found in the references.8

Briefly, the SFG intensity is given by

I ( SFG ) [L( SFG ) e( SFG )] ( 2 ) : [L(vis ) e(vis )][L( IR ) e( IR )] I (vis ) I ( IR )


2

(6.1)

where L(i ) is the tensorial Fresnel coefficient, and e(i ) is the unit polarization vector,

( 2 ) is the surface nonlinear susceptibility tensor, and I (vis ) and I ( IR ) are the

intensities of incident visible and IR beams, respectively. The surface nonlinear

(2)
susceptibility ijk can be expressed as

Aq ,ijk
ijk( 2 ) = NR ,ijk +
(2)
(6.2)
q IR q +i q

where NR
(2)
,ijk describes the nonresonant contribution, and Aq ,ijk , q , and q are the

amplitude, frequency, and damping constant of the qth vibrational mode, respectively.

The amplitude Aq ,ijk in the lab coordinate is related to the molecular hyperpolarizability

q,lmn in the molecular coordinates:

76
^ ^ ^ ^ ^ ^
Aq ,ijk = n S
l ,m ,n
q ,lmn (i l )( j m )( k n ) (6.3)

where n S is the surface number density, and the angular brackets refers to an average

over the molecular orientation. If the orientation of the molecule on a surface is not

strongly dependent on the coverage, the amplitude Aq ,ijk is proportional to the surface

density n S . In the current study, the orientation of toluene was verified using the ratio of

Appp
the 20a peak in ssp and ppp configurations.153 The amplitude ratio was found to be
Assp

~ 0.42 and independent of the toluene coverage. Therefore, it is feasible to correlate the

amplitude of the 20a peak to the surface number density of toluene on silica. The SFG

spectra in Figures 6.3-6.5 were fitted using Equation (6.1) and (6.2) to obtain the

amplitude Aq ,ijk .

Calculating the Fresnel coefficients described in Equation (6.1) requires the

refractive index of the mixture nmix . In general, nmix changes with the mixture

composition. The refractive index of a mixture follows a mixture rule167

nmix = 1n1 + 2 n2 (6.4)

where i and ni are the volume fraction and the refractive index of component i ,

respectively. Although small deviations from Equation (6.4) have been reported,168 the

small deviations are insignificant for the current study. The refractive indices for toluene,

pentane, heptane and tetradecane are 1.4963, 1.357, 1.38, and 1.428, respectively.169 The

amplitude Aq ,ijk was then calibrated using Equation (6.1) and (6.4) so that Aq ,ijk is

nS
proportional to the surface number density n s . The toluene surface coverage c S
,
nmax

77
S
where nmax is the maximum density adsorbed on the surface when only toluene is

presented in the solution, was then derived using the amplitude of the 20a peak. Figure

6.6 shows the toluene surface coverage as a function of the bulk mole fraction.

Figure 6.6. Adsorption isotherms of toluene on silica for binary mixtures of pentane

toluene (), heptanetoluene (), and tetradecanetoluene (). The solid curves are

fitting curves using Equation (6.8).

As shown in Figure 6.6, toluene competes favorably against pentane but the

advantage decreases as the chain length of alkane increases. This is consistent with the

conclusion that alkanes on average lie flat on the silica surface. In this geometry, the

molar adsorption energy of alkane increases as the chain length increases. To gain better

insight into the competitive adsorption process, a theoretical model is needed. The well-

known Langmuir equation is not a good description for the adsorption isotherm at

liquid/solid interface because there is no empty site at a liquid/solid interface. A number

78
of theories have been developed for the adsorption isotherm at binary liquid/solid

interfaces over the complete mole fraction range.41,42,170-173 It has been shown that, for

adsorption on a homogeneous surface from an ideal miscible binary liquid with

component 1 and 2, the surface mole fraction of component 1 can be written as42

K1 x1B
x1S = (6.5)
1 + ( K1 1) x1B

with

( 1S 1B ) ( 2S 2B ) a 1 a 2
K1 = exp = exp (6.6)
RT RT

where x1B and x 2B are the bulk mole fractions for component 1 and 2, respectively, x1S

and x 2S are the surface mole fractions for component 1 and 2, respectively, R is the gas

constant, T is the temperature, iB and iS are the chemical potential (or partial molar

Gibbs free energy) of component i in its standard state for the bulk and surface,

respectively, and a i iS iB is the chemical potential change associated with the

adsorption of component i . The above expression was derived by Everett171, and it is

often called the Everett isotherm.

If component 1 is a molecule strongly adsorbed on the surface, or

( a 1 a 2 ) >> RT , one gets K 1 >> 1 . In this case, Equation (6.5) can be

approximated by

K1 x1B
x1S = (6.7)
1 + K1 x1B

This expression is analogous to the Langmuir equation. However, the Langmuir equation

is a good approximation only for a strongly adsorbed component. As the adsorption free

79
energies of toluene and alkanes are comparable, this approximation is not valid for the

current study. The Everett isotherm will be used in the following analysis.

The amplitude of SFG peaks, as described in Equation (6.3), measures the surface

number density n S , instead of the surface mole fraction x1S shown in Equation (6.5).

Therefore, it is desirable to express Equation (6.5) in terms of surface coverage c .

Equation (6.5) can be rewritten as42

n1S K1 x1B
c = (6.8)
S
nmax,1 1 + ( K1 1) x1B

with

S
n max,
S
2
(6.9)
n max,1

Equation (6.8) and (6.9) indicate that the adsorption isotherm is governed by the values of

and K1 . The value of describes the relative footprint on the surface for component 1

and 2, and the value of K1 is determined by the adsorption free energy difference

between the two components.

The Everett isotherm as described in Equation (6.8) allows a more quantitative

analysis for the meassured isotherms in Figure 6.6. The values of and K1 are coupled

in Equation (6.8). Without additional information, the measured adsorption isotherms in

Figure 6.6 are not sensitive to the individual value of and K1 . Using K1 as a single

parameter to fit the measured adsorption isotherms in Figure 6.6, we obtained the best fits

with K1 = 1.69, 0.659, and 0.296 for pentane, heptane and tetradecane mixtures,

respectively. To estimate K1 , further assumptions must be made for . The value of ,

80
S
n max,
defined as S
2
, is mainly determined by the sizes of the molecules because a
n max,1

S
larger molecule has a larger footprint on the surface and a smaller value of n max . The

space that a molecule occupies can be estimated using the molar volume. The molar

volumes Vm of toluene, pentane, heptane and tetradecane are 106.29, 115.26, 146.51, and

2/3
V
260.3 mL/mol. To a first approximation, the value of should scale as m , 2 , which
Vm ,1

corresponds to a random adsorption geometry on silica. With this assumption, the values

of are 0.95, 0.81 and 0.55 for toluene-pentane, toluene-heptane, and toluene-

tetradecane mixtures, respectively. However, molecules adsorbed on a surface have a

preferred orientation, and the values need to be corrected with an orientation factor.

