experimental and computational studies of the unimolecular

175
University of Wollongong Thesis Collections University of Wollongong Thesis Collection University of Wollongong Year Experimental and computational studies of the unimolecular rearrangements of sulphonated azo dyes and phenoxide anions in the gas-phase Aravind Ramachandran University of Wollongong Ramachandran, Aravind, Experimental and computational studies of the unimolecular rearrangements of sulphonated azo dyes and phenoxide anions in the gas-phase, Master of Science - Research thesis, School of Chemistry, Faculty of Science, University of Wollongong, 2008. http://ro.uow.edu.au/theses/2625 This paper is posted at Research Online.

Upload: others

Post on 16-Apr-2022

2 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Experimental and computational studies of the unimolecular

University of Wollongong Thesis Collections

University of Wollongong Thesis Collection

University of Wollongong Year

Experimental and computational studies

of the unimolecular rearrangements of

sulphonated azo dyes and phenoxide

anions in the gas-phase

Aravind RamachandranUniversity of Wollongong

Ramachandran, Aravind, Experimental and computational studies of the unimolecularrearrangements of sulphonated azo dyes and phenoxide anions in the gas-phase, Master ofScience - Research thesis, School of Chemistry, Faculty of Science, University of Wollongong,2008. http://ro.uow.edu.au/theses/2625

This paper is posted at Research Online.

Page 2: Experimental and computational studies of the unimolecular
Page 3: Experimental and computational studies of the unimolecular

Experimental and Computational Studies of

the Unimolecular Rearrangements of

Sulphonated Azo Dyes and Phenoxide

Anions in the Gas-Phase

A thesis submitted in fulfilment of the requirements for the award of the degree

MASTER OF SCIENCE (RESEARCH)

from

UNIVERSITY OF WOLLONGONG

by

Aravind Ramachandran

School of Chemistry, Faculty of Science

2008

Page 4: Experimental and computational studies of the unimolecular

CERTIFICATION

I, Aravind Ramachandran, declare that this thesis, submitted in fulfilment of the requirements for the award of Master of Science (Research), in the Faculty of Science, University of Wollongong, is wholly my work unless otherwise referenced or acknowledged. The document has not been submitted for qualifications at any other academic institutions. Aravind Ramachandran April 14, 2008

Page 5: Experimental and computational studies of the unimolecular

i

TABLE OF CONTENTS TABLE OF CONTENTS_________________________________________________________ I

LIST OF FIGURES____________________________________________________________ III

LIST OF TABLES______________________________________________________________ V

LIST OF SCHEMES ___________________________________________________________ VI

ABSTRACT ________________________________________________________________ VIII

ACKNOWLEDGEMENT ________________________________________________________ X

CHAPTER ONE: INTRODUCTION ________________________________________________ 1

1.1 IONIZATION TECHNIQUES _____________________________________________________ 2

1.2 TANDEM MASS SPECTROMETERS _______________________________________________ 7

1.3 ELUCIDATION OF GAS PHASE FRAGMENTATION MECHANISMS BY MASS SPECTROMETRY AND COMPUTATIONAL STUDY _______________________________________________________ 11

1.4 IDENTIFICATION OF TWO UNUSUAL ANION REARRANGEMENT IN THE GAS- PHASE____________ 18

1.5 COMPUTATIONAL METHODS__________________________________________________ 20

1.5.1 Basis set ___________________________________________________________ 21

1.5.2 Theoretical model ____________________________________________________ 23

REFERENCES FOR CHAPTER ONE_________________________________________________ 25

CHAPTER TWO: EVIDENCE OF AN INTRAMOLECULAR NUCLEOPHILIC AROMATIC SUBSTITUTION: AN EXPERIMENTAL AND THEORETICAL STUDY OF THE GAS-PHASE REARRANGEMENT OF AZO DYES______________________________________________ 28

ABSTRACT _________________________________________________________________ 28

2.1 INTRODUCTION ___________________________________________________________ 30

2.2 RESULTS AND DISCUSSION __________________________________________________ 34

Page 6: Experimental and computational studies of the unimolecular

ii

2.2.1 Collision induced dissociation (CID) of deprotonated azo dyes _________________ 34

2.2.2 Solution phase labelling _______________________________________________ 38

2.2.3 Evidence towards the mechanism for N2 loss ______________________________ 41

2.2.3.1 Positive ion experiment ___________________________________________________ 41

2.2.3.2 Dianion Experiment ______________________________________________________ 42

2.2.3.3 Authentic product experiment ______________________________________________ 44

2.2.3.4 Change of nucleophile____________________________________________________ 50

2.2.3.5 CID of Substituted analogues ______________________________________________ 52

2.2.3.6 Electronic structure calculation _____________________________________________ 54

2.3 CONCLUSION ____________________________________________________________ 70

2.4 EXPERIMENTAL___________________________________________________________ 71

2.4.1 Mass spectrometry ___________________________________________________ 71

2.4.2 Synthesis of azo compounds ___________________________________________ 71

2.4.3 Synthesis of authentic amine compounds _________________________________ 72

2.4.4 Calculations ________________________________________________________ 72

REFERENCES FOR CHAPTER TWO ________________________________________________ 74

CHAPTER THREE: COMPUTATIONAL INVESTIGATION OF THE REARRANGEMENT AND FRAGMENTATION OF PHENOXIDE ANION IN THE GAS-PHASE _____________________ 76

ABSTRACT _________________________________________________________________ 76

3.1 INTRODUCTION ___________________________________________________________ 78

3.2 MATERIALS AND METHODS __________________________________________________ 85

3.3 RESULTS AND DISCUSSION __________________________________________________ 86

3.4 CONCLUSION ___________________________________________________________ 108

REFERENCES FOR CHAPTER THREE ______________________________________________ 109

APPENDIX-1 _______________________________________________________________ 111

APPENDIX-2 _______________________________________________________________ 112

APPENDIX-3 _______________________________________________________________ 153

Page 7: Experimental and computational studies of the unimolecular

iii

LIST OF FIGURES Fig-1.1: Components of a basic mass spectrometer ___________________________________ 2

Fig-1.2: Schematic of the electrospray ionization process ______________________________ 5

Fig-1.3: Components of a tandem in space mass spectrometer__________________________ 8

