<<

ULTRAFAST PHOTOEXCITATION STUDIES OF CONCENTRATED SOLUTIONS OF HALIDES

Udaya Indike Rodrigo

A Thesis

Submitted to the Graduate College of Bowling Green State University in partial fulfillment of the requirements for the degree of

MASTER OF SCIENCE

December 2006

Committee:

Michael Rodgers, Adviser

Thomas Kinstle

John Cable.

ii

ABSTRACT Michael A.J. Rodgers, Adviser

The primary photochemical processes that occur via a two photon mechanism in pure water

during high intensity excitation with a femtosecond pulse at 267 nm were studied by using a white

light continuum as probe pulse. Two photon femtosecond laser studies of the excitation and

relaxation of pure water showed that, geminate lifetimes were comparable with reported data but the

recorded absorption spectrum was blue shifted.

Experiments carried out for highly concentrated compounds with different anion

indicated that these compounds produced their maximum transient absorption between the

wavelengths of 550nm to 700nm. The time profiles of these compounds fell into two major

categories. LiI and LiOH showed different kinetic profiles from the others by the initial presence of

a short lived species along with longer decay kinetics, where as LiCl, LiBr and LiClO4 shows typical decay kinetics of long lived species only. The fitted experimental data on single exponential function for absorption kinetics of LiCl showed that the absorption rise time falls into same time regime as

reported for the purported three body complex of NaCl. This leads to the thought that LiCl may

follow the same mechanism.

The concentration dependence study of LiCl showed its absorption maxima were red shifted

as the concentration was changed from high to low. The kinetic data revealed that rate of absorption

for low concentrated solution of LiCl is higher than that of high concentrated LiCl solution. The

comparison of LiCl with KCl and NaCl showed that there is a significant contribution from the

cation for higher concentrated solution to alter their kinetic and transient absorption. But in the case

of low concentrated solutions that effect is not significant.

iii

To my mother, late father, all my teachers and friends

iv

ACKNOWLEDGEMENT

I would like to express my sincere gratitude to my research Adviser, Dr. Michael Rodgers for

accepting of me as a master’s research student, and providing continuous guidance and close

supervision throughout this research work. Also, I would like to express my deep appreciation to

Dr. Paul Endres, for his continuous encouragement and guidance during the project. My sincere

thanks are also due to Dr. Eugene Danilov, who was the research coordinator, for teaching me

how to use facilities of Ohio Laboratory for Kinetic Spectrometry research laboratory.

I am grateful to, Department of Chemistry, Bowling Green State University, giving me

unrestricted laboratory facilities during day and night. I am also grateful to the technical and office

staff of the Department of Chemistry, for their assistance given to me during my technical and

administrative issues.

I take this opportunity to thank my colleagues, who shared good and bad moments with

me during my research career at bowling green. Finally, I am grateful to the Ohio Laboratory for

Kinetic Spectrometry for instrumental support to make this research project a success.

v

TABLE OF CONTENTS

Chapter 1 INTRODUCTION...... 1

1.1 Solvated ...... 1

1.2 Method of Formation of Solvated Electrons...... 2

1.3 Hydrated Electrons ...... 4

1.4 Structure of Hydrated Electrons ...... 6

1.5 Two photon excitation and femtosecond lasers ...... 9

Primary Events during the Excitation of Water ...... 10

1.6 Metal-solvent systems...... 12

1.7 Ultrafast Dynamics of Aqueous ...... 15

1.8 Flash Photolysis of Halides...... 17

1.9 Photolysis of Halides in Aqueous Solution ...... 18

Chapter 2 MATERIALS AND METHODS...... 20

2.1 Materials...... 20

2.2 Experimental Section...... 21

Instrumentation and Method...... 21

UV- Visible Absorption Spectroscopy...... 21

Femtosecond Transient Absorption Spectroscopy...... 21

Chapter 3 RESULTS...... 24

3.1 Pure Water...... 24

3.2 Concentrated Alkali Metal Halide Solutions ...... 28

3.3 LiCl Solutions; Effect of Concentration ...... 34

Chapter 4 DISCUSSION AND CONCLUSION ...... 40

References ...... 44 vi

LIST OF FIGURES

Figure 1.1 The structure of the nearest solvation shell of hydrated ...... 6 in glassy water

Figure 1.2 Shows the lowest electronic transition in the hydrated electron ...... 7 (a) Absorption spectrum. The smooth solid curve shows experimentally measured absorption at room temperature. Squares depict the result of quantum molecular simulations The dashed curves correspond to the individual absorption components originating from three non-degenerate s–p transitions. (b) Electronic wave function plots for typical ground, s-like, state and lowest three excited, p-like, states

Figure 1.3 Model of primary events in pure liquid water at two...... 10 photon excitation.

Figure 2.1 Experimental layout of the transient absorption ...... 22 Spectrometer.

Figure 3.1 Transient absorption spectrum of distilled deionized water...... 25

(at 80ps delay time) in 0.09mm fluoride(CaF2) thin window cell.

Figure 3.2 Transient absorption spectrum of distilled deionized...... 25 water (at 114ps delay time) in 2mm quartz cuvette.

Figure 3.3 Transient absorption kinetic data for a 0.09 mm cell of...... 26 neat water excited with 267nm femtosecond pulses (maximum absorption at 650 nm wavelength.).

Figure 3.4 Plot of Transient Absorption (ΔA) vs Wavelength (nm) of ...... 29 different lithium compound(in 2 mm flow through cuvette) excited with 267nm femtosecond pulses at room temperature.

Figure 3.5 Kinetic plots of data for different lithium compounds (in...... 30 mm flow through cuvette) excited with 267nm femtosecond pulses (at their respective maximum absorption wavelength). vii

Figure 3.6 Plot of transient absorption kinetic data for different ...... 31 lithium compounds (in 2 mm flow through cuvette) excited with 267nm femtosecond pulses(at their respective maximum absorption wavelength).

Figure 3.7 Plot of Absorption vs. Wavelength for LiCl before...... 32 and after the experiment.

Figure 3.8 Transient absorption spectra of LiCl at different ...... 34 concentrations (in 2 mm flow through cuvette) excited with 267nm femtosecond pulses at room temperature.

Figure 3.9 Kinetic plots of highest (a) and lowest (b) concentrated...... 36 solutions of LiCl in deionized water (in 2 mm flow through cuvette) excited with 267nm femtosecond pulses and solid line is the two-exponential decay fit for the experimental data.

Figure 3.10 Transient absorption spectra of highest concentrated...... 37 solutions KCl, NaCl and LiCl in deionized water (in mm flow through cuvette) excited with 267nm femtosecond pulses at room temperature.

Figure 3.11 Transient absorption spectra of one molar (1M) solutions...... 38 KCl, NaCl and LiCl in deionized water (in 2 mm flow through cuvette excited with 267nm femtosecond pulses at room temperature.

Figure 3.12 Kinetic plots of highest concentrated solutions of LiCl, KCl ...... 39 and NaCl. (in 2 mm flow through cuvette) excited with 267nm femtosecond pulses viii

LIST OF TABLES

Table 3.1 Table of lithium compounds with their respective...... 28 highest concentrations.

Table 3.2 First order fitting parameters for the transient...... 31 absorption kinetic data.