As described above, the tilting angle of toluene with respect to the surface normal was

estimated at ~ 25, and the alkanes mostly lie flat on the silica surface with ~ 90.

With a preferred orientation, the average footprint of molecules on the surface scales as

sin oriented
where the angle brackets denote the orientational average
sin random

f ( ) sin d with f ( ) denoting the orientational distribution function. Assuming


f ( ) d
the orientational distribution function is a delta function with = 25 for toluene and =

90 for alkane, the values of become 0.40, 0.34 and 0.23 for toluene-pentane, toluene-

heptane, and toluene-tetradecane mixtures, respectively. Then the best fit for K1 were

obtained with a toluene a pen tan e = 3.4 0.3 kJ/mol, a toluene a hep tan e = 1.8

81
0.3 kJ/mol, and a toluene a tetradecane = 0.84 0.3 kJ/mol. The best-fit curves are

shown in Figure 6.6. In all cases, toluene competes favorably against the alkanes for the

adsorption on silica. However, the adsorption free energy of alkane increases as the chain

length increases.

82
6.4 Conclusions

IR-visible sum frequency vibrational spectroscopy was applied to study the

competitive adsorption of toluene and n-alkanes at binary solution/silica interfaces. The

surface coverage of toluene on silica for toluene-pentane, toluene-heptane, and toluene-

tetradecane mixtures were obtained using the measured SFG peaks intensity of toluene.

The competitive adsorption processes are reversible, and the surface coverage of toluene

only depends on the toluene molar fraction in the binary mixture, not on the history of the

mixture in contact with the silica. The measured adsorption isotherms fitted well with the

Everett isotherm over the complete mole fraction range. The estimated molar adsorption

free energy of toluene is 3.4 0.3 kJ/mol, 1.8 0.3 kJ/mol, and 0.84 0.3 kJ/mol higher

than pentane, heptanes, and tetradecane, respectively. Overall, toluene competes

favorably on silica against the alkanes, and the molar adsorption free energy difference

between toluene and alkane decreases as the chain length of the alkane increases.

83
Chapter 7

Effect of Interfacial Water Content on Bitumen

Liberation from Silica and Mica Surfaces*

* A version of this chapter will be submitted for publication. Yang, Z.; Bailey, G.; Gray, M. R.; Chou, K.
C. Effect of Interfacial Water Content on Bitumen Liberation from Silica and Mica Surfaces. 2011

84
7.1 Introduction

Bitumen liberation from mineral surfaces is the prerequisite for a bitumen extraction

process.44 Understanding the mechanisms of oil displacement by water on minerals is

important for oil sand recovery. Many investigations on bitumen displacement and

detachment from the solid surface in the presence of water containing salt, surfactants and

clays at different temperature and pH have been completed using dynamic contact angle

measurements.45-48,174 Dynamic contact angle measurements provide macroscopic

information such as rate of liberation, the dynamic, and the equilibrium contact angles in the

bitumen liberation processes. Previous studies have also suggested that the interfacial water

content between mineral surfaces and bitumen is one of the most important factors affecting

bitumen recovery.49 However, the detailed mechanism remains poorly understood. To study

the effect of interfacial water on bitumen liberation both macroscopic (liberation rate,

dynamic and static contact angle) and microscopic (water amount at mineral surfaces under

different relative humidity, structures of water molecules adsorbed on mineral surfaces)

information is needed. In this chapter, we investigate the effect of the interfacial water

content on bitumen liberation from silica and mica surfaces in an aqueous environment using

both sum frequency generation (SFG) vibrational spectroscopy and dynamic contact angle

measurements. The objective of the current study is to provide a molecular-level

understanding for the effect of interfacial water on the liberation process.

85
7.2 Experimental Section

7.2.1. Sum frequency generation spectroscopy measurements

The visible and tunable IR laser beams for SFG vibrational spectroscopy were

obtained from a Nd:YAG (yttrium aluminum garnet) laser with output wavelength of 1064

nm (30 ps, 40 mJ/pulse, and 10 Hz). The laser was used to generate a second harmonic beam

at 532 nm in a KTiOPO4 (KTP) crystal. The tunable IR beam was produced by difference

frequency mixing of the 1064 nm beam with the output of a KTP optical parametric

generator/amplifier (OPG/OPA) pumped by the 532 nm beam. The 532 nm and IR beams

were overlapped, both spatially and temporally, on the silica/vapor interface, as shown in

Figure 7.1. The laser fluence was approximately 2 mJ/cm2 per pulse for the visible beam and

5 mJ/cm2 per pulse for the IR beam. The polarizations of the beams were s-, s-, and p-

polarized for SFG, visible, and IR, respectively. The SFG intensity was detected by a

photomultiplier tube after spatial filtering by an aperture, and spectral filtering by a bandpass

filter. The SFG intensity was normalized against that from a z-cut quartz. Each spectrum

shown in the current study is an average of 5 scans in a 10 cm-1 step, and a scan represents an

average of 50 laser shots per step.

Fused silica plates, with a thickness of 3mm, were cleaned thoroughly with a

commercial cleaning agent (extran AP12). It was then immersed in a mixture of sulfuric acid

(98%) and nochromix reagent (GODAX Laboratories, Inc.) for 24 hours, followed by rinsing

in pure water (resistivity > 18.2 Mcm, Millipore) and finally drying at 100C for 2h. Then it

was put on the top of a chamber which was made of Teflon and cleaned thoroughly with a

50/50 (v/v) HNO3/H2SO4 mixture. The humidity in the chamber was controlled by adjusting

the flow rate of N2 gas bubbling through a container filled with water and was measured with

86
a hygrometer (model: Omega RH82). The humidity was adjusted from 1% relative humidity

(RH) to 99% RH with an accuracy of 1% RH.

Figure 7.1. Schematic layout of the spectroscopic setup. The frequency of the visible beam

was fixed at 532 nm, and the frequency of the IR beam was tunable. The 532 nm and IR

beams were overlapped both spatially and temporally at the bottom surface of silica plate.