Fig-1.4: Cross section of a linear ion trap __________________________________________ 10

Fig-1.5: The MS/MS spectrum of the 3-methyl- 2,3 epoxybutoxide anion__________________ 15

Fig-1.6: MS/MS spectrum of the 2,2-dimethyloxetan-3-olate ___________________________ 16

Fig-1.7: Reaction coordinate diagram of the Payne rearrangement ______________________ 17

Fig-2.1: Range of azo compounds________________________________________________ 32

Fig-2.2: CID Spectrum of the azodye 4-amino-3-(phenyldiazenyl)naphthalene-1-sulfonic acid as

shown in Fig-2.1(a) _______________________________________________________ 35

Fig-2.3: CID mass spectrum of the mass selected product ion at m/z 262 _________________ 37

Fig-2.4: CID mass spectrum of the mass selected product ion at m/z 298 _________________ 38

Fig-2.5: MS spectrum of deuterium labeled sample __________________________________ 39

Fig-2.6: CID mass spectrum of completely deuterium exchanged sample _________________ 40

Fig-2.7: CID mass spectrum of the positive ions of the azodye 4-amino-3-(phenyldiazenyl)

naphthalene-1-sulfonic acid as in (Fig-2.1(a))___________________________________ 41

Fig-2.8: CID mass spectrum of [M-2H]2- ions from the azo dye 4-amino-3-((4sulfophenyl)diazenyl)

naphthalene-1-sulfonic acid as in Fig-2.1(e) ____________________________________ 43

Fig-2.9: Comparision of MS/MS spectra of the secondary amine product with that of the MS3

spectrum of the azo dye anion, 4-amino-3-(phenyldiazenyl)naphthalene-1-sulfonate (Fig-

1(a)) ___________________________________________________________________ 46

Fig-2.10: Comparision of MS/MS spectra of the amine product with that of the MS3 spectrum of

the dianions from 4-amino-3-((4-sulfophenyl)diazenyl)naphthalene-1-sulfonic acid (Fig-

2.1(e))__________________________________________________________________ 47

Fig-2.11: Comparision of MS/MS spectra of the monoanions from disubstituted amine product

with that of the MS3 spectrum of the monoanions from 4-amino-3-((4-sulfophenyl)diazenyl)

naphthalene-1-sulfonic acid (Fig-2.1(e)) _______________________________________ 50

Page 8: Experimental and computational studies of the unimolecular

iv

Fig-2.12: CID mass spectrum of negative ions from 4-hydroxy-(phenyldiazenyl) naphthalene-1-

sulfonic acid (Fig-1(f)) _____________________________________________________ 51

Fig-2.13: CID mass spectrum of negative ions from 4-amino-3-((4-nitrophenyl)diazenyl)

naphthalene-1-sulfonic acid (Fig-1(c)) ________________________________________ 53

Fig-2.14: CID mass spectrum of negative ions from 4-amino-3-((4-methoxyphenyl)diazenyl)

naphthalene-1-sulfonic acid (Fig-1(b)) ________________________________________ 54

Fig-2.15: Structures of the Tautomer (1M1) and Meisenheimer transition state (TS1) on the 4-

amino-3-(phenyldiazenyl)benzene-1-sulfonate potential energy surface optimized at

B3LYP/6-31+G(d) level of theory_____________________________________________ 57

Fig-2.16: Reaction coordinate diagram for the intramolecular rearrangement of the model diazo

anion, 4-amino-3-(phenyldiazenyl)benzene-1-sulfonate model system calculated at

B3LYP/6-31+G(d) level of theory_____________________________________________ 60

Fig-2.17: Reaction coordinate diagram for the intramolecular rearrangement of the 4-hydroxy-

(phenyldiazenyl)benzene-1-sulfonate model system calculated at B3LYP/6-31+G(d) level of

theory __________________________________________________________________ 66

Fig-3.1: CID Mass Spectrum of the labeled ethoxide anion (Ph16O(CH2)218O-) _____________ 83

Fig-3.2: Reaction coordinate diagram for the fragmentation of the phenoxide anion in the gas-

phase calculated at B3LYP/6-311++G(d,p) level of theory _________________________ 88

Fig-3.3: Structures of the stationary points on the phenoxide potential energy surface optimized

at B3LYP/6-311++G(d,p) level of theory _______________________________________ 91

Fig-3.4: Potential energy diagram showing the involvement of benzene-oxide and oxepin

structures in the fragmentation of phenoxide calculated at B3LYP/6-311++G(d,p) level of

theory __________________________________________________________________ 95

Fig-3.5: Structures of the stationary points on the benzene-oxide and oxepin potential energy

surface optimized at B3LYP/6-311++G(d,p) level of theory ________________________ 98

Fig-A_2.1:Structures of the stationary points on the 4-amino-3(phenyldiazenyl)benzenesulfonate

potential energy surface optimized at B3LYP/6-311++G(d,p) level of theory __________ 130

Page 9: Experimental and computational studies of the unimolecular

v

LIST OF TABLES Table-2.1: Optimized stationary points calculated for the loss of N2 from 4-amino-3-

(phenyldiazenyl)benzenesulfonate anion ______________________________________ 61

Table-3.1: Optimized stationary points calculated for the loss of CO from phenoxide anion __ 107

Table-A_1.1: Tandem mass spectra of [M-H+]- Ions from azo dyes anions para substituted phenyl

and sulphonic acid analogues. _____________________________________________ 111

Table-A_2.1: The Cartesian coordinates for all the stationary points for the fragmentation of 4-

amino-3(phenyldiazenyl)benzenesulfonte calculated at B3LYP/6-31+G(d) level as illustrated

in Figure-2.16. __________________________________________________________ 112

Table-A_2.2: Optimized stationary points calculated for the loss of N2 from 4-hydroxy-3-

(phenyldiazenyl)benzenesulfonate Anion _____________________________________ 131

Table-A_2.3: Optimized stationary points calculated for the loss of N2 from (4-(4-

sulfonatophenylamino) benzene-1-sulfonic acid) _______________________________ 132