Table 3.3 Fitting parameters of highest concentrated solutions ...... 39 of LiCl, KCl and NaCl. 1

CHAPTER 1: INTRODUCTION

1.1 Solvated Electrons

Upon ejection of an electron into a liquid, it can be captured and relocated in a potential

energy well formed by neighboring molecules of the liquid. Such an electron is known as solvated

electron. Solvated electrons were first observed in liquid in 1864 (1); since then the study

of excess, or solvated, electrons in liquids has attracted much interest in the fields of chemistry,

physics and biology. The existence of such electrons in water (known as hydrated electrons) was first

postulated independently in 1952 by Stein (1, 2) and Platzman (1, 3). After decades of research and indirect evidence, the hydrated electron was finally observed in 1962 by Boag and Hart (4), who used visible and near infrared (NIR) absorption spectrometry in an electron pulse radiolysis experiment in water. Subsequently observations of solvated electrons produced by ionization radiation were expanded to include other solvents such as ethers, hydrocarbons and metal (alkali and alkali earth) solutions (5).

This newly developed capability of electron pulse radiolysis for studying the solvated electron in a wide variety of solvents created a new surge of interest in the whole field. A variety of

different approaches and combinations have been employed which has resulted in hundreds of absolute rate constants of solvated electron reactions as well as many physical properties of the solvated electron (5). There have also been developments in the study and understanding of metal ion-solvent systems. These discoveries have led to concomitant growth in theoretical research in the physical and chemical properties of the solvated electron.

2

1.2 Method of Formation of Solvated Electrons

Radiolysis is the most common way for the production of solvated electrons. Radiation- induced (or laser) ionization of liquid phase chemical species generates an excess electron in the delocalized conduction band; the electron subsequently undergoes rapid localization in a micro- cavity that exists among the solvent molecules (1,6). The localized electron undergoes nuclear relaxation and becomes an “equilibrated solvated (hydrated) electron”.

There are two factors that can affect the solvation process of the free electron in solution

1. There should be a physical interaction existing between solvent molecule and the electron

which is trapped in the potential energy well.

2. The chemical reactivity between the solvent and the electron must be low enough to permit

time for the solvation to occur.

The solvated electrons that form during radiolysis can react with other transient products formed during the process (5). The hydrated electron was the first of the radiolytically produced species to be observed directly through its visible/NIR optical absorption in a pulse radiolysis experiment (5, 6). Since reactions of solvated electrons have been studied more extensively in water and other solvents, it important to first discuss the processes during aqueous radiolysis.

Whenever a charged particle travels very rapidly (keV → MeV) through a solution, it tends to ionize and excite molecules along its path. Secondary electrons produced by such a process are lower energy and may cause further ionization and excitation in small volumes. Most primary products of radiolysis are inhomogeneously distributed along the primary track of the electron in these small volumes (spurs). A faster moving electron can produce spurs that are widely separated by several hundreds of nanometers. Heavier particles such as -particles, which are in the same energy

3 range, travel much more slowly and lose more energy per unit path length. Therefore they produce an essentially continuous track of primary products.

These primary species are produced in times less than 10 -11 sec for water (1, 5). Since these

primary species are highly reactive and have high local concentrations, some of them react with each

other to produce secondary products as they diffuse out of the spurs to form a homogeneous

solution. According to the literature the complete spur reaction duration is about 10 -7 sec for

water (1, 5).

4

1.3 Hydrated Electrons

The high degree of interest in the hydrated electron from both theoretical and experimental points of view is due to the importance of the hydrated electron as the transient species in aqueous

systems. Photosynthesis, charge transport through bio-membranes and long distance charge

transport in nerves are important examples (7-9). Also, the hydrated electron is generally the key

intermediate in and electrochemistry (4, 10-13). Unlike free electrons that are

delocalized, electrons in polar solvents become self-trapped because of their interactions with the

solvent environment. Owing to strong solute–solvent coupling, the evolution of the electronic

structure is completely determined by the rearrangement of the solvent molecules. Therefore a

detailed study of hydrated electrons is particularly interesting from the point of view of the solvent

which is involved in the process of experiment.

Being the natural environment of such important biological molecules as nucleic acids and

peptides, water is present in all living organisms (1, 14-16). Therefore a detailed understanding of

charge transfer reactions initiated by light absorption in water is of significant importance (14, 17).

The study of solvated electrons is also very interesting from the point of view of the solvent

involved. Among all solvents, water occupies an extraordinary position regarding its specific role in

nature, and as such it may influence the results of many reactions because of its large dipole moment

and its strong hydrogen bonding tendency (1).

The structural and dynamical properties of water are of long standing interest in science. As

a result many theoretical and experimental studies have been carried out to study the properties of

water as a solvent as well as its solvation dynamics (1, 17). The detailed understanding of

solute – solvent interaction has a number of practical implications including the dynamics of

chemical reactions. All chemical reactions involve the rearrangement of electrons and this is affected

5 by the motion of surrounding solvent molecules which are coupled to the reactant’s energy levels.

The time scale of reactions where the solvent acts to stabilize the new charge distribution of the reacting species can determine how fast the particular reaction crosses into its transition state.

During the last several decades, molecular dynamic simulations and ultrafast studies on dye solutions

have been used to obtain the basic picture of the solvation process and the relevant time scales (18-

23). Because of most of these dye solution in water showed the initial solvation processes to be

exceptionally fast and the studies lacked time resolution, many important aspects of early dynamics

remain unexplored and unresolved (24).

The other motivation for a detailed study of the hydrated electron is the fact that this species

is ideally suited for quantum molecular dynamics stimulation in the liquid phase. In this regard it is

important to understand the extent of the quantum-mechanical character of the electron interaction with its nearest neighboring water molecules. Since there is no internal degree of freedom in the electron itself, the hydrated electron is ideal for verifying the model potential that describes the interaction between the molecules of liquid water.

6

1.4 Structure of Hydrated Electrons

Numerous computational studies have been performed to investigate the quantum mechanical status of the hydrated electron and its surrounding microscopic structure. The structure of this species is revealed in an electronic-spin-echo study by L. Kevan (25). According to his study, it was shown that each electron is surrounded by six water molecules each with an OH bond directed towards the electron as shown in the figure (1.1). This idea is confirmed by recent computational studies on hydrated (solvated) electron in water, which show that the first solvation shell is composed of approximately six water molecules (26, 27). There are other hypotheses which

Figure 1.1: The structure of the nearest solvation shell of hydrated electron in glassy water (adapted from Ref [1, 25] )

suggest that the electron might be attached closer to one of the “dangling protons” that are not involved in the hydrogen bonding of the molecule that forms the solvent cage (28). Therefore the details of the exact structure are still under discussion.

7

The localization of a hydrated electron in the solvent cavity gives rise to bound eigenstates.

These eigenstates are broadened since they are modulated by the coupling to the fluctuations of surrounding water molecules. It has been reported that the high sensitivity of the electronic states of the hydrated electron to the aqueous environment results an intense broad electronic absorption spectrum with a peak at 720nm. Also molecular dynamic stimulations (29) and computational studies have shown that the lowest energy eigenstate of the hydrated electron is nearly spherical and corresponds to an s-like state and the first excited state was found to consist of three non-degenerate p-like orbitals.

Figure 1.2 Shows the lowest electronic transition in the hydrated electron.(a) Absorption spectrum. The smooth solid curve shows experimentally measured absorption at room temperature. Squares depict the result of quantum molecular simulations (adapted from ref [1, 30]).The dashed curves correspond to the individual absorption components originating from three non-degenerate s–p transitions. (b) Electronic wave function plots for typical ground, s-like, state and lowest three excited, p-like, states (reproduced from ref. [1, 31]).

The existence of the s-state indicates that on average the potential energy well created by the molecules surrounding the electron is close to spherical. However, the dynamic nature of liquid

8 water causes some asymmetries, therefore the potential energy surface does not have a perfect spherical shape and the excited states were considered as three non-degenerate p-states.