7.2.2. Dynamic and equilibrium contact angle measurements

The dynamic and equilibrium contact angle measurements were conducted on three

different substrates: silica plate, rinsed muscovite mica sheet and freshly cleaved muscovite

mica sheet. The silica plate was cleaned in the same way as described above. The mica sheet

was cleaved on both sides, followed by rinsing in pure water (resistivity > 18.2 Mcm,

Millipore) and blow-drying with nitrogen gas. Experiments were performed on freshly

cleaved muscovite mica sheet, while the mica sheet was just cleaved on both sides. In each

contact angle measurement the substrate was placed into a glass container with a top cover

for 30 min. The relative humidity in the container was also adjusted in the same way as

above. After that 11 drops (~0.3ml) of diluted bitumen (Vtoluene/Vbitume = 3/7) were placed on

the surface of substrate and kept in the glass container for 2 min to let the three phases

(substrate, diluted bitumen and vapor) reach equilibrium. The bitumen became a thin layer

and formed a disk. Then the substrate with a bitumen layer on the top was quickly placed

87
into a chamber of water. The temperature of water in the chamber was maintained at 25 0.5

C by a hot plate for all experiments. And the depth of water was always kept at 1.5 cm. The

bitumen displaced by water contracted uniformly along the inward radial direction and

finally the thin layer of bitumen became a droplet and the equilibrium condition was reached.

For all RH values, the bitumen droplet remained attached on the silica plate and rinsed mica

sheet, but left the surface of freshly cleaved mica sheet immediately. The whole recession

process was recorded by a Nikon D90 camera with a 60mm macro lens.

88
7.3 Results and Analysis

7.3.1. Sum frequency generation spectroscopy experiments

Figure 7.2 shows the SFG vibrational spectra of the water molecules on silica under

various degrees of relative humidity. The spectrum is similar to the SFG spectrum of the

water/air interface and it exhibits three peaks at 3200cm-1, 3400cm-1 and 3700cm-1.10 The

two peaks at 3200cm-1 and 3400cm-1 are associated with an ordered and disordered

hydrogen-bonding network. They have been called ice-like and water-like peaks

respectively because ice has an IR adsorption at 3200cm-1 and water has an IR adsorption at

3400cm-1.10,20,80 The sharp peak at 3700cm-1 corresponds to the O-H stretch of the non-

hydrogen-bonded O-H groups of water which is also known as the free O-H stretch.10 The

SFG intensities of ice-like and water-like peaks are very small in the spectrum with 1% RH

and is indicative of few water molecules adsorbed on the silica surface at 1% RH. However,

there is a relatively large and sharp peak at 3750cm-1. This peak originates from the O-H

stretch of SiOH groups on silica surface.175 At 28% RH O-H stretch modes associated with

the hydrogen-bonded O-H groups can be observed in the spectrum. The peak of the SiOH O-

H stretch becomes weaker because of the formation of hydrogen bonds between SiOH and

adsorbed water molecules. Another sharp peak at 3700cm-1 corresponding to the free O-H

stretch of water begins to emerge. As the relative humidity further increases, the intensity of

the water peaks becomes larger, indicating that more water molecules are adsorbing onto the

silica surfaces. This result is similar to previous SFG experiments done on the mica/D2O

vapour interface except that a 2D water structure formed on the mica surface at 90% RH.114

89
Figure 7.2. SFG vibrational spectra of water adsorbed on silica surface as a function of

relative humidity (RH).

7.3.2. Dynamic and equilibrium contact angle study

Figure 7.3 shows a sequence of frames extracted from a video. The pictures show the

shape of a bitumen droplet on the silica plate when exposed to water. The recorded videos

have 24 frames per second. Each frame was extracted and analyzed with a home-made

Matlab program to obtain the dynamic contact angle. The dynamic contact angles were

plotted as a function of time, as shown in Figure 7.4. For all videos, the time axis has been

reset as zero when the contact angle equals to 60.

90
Figure 7.3. Sequential images extracted from a video of bitumen liberation from silica at

time equal to (a) 0 , (b) 20, (c) 80 sec.

Figure 7.4 (a) shows the dynamic contact angle for bitumen on the silica surface as a

function of time with various relative humidity. In general, the dynamic contact angles

increase with time until they reach equilibrium values, which are the static contact angles.

The value of the static contact angle is an indication of the tendency for the bitumen droplet

to detach from the mineral surface. As shown in Figure 7.4 (a), when RH equals to 0%, 25%

and 46%, the changing rate of dynamic contact angle is basically the same. However, when

the RH increased to 70%, the changing rate was significantly larger. When the RH increased

to 90%, the changing rate was very similar to that at RH = 70%. It can also be observed that

under different RH the static contact angles of bitumen on the silica surface are different.

The static contact angle is 115 when RH is equal to 0%, 25% and 46%, while it is 122

when RH is equal to 70% and 90%. The results suggest that the increase of water content on

the silica surface caused an increase in the changing rate of dynamic contact angle and an

increase of static contact angle of bitumen, although the effect of water content is small for

RH up to 46%.

91
(a)

(b)

(c)

Figure 7.4 Measured dynamic contact angles for bitumen on (a) silica, (b) rinsed mica, and

(c) cleaved mica.

92
Figure 7.4 (b) shows the dynamic contact angle for bitumen on the rinsed mica

surface as a function of time with various relative humidity. Similarly, both the changing rate

of dynamic contact angle of bitumen and the static contact angle of bitumen on rinsed mica

surface increased with RH. However, the time scale is much shorter compared to the silica

surface. Furthermore, the static contact angles on the rinsed mica surface are larger

compared to silica. A larger static contact angle indicates a higher tendency for bitumen to

detach from the surface. On the rinsed mica surface, the increase of changing rate of

dynamic contact angle due to the increase of interfacial water content was obvious at RH =

50%, unlike the silica experiment in which this increase was observed when RH is as high as

70%.

Figure 7.4 (c) shows the time dependence of the dynamic contact angle of bitumen on

the freshly cleaved mica surface with various RH. The dynamic contact angles reached

equilibrium in several seconds, which is an order of magnitude faster than those on rinsed

mica. As the RH increases, it is observed that the changing rates of dynamic contact angle

have little difference, and so do the static contact angles.

Based on the experimental data, it can be concluded that more water content on the

silica and rinsed mica surface help the liberation of bitumen and increased the rate of the

process. However, the properties of the mineral surface play a more critical role in the

liberation process. The changing rate of dynamic contact angle showed an order-of-

magnitude difference for mica and silica. On three different mineral surfaces the changing

rate of dynamic contact angle of bitumen is: cleaved mica > rinsed mica > silica.