Table-A_2.4: The Cartesian coordinates for all the stationary points for the fragmentation of 4-

amino-3(phenyldiazenyl)benzenesulfonte calculated at B3LYP/6-31+G(d) level as illustrated

in Figure-2.17. __________________________________________________________ 133

Table-A_2.5: The Cartesian coordinates for all the stationary points for the fragmentation of (4-

(4-sulfonatophenylamino) benzene-1-sulfonic acid) calculated at B3LYP/6-31+G(d) level as

illustrated in Figure-2.17. _________________________________________________ 147

Table-A_3.1: The Cartesian coordinates for all the stationary points for the fragmentation of

phenoxide anion calculated at B3LYP/6-31+G(d) level as illustrated in Figure-3.3. ____ 153

Page 10: Experimental and computational studies of the unimolecular

vi

LIST OF SCHEMES Scheme-1.1: Payne rearrangement_______________________________________________ 14

Scheme-1.2: Payne rearrangement in an unsymmetrical system________________________ 14

Scheme-1.3: Alternative cyclization to oxetane system _______________________________ 14

Scheme-1.4: Fragmentation mechanism from the two isomeric epoxides _________________ 16

Scheme-1.5: Loss of nitrogen from azodye anions. __________________________________ 18

Scheme-1.6: Fragmentation of phenoxide anions____________________________________ 19

Scheme-1.7: Evidence of benzene-oxide and oxepin structures instead of phenoxide anion.__ 19

Scheme-2.1: Proposed dissociation-recombination reaction mechanism__________________ 33

Scheme-2.2: Proposed intramolecular nucleophilic aromatic substitution reaction. __________ 33

Scheme-2.3: Nucleophilic aromatic substitution reaction in the gas-phase ________________ 34

Scheme-2.4: Fragmentation pathways for major ions in the CID spectrum of the azodye 4-amino-

3-(phenyldiazenyl)naphthalene-1-sulfonate (Fig-2.1(a))___________________________ 36

Scheme-2.5: Azo-hydrazone type fragmentation ____________________________________ 36

Scheme-2.6: Resonance contributing structures of the secondary amine _________________ 45

Scheme-2.7: Substitution nucleophilic mechanism in the monoanions of 4-amino-3-((4-

sulfophenyl)diazenyl)naphthalene-1-sulfonate __________________________________ 49

Scheme-2.8: Aromatic substitution reaction involving tautomeric structure ________________ 55

Scheme-2.9: Resonance contributing structure of the tautomer_________________________ 56

Scheme-2.10: Calculated reaction mechanism-1 for 4-amino-3-(phenyldiazenyl) naphthalene-1-

sulfonate _______________________________________________________________ 58

Scheme-2.11: Calculated reaction mechanism-2 for 4-amino-3-(phenyldiazenyl) naphthalene-1-

sulfonate _______________________________________________________________ 59

Scheme-2.12: Calculated reaction mechanism-1 for 4-hydroxy-3(phenyldiazenyl) naphthalene-1-

sulfonate _______________________________________________________________ 64

Scheme-2.13: Calculated reaction mechanism-2 for 4-hydroxy-3(phenyldiazenyl) naphthalene-1-

sulfonate _______________________________________________________________ 65

Scheme-2.14: Preliminary calculation on the dianion system ___________________________ 69

Page 11: Experimental and computational studies of the unimolecular

vii

Scheme-3.1: Binkley and coworker’s mechanism of phenoxide fragmentation _____________ 81

Scheme-3.2: Smiles rearrangement in the Gas-Phase________________________________ 81

Scheme-3.3: Fragmentation of phenoxy ethoxide anion_______________________________ 82

Scheme-3.4: Fragmentation of the labeled phenoxy ethoxide anion (Ph16O(CH2)218O-) ______ 83

Scheme-3.5: Fragmentation of the 13C labeled phenoxy ethoxide anion (PhO(CH2)2O-) ______ 84

Scheme-3.6: Fragmentation of Perbenzoate anion___________________________________ 85

Scheme-3.7: Proposed fragmentation pathway for phenoxide decomposition ______________ 86

Scheme-3.8: Resonance contributing structure of the intermediate IM3 __________________ 92

Scheme-3.9: Calculated fragmentation mechanism for phenoxide anion__________________ 93

Scheme-3.10: Calculated pathway of phenoxide fragmentation _________________________ 94

Scheme-3.11: Calculated pathway for oxepin fragmentation ___________________________ 99

Scheme-3.12: Resonance contributing structure of the intermediate IM5 ________________ 100

Scheme-3.13: Resonance structures of methanolate anion ___________________________ 101

Scheme-3.14: Rearrangement of the oxepin to benzene-oxide anion ___________________ 102

Scheme-3.15: Fragmentation pathway for the loss of CHO. ___________________________ 103

Scheme-3.16: Possible rearrangement between ketene and oxepin intermediates _________ 104

Scheme-3.17: Rearrangement of the labeled perbenzoate anion through initial nucleophilic

attack at the ortho position_________________________________________________ 105

Scheme-3.18: Rearrangement of the labeled perbenzoate anion through initial nucleophilic

attack at the ipso position _________________________________________________ 106

Page 12: Experimental and computational studies of the unimolecular

viii

ABSTRACT

The tandem mass spectrometer is an ideal tool to probe unimolecular

reactions of ions. Of particular interest are reactions involving skeletal rearrangement of

ions prior to dissociation. Given that such unimolecular reactions occur in the absence of

complicating factors such as solvent and counter ions, within the vacuum environment of

the mass spectrometer, computational methods employing molecular orbital and density

functional theories are ideally suited to examine the reaction mechanisms.

A surprising rearrangement was identified by electrospray ionization

tandem mass spectrometry of anions of azo dyes and the rearrangement was found to

effect a loss of the azo moiety bridging aromatic rings as nitrogen. Even though the

fragmentation reaction was previously reported, we are unaware of any conclusive

mechanistic study. In the present thesis, by combination of tandem mass spectrometry

and computational methods, we have identified that the rearrangement proceeds via an

initial tautomerization, followed by nucleophilic aromatic substitution reaction (Scheme-

1).