Numerous attempts have been made to fit the experimental spectra with various line shapes and superposition of lines (1, 32, 33). Over the last three decades, there are a few outstanding problems about homogeneous / inhomogeneous broadening in the optical absorption spectrum of the excess electron in the water (and other fluids) as well as the explanation of asymmetry and extraordinary spectral width (34-38). Computational and computer simulation studies have provided a fruitful path to address this issue. Computer simulation studies have been used to pictorially describe the shapes of the spectra. The absorption spectrum produced in these computational studies was used to discuss the superposition of three non- degenerate s-p transitions and contributions of the transition from the higher delocalized states (29). Interestingly, it also revealed the charge distribution upon the promotion to the p-state. These models were used to study the dynamic behavior of the hydrated electron and to modulate the energy of relaxation after the instantaneous s-p excitation (26, 27). Also most of these studies show that the solvation dynamics are bimodal and the initial decay is responsible for half of the total energy of relaxation. This relaxations initially occurs at 10-25 fs time scale and followed by a slower decay around 130-250 fs (1, 26). The slower decay appeared to be associated with the diffusional motion of water molecules into and out of the first solvation shell. The microscopic nature of the first initial decay is still under considerable debate.

9

1.5 Two photon excitation and femtosecond lasers

In the past two decades increasing attention has be been paid towards a detailed understanding of charge transfer reactions initiated in water by the absorption of light. According to the literature it is known that intense light at λ > 190 nm is able to excite a water molecule via a two photon absorption (TPA) mechanism and initiate its chemical decomposition (14). It was observed that high intensity picosecond pulse UV excitation at λ = 266 nm can cause the water molecule to undergo two-photon absorption with subsequent ionization and dissociation (14 -16).

- + H2O + 2hν→ H2O* ~~>eaq , H3O , OH, H

It is thus possible to investigate the reaction dynamics of the electron in pure water using

high intensity femtosecond laser systems. In this regard most experimental groups use pump and

probe techniques with two photon excitation in the UV and probing in the visible and near IR

ranges. They have studied electron trapping, solvation and geminate recombination with time

resolution down to 20 fs (14, 16). In many of these investigations a colliding pulse mode–locked

dye laser with second harmonic generation was employed which yields femtosecond UV pulses with

corresponding two photon energies (14, 15). Later a few groups directed their studies towards

femtosecond kinetic spectroscopy with pumping at 282 nm and probing in the visible, near IR and

the ultraviolet ranges. This allowed them to monitor the two photon absorption process in time with

subsequent measurement of the UV pump pulse duration. This leads to the determination of the

two photon absorption coefficients and quantum yields of hydrated electrons in water.

10

Primary Events during the Excitation of Water

*(1) * H2O H2O

τ tr - e wet

τ hyd

- + e eq, H30 ,OH ,H

T j geminate and volume recombination

H2O

Figure 1.3 Model of primary events in pure liquid water at two photon excitation (14)

With the absorption of two light quanta (λ= 267 nm) a water molecule acquires an energy of

9.2 eV, more than sufficient to promote the water molecule to the ionization and dissociation threshold energies represented in the scheme above(14,16). Therefore as a result a water molecule may either undergo ionization or dissociation. the ionization channel as shown in the figure 3 (14).

* Two-photon excitation promotes the water molecule to an excited state of H2O . During the

*(1) ionization channel, it proceeds via H2O and includes the processes of electron detachment and

localization (with time constant τ tr ) forming the so-called wet electron in the pre-hydrated state

- (e wet). This is followed by hydration with characteristic time τ hyd producing the hydrated electron

- + (e eq).Then the geminate recombination with positive ion H2O (5) occurs inside a solvent cage with

appropriate geminate recombination time. that escape the solvent cage will undergo volume

recombination (kinetically second order) over a longer time.

11

Other researchers have used alternative approaches to carry out the time resolved studies of fluids rather than the use of conventional time resolved studies of hydrated electrons generated by multi photon ionization of neat water and the observation of the transient absorption of super continuum probe. Recently, a few group of researchers tried to excite the hydrated electron, which is already in the equilibrated form, from its ground s-state to the p-state using a short pulse and the resulting solvation dynamics was probed as a function of time with another delayed pulse (39). By following this route researchers were able to explain how these excited electrons relax to the ground state via not yet equilibrated “hot” ground state, state where the solvation took place before relaxation back to the ground state, by using the so called three-state model (40).

12

1.6 Metal-solvent systems

Because of substantial reactivity of the secondary products which are produced in the spurs during the ionization process, researchers have tended to examine other possibilities such as competition from cations and anions by introducing them into the system. Most have focused their interest on metal ions in solutions with different concentrations. In this regard the nature of metal ions, especially alkali and alkaline earth, in different solvents and properties of such solutions have been extensively studied.

The solubility of alkali and alkaline earth metals in liquid ammonia is well known and also it is known that in dilute solutions one finds mainly ammoniated metal ions and ammoniated electrons which are independent from each other. As the concentration of the metal ions in the solutions increased the association becomes appreciable and the solutions tend to show the properties of

liquid metal. Therefore initial concerns were drawn towards the dilute solutions in which the

solvated electrons predominant. During such experiments, two absorption bands are found in metal-

ammonia solutions. One of these found in the infrared region and is nearly independent of the metal

solute and it is assigned to the solvated electron. This absorption peak can also be found in the

solutions in which the metal ions are absent. The other band is at higher energy is due to the metal

ion and the position of the peak depends on the metal. Several groups have examined the solubilities

and the properties of the absorption spectra of metal ions in different solvents (5). It has been

concluded that photochemical formation of solvated electrons should be possible at least in the

more polar solvents where it solvates with suitable absorption bands (due to its charge transfer to

the solvent).

At the same time it was found that aqueous solutions of inorganic ions such as Cl−, Br−, I-−

− − − OH and CN can be photoionized to give eaq . Also, there are a few organic solvents with aromatic

ring structure (with low π → π* energies) capable of producing intermediates that can release an

13 electron to the solvent on flash photolysis (5). Later, quantum yields were determined, tabulated and interpreted further according to the yield of radicals escaping cage recombination in aqueous halide

− solutions. It was found that the initial yield of (I aq +eaq ) cage is higher than the yield shown in the data tables (5), presumably because I − has a substantial fraction recombining geminately. Since then people have been studying and reviewing the process of photolysis of simple inorganic anions in aqueous solution and the spectral nature of the charge transfer to solvent as well as to check whether there is a reasonable photochemical efficiency for hydrated electron production (5).

The detailed studies on the absorption spectra of dilute solutions of alkali metals in ammonia

and some organic solvents are revealing and have resulted in the characterization of the absorption

spectrum of the solvated electron (1, 5, 41). In a liquid consisting of polar molecules with very low

(or no) electron affinity, such as water and alcohols, the loss of an electron from a photoexcited

anion frequently involves a short lived mediating state which is unique to the particular polar liquid.

In a charge transfer to solvent state (CTTS), the excess charge resides in a diffuse orbital that

protrudes from the anion into a cavity which consists of solvent molecules. (This diffuse orbital of

anions such as halide and hydroxide, where they promote electron from its p-orbital has the primary

s-like character). It is found that this type of state is stabilized by the pre-existing orientation of the

solvent dipoles and electrostatic attraction is known to stabilize solvated electrons in the bulk liquid.

It is also known that a fully detached, thermally relaxed electron is formed as the CTTS state

dissociates and the resulting species equilibrates with the solvent (1, 5, 41). Later it was shown that a

large fraction of solvated electrons reside in close proximity to their geminate partner and weakly

interact with it by means of an attractive potential; they are essentially indistinguishable from the

bulk species

( 1, 5, 41). It is still uncertain what fractions of those close pairs are in actual contact or separated by

14 a solvent molecule, or what is the barrier between these forms. The Schwartz group interpreted their observation in terms of populations of contact and solvent-separated pairs

(5, 41). They interpreted their observation as being due to both types of these pairs, generated in the photoexcitation of small inorganic anions and corresponding branching ratio of some of these anions upon excitation energy (5).