To explain the huge differences among the changing rates of dynamic contact angle

on three different minerals, the surface charge of the mineral surfaces and the formation of

93
hydrogen bond between the mineral surfaces and bitumen should be considered. On one

hand, bitumen does not carry any net charge, but if there are some net charges on the mineral

surfaces, then the attractive electrostatic force between bitumen and the mineral surfaces will

occur. The attractive force will slow down the liberation of bitumen from the mineral

surfaces when they are exposed to an aqueous environment. On the other hand, as a complex

mixture, bitumen bears various types of natural surfactants on its surface, such as R-NH2 and

R-COOH.176,177 These functional groups can form hydrogen bonds with the mineral

surfaces,178 and the hydrogen bonds will also decrease the liberation rate of bitumen in water.

So the larger the surface charge density is on mineral surfaces and the more hydrogen bonds

can form between bitumen and mineral surfaces, the slower the liberation of bitumen will be.

The functional groups on the silica surface are basically silanol groups and the

surface density of silanol groups is estimated to be 5 1014 cm-2.76,83 These silanol groups

have two forms: -SiOH and SiO-. The SiO- groups create a charge density of ~0.05 e- per

nm2 at pH = 6. Both the -SiOH and SiO- can form hydrogen bond with the functional groups

on bitumen surface. The surface properties of mica are different from silica. The negative

surface charges resulted from O- are neutralized by K+ after the mica is cleaved. However,

after rinsing with water, 1% of K+ on the mica surface dissolves and the mica surface will

carry negative charges with a density of ~0.015 e- per nm2 which is smaller than that on silica

surface.179 These O- groups can also form hydrogen bonds with bitumen, but the density of

hydrogen bonds between mica and bitumen is smaller than that between silica and bitumen.

So this explains why bitumen on the mica surface has a larger liberation rate compared to

silica. As for freshly cleaved mica, there is neither surface charge on the surface nor

hydrogen bond formation between mica and bitumen, thus the force between cleaved mica

94
and bitumen should be the smallest and it explains why bitumen on mica has the largest

liberation rate in our experiments performed on three different mineral surfaces.

The surface charge density and hydrogen bond formation can also explain the

phenomenon that water content on silica and rinsed mica surfaces increases the bitumen

dynamic contact angle changing rate. In the silica experiment, for example, silica has a

negative surface charge and the silanol groups can form hydrogen bonds with bitumen. At

RH = 0% bitumen contacts the silica surface directly, so the attractive electrostatic force

should be largest and the number of hydrogen bonds between bitumen and silanol groups

should also be largest. As the RH increases, the water content on the silica surface increases.

The negative charge on the silica surface will be partially shielded by water molecules so the

attractive force will decrease. Because of the existence of water molecules on the silica

surface, bitumen interact less with the silanol groups. Therefore, the number of hydrogen

bonds will also decrease. Consequently, the water content on the silica surface will help the

liberation of bitumen. The explanation for the rinsed mica experiments is similar to that for

silica. For the cleaved mica surface, there is no surface charge on the surface, and no

hydrogen bond formation between cleaved mica and bitumen. Therefore, the water content

has little effect on the bitumen liberation rate, and this is consistent with what we observed in

the experiment.

95
7.4 Conclusions

SFG vibrational spectroscopy and dynamic contact angle measurements experiments

were applied to study the effect of interfacial water content on bitumen liberation from

mineral surfaces in water. The time dependence of the dynamic contact angle of bitumen on

the silica, freshly cleaved mica, and rinsed mica were compared with different water content

on mineral surfaces. The bitumen liberation rate increases with higher water content on the

mineral surfaces for silica and rinsed mica. At the same water content, the rate of bitumen

liberation on different mineral surfaces is: freshly cleaved mica > rinsed mica > silica. To

explain the differences in bitumen liberation on the three minerals and to understand the role

of water content on mineral surfaces on bitumen liberation, the surface charge density of the

mineral surfaces and the formation of hydrogen bond between the mineral surfaces and

bitumen were considered.

96
Chapter 8

Conclusion

97
This thesis presents studies of petrolic, edaphic and atmospheric relevant

liquid/mineral interfaces using SFG, which include water structures at aqueous salt

solutions/silica interfaces, solvent/silica interfaces, and sulphuric acid solution/mica

interfaces, effects of interfacial water content on bitumen liberation from silica and mica

surfaces, and competitive adsorption of toluene and n-alkanes on silica interfaces.

The cations play a key role in perturbing the hydrogen-bond network at the

water/silica interfaces as they interact with the silica surface via electrostatic interaction

(Chapter 3). Significant perturbations of the interfacial water structures were observed

with a 1x10-4 M NaCl solution. Different alkali cation species produce different degrees

of perturbation in the order: K+ > Li+ > Na+. This order can be explained by considering

the electrostatic interaction between the cations and silica surfaces and the effective ionic

radii of the cations. Our current SFG setup cannot provide the information on the

orientation of water molecules. Phase-sensitive SFG vibrational spectroscopy is needed

for further experiments to study how the orientation of water molecules responds to the

concentration changes of alkali cations.

Water molecules adsorbed at the toluene/silica interface form a highly H-bonded

layer with no detectable free OHs (Chapter 4). The water layer without free OHs showed

resistance against further adsorption of water molecules. However, this special water

structure was not observed at heptane/silica interfaces, at which free OHs were always

observed. The experimental data showed that the interactions between solvents and water

molecules can significantly change the interfacial water properties. However, the reason

why toluene aided the formation of water-resistant structure is still unclear. More

98
experiments with different solvents and minerals may reveal the mechanism for the

formation of this water-resistant water layer.

The structures of water on the mica surfaces in the presence of H2SO4 with

atmospheric relevant concentrations were studied (Chapter 5). We found that the ordered

water structure disappeared when the concentration of H2SO4 reached 5 mol/L. The

observed phenomenon was a combined effect of the decreased mica surface potential at

low pH, the adsorption of sulfates on mica, and the lack of free water molecules in highly

concentrated acidic solution. The results have offered a microscopic understanding for

why H2SO4-coated mineral surfaces are poor ice nuclei. The good ice nucleation ability

above liquid water saturation is correlated with the presence of structured water

suggesting that structured water at the interface may be needed for efficient

heterogeneous ice nucleation. More experiments could be done to confirm this

explanation. Water structures on efficient ice nuclei minerals (such as kaolinite and

muscovite) and poor ice nuclei minerals (such as quartz and calcite) can be studied by

SFG. If the proposed mechanism is correct, water will form ordered structure on efficient

ice nuclei minerals, while disordered water structure will be found on poor ice nuclei

minerals. However, some minerals are not flat or transparent. In this case, SFG signal

will be scattered, and the signal lost could be a potential issue.