NH2

SO3

NN

SO3

NHN

HN

SO3

HN NHN

SO3

NHN N

H

SO3

HN-N2

H

SO3

HN

Scheme-1

For the past two decades, phenoxide anions were reported to undergo

unimolecular fragmentation resulting in the loss of CO. The present thesis presents an

electronic structure calculation study on this unimolecular fragmentation, where in it was

Page 13: Experimental and computational studies of the unimolecular

ix

identified that the loss of CO occurs via reaction pathways involving ketene like

intermediates and transition states (Scheme-2).

O CO

CO + CO

H HCO

CO

CO

+ CO

HC

OC

O

C

OCO

H

Scheme-2

Page 14: Experimental and computational studies of the unimolecular

x

ACKNOWLEDGEMENT 1. I would like to thank my supervisor Dr Stephen Blanksby, not only for his guidance

and supervision, but also for supporting me through my tough times of my University life.

2. I would also like to thank all the lab members for their support through my degree.

3. Finally, I would like to thank my Mom and Dad for support, guidance and

encouragement in times of uncertainty.

Page 15: Experimental and computational studies of the unimolecular

1

CHAPTER ONE: INTRODUCTION

Mass spectrometry is the branch of science which deals with the study of

gaseous ions, with or without fragmentation, which are characterised by their mass to

charge ratio (m/z) and their relative abundances. The technique evolved from the

discovery of positively charged rays by Goldstein in 1886, which was followed by Wein

(1898) studying their electric and magnetic properties.1 In the early 1900’s J. J Thompson

built his parabola mass spectrograph to measure the charge to mass ratio (z/m) for several

ionic species.2 This was followed by Aston’s work, to build instruments to obtain

accurate measurements of the ratios of the stable isotopes of many known elements.3 The

first commercial mass spectrometer dedicated to elucidate organic structure in the

petroleum industry was built in 1934.4 From these beginnings, mass spectrometers were

used for mass separation and to elucidate organic structures.

All mass spectrometers consist of three basic components, namely, the ion

source, the mass analyser and the ion detector (Fig-1.1). The role of the ion source is to

vaporize the analyte in the vacuum environment of the mass spectrometer, and to convert

it into an ionised form. After ions are formed in the source, they are accelerated into the

mass analyser where they are separated in vacuum according to their mass to charge ratio

through the use of electric and/or magnetic fields. Finally, the ions are passed into an ion

detector where the ions are destroyed in a process generating electric current that is

amplified and recorded. Correlation of the position of the mass-analyser and the detection

of ion current yields the mass spectrum.5

Page 16: Experimental and computational studies of the unimolecular

2

Fig-1.1: Components of a basic mass spectrometer

Although sample is consumed destructively, the technique is very

sensitive and only trace amount of materials are needed in the analysis. The emergence of

the tandem mass spectrometer offered chemists an opportunity to observe and study the

unimolecular fragmentations of a mass-selected ion. Such studies are of fundamental

importance to understanding the intrinsic reactivity of ions in the absence of solvent and

furthermore provide a characteristic “fingerprint” for structure elucidation in analytical

applications.

1.1 Ionization techniques

The traditional method of ion production in mass spectrometry is electron

ionization (EI) in which the gaseous sample molecules are bombarded with a fast (70 eV)

beam of electrons. Chemical ionization (CI) mass spectrometry, since its introduction by

Munson and Field, has also become a widely used technique.6 CI is related to EI except

that ionization of a reagent gas occurs first rather than direct ionization of the sample

molecule. This is followed by transfer of charge to the sample by various chemical

processes.6

In a typical CI mass spectrometer, the reagent gas (M) is subjected to

high-energy 70 eV electron impact at high pressures. In this process a radical cation (M+.)

is formed (Eq-1).

M M+ e + 2e(70eV) (1)

Page 17: Experimental and computational studies of the unimolecular

3

The most commonly used reagent gas is methane and its electron

bombardment at high pressures forms the methane radical cation (Eq-2). The CH4+. then

reacts rapidly with another molecule of methane to form CH5+ and CH3

. (Eq-3). CH5+ is

an extremely powerful proton donor and will react to protonate most molecular analytes

(A), where A is any organic compound (Eq-4).6

CH4 + e CH4 2e+ (2)

CH4 + CH3 + CH5CH4 (3)

CH5 + A AH + CH4 (4)

Chemical ionization is widely used to produce positive ions, but can also

be used to form negative ions in the gas-phase. However, under negative ion CI

conditions, the reagent gas has three important functions. Firstly, the reagent gas serves to

thermalize the electrons by elastic scattering and dissociative ionization process.

Secondly, the reagent gas should capture the thermal electrons by either electron capture

or dissociative electron capture process. Thirdly, the reagent gas should transfer the

negative charge to the analyte molecule by a variety of chemical reactions such as proton

transfer, H2+. abstraction, charge exchange, or nucleophilic displacement.7

The most popular reagent gas used for negative ion CI is a mixture of

methane and nitrous oxide. The thermal electrons produced by electron bombardment of

a mixture of methane and nitrous oxide are captured by N2O to produce O-. (Eq-5). The

atomic oxygen anion radical reacts with CH4 to form hydroxide ion and methyl radical

(Eq-6).

Page 18: Experimental and computational studies of the unimolecular

4

RCOO

RCH3COO

O

CH3

C

O

H3C R

+

++

N2O + e N2 + O (5)

O + CH4 OH + CH3 (6)

Hydroxide, being a strong gas-phase base, abstract protons from a wide

range of compounds such as carboxylic acids, alcohols, thiols, ketones and amino acids.8

The other known method is to capture thermal electrons by using a

mixture of nitrous oxide and a non-reactive buffer gas such as N2. This technique

produces the atomic oxygen radical anion that can ionize carbonyl compounds by

nucleophilic displacement (Eq-7, Eq- 8).9

(7) (8)

The CI methods described here provide access to a range of molecular

ions causing very little fragmentation. Hence CI is popular as a technique to gain

molecular mass information.10

After the introduction of electrospray ionization (ESI) by Fenn and co-

workers, ESI has become another popularly used soft ionization method. The method

ionizes samples at atmospheric pressure and has proven to be the best method to

accomplish ionization of bio-molecules as it allows for large, non-volatile molecules to

be analysed directly from the liquid phase.11 This is not possible in a CI source, because

the analytes must be first vapourised, which could cause decomposition of bio-molecules.