In the 1960’s it was found that the absorption spectrum of the hydrated electron is modified in concentrated solutions of ammonia (4, 41) where it shifts its absorption spectrum of the hydrated

electron to longer wavelength. This aroused the curiosity and created the interest to investigate the

− effect of other solutes on the absorption spectrum of e aq . This led researchers to think about the

generation of solvated electron in concentrated aqueous electrolytes. In contrast to the case of

− ammonia, changing the solvents affected the absorption spectrum of e aq shifting its position to

shorter or longer wavelength (4, 42, 43).

Anbar and Hart were the first to report the properties of electron pulse-generated solvated

electrons in concentrated aqueous electrolytes (43). During their experiment, the absorption spectra

of solvated electrons in ethylenediamine and in concentrated aqueous solutions were measured by

determining the absorption spectra of the transients formed on pulsed radiolysis. They found that

the maximum of the absorption band of the hydrated electron shifted from 720 nm in pure water is

to shorter wavelengths in concentrated solutions of MgCl2, KF, NaOH, KOH, NaCLO4 and LiCl.

− Further they found that the variation of the rate constant of e aq in different salt solution differed

significantly from the corresponding rate constant in water (42, 43).

15

1.7 Ultrafast Dynamics of Aqueous Hydroxide

Photoinduced detachment of an electron from aqueous hydroxide anion is one of the less studied CTTS reactions in this field (41). Photochemically it yields a geminate pair of hydroxyl

. ─ (OH ) and hydrated electron ( eaq ).

─ ─ OH → OH + eaq ………………….( 1 )

Although the structure of the aqueous hydroxide is simply written as OH ─, it is a strongly bonded

structure where on average 3-4 water molecules are bound to each other though atoms

leading to a strong solvent ordering. Also there is a rapid proton transfer along the hydrogen

bonding network. Therefore these structural and dynamical properties of aqueous hydroxide

solution offer a new opportunity for the current research (41, 44-46).

Species formed during the photoexcitation of OH − rapidly recombine both geminately and

in the bulk to either reform the parent anion or by deprotonation of the hydroxyl radical with parent

anion to yield the O− radical anion.

The possible reactions in the aqueous hydroxide solution are

─ ─ OH + eaq OH ……………… (a)

─ − OH + OH O +H2O …...…………. (b)

− ─ ─ O + eaq + H2O 2 OH ...…………… (c)

It was found that reaction (a) is one of the fastest reactions for the hydrated electron and reaction (c)

also happens nearly as rapid as the OH radical in reaction (a).

16

Later, researchers from Argonne National Laboratory and Department of Chemistry at

University of Southern California have studied the charge transfer to solvent for hydroxide ions (41).

They used either monophotonic or biphotonic excitation of hydroxide anion in aqueous solution by means of pump-probe ultrafast laser spectroscopy. They studied the transient absorption of

hydrated electrons as function of anion (hydroxide) concentration and temperature. They were able

to observe that the geminate decay kinetics are bimodal, with a fast exponential component and a

slower power tail due to diffusive escape of the electron. They also found that for biphotonic

excitation the fraction of escaped electrons is twice that for the monophotonic excitation because of

the broadening of the electronic distribution. According to their observation, biphotonic electron

detachment is very insignificant and it shows no concentration dependence for the time profile of

concentration of the solvated electron between hydroxide concentrations of 10mM to 10M. Further,

it also shows that at higher temperatures the escape fraction of electrons increases and the

recombination and diffusion controlled dissociation of the close pair become faster (41).

17

1.8 Flash Photolysis of Halides

Although the halogens (halides and pseudo-halides anions) are well known as efficient electron acceptors in charge transfer complexes with various organic and inorganic substrates, they can also act as electron donors in many ion pairs (47). This dual property accounts for the existence

of poly-halide radical anions through the interaction of halogen atom and anion.

The existence of polyhalide radical anions as an intermediate in many redox reactions was

postulated many years ago (47-49) but the developments in flash photolysis and pulse radiolysis give

a new ability to find direct evidence for these species. In conventional flash photolysis

polychromatic light is used and the resulting time resolution is of the order of few microseconds

which limit the ability to study the reactivity of the radical. Recent developments in pulsed laser

photolysis allow researchers to use monochromatic excitation and to reduce of the period of

observation to sub - picoseconds.

There is no corresponding problem regarding the technique of pulse radiolysis since the

electrons are absorbed by the solvent. Most of the studies that have been conducted so far concern

the polyhalide radical anion in polar solvents such as water and various alcohols (47). Other low

polarity solvents have been used to a lesser extent since they do not favor the solvation of charged

species. Researchers ultimately found that whatever the experimental method used to produced

polyhalide radicals, in nearly all cases the primary photochemical or radiochemical processes

corresponds to the formation of a halogen atom or pseudo-halogen followed by a dark equilibrium

reaction with a halide or a polyhalide present in the medium (47).

18

1.9 Photolysis of Halides in Aqueous Solution

Alkali and alkali earth halides in aqueous solutions have been studied on numerous occasions using conventional flash photolysis (47, 50-52). Grossweiner and Matheson were the first people to observe the optical transient absorption in the near UV due to the unstable species produced by the

conventional flash photolysis of KCl, KBr and KI (42, 45). During their experiment they found

absorption maximums at 340, 350 and 370 nm respectively and assigned the transient absorption to

. ─ . ─ . ─ . ─ Cl2 , Br2 and I2 . They were able to carry out detailed investigation which revealed that both I2

. ─ and Br2 show some concentration dependence of their respective decay time . It is found that in

. ─ the case of I2 the decay kinetics was second order and became faster when the concentration of

. ─ iodide ion decreases. By contrast Br2 shows the same decay kinetics but the rate constant decreases

as concentration of bromide ion increases. Further experiments have confirmed their decay kinetics

and UV band assignments, and revealed a new absorption in the red spectral region which was

. ─ assigned to X2 .

The simplest mechanism proposed for the production of radical anion is

─ ─* . X X X + eaq …………………… (d)

. ─ . ─ X + X X2 ……………………. (f)

and it was assumed that formation of photo induced X. is due to electron transfer from halide anion

to the solvent( 47). Also they proposed decay processes to explain the dependence of radical anion

disappearance on the halide ion concentration. To extend the scope of the proceeding

investigations, researchers have studied various inorganic compounds containing halide anions in

order to understand the electron solvation in water (51). It was found that in all cases transient

absorptions were in the UV region and show similar kinetics confirming the assignments and

mechanism mentioned above. Further, there were some experiments carried out with mixtures of

19 halide anions but there is no evidence of formation of interhalide radical anions. Up to now people have been making numerous attempts to study these so called I ─ /I . ─ system in aqueous and non

aqueous solvents. Later research is directed towards different solvents, various types of halide such

as trihalide and so on.

A paper by M. Anbar and Edwin J. Hart in 1968 (43), was found to be one of the few studies in

the area of solvated electrons in highly concentrated electrolytes. In it they discussed how solvent

and solute affect the absorption spectrum (43). Other researchers found that ultra fast dynamics of

the electrolyte solution offers opportunities for some direct observation of elementary oxidation and

reduction reactions with charged reactants (54). It was also found that short time dynamics of charge

transfer process between and electron donor and acceptor (cationic) could be influenced by non-

equilibrium electronic configurations taking place in solvent bridged ion

pairs (54, 55). When the solute-solvent caging effects are strong, early cage back geminate

recombination may take place which is the case with the transient charge transfer to solvent state of

aqueous halide ions (54, 56).