Besides water/solid interfaces, competitive adsorption of toluene and n-alkanes at

binary solution/silica interfaces was studied (Chapter 6). The surface coverage of toluene

on silica for toluene-pentane, toluene-heptane, and toluene-tetradecane mixtures was

obtained using the measured SFG peaks intensity of toluene. The competitive adsorption

processes are reversible, and the surface coverage of toluene only depends on the toluene

99
molar fraction in the binary mixture, not on the history of the mixture in contact with the

silica. The measured adsorption isotherms fitted well with the Everett isotherm over the

complete mole fraction range. Overall, toluene competes favorably on silica against the

alkanes, and the molar adsorption free energy difference between toluene and alkane

decreases as the chain length of the alkane increases. Theoretically the surface coverage

of any molecule on a solid surface accessible by light can be measured with a similar

method as long as the molecule has a unique vibration peak. In reality most organic

molecules have similar vibration peaks. However, if the target molecule can be

deuterated, this method will still work.

SFG and dynamic contact angle measurements experiments were applied to study

the effect of relative humidity (RH) on bitumen displacement by an aqueous phase on

mineral surfaces (Chapter 7). The time variation of the dynamic contact angle of bitumen

on the silica plate, freshly cleaved mica pieces, rinsed mica pieces were compared with

different water content on mineral surfaces. The bitumen displacement rate increases with

higher water content on mineral surfaces. At the same water content, the rate of bitumen

displacement on different mineral surfaces is: freshly cleaved mica > rinsed mica > silica.

To explain the differences in bitumen displacement on the three minerals and to

understand the role of water content on bitumen displacement on mineral surfaces, the

surface charge density of mineral surfaces and the formation of hydrogen bonds between

the mineral surfaces and bitumen should be considered. In our current setup the visible

beam is 532 nm which can stimulate our bitumen samples to give off strong fluorescence.

Therefore, it is impossible to detect the SFG signal from water molecules between

minerals and bitumen. In the future study, the visible beam with a longer wave length can

100
be used to minimize the fluorescence from bitumen, then information of water molecules

between minerals and bitumen can be gathered.

101
References
(1) Adamson, A. W.; Gast, A. P. Physical chemistry of surfaces, 6th ed.;
Wiley: New York, 1997.
(2) Evans, R.; Marconi, U. M. B. J. Chem. Phys. 1987, 86, 7138.
(3) Israelachvili, J.; Wennerstrom, H. Nature 1996, 379, 219.
(4) Cerda, J.; Michaelides, A.; Bocquet, M. L.; Feibelman, P. J.; Mitsui, T.;
Rose, M.; Fomin, E.; Salmeron, M. Phys. Rev. Lett. 2004, 93.
(5) Mitsui, T.; Rose, M. K.; Fomin, E.; Ogletree, D. F.; Salmeron, M. Science
2002, 297, 1850.
(6) Tyrrell, J. W. G.; Attard, P. Phys. Rev. Lett. 2001, 87.
(7) Shen, Y. R. Nature 1989, 337, 519.
(8) Miranda, P. B.; Shen, Y. R. J. Phys. Chem. B 1999, 103, 3292.
(9) Eisenthal, K. B. Chem. Rev. 1996, 96, 1343.
(10) Richmond, G. L. Chem. Rev. 2002, 102, 2693.
(11) Du, Q.; Freysz, E.; Shen, Y. R. Science 1994, 264, 826.
(12) Fisk, J. D.; Batten, R.; Jones, G.; O'Reilly, J. P.; Shaw, A. M. J. Phys.
Chem. B 2005, 109, 14475.
(13) Gurau, M. C.; Kim, G.; Lim, S. M.; Albertorio, F.; Fleisher, H. C.; Cremer,
P. S. Chemphyschem 2003, 4, 1231.
(14) Hassanali, A. A.; Singer, S. J. J. Phys. Chem. B 2007, 111, 11181.
(15) Helms, C. R.; Deal, B. E.; Electrochemical Society. Electronics Division.;
Electrochemical Society. Dielectric Science and Technology Division. The Physics and
chemistry of SiO b2 s and the Si-SiO b2 s interface 2; Plenum Press: New York, 1993.
(16) Kim, J.; Kim, G.; Cremer, P. S. Langmuir 2001, 17, 7255.
(17) Kim, J.; Kim, G.; Cremer, P. S. J. Am. Chem. Soc. 2002, 124, 8751.
(18) Konek, C. T.; Musorrafiti, M. J.; Al-Abadleh, H. A.; Bertin, P. A.; Nguyen,
S. T.; Geiger, F. M. J. Am. Chem. Soc. 2004, 126, 11754.
(19) O'Reilly, J. P.; Butts, C. P.; I'Anson, I. A.; Shaw, A. M. J. Am. Chem. Soc.
2005, 127, 1632.
(20) Shen, Y. R.; Ostroverkhov, V. Chem. Rev. 2006, 106, 1140.
(21) Blokzijl, W.; Engberts, J. B. F. N. Angew. Chem. Int. Ed. 1993, 32, 1545.
(22) Sefler, G. A.; Du, Q.; Miranda, P. B.; Shen, Y. R. Chem. Phys. Lett. 1995,
235, 347.
(23) Miranda, P. B.; Pflumio, V.; Saijo, H.; Shen, Y. R. J. Am. Chem. Soc.
1998, 120, 12092.
(24) Motokura, K.; Tada, M.; Iwasawa, Y. J. Am. Chem. Soc. 2007, 129, 9540.
(25) Brindza, M. R.; Walker, R. A. J. Am. Chem. Soc. 2009, 131, 6207.
(26) Brindza, M. R.; Ding, F.; Fourkas, J. T.; Walker, R. A. J. Chem. Phys.
2010, 132.
(27) Steel, W. H.; Walker, R. A. Nature 2003, 424, 296.
(28) Acevedo, S.; Ranaudo, M. A.; Garcia, C.; Castillo, J.; Fernandez, A.
Energy Fuels 2003, 17, 257.
(29) Arabshahi, S. H.; Ackerson, M. D.; Rye, W. C.; Babcock, R. E. Chem.
Eng. Commun. 1990, 89, 195.