In ESI-MS (Fig-1.2), the analyte solution is pumped at a very low flow

rate through a steel capillary, which is maintained at a very high positive or negative

Page 19: Experimental and computational studies of the unimolecular

5

voltage, typically 3-5 kV. The resulting field at the tip of the capillary charges the surface

of the emerging liquid. As a result, the liquid protrudes from the capillary tip forming

what is known as a ‘Taylor cone’ (Fig-1.2).12, 13 Droplets from the Taylor cone detach

from the surface when the Coulombic repulsion of the surface is equal to the surface

tension of the solution.14 Molecular ions are formed from these droplets by one of the two

proposed mechanisms, namely, (a) the Coulomb fission mechanism or (b) the ion

evaporation mechanism. The Coulomb fission mechanism assumes that increased charge

density due to solvent evaporation causes large droplets to divide into smaller and smaller

droplets until eventually only charged molecules remain. By contrast, the ion evaporation

mechanism, assumes that the increased charge density that results from solvent

evaporation eventually causes Coulombic repulsion to overcome the droplet surface

tension, resulting in a release of ions directly from the droplet surfaces.12 Regardless of

the formation mechanisms, ESI generates vapour phase ions that can be analysed

according to mass to charge ratio (m/z).

Fig-1.2: Schematic of the electrospray ionization process.12

Page 20: Experimental and computational studies of the unimolecular

6

There are four major mechanisms by which the analytes becomes charged

and get discharged as ions during ESI. Such charging can occur either in solution or in

the gas-phase (Taylor cone). They are: (a) ionization in solution, (b) ionization in the gas-

phase, (c) ionization through electro-chemical oxidation and reduction, and (d)

ionization/interaction in solution and gas-phase.12 (a) Ionization in solution is the primary

method by which analytes are ionized during ESI. In general, analytes are ionized in

solution by addition of a base or an acid. Organic and biomolecules possessing basic

sites are protonated in solution by addition of an acid to form their respective [M+H]+

ions, while analytes with acidic moieties are deprotonated by addition of a base to form

their respective [M-H]- ions . The analytes that already exist as ions in solution separate

from counter ions during the formation of Taylor cone and are eventually transferred to

the gas phase by charge separation and/or ion evaporation mechanism.12 (b) ESI is known

to produce, large quantities of charged solvent/analytes in the gas-phase. Sometimes,

analytes uncharged in solution can undergo gas-phase interaction, such as a simple proton

transfer with the charged solvent molecules during the formation of Taylor-cone. In this

way, analytes that evaporate from the droplet as neutrals can become charged through

gas-phase interactions and eventually form ions by charge separation and/or ion

evaporation mechanism.12 (c) Electro-chemical reactions can convert an uncharged

analyte into a charged form by achieving charge balance via an oxidation reaction in the

capillary, in positive ion mode, and reduction reaction in negative ion mode. The analyte

best suited, is one that is easily oxidized or reduced, to form a solution stable ionic

species in the capillary. The stable ions formed, would then be transferred to the gas-

phase by charge separation or/and ion evaporation mechanism.12 One such example of

Page 21: Experimental and computational studies of the unimolecular

7

analyte that is known to ionize efficiently by electro-chemical ionization in the ESI

source are the ferrocene derivatives of a variety of alcohols. They were found to form the

corresponding ferrocinium cation of the various alcohols in the gas-phase.15 Polar

analytes that do not contain acidic or basic groups can be charged by preliminary

interaction in solution followed by ionization in the gas-phase by charge separation

or/and ion evaporation mechanism. One such example is the formations of stable ions by

adduct formation. Cole and Zhu have shown that the chlorinated adduct of aniline is

formed when chlorinated solvents such as chloroform are used in the ESI process. This

method allows for successful negative ion ESI detection of molecules that lack an acidic

site.16 Similarly, addition of sodium, lithium, ammonium and potassium salts to samples

with weakly basic or polar, neutral samples aid in the formation of positive ions by cation

attachment in the ESI process.12, 17

1.2 Tandem mass spectrometers

The soft ionization techniques such as chemical ionization and

electrospray ionization generally produce molecular species with low internal energy and

hence show little or no fragmentation. An effort to increase the fragment ion information,

to better characterize the structure of analytes, saw the introduction and the use of the

tandem mass spectrometer. The tandem mass spectrometer is similar to a normal mass

spectrometer, except that it has two mass analysers, with a dissociation region in-between

(Fig-1.3). The dissociation region, is the region where ions may be excited energetically

using a number of different methods namely: (a) Collision Induced Dissociation or (CID),

in which the ions are excited by collisions with neutral atoms or molecule, (b) Photon

Induced Dissociation or (PID), in which the ions are activated by absorption of photons,

Page 22: Experimental and computational studies of the unimolecular

8

(c) Surface Induced Dissociation or (SID), in which the ions are excited by collisions

with solid or viscous liquid, and (d) Electron Capture Dissociation or (ECD), in which the

ions are fragmented by dissociative recombination with thermal electrons.18, 19 Each of

these methods leads to the production of charged and neutral fragments through bond

cleavage in the molecular ions.

In analytical applications, the tandem mass spectrometer may be used to

selectively detect an analyte of interest from complex mixtures. The first mass analyser,

transmits only ions of a particular m/z ratio into the dissociation region, where the

fragmentation of ions occur. The second mass analyser is scanned to pass, in turn, the

products of dissociation along with the molecular ion onto the detector. This technique is

routinely used to selectively detect and identify analytes of interest from complex

mixtures by simultaneous identification of fragment ions together with the molecular ion.

Since two mass analysis steps are involved, tandem mass spectrometry is often referred to

as MS/MS or MS2 (Fig-1.3).