Therefore attention was drawn toward high concentrated alkali and alkali earth metal solutions

with different anions. Thus, set out to examine how such solutes could affect the behavior of the

hydrated electron, specifically the kinetics and spectral shifts (change in absorption spectra).

20

CHAPTER 2: MATERIALS AND METHODS

2.1 Materials

Lithium Hydroxide (98%, Sigma-Aldrich, Reagent grade), Sodium Chloride (>99.5%, Fluka

& Fisher, Reagent grade), Potassium chloride (>99.5%, Fluka & Fisher, Reagent grade), Lithium

Chloride (99%, Sigma-Aldrich, Sigma ultra minimum), Lithium Bromide, Anhydrous (99%,

Matheson Coleman & Bell, Reagent grade), Lithium Perchlorate (99.99%, Aldrich, ReagentPlus

Grade) were used as received to prepare experimental solutions.

A micro pump (Micropump, Inc) and Teflon tubing, 2mm Quartz flow though cuvettes

(Uvonic Instruments, Inc), 10mm Quartz Flourometer cells (Sterna cell, Inc) and a calcium fluoride

(CaF2) thin window cell were used during the experiments. All the sample solutions were prepared in

deionized water with a pH of 6.70.

21

2.2 Experimental Section

Instrumentation and Method

UV- Visible Absorption Spectroscopy

A Varian Cary 50 Bio (Varian Corporation) single beam spectrophotometer was used to record the

UV-visible electronic spectra of the ground state species. All spectra were recorded at room temperature (293K-295K), using 10mm path length quartz cells.

Femtosecond Transient Absorption Spectroscopy

All the transient absorption experiments were carried out in the Ohio Laboratory for Kinetic

Spectrometry at Bowling Green State University. The experimental setup of the transient absorption

spectrometer (shown in Fig. 2.1) was used. It employs a Mai Tai diode-pumped mode-locked Ti:

Sapphire laser which generates pulses of 60 fs duration at 80 MHz repetition rate with average

power of 700 mW. These pulses are used to seed a Ti: Sapphire amplifier pumped by a Q-switched

Nd: YLF laser. The fundamental output of the Ti: Sapphire laser (Hurricane, Spectra Physics) is

1W at 800 nm generated as 100-fs pulses in a 1-kHz train. This output was divided into two parts:

one was used to excite the sample and the other was employed for generation of the probe light for

monitoring the optical absorption of the excited molecules generated by the pump pulse.

For pumping, 92% of the 800nm laser output was sent through a X3 telescope onto a second harmonic generator (SHG) then to a BBO crystal (2 mm length, cut at θ =32o for type I phase

matching), where it provided vertically polarized pulses at 267nm. (The typical energy conversion

efficiency of BBO crystal is 25%)

22

23

A white light continuum was used as the probe pulse. Prior to continuum generation the 800 nm beam goes to an optical variable delay stage, which provides an experimental time window of 1.6 ns with a step resolution of 6.6 fs. An iris diaphragm is used to modulate the beam cross-section to get the most stable white light. In order to avoid dipole-dipole interaction between the probe pulse and transition dipoles of the excited molecules, the polarizer was set up at the magic angle (54.7o) with respect to the pump beam polarization. A half-wave plate was used in front of the polarizer to adjust the intensity of the beam and a spherical (f = 10 cm) mirror was used to focus the beam on a

3 mm thick sapphire plate where the white light continuum was generated. Another spherical

(f = 5 cm) mirror was used after the sapphire plate to collimate and focus the white light beam onto the sample cell. The pump and probe beams overlap in the sample and the white light experiences an absorption dependence on the presence of the pump. A 750 nm short-pass filter was used to eliminate the remaining fundamental beam and a neutral density filter was used to adjust the intensity of the probe light. An f = 15 cm lens focused the white light beam on to a 400 μm fiber optic cable which serves as the input into a CCD based spectrometer for time- resolved spectral information (380-750 nm).

24

CHAPTER 3: RESULTS

3.1 Pure Water

The initial stage of the experiment studied the photo detachment of electrons from water

using two-photon excitation. Various types of laboratory available waters were tried and finally

came to a conclusion that deionized water was the best option. Therefore, throughout our

experiments distilled deionized water was used. A calcium fluoride (CaF2) thin window cell with

optical path of 0.09mm was used throughout the experiment. During the measurement the water cell was continuously monitored and kept perpendicular to the pump beam in order to avoid any air bubbles due to undesirable thermal effects caused by the pump beam.

A first attempt was to reproduce the absorption spectrum of the hydrated electron in water which has maximum absorption at 720nm (38). The absorption spectrum obtained by irradiating pure (distilled deionized) water with 267nm pulses is shown in the figure 3.1; the maximum

absorption occurred at 650nm. The difference in the reported value and that recorded here may be

due to the calcium fluoride (CaF2) window cell reacting in some manner with the solvent water. To

check this, the experiment was repeated using a 2 mm quartz cuvette. Figure 3.2 shows the

maximum absorption at 665 nm and another smaller absorption at around 720 nm.

25

0.05

0.04

0.03

0.02

0.01 Transient Absorption delta A (a.u.) A delta Absorption Transient

0.00 450 500 550 600 650 700 750 Wavelength / nm

Figure 3.1: Transient absorption spectrum of distilled deionized water (at 80ps delay time) in

0.09mm calcium fluoride (CaF2) thin window cell.

0.05

0.04

0.03

0.02

Transient Absorption delta A (a.u.) A delta Absorption Transient 0.01

0.00

450 500 550 600 650 700 750 Wavelength / nm

Figure 3.2: Transient absorption spectrum of distilled deionized water (at 114ps delay time) in 2mm quartz cuvette.

26

Figure 3.3 shows the time profile of the absorption at 650 nm. The decay of the signal was clearly biphasic

0.030

0.025

0.020

0.015 delta A delta 0.010

0.005

0.000

-0.005 0 1020304050607080 Time/ps

Figure 3.3 Transient absorption kinetic data for a 0.09 mm cell of neat water excited with 267nm femtosecond pulses (maximum absorption at 650 nm) Data: Data6_B Model: ExpDec2 The signal rose to the maximum value in close to ~3 ps and then decayed over a 20 ps to give a Chi^2/DoF = 1.4736E-7 long lived absorption change. According to theR^2 literature = 0.97998 this picosecond relaxation is due to the

y0 0.01455+ ±0.00692 geminate recombination of the hydrated electronA1 with 0.0074 the ±0.00064 H30 ion (12). t1 7.66179 ±1.40068 A2 0.00868 ±0.00606 t2 123.51478 ±153.13918

27

We evaluated the lifetime of the geminate recombination process at 267 nm pump by fitting the data to a double exponential decay. It was found that geminate recombination lifetime of water was 7.6 ± 1.4 ps. The literature records geminate recombination life times between 2.7 ± 5 ps and

12.2 ± 2.2 ps (12). The observed lifetime for the geminate process falls into the range in the literature.

28

3.2 Concentrated Alkali Metal Halide Solutions

To investigate the behavior of transient species in highly concentrated aqueous ionic halide

solutions, lithium compounds with different anions such as lithium hydroxide(LiOH) , lithium

chloride ( LiCl), lithium bromide( LiBr), lithium perchlorate ( LiClO4) and lithium iodide (LiI) were

selected. These lithium compounds were used to prepare the highest concentrated solution that

could be prepared.