102
(30) Grandel, S.; Dahmke, A. Biodegradation 2004, 15, 371.
(31) Hubert, A.; Wenzel, K. D.; Manz, M.; Weissflog, L.; Engewald, W.;
Schuurmann, G. Anal. Chem. 2000, 72, 1294.
(32) Kenchington, J. M.; Phillips, C. R. Energy Sources 1981, 5, 317.
(33) Khaitan, S.; Kalainesan, S.; Erickson, L. E.; Kulakow, P.; Martin, S.;
Karthikeyan, R.; Hutchinson, S. L. L.; Davis, L. C.; Illangasekare, T. H.; Ng'oma, C.
Environ. Prog. 2006, 25, 20.
(34) Long, Y. C.; Dabros, T.; Hamza, H. Fuel 2004, 83, 823.
(35) Luo, Y.; Zou, L.; Hu, E. Environ. Technol. 2006, 27, 359.
(36) Saim, N.; Dean, J. R.; Abdullah, M. P.; Zakaria, Z. J. Chromatogr. A 1997,
791, 361.
(37) Scullion, J. Naturwissenschaften 2006, 93, 51.
(38) Stasiuk, E. N.; Schramm, L. L. Fuel Process. Technol. 2001, 73, 95.
(39) Fowkes, F. M. Ind. Eng. Chem 1964, 56, 40.
(40) Blackburn, A.; Kipling, J. J. J. Chem. Soc. 1954, 3819.
(41) Everett, D. H. Pure Appl. Chem. 1986, 58, 967.
(42) Oscik, J. Adsorption; Chichester: New York, 1982.
(43) Burrowes, A.; Marsh, R.; Ramdin, N.; Evans, C. Alberta Energy and
Utilities Board 2007.
(44) Clark, K. A.; Pasternak, S. D. Ind. Eng. Chem. 1932, 24, 1410.
(45) Basu, S.; Kanda, W. C.; Nandakumar, K.; Masliyah, J. H. Ind. Eng. Chem.
Res. 1998, 37, 959.
(46) Basu, S.; Nandakumar, K.; Lawrence, S.; Masliyah, J. Fuel 2004, 83, 17.
(47) Basu, S.; Nandakumar, K.; Masliyah, J. H. J. Colloid Interface Sci. 1996,
182, 82.
(48) Basu, S.; Nandakumar, K.; Masliyah, J. H. Colloids Surf., A 1998, 136, 71.
(49) Dang-Vu, T.; Jha, R.; Wu, S. Y.; Tannant, D. D.; Masliyah, J.; Xu, Z. H.
Energy Fuels 2009, 23, 2628.
(50) Zhu, X. D.; Suhr, H.; Shen, Y. R. Phys. Rev. B 1987, 35, 3047.
(51) Lambert, A. G.; Davies, P. B.; Neivandt, D. J. Appl. Spectrosc. Rev. 2005,
40, 103.
(52) Zhuang, X.; Miranda, P. B.; Kim, D.; Shen, Y. R. Phys. Rev. B 1999, 59,
12632.
(53) Du, Q.; Superfine, R.; Freysz, E.; Shen, Y. R. Phys. Rev. Lett. 1993, 70,
2313.
(54) Shen, Y. R. The principles of nonlinear optics; J. Wiley: New York, 1984.
(55) Casson, B. D.; Bain, C. D. Langmuir 1997, 13, 5465.
(56) Hirose, C.; Akamatsu, N.; Domen, K. J. Chem. Phys. 1992, 96, 997.
(57) Zhang, D.; Gutow, J.; Eisenthal, K. B. J. Phys. Chem. 1994, 98, 13729.
(58) Frenkel, J. Kinetic Theory of Liquids; Dover: New York, 1955.
(59) Lee, C. Y.; McCammon, J. A.; Rossky, P. J. J. Chem. Phys. 1984, 80,
4448.
(60) Rossky, P. J.; Lee, S. H. Chemica Scripta 1989, 29A, 93.
(61) Lee, S. H.; Rossky, P. J. J. Chem. Phys. 1994, 100, 3334.
(62) Jedlovszky, P.; Vincze, A.; Horvai, G. J. Mol. Liq. 2004, 109, 99.
(63) Jedlovszky, P.; Vincze, A.; Horvai, G. PCCP 2004, 6, 1874.

103
(64) Puibasset, J.; Pellenq, R. J. M. J. Chem. Phys. 2003, 119, 9226.
(65) Puibasset, J.; Pellenq, R. J. M. PCCP 2004, 6, 1933.
(66) Benjamin, I. J. Chem. Phys. 2004, 121, 10223.
(67) Toney, M. F.; Howard, J. N.; Richer, J.; Borges, G. L.; Gordon, J. G.;
Melroy, O. R.; Wiesler, D. G.; Yee, D.; Sorensen, L. B. Nature 1994, 368, 444.
(68) Reedijk, M. F.; Arsic, J.; Hollander, F. F. A.; de Vries, S. A.; Vlieg, E.
Phys. Rev. Lett. 2003, 90.
(69) Chu, Y. S.; Lister, T. E.; Cullen, W. G.; You, H.; Nagy, Z. Phys. Rev. Lett.
2001, 86, 3364.
(70) Ruan, C. Y.; Lobastov, V. A.; Vigliotti, F.; Chen, S. Y.; Zewail, A. H.
Science 2004, 304, 80.
(71) Ong, S. W.; Zhao, X. L.; Eisenthal, K. B. Chem. Phys. Lett. 1992, 191,
327.
(72) Du, Q.; Freysz, E.; Shen, Y. R. Chem. Phys. Lett. 1994, 72, 238.
(73) Ostroverkhov, V.; Waychunas, G. A.; Shen, Y. R. Phys. Rev. Lett. 2005,
94.
(74) Liu, D. F.; Ma, G.; Levering, L. M.; Allen, H. C. J. Phys. Chem. B 2004,
108, 2252.
(75) Raymond, E. A.; Richmond, G. L. J. Phys. Chem. B 2004, 108, 5051.
(76) Iler, R. K. The Chemistry of Silica; Wiley: New York, 1979.
(77) Li, I.; Bandara, J.; Shultz, M. J. Langmuir 2004, 20, 10474.
(78) Kendall, J. J. Am. Chem. Soc. 1916, 38, 2460.
(79) Kendall, J. J. Am. Chem. Soc. 1916, 38, 1480.
(80) Schrodle, S.; Richmond, G. L. J. Phys. D: Appl. Phys. 2008, 41, 14.
(81) Bousse, L.; Meindl, J. D. Acs Symposium Series 1986, 323, 79.
(82) Sahai, N.; Sverjensky, D. A. Geochim. Cosmochim. Acta 1997, 61, 2827.
(83) The Colloid Chemistry of Silica; Bergna, H. E., Ed.; American Chemistry
Society: Washington, DC, 1994.
(84) Ostroverkhov, V.; Waychunas, G. A.; Shen, Y. R. Chem. Phys. Lett. 2004,
386, 144.
(85) Kosmulski, M. Chemical Properties of Material Surfaces; CRC Press,
2001.
(86) Dove, P. M.; Craven, C. M. Geochim. Cosmochim. Acta 2005, 69, 4963.
(87) Kielland, J. J. Am. Chem. Soc. 1937, 59, 1675.
(88) Malati, M. A.; Estefan, S. F. J. Colloid Interface Sci. 1966, 22, 306.
(89) Malati, M. A. Surf. Coat. Technol. 1987, 30, 317.
(90) Kosmulski, M. Ber. Bunsen-Ges. Phys. Chem 1994, 98, 1062.
(91) Kosmulski, M. Colloids Surf., A 1994, 83, 237.
(92) Perera, L.; Berkowitz, M. L. J. Chem. Phys. 1991, 95, 1954.
(93) Lee, H. M.; Kim, J.; Lee, S.; Mhin, B. J.; Kim, K. S. J. Chem. Phys. 1999,
111, 3995.
(94) Bock, C. W.; Markham, G. D.; Katz, A. K.; Glusker, J. P. Theor. Chem.
Acc. 2006, 115, 100.
(95) Thomas, A. S.; Elcock, A. H. J. Am. Chem. Soc. 2007, 129, 14887.
(96) Morais-Cabral, J. H.; Zhou, Y. F.; MacKinnon, R. Nature 2001, 414, 37.