Fig-1.3: Components of a tandem in space mass spectrometer (Redrawn from ref)20

The most common method of adding energy to the ions is by collisional

activation. In this process, the mass selected ion collides with a neutral atom or molecule,

in a dissociation region. The overall process occurs in two steps. The first step is very fast

and corresponds to the collision between the precursors ions (mp+) and the target gas (n)

bringing the ions into the excited state (mp+*) (Eq-9). The second step is the unimolecular

Page 23: Experimental and computational studies of the unimolecular

9

decomposition of the activated ions to form charged (mf+) and the neutral (mn) fragments

(Eq-10). 19, 21

mp+ + N → mp

+* (9) mp

+* → mf

+ + mn (10)

The choice of the collision gas is important in order to prevent the

reactions of ions with the target gas and also to prevent significant deflections from their

trajectories. For these reasons, He, Ar or N2 are typically used as the collision gas.21

Tandem mass spectrometry was originally developed using electric and

magnetic sector mass spectrometers and was later expanded to include arrangements of

mass/energy analysers, such as multiple quadrupoles22, sectors20, and time-of flight.23

These are called tandem-in-space mass spectrometers (Fig-1.3) since primary and

secondary mass analyses are performed by two analysers that are coupled in series.21

The concept was further extended to tandem-in-time mass spectrometers.

One such example of tandem-in-time mass spectrometer is an ion-trap mass spectrometer.

An ion trap is a quadrupole mass analyser, and can be conceptually thought of a closed

loop of radiofrequency that helps in the storage of ions. The quadrupole mass analyser

can be 3-dimensional or 2-dimensional (linear) in space (Fig-1.4). The mass analyser is

often filled with buffer gas such as helium at an appropriate pressure to reduce the loss of

ions by ion repulsions.5, 24

Page 24: Experimental and computational studies of the unimolecular

10

Fig-1.4: Cross section of a linear ion trap

Four major processes occur within the mass analyser that contributes to

performing a successful tandem mass spectrometry experiment with an ion-trap. These

are (i) storage of ions, (ii) isolation of ion of interest: ions of interest can be isolated

according to their mass to charge (m/z) ratio, by expelling all other ions by applying a

destabilizing radio frequency, (iii) collision-induced dissociation (CID) of mass selected

ions: energy can be imparted to the mass-selected ions by collisions with the helium

buffer gas to form fragment ions, and (iv) detection of precursor and fragment ions.24

Additionally, the analyzer can also isolate a particular fragment ion arising

from the dissociation of the precursor ion. The mass-selected ions can be further activated

to undergo fragmentations. These types of experiments are popularly known as MSn

experiments where n represents the number of mass selection and dissociation

experiments. These experiments further help in characterization of the structure of the

fragment ions.24

Page 25: Experimental and computational studies of the unimolecular

11

1.3 Elucidation of gas phase fragmentation mechanisms by mass spectrometry and computational study

The soft ionization techniques such as chemical ionization and

electrospray ionization generally produce molecular species with an even number of

electrons, most often by addition or abstraction of a proton, thus producing positive or

negative ions in the gas-phase. Most often, fragmentation of ions involve simple

heterolytic bond cleavages leading to the formation of a stable fragment ion and a neutral.

For example, heterolytic cleavages of a single bond can occur in a

molecular cation forming a smaller cation and displacing a stable neutral (Eq-11). The

initiation of the cleavage depends on the ability of the heteroatom Y to attract the electron

pair. In general, halogens cleave more readily followed by oxygen, sulphur, nitrogen and

carbon. With reference to neutral formation, stability of neutral formed depends on the

ease with which it can leave.25, 26

R YH R + YH (11)

In some instances, cleavages through cyclization displacement are

common in the positive ions. Such cleavages require stabilization of the charge by distant

groups in the same ion and the cleavage occurs via a nucleophilic substitution type

mechanism.25, 26 For example, in Equation-12 the oxygen atom, displaces the cyclic

amino group by a nucleophilic substitution type mechanism forming a stable six member

cyclic ion and a stable six member unsaturated amine.25

(12)

CHN

ON

HN

O+

N

Page 26: Experimental and computational studies of the unimolecular

12

The fragmentation of even electron negative ions also follows a number

of predictable rules. In negative ions, simple heterolytic cleavage leading to the

formation of an anion-neutral complex is thought to be an important mechanistic step.

The anion-neutral complex may dissociate directly or undergo ion-molecule chemistry

prior to dissociation.27 For example, collisional activation of the acetate anion leads to

the formation of methyl anion via loss of carbon dioxide. The fragmentation

mechanism would involve a simple heterolytic initial formation of anion-neutral

complex between the methyl anion and carbon dioxide. The anion-neutral complex

then dissociates to form methyl anion via a loss of CO2 ( Eq-13).28

(13)

With reference to the ion-molecule chemistry within the anion-neutral

complex, the anion part of the complex can also undergo proton transfer or nucleophilic

substitution reaction with the neutral, to form a new anion-neutral complex, which would

be followed by dissociation.27 Proton transfer reactions within the anion-complex are

widely observed. For example, collisional activation of the ethoxide anion leads to the

formation of acetaldehyde enolate anion via the loss of H2. The loss of H2 from the

ethoxide anion, would involve the formation of an ion-neutral complex between hydride

ion and acetaldehyde. The hydride ion, then deprotonates the neutral portion to form the

second anion-neutral complex between acetaldehyde enolate anion and H2. The resulting

anion-neutral complex undergoes direct dissociation, driven by the negative charge to

form the enolate anion and H2 neutral (Eq-14).29

H3C CO2 CH3 CO2 CH3 + CO2

Page 27: Experimental and computational studies of the unimolecular

13

(14)

In some instances, nucleophilic substitution chemistry can also occur

within the anion-neutral complex. For example, collisional activation of N-ethyl

ethanamide, results in the formation of NCO-, via loss of propane. The loss of propane

from the N-ethyl ethanamide, would involve the initial formation of ion-neutral complex

between methyl anion and ethyl isocyanate. The methyl anion then acts as a nucleophile,

displacing NCO- from the ion-neutral complex, resulting in the production of propane

and the isocyanate anion (Eq-15).27

(15)

In some instances, where simple cleavage is energetically unfavourable,

the ion undergoes skeletal rearrangement prior to fragmentation. The skeletal

rearrangement sometimes shares similar mechanistic character to some of the named

rearrangements that would occur in the condensed phase for the same ion. The course of

the reaction in the condensed phase however often depends on the nature of the solvent

and the counter ions. Such factors, could directly affect the nature of the reacting species.