Table 3.1: Table of lithium compounds with their respective highest concentrations

Lithium Compound Concentration (Mol/ L)

Lithium Hydroxide(LiOH) 3

Lithium Chloride( LiCl) 15

Lithium Bromide( LiBr) 15

Lithium Perchlorate ( LiClO4) 10

Lithium Iodide (LiI) 3

Transient absorption spectra were recorded for these lithium compounds at the

concentrations as shown in table 3.2. A 267 nm pump beam was used which produced 4 mJ/s

energy light on the sample as shown in the figure 3.4. A 2 mm quartz flow through cell was

employed in order to minimize buildup of photo products. The flow rate was of 2 ml/s. The

excitation of LiCl, LiBr and LiClO4 gave the highest absorption with well defined absorption

maxima; other solutions such as LiI and LiOH gave absorption spectra similar to that seen for water. LiCl at 15 M concentration gave an absorption maximum near 590 nm and 15M LiBr near

640 nm (figure 3.4). LiCl showed an absorption maximum around the same wavelength (590nm) as reported by Anbar and Hart in 1968 (38).

29

water

10M LiClO4 3M LiOH 15M LiCl 0.04 3M LiI 15M LiBr

0.02 delta A (a.u.)

0.00 500 550 600 650 700 750 Wavelength / nm

Figure 3.4 Plot of Transient Absorption (ΔA) vs Wavelength (nm) of different lithium compound (in 2 mm flow through cuvette) excited with 267nm femtosecond pulses room temperature.

Figure 3.5 shows the time profiles obtained by 267nm pump excitation of the above lithium compounds (water is included, for comparison). Qualitatively the kinetics exhibit two major regimes;

during the first few picoseconds there is a fast evolution of the photoinduced absorption followed

by a decay. The two species LiI and LiOH show different decay kinetics than the other species.

Those two traces showed the presence of a short lived (few ps) species (similar to H2O), whereas

other lithium compounds, such as LiCl, LiBr and LiClO4, showed decay kinetics in the hundreds of

ps range. This type of spectral evolution in the first 2-3 ps was obtained in all the experiments on

ionization of neat water (36).

30

Figure 3.6 shows the first 3 ps of the kinetic profiles, focusing on the signal rise. The data were fitted to the first order exponential rise using Origin software and the parameters are collected in table 3.2.

water

10M LiClO4 3M LiOH 15M LiCl 0.07 3M LiI 15M LiBr 0.06

0.05

0.04

0.03

0.02

Transiant Absorpton delta A (a.u.) 0.01

0.00 012345678910 Time / ps

Figure 3.5 Kinetic plots of data for different lithium compounds (in 2 mm flow through cuvette) excited with 267nm femtosecond pulses (at their respective maximum absorption wavelength).

31

0.06 water

10M LiClO4 3M LiOH 0.05 15M LiCl 3M LiI 15M LiBr 0.04

0.03

0.02

0.01 Transiant Absorpton delta (a.u.) A

0.00 0.0 0.5 1.0 1.5 2.0 2.5 3.0 Time / ps

Figure 3.6 Plot of transient absorption kinetic data for different lithium compounds (in 2 mm flow through cuvette) excited with 267nm femtosecond pulses (at their respective maximum absorption wavelength).

Table 3.2: First order fitting parameters for the transient absorption kinetic data

Compound Probing Wavelength Tab ( nm ) (ps) water 723 1.96 ± 0.47

LiCl 590 0.90 ± 0.07

LiBr 641 0.90 ± 0.10 LiI 658 1.42 ± 0.13

LiClO4 650 2.87 ± 0.76

LiOH 693 1.56 ± 4.67

32

Preliminary work showed a significant permanent difference in the ground state UV spectrum of LiCl solution before and after the radiation as shown the figure 3.7. but the other

1.0

before after

0.5 Absorption

0.0 200 300 400 500 600 700 800 Wavelength / nm

Figure 3.7 Plot of Absorption vs. Wavelength for LiCl before and after the experiment.

samples did not show such an effect. The UV spectrum, as in the figure 3.7, recorded after the

transient experiment showed a new peak at around 256 nm. This gives us clear evidence of

formation of a new species during transient experiment. According to the literature an absorption

− peak around this wavelength may be due to the formation of polychloride (Cl3 ) ions.

33

Reactions that are possible in the aqueous solution of LiCl in the formation of chlorine gas and polychloride ion included:

LiCl Li+ + Cl−

Cl− h ν Cl. + e−

. . Cl + Cl Cl2

. − .− Cl + Cl Cl2

.− . − Cl2 + Cl Cl3

.− .− − − Cl2 + Cl2 Cl3 + Cl

34

3.3. LiCl Solutions; Effect of Concentration

Experiments were carried out with a series of LiCl Solutions at different concentrations. The series of solutions with six different concentrations of LiCl were prepared using distilled deionized water. Transient absorption spectroscopic data (figure 3.7) were recorded along with the ground state absorption spectroscopic data (before and after the experiment). The kinetic data at the maximum absorption of the transient absorption spectrums was recorded at each different concentration.

1.0M LiCl 2.5M LICl 1.0 5.0M LiCl 7.5M LiCl 10.0M LiCl 15.0M LiCl

0.5 Normalized OD(a.u) Normalized

0.0 450 500 550 600 650 700 750 800 Time / ps

Figure 3.8 Transient absorption spectra of LiCl at different concentrations (in 2 mm flow through cuvette) excited with 267nm femtosecond pulses at room temperature.

35

The transient absorption spectra of LiCl show (figure 3.7) that the absorption maxima depend upon the concentration. With increasing concentration the absorption maxima are shifted towards shorter wavelengths. Thus 15M LiCl shows a maximum around 590 nm while 1M LiCl shows a maximum around 650nm. Thus, as the concentration of the LiCl decreased, the absorption maxima were red shifted and approaches that of the hydrated electron in pure water (720 nm).

Time profiles at the absorption maxima were normalized and fitted to a two exponential decay function. The fitted values for absorption and decay times of the high and low concentrated

LiCl solutions were compared. This shows two-photon photoexcitation yield different absorption and decay kinetics for high and low concentrated solutions at nearly the same total excitation energy.

These results show a slower rate for absorption growth and decay for the high concentration samples and a higher rate for low concentration samples (Figure 3.8).

36

Kinetic Plot for 15M LiCl at~590 nm (a)

1.0

0.9

0.8 Data: Data1_B Model: ExpDec2 0.7 Chi^2/DoF = 0.00095 R^2 = 0.96865 0.6 y0 0.83352±0.16607 A1 -2.09872 ±0.06145 t1 0.9549 ±0.07619 0.5 A2 0.2446 ±0.07819 t2 8.51945±13.11733 Normalized delta (a.u) A 0.4

0.3

-10123456789101112 Time / ps

Kinetic Plot of 1M LiCl at ~646nm (b) 1.0

0.9

0.8

Data: Data2_B 0.7 Model: ExpDec2

Chi^2/DoF = 0.00137 R^2 = 0.93252 0.6 y0 0.86401±0.01689 A1 -3.31737 ±0.62238 0.5 t1 0.88451±0.18038 A2 0.64752±0.87095 Normalized delta (a.u.) A t2 2.14072±1.26569 0.4

0.3 -10123456789101112 Time / ps

Figure 3.9 Kinetic plots of highest (a) and lowest (b) concentrated solutions of LiCl in deionized water (in 2 mm flow through cuvette) excited with 267nm femtosecond pulses and solid line is the two-exponential decay fit for the experimental data.

37

The transient absorption and kinetic data of other highly concentrated alkali halides were compared to check whether there was an effect on cation. Highly concentrated solutions of NaCl and KCl were used in this regard. The highest concentrations that could be achieved with NaCl and

KCl were 5M and 3M respectively. The variation in the transient absorption spectrum in high concentrated solution was compared with 15M LiCl solution as in the figure 3.9.