104
(97) Zhou, Y. F.; Morais-Cabral, J. H.; Kaufman, A.; MacKinnon, R. Nature
2001, 414, 43.
(98) Chuang, I. S.; Maciel, G. E. J. Phys. Chem. B 1997, 101, 3052.
(99) Liu, J. J.; Xu, Z. H.; Masliyah, J. Energy Fuels 2005, 19, 2056.
(100) Scatena, L. F.; Brown, M. G.; Richmond, G. L. Science 2001, 292, 908.
(101) Taylor, D. R.; Maher, K. J. Chromatogr. Sci. 1992, 30, 67.
(102) Crego, A. L.; Gonzalez, A.; Marina, M. L. Crit. Rev. Anal. Chem. 1996,
26, 261.
(103) Vissers, J. P. C.; Claessens, H. A.; Cramers, C. A. J. Chromatogr. A 1997,
779, 1.
(104) Vervoort, R. J. M.; Debets, A. J. J.; Claessens, H. A.; Cramers, C. A.; de
Jong, G. J. J. Chromatogr. A 2000, 897, 1.
(105) Yang, Z.; Li, Q. F.; Hua, R.; Gray, M. R.; Chou, K. C. J. Phys. Chem. C
2009, 113, 20355.
(106) Bittrich, H. J.; Gedan, H.; Feix, G. Z. Phys. Chem. 1979, 260, 1009.
(107) Polak, J.; Lu, B. C. Y. Can. J. Chem. 1973, 51, 4018.
(108) Schatzberg, P. J. Phys. Chem. 1963, 67, 776.
(109) Glasoe, P. K.; Schultz, S. D. J. Chem. Eng. Data 1972, 17, 66.
(110) Gregory, M. D.; Christia.Sd; Affsprun.He. J. Phys. Chem. 1967, 71, 2283.
(111) Kimmel, G. A.; Petrik, N. G.; Dohnalek, Z.; Kay, B. D. Phys. Rev. Lett.
2005, 95, 166102.
(112) Hu, J.; Xiao, X. D.; Ogletree, D. F.; Salmeron, M. Science 1995, 268, 267.
(113) Odelius, M.; Bernasconi, M.; Parrinello, M. Phys. Rev. Lett. 1997, 78,
2855.
(114) Miranda, P. B.; Xu, L.; Shen, Y. R.; Salmeron, M. Phys. Rev. Lett. 1998,
81, 5876.
(115) Pruppacher, H. R.; Klett, J. D. Microphysics of clouds and precipitation,
2nd rev. and enl. ed.; Kluwer Academic Publishers: Dordrecht ; Boston, 1997.
(116) DeMott, P. J.; Sassen, K.; Poellot, M. R.; Baumgardner, D.; Rogers, D. C.;
Brooks, S. D.; Prenni, A. J.; Kreidenweis, S. M. Geophys. Res. Lett. 2003, 30.
(117) Hung, H. M.; Malinowski, A.; Martin, S. T. J. Phys. Chem. A 2003, 107,
1296.
(118) Archuleta, C. M.; DeMott, P. J.; Kreidenweis, S. M. Atmos. Chem. Phys.
2005, 5, 2617.
(119) Mangold, A.; Wagner, R.; Saathoff, H.; Schurath, U.; Giesemann, C.;
Ebert, V.; Kramer, M.; Mohler, O. Meteorol. Z. 2005, 14, 485.
(120) Sassen, K. Geophys. Res. Lett. 2002, 29.
(121) Sassen, K. Nature 2005, 434, 456.
(122) DeMott, P. J.; Cziczo, D. J.; Prenni, A. J.; Murphy, D. M.; Kreidenweis, S.
M.; Thomson, D. S.; Borys, R.; Rogers, D. C. PNAS 2003, 100, 14655.
(123) Sassen, K.; DeMott, P. J.; Prospero, J. M.; Poellot, M. R. Geophys. Res.
Lett. 2003, 30.
(124) Toon, O. B. Nature 2003, 424, 623.
(125) Twohy, C. H.; Poellot, M. R. Atmos. Chem. Phys. 2005, 5, 2289.
(126) Wiacek, A.; Peter, T. Geophys. Res. Lett. 2009, 36.