Whereas in the gas-phase phase, the reaction occurs in the absence of the solvent and

counter ion effects thus providing insight into the intrinsic reactivity of ions.30

Characterization of molecular rearrangements most often requires a

combination of mass spectrometric and computational techniques. One such example is

H H2 H2C C H2 + H2C CH

OH3C CH O

HH

OH2C

H

CHO

H3C CO

N CH2CH3 CH3 C N CH2CH3 C N + C3H8OO

Page 28: Experimental and computational studies of the unimolecular

14

the study of Payne rearrangement in the gas-phase by Bowie and co-workers.31 The

Payne rearrangement in the condensed phase involves an intramolecular cyclisation of

the alkoxide anion at the carbon of an epoxide ring to open the epoxy ring and form a

new epoxide ring (Scheme-1.1).32

OO

OO

Scheme-1.1: Payne rearrangement

In the condensed phase, the rearrangement is known to be affected by

three important features: (a) In an unsymmetrical system, such as methyl substituted 2,3

epoxy propoxide anion, the predominant product is normally that which has the more

substituted epoxide (Scheme-1.2).33, 34

OO

OO

A B

Scheme-1.2: Payne rearrangement in an unsymmetrical system

(b) The reaction takes place under the influence of the base, a protic solvent with a high

dielectric constant. (c) In the condensed phase, there is no indication of alternative

cyclisation to form a four membered oxetane system (Scheme-1.3).33, 34

OO

OO

O

O

A B

C

Scheme-1.3: Alternative cyclization to oxetane system

Page 29: Experimental and computational studies of the unimolecular

15

In order to verify the occurrence of the Payne rearrangement in the gas-

phase, the Payne rearrangement product of the 2,3-epoxy propoxide anion was

synthesized and was subjected to CID. The MS/MS spectrum of the synthesized Payne

rearrangement product matched with that of dimethyl substituted 2,3-epoxy propoxide

anion (Fig-1.5), thus validating the occurrence of the Payne rearrangement.31

Fig-1.5: The MS/MS spectrum of the 3-methyl- 2,3 epoxybutoxide anion 31 To further investigate the possibility of occurrence of the competing

mechanism (Scheme-1.3), the oxetane intermediate was synthesized independently and

was subjected to Collision-Induced Dissociation. The spectrum obtained matched with

that of the 3-methyl-2,3-epoxybutoxide anion within the experimental uncertainty (Fig-

1.6).31

Page 30: Experimental and computational studies of the unimolecular

16

Fig-1.6: MS/MS spectrum of the 2,2-dimethyloxetan-3-olate 31

The result suggests that the Payne rearrangement occurs in the gas-phase.

Moreover the competing oxetane mechanism also occurs in the gas-phase, with the three

structures A, B, C (Scheme-1.3) equilibrating in the gas-phase. 31

From the data it was suggested that the fragment ions at m/z 71 and m/z 43

(Fig-1.5) correspond to the fragment ions from the two epoxides. (Scheme-1.4) Since the

ion at m/z 71 corresponding to the loss of CH2O is the base peak (Fig-1.5), it was

concluded that the more substituted epoxide is the predominant product from the Payne

rearrangement. 31

OO

O

H+ CH2O

m/z 71

OO CH2CHO + CH3COCH3

m/z 43

Scheme-1.4: Fragmentation mechanism from the two isomeric epoxides31

To add to the experimental work, a computational study was conducted to

better understand the energetics of the Payne rearrangement. Figure-1.7 illustrates the

reaction coordinate diagram for the rearrangement with respect to the methyl substituted

Page 31: Experimental and computational studies of the unimolecular

17

2,3 epoxy propoxide anion. In the calculated mechanism, the Payne rearrangement of the

epoxide anion (A) to its isomer (B) occurs over a barrier of 33 kJ mol-1. The result clearly

suggests that the barrier to Payne rearrangement is low and is easily accessible.

Surprisingly, the isomeric propoxide B is 14 kJ mol-1 lower than the parent proponoxide

anion. The result suggests that the structure B is more energetically favourable than the

more substituted isomer A. This contradicts the experimental data and the solution phase

observation but is not significant overall as these are not in equilibrium condition.31

Furthermore the calculations suggest that the rearrangement to oxetane intermediate is

feasible from both the two isomeric epoxides A and B. However, the rearrangement to

the oxetane intermediate requires three-fold more energy (117 kJ mol-1 from A) than that

required for the Payne rearrangement.31

O0

-14

O

O

O

O

33

O

O O

O

O

A

B

CC

-43-43

O

O117

O

O

117

TS1

TS2TS3

Fig-1.7: Reaction coordinate diagram of the Payne rearrangement. All energies are in kJ mol-1. 31

Page 32: Experimental and computational studies of the unimolecular

18

1.4 Identification of two unusual anion rearrangement in the gas- phase

In this thesis, we have identified two unusual anion rearrangements in the gas-phase, one with the azodye anions and other with the phenoxide anions. (i) In anions of azodyes an unusual fragment ion corresponding to the loss of 28 Da

was observed. Both Richardson and co-workersa and Gaskall and co-workersb in their

studies of the anions of azodyes have also reported the loss of 28 Da. The loss was

proposed to be the loss of N2 from the azodyes but neither group has proposed a

mechanism to account for this observation. What is so interesting about this

fragmentation is that the azo moiety bridging the two aromatic rings is lost as nitrogen,

leaving the two aromatic rings intact (Scheme-1.5). By a combination of previously

described mass spectrometric technique together with complementary computational

study we elucidate the reaction mechanism for this unusual fragmentation

SO3

NH2

NN

CID

-N2

m/z 326

m/z 298

Scheme-1.5: Loss of nitrogen from azodye anions. (ii) In fragmentations of the anions of phenoxide, product ions consistent with the

loss of CO were observed. Even though mass spectrometric study was conducted by both

a Richardson, S. D.; McGuire, J. M.; Thructon, A. D.; Baughman, G. L., Org. Mass Spectrom. 1992, 27, 289. b Sullivan, A. G.; Gaskell, S. J., Rapid Commun. Mass Spectrom. 1997, 11, 803

Page 33: Experimental and computational studies of the unimolecular

19

Binkley and co-workersc and Bowie and co-workersd to elucidate the reaction mechanism

(Scheme-1.6), we are unaware of any complementary computational study.