0.15 3M KCl 5M NaCl 15M LiCl

0.10

0.05 Transient Absoption OD (a. u.) (a. OD Absoption Transient

0.00 500 600 700 800 Wavelength / nm

Figure 3.10 Transient absorption spectra of highest concentrated solutions KCl, NaCl and LiCl in deionized water (in 2 mm flow through cuvette) excited with 267nm femtosecond pulses at room temperature.

38

The high concentration solutions showed a significant difference in absorption maxima of

KCl and NaCl from LiCl (figure 3.9). Solutions of those salts at 1M did not show this behavior

(figure 3.10). The lower concentration solutions (1M) of all the above alkali halide showed their maximum absorption close to the same wavelength.

1.2 KCl Nacl LiCl 1.0

0.8

0.6

0.4 Normalized OD(a.u.)

0.2

0.0 500 600 700 800 Wavelength(nm)

Figure 3.11 Transient absorption spectra of one molar (1M) solutions KCl, NaCl and LiCl in deionized water (in 2 mm flow through cuvette) excited with 267nm femtosecond pulses at room temperature.

Thus, low concentrated alkali halide solutions show transient absorption maxima no matter what cation is present in the solution. The kinetic data of the highest concentrated solutions of those

39 three alkali halide was recorded at the respective maxima of their transient absorption spectra

1.1

1.0

0.9

0.8

0.7

0.6 3M KCl 5M NaCl 0.5 15M LiCl

Normalized delta (a.u.) A 0.4

0.3

0.2 -10123456789101112 Time(ps)

Figure 3.12 Kinetic plots of highest concentrated solutions of LiCl, KCl and NaCl. (in 2 mm flow through cuvette) excited with 267nm femtosecond pulses.

(Figure 3.11) and fitted to a two exponential decay function. The fitting parameters of LiCl, KCl and

NaCl are shown in the table3.3.

Table 3.3 Fitting parameters of highest concentrated solutions of LiCl, KCl and NaCl

Compound λ(nm) A1 t1(ps) A2 t2(ps)

3M KCl 638 -2.54±0.35 0.80±0.13 0.30±0.10 7.83±0.10

5M NaCl 637 -2.38±0.35 0.86±0.18 0.27±0.07 7.58±17.15

15M LiCl 590 2.09±0.06 0.95±0.07 0.24±0.07 8.51±13.11

40

CHAPTER 4: DISCUSSION

Femtosecond laser studies of the excitation and relaxation in pure water (deionized) have

been carried out at 267 nm pumping with two-photon absorption. The quantum yield of hydrated

electron at 267nm pump was calculated using the two photon excitation confirmed at 282nm. The

value for β (β= (1.9 ± 0.5) x 10-12 m/W) was known here and assumed that value for β at 267nm does not change significantly with wavelength. The measured dynamics of geminated recombination for excitation at 267 nm were comparable to the result obtained by the other groups (7, 42, 43) who used different excitation wavelength. However it was not possible to reproduce exactly the transient absorption spectrum of the hydrated electron in pure water reported in earlier. It is not clear why it is so but it may be connected with the orientation of the probe beam as it entered the fiber optic cable leading to the spectrograph.

Two photon induced detachment of electrons from highly concentrated electrolyte solutions have been studied on the femtosecond time scale, as reported in the chapter 3. It was found that high concentrated solution of lithium compounds with different anion such as hydroxide (LiOH) ,

chloride ( LiCl), bromide( LiBr), perchlorate ( LiClO4) and iodide (LiI) all generated hydrated

electrons. It is impossible to prepare the same high concentration solution for each compound;

therefore the highest possible concentration solutions were prepared. These compounds produced

their transient absorption maxima between the wavelength of 550nm and 700nm. The kinetics for

these five different compounds fell into two major categories. Specifically LiI and LiOH differed

41 from the others by the presence of a short lived species and then showed longer decay kinetics,

whereas other lithium compounds, such as LiCl, LiBr and LiClO4, shows typical decay kinetics of

long lived species.

Recently it was observed that the near-IR absorption associated with a transient

electron-Cl atom pair completely disappears in less than 2ps in NaCl aqueous solution pumped by

two 310 nm photons (54). A semi-quantum molecular dynamic simulation study of the

─ − 3P Æ 4S transitions of aqueous Cl , proposed that the transient (Cl….e )aq pair can be likened to

an excess electron in the solvent shell of chlorine atom (54, 57). This preceeds the complete

detachment of an electron from excited halide ion yielding a long lived 1s-like hydrated electron

ground state (54, 57). This description of NaCl appears to be valid the current experimental kinetic

data for LiCl as shown in figure 3.5. The model postulates that, a three body complex may form,

comprised of halide (X), electron (e) and the cation (Y+) denoted as { X…e…Y+}. This complex

shows complete electron detachment occurring in the vicinity of an aqueous cation. It was found

that for the aqueous solution of NaCl, this electronic process occurs with the characteristic time of

850fs and yields a significant contribution to a polaron-like state{ Y+: e─} (54).

The first order exponential fit of transient absorption data of 15 M LiCl shows its absorption

process with time of 955±76 fs as in the table 3.2 which compares favorably with the reported data

(850 fs) for the electronic process in NaCl (54). It is thus likely that LiCl follows the same

mechanistic process during the formation of the transient absorption signal.

The concentration dependence study of photo-electron detachment in LiCl shows

significant variation in transient absorption spectra and kinetics. The absorption maxima are red

shifted as the salt concentration was decreased. The kinetic data revealed that rate of formation of

the absorption for low concentrated solution of LiCl is higher than that of high concentrated LiCl

solution. This may be due to the higher viscosity (and density) of the concentrated solutions

42 compared to pure water. Based on Stokes law, one can expect to observe slower electron diffusion in highly concentrated solution which could slow down the decay kinetics. Another factor, well known in electrochemistry, is that dragging by the slowly moving ionic atmosphere could slow down the rapidly migrating electron. These interactions mainly affect the geminate partners that eventually reduced the rate of decay path in the concentrated solution.

Further experiments were carried out to compare transient data of lithium chloride with other alkali halides such as potassium chloride (KCl) and sodium chloride (NaCl). The fitted parameters in the table 3.3 shows that NaCl, KCl absorption and decay kinetics happen at a higher rate and time profiles were fell into same time regime but that for LiCl happen in a slower rate.

Therefore it can propose that there could be an affect of cation in the transient absorption and the kinetics of high concentrated solutions alkali halides. During the kinetic analysis of low concentrated solution of all above showed similar absorption kinetics but decay kinetics varies from one to another.

Conclusions

Two photon femtosecond laser studies of the excitation and relaxation of pure water showed that, geminate lifetimes were comparable with reported data but the recorded absorption spectrum was blue shifted.

Experiments carried out for five different highly concentrated lithium compounds with different anion indicated that these compounds produced their maximum transient absorption between the wavelengths of 550nm to 700nm. The time profiles relevant to these compounds s fell into two major categories. LiI and LiOH showed different kinetic profiles from the others by the initial presence of a short lived species along with longer decay kinetics, whereas other lithium

compounds, such as LiCl, LiBr and LiClO4, shows typical decay kinetics of long lived species only.

The experimental data for absorption kinetics of LiCl fitted by single exponential function showed

43 that the absorption rise time falls into same time regime as reported for the purported three body complex of NaCl. This leads to the thought that LiCl may follow the same mechanism.

The concentration dependence study of LiCl showed significant change in its transient absorption spectra and kinetics. The absorption maxima red shifted as the concentration was changed from high to low. The kinetic data revealed that rate of absorption for low concentrated solution of LiCl is higher than that of high concentrated LiCl solution.