105
(127) Sullivan, R. C.; Guazzotti, S. A.; Sodeman, D. A.; Prather, K. A. Atmos.
Chem. Phys. 2007, 7, 1213.
(128) Cziczo, D. J.; Froyd, K. D.; Gallavardin, S. J.; Moehler, O.; Benz, S.;
Saathoff, H.; Murphy, D. M. Environ. Res. Lett. 2009, 4.
(129) Eastwood, M. L.; Cremel, S.; Wheeler, M.; Murray, B. J.; Girard, E.;
Bertram, A. K. Geophys. Res. Lett. 2009, 36.
(130) Gallavardin, S. J.; Froyd, K. D.; Lohmann, U.; Moehler, O.; Murphy, D.
M.; Cziczo, D. J. Aerosol Sci. Technol. 2008, 42, 773.
(131) Knopf, D. A.; Koop, T. J. Geophys. Res. Atmos. 2006, 111.
(132) Chernoff, D. I.; Bertram, A. K. J. Geophys. Res. Atmos. 2010, 115.
(133) Sullivan, R. C.; Petters, M. D.; DeMott, P. J.; Kreidenweis, S. M.; Wex,
H.; Niedermeier, D.; Hartmann, S.; Clauss, T.; Stratmann, F.; Reitz, P.; Schneider, J.;
Sierau, B. Atmos. Chem. Phys. 2010, 10, 11471.
(134) Zimmermann, F.; Ebert, M.; Worringen, A.; Schutz, L.; Weinbruch, S.
Atmos. Environ. 2007, 41, 8219.
(135) Eastwood, M. L.; Cremel, S.; Gehrke, C.; Girard, E.; Bertram, A. K. J.
Geophys. Res. Atmos. 2008, 113.
(136) Clegg, S. L.; Brimblecombe, P.; Wexler, A. S. J. Phys. Chem. A 1998, 102,
2137.
(137) Glaccum, R. A.; Prospero, J. M. Mar. Geol. 1980, 37, 295.
(138) Serratosa, J. M.; Bradley, W. F. Nature 1958, 181, 111.
(139) Stamenkovic, V.; Chou, K. C.; Somorjai, G. A.; Ross, P. N.; Markovic, N.
M. J. Phys. Chem. B 2005, 109, 678.
(140) Gragson, D. E.; Richmond, G. L. J. Am. Chem. Soc. 1998, 120, 366.
(141) Watanabe, M.; Brodsky, A. M.; Reinhardt, W. P. J. Phys. Chem. 1991, 95,
4593.
(142) Gavish, M.; Wang, J. L.; Eisenstein, M.; Lahav, M.; Leiserowitz, L.
Science 1992, 256, 815.
(143) Xu, L.; Lio, A.; Hu, J.; Ogletree, D. F.; Salmeron, M. J. Phys. Chem. B
1998, 102, 540.
(144) Claesson, P. M.; Herder, P.; Stenius, P.; Eriksson, J. C.; Pashley, R. M. J.
Colloid Interface Sci. 1986, 109, 31.
(145) Herder, P. C.; Claesson, P. M.; Herder, C. E. J. Colloid Interface Sci. 1987,
119, 155.
(146) Scales, P. J.; Grieser, F.; Healy, T. W. Langmuir 1990, 6, 582.
(147) Lyons, J. S.; Furlong, D. N.; Healy, T. W. Aust. J. Chem. 1981, 34, 1177h.
(148) Chao, T. T.; Harward, M. E.; Fang, S. C. Soil Sci. Soc. Am. Proc. 1964, 28,
632.
(149) Parfitt, R. L. Adv. Agron. 1978, 30, 1.
(150) Zhang, G. Y.; Zhang, X. N.; Yu, T. R. J. Soil Sci. 1987, 38, 29.
(151) Courchesne, F.; Hendershot, W. H. Can. J. Soil Sci. 1989, 69, 337.
(152) Cannon, W. R.; Pettitt, B. M.; McCammon, J. A. J. Phys. Chem. 1994, 98,
6225.
(153) Zhang, L. N.; Liu, W. T.; Shen, Y. R.; Cahill, D. G. J. Phys. Chem. C
2007, 111, 2069.

106
(154) Gracias, D. H.; Chen, Z.; Shen, Y. R.; Somorjai, G. A. Acc. Chem. Res.
1999, 32, 930.
(155) Shultz, M. J.; Schnitzer, C.; Simonelli, D.; Baldelli, S. Int. Rev. Phys.
Chem. 2000, 19, 123.
(156) Chen, Z.; Shen, Y. R.; Somorjai, G. A. Annu. Rev. Phys. Chem. 2002, 53,
437.
(157) Morita, A.; Hynes, J. T. Chem. Phys. 2000, 258, 371.
(158) Blum, L.; Huckaby, D. A. J. Chem. Phys. 1991, 94, 6887.
(159) Zhu, B. Y.; Gu, T. R. Adv. Colloid Interface Sci. 1991, 37, 1.
(160) Chang, C. H.; Franses, E. I. Colloids Surf., A 1995, 100, 1.
(161) Zhang, Z. H.; Tsuyumoto, I.; Takahashi, S.; Kitamori, T.; Sawada, T. J.
Phys. Chem. A 1997, 101, 4163.
(162) Cosman, N. P.; Roscoe, S. G. Langmuir 2004, 20, 1711.
(163) Mohan, D.; Pittman, C. U.; Steele, P. H. J. Colloid Interface Sci. 2006,
297, 489.
(164) Hommel, E. L.; Allen, H. C. Analyst 2003, 128, 750.
(165) Choi, P. private communication, 2009.
(166) Macphail, R. A.; Strauss, H. L.; Snyder, R. G.; Elliger, C. A. J. Phys.
Chem. 1984, 88, 334.
(167) Heller, W. J. Phys. Chem. 1965, 69, 1123.
(168) Brocos, P.; Pineiro, A.; Bravo, R.; Amigo, A. PCCP 2003, 5, 550.
(169) Dean, J. A.; Lange, N. A. Lange's handbook of chemistry, 15th ed.;
McGraw-Hill: New York :, 1999.
(170) Fu, Y.; Hansen, R. S.; Bartell, F. E. J. Phys. Colloid Chem. 1948, 52, 374.
(171) Everett, D. H.; Munn, R. J. Trans. Faraday Soc. 1964, 60, 1951.
(172) Ash, S. G.; Everett, D. H.; Findeneg.Gh. Trans. Faraday Soc. 1970, 66,
708.
(173) Ash, S. G.; Bown, R.; Everett, D. H. J. Chem. Thermodyn. 1973, 5, 239.
(174) Basu, S.; Nandakumar, K.; Masliyah, J. H. J. Colloid Interface Sci. 1998,
205, 201.
(175) Desyatov, I. V.; Paukshtis, E. A.; Mashkina, A. V. React. Kinet. Catal.
Lett. 1990, 41, 161.
(176) Liu, J. J.; Zhou, Z.; Xu, Z. H. Ind. Eng. Chem. Res. 2002, 41, 52.
(177) Liu, J. J.; Xu, Z. H.; Masliyah, J. Langmuir 2003, 19, 3911.
(178) Smith, C.; Chatergoon, L.; Whiting, R. Analyst 1996, 121, 373.
(179) Pashley, R. M. J. Colloid Interface Sci. 1981, 80, 153.

107

You might also like