OC

O CO

+ CO

A B C D

Scheme-1.6: Fragmentation of phenoxide anions

Furthermore, in a recent study by Blanksby and co-workers e on the

dissociation of perbenzoate anions fragment ions corresponding to the formation of

phenoxide anion were observed following the loss of CO2. Somewhat surprisingly,

computational studies support the formation of benzene-oxide and oxepin structure as

product ions, instead of the phenoxide anion (Scheme-1.7).The result might suggest that

the benzene-oxide and oxepin intermediates may be linked via facile rearrangement to

phenoxide anion in the gas-phase. In this thesis we would elucidate the possible

rearrangement mechanism of benzene-oxide and oxepin structure to phenoxide by

computational techniques.

OOO

ortho

ipso

OO

O OO

O OCO2

OO

OOOO OC

O

O

CO2O

Scheme-1.7: Evidence of benzene-oxide and oxepin structures instead of phenoxide anion. c Binkley, R. W.; Fletcher, T. W.; Winnik, W., J. Org. Chem. 1992, 57, 5507 d Eichinger, P. C. H.; Bowie, J. H.; Hayes, R. N., J. Am. Chem. Soc. 1989, 111, 4224 e Harman, D. G.; Ramachandran, A.; Gracanin, M.; Blanksby, S. J., J. Org. Chem. 2006, 71, 7996

Page 34: Experimental and computational studies of the unimolecular

20

1.5 Computational methods

Computational methods as such have evolved as a complementary

technique for probing the mechanisms of ionic fragmentations and rearrangements in the

gas-phase. Given that reactions are observed in the vacuum environment of the mass

spectrometer, the experimental setup is quite close to a truly isolated system, which is the

primary requirement of molecular orbital or density functional theory calculations. The

combination of mass spectrometry techniques with theoretical calculations allows: (i)

characterization of structures, in which the traditional bonding rules are violated, (ii)

prediction of potentially stable isomers and evaluation of their stability, (iii) formulation

of potential reaction pathways, evaluation of heats of formation of intermediates not

amenable by experimental techniques, (iv) estimation of gas-phase acidity or basicity for

different acids.35

The theoretical models used in the gas phase chemistry are based on ab

initio molecular orbital theory. The ab initio molecular orbital theory is based on the

fundamental laws of quantum mechanics and requires no experimental parameters. The

computations are solely based on the laws of quantum mechanics and on the values of

small number of physical constants such as speed of light, charge and mass of electrons

and nuclei and Planck’s constant. Quantum mechanics states that energy and other related

properties of the molecule may be obtained by solving the Schrödinger equation ( Eq-16).

36, 37

ĤΨ= EΨ (16)

Page 35: Experimental and computational studies of the unimolecular

21

Here, Ĥ is the Hamiltonian operator, which when applied to the wave

function Ψ, gives the electronic energy E of the wave function. The ab initio methods,

compute solutions to the equation using a series of rigorous mathematical approximations.

The two most important approximations used by ab initio methods to solve the

Schrödinger equation are the basis set and the theoretical model.36, 37

1.5.1 Basis set

The basis set is a mathematical representation of the molecular orbitals

within the molecule. The basis set can be interpreted as restricting each electron to a

particular region of space.36 GAUSSIAN 0338

offers a number of pre-defined basis sets,

which approximate the molecular electronic wave function, using a linear combination of

atomic orbitals. The basis set assigns a group of basis functions to each atom within the

molecule to approximate its orbitals. The number and types of basis functions that they

contain classify basis sets. These basis functions themselves are composed of a linear

combination of Gaussian functions. Such basis functions are called contracted function

and the component Gaussian functions are referred to as primitives.36 Minimal basis set

such as STO-3G use fixed-size atomic-type orbitals and approximate all orbitals to be of

the same shape.39,40 However, such approximations are often not adequate. There are

several types of extended basis sets, namely split valence basis set, polarized basis set and

diffuse functions, which consider the higher orbitals of the molecule and accounts for size

and shape of molecular charge distributions.

The split valence basis sets, increases the size of the orbital generated, by

increasing the number of basis functions per atom.36 An example of split valence basis set

is 6-31G, which uses two sizes of basis function for each valence orbital. It uses one

Page 36: Experimental and computational studies of the unimolecular

22

Gaussian type function for inner shell and for its outer shell uses two basis functions, one

is a linear combination of three primitives and other is a diffuse primitive.36, 41 Similarly,

triple split valence basis sets, such as 6-311G, use three sizes of contracted functions for

each valence orbital type.42 Hence, split valence basis sets allow orbitals to change in size,

but not to change shape.

Polarised basis sets remove this limitation by adding orbitals with angular

momentum beyond what is required for the ground state to the description of each

atom.36, 40 Polarised basis sets acknowledge and account for the fact that, when atoms are

brought close to each other, their charge distribution causes polarization effect i.e. the

positive charge is drawn to one side while the negative charge is drawn to the other side,

thus resulting in distorting the shape of the atomic orbitals. This results in the ‘s’ orbital

sharing the characteristics of the ‘p’ orbitals and the ‘p’ orbital to share the characteristics

of the ‘d’ orbitals. Hence the use of a polarised basis set results in a better approximation

than a split valence basis set alone, which treated atomic orbitals as existing only as

‘s’, ’p’, ’d’ etc.40 Polarised basis sets are represented by (d) or (p), e.g. 6-31G(d) and 6-

31G(d,p). In 6-31G(d) basis set, d functions are added to the heavy atoms, whereas in 6-

31G(d,p), p functions are added to the s orbitals of hydrogen atoms in addition to d

functions on heavy atoms.36, 40

Diffuse functions are large-size versions of s and p type function and

hence allow the molecular orbital to occupy a larger region in space.36, 43 The diffuse

functions are important for systems where electrons are relatively far away from the

nucleus, e.g., molecules with lone pair of electrons, anions and other systems with

significant negative charge, systems in their excited states, systems with low ionization

Page 37: Experimental and computational studies of the unimolecular

23

potentials, and description of absolute acidities. Diffuse functions are designated with a

‘+’ e.g. 6- 31+G(d), 6-31++G(d,p). In 6-31+G(d), diffuse functions are added to the

heavy atoms whereas in 6-31++G(d,p), diffuse functions are added to the hydrogen atoms

as well.36, 43

1.5.2 Theoretical model

A hybrid method, ‘Becke’s three-parameter non