The comparison of LiCl with KCl and NaCl showed that there is a significant contribution from the cation for higher concentrated solution to alter their kinetic and transient absorption. But in the case of low concentrated solutions that effect is minimal.

44

REFERENCE

1. Baltuska, A; Dissertation, 2000, University of Groningen (Hydrated electron dynamics explored with

5-fs optical pulses);W. Weyl, Pogg. Ann. 1864, 123, 350.

2. Stein, G. Diss. Faraday Soc., 1952, 12, 227.

3. Platzman, R. L. Natl. Res. Coun. Publ., 1953, 305, 34.

4. Hart, E. J.; Boag, J. W. J. Am. Chem. Soc., 1962, 84, 4090.

5. Matheson, M.S.; in Physical Chemistry, An Advanced Treatise, Eyring, H. Ed.; Academic press:

New York, 1967–75; Vol. III , Chapter 10, 533-628

6. Maronceli, M.; Flemming, G. R. J. Chem. Phys., 1988, 89, 5044.

7. Stowell, M. H. B.; McPhillips, T. M.; Rees, D. C.; Soltis, S. M.; Abresch, E.; Feher, G. Science,

1997, 276, 812.

8. Steinberg-Yfrach, G.; Liddell, P. A.; Hung, S.C.; Moore, A. L.; Gust, D.; Moore T. A. Nature,

1997, 385, 239.

9. Schindelin, H.; Kisker, C.; Schlessman, J. L.; Howard, J. B.; and Rees, D. C. Nature, 1997, 387,

370.

10. Borja, M.; Dutta, P. K. Nature,1993,362, 43.

11. Sykora, M.; Kincaid, J. R. Nature, 1997, 387, 162.

12. Khaselev, O.; Turner, J. A. Science, 1998, 280, 425.

13. Bach, U.; Lupo, D.; Comte, P.; Moser, J. E.; Weissörtel, F.; Salbeck, J., Spreitzer, H.; Grätzel,

M. Nature, 1998, 395, 583.

14. Reuther, A.; Laubereau, A.; Nikogosyan, D. N. J. Phys. Chem., 1996, 100, 16794.

15. Gauduel, Y.; Pommeret, S.; Migus, A.;Antonetti, A. J. Phys. Chem. , 1989, 93, 3880.

16. Mizouno, M.; Yamaguchi, S.; Tahara, T. J. Phys. Chem., 2005, 109, 5257.

45

17. Garrett et al.Chem. Rev. 2005, 105, 355-389.

18. De Boeij, W. P.; Pshenichnikov, M. S.; and Wiersma, D. A. J. phys. Chem, 1996, 100, 11806.

19. M. Maroncelli, J. Mol. Liq. 1993, 57, 1.

20. De Silvestri, S.; Weiner, A.M.; Fujimoto, J. G.; Ippen, E. P. Chem. Phys. Lett., 1984,112, 195.

21. Becker, P. C.; Fragnito, H. L.; Bigot, J.-Y.; Brito Cruz, C. H.; Fork, R. L.; Shank, C. V. Phys.Rev.

Lett., 1989, 63, 505 .

22. Nibbering, E. T. J.; Wiersma, D. A.; Duppen, K. Phys. Rev. Lett , 1991, 66, 2464.

23. Passino, S. A.; Nagasawa, Y.; Flemming, G. R. Phys. Rev. Lett , 1997, 107 , 6094

24. Jimenez, R;. Flemming, G. R.; Kumar, P. V.; Maroncelli, M. .Nature , 1994, 369, 471.

25. Kevan, L; .Acc. Chem. Res., 1981, 14, 138.

26. Schwartz B. J.; Rossky, P. J.; J. Chem. Phys. 1994,101, 6902.

27. Staib A.; Borgis, D. J. Chem. Phys. 1995, 103, 2642 .

28. Kim, K. S.; Park, I.; Lee, S.; Cho, K.; Lee, J. Y.; Kim, J.; Joannopoulos, J. D. Phys. Rev. Lett.,

1996, 76, 956.

29. Rossky, P. J.; Schnitker, J. J. Phys. Chem., 1988, 92, 4277.

30. Schwartz, B. J.; Rossky, P. J., J. Chem. Phys. 1994,101, 6902.

31. Schnitker, J.; Motakabbir, K.;. Rossky, P. J.; Friesner, R., Phys. Rev. Lett., 1988, 60, 456.

32. Lugo, R.; Delahay, P. J. Chem. Phys., 1972, 57, 2122.

33. Assel, M.; Laenen, R.; Laubereau, A. J. Phys. Chem. A, 1998, 102, 2256.

34. Katner, N. R. Electron-Solvent und Anion Solvent Interactions; Kevan, L., Webster, B.C., Eds.;

Elsevier: Amsterdam, 1976, Chapter 1.

35. Bartczak, W. M.; Hilczer, M.; Kroh, J. J. Phys. Chem , 1987, 91, 3834.

36. Banerjee, A.; Simons, J. J. Chem. Phys. 1978, 68, 415.

46

37. Carmichael, I. J. Phys. Chem., 1980, 84, 1076.

38. Kajiwara, T.; Funabashi, K.; Naleway, C. Phys. Rev. A , 1972, 6, 808 .

39. Reid, P. J.; Silva, C.; Walhout, P. K.; Barbara, P. F. Chem. Phys. Lett. , 1994, 228, 658.

40. Assel, M.; Laenen, R.; Laubereau, A. J. Phys. Chem. A, 1998,102, 2256.

41. Crowell, R.A; Lian, R.; Shkrob, L.A.; Bartels, D.M.; Chen X.; Bradforth, S.E. J. Chem. Phys.,

2004, 120, 11712-11725.

42. Boag, J.W.; Hart, E.J. Nature, 1963, 197, 45.

43. Hart, E.J.; Anber ,M. J. Phys. Chem., 1965, 69, 1244-1247.

44. Robertson, W.H.; Diken, E.G.; Price. E.A.; Shin, J.W.; Johnson, M.A. science, 2003, 299, 1367.

45. Masamura, M. J. Chem. Phys., 2002, 117, 5257.

46. Nienhuys, H.K.; Lock, A.J.; Van Santen, R.A.; Bakker, H.J. J. Chem. Phys., 2002, 117, 8021.

47. De Violet, Ph. F.; Review of Chemical Intermediates; Verlag Chemie International, Inc.: France, 4,

121-169, 1981.

48. Taubu, H. J.Am.Chem.Soc., 1946, 68,611.

49. Cater, P. R.; Davidson, N. J. Phys. Chem., 1952, 56, 877.

50. Grossweiner, L.I; Matheson, M.S. J. Phys. Chem., 1957, 61, 1089.

51. Matheson, M.S.; Mulac, W.A.; Rabani, J. J. Phys. Chem. 1963, 67, 2613.

52. Grossweiner, L.I; Swenson, G.W.; Zwicker, E.F.; Science , 1964, 141, 919.

53. Kimura, Y.; Alfano, J. C.; Walhout, P.K.; Barbara, P. F. J. Phys. Chem., 1994, 98, 3455.

54. Brozek-Pluska, B.; Gliger, D.; Hallou, A.; Malka, V.; Gauduel, Y. A.; J. Radiat. Phys, Chem.,

2005, 72, 149-157.

55. Rey, R.; Guardia, E.; J. Phys. Chem.., 1992, 96, 4712-4718.

56. Sprik, M.; Klein, M. L.; Watanabe, K.; J. Phys. Chem.., 1990, 94, 6483-6488.

57. Staib, A.; Borgis, D.; J. Chem. Phys., 1996, 104, 9027-9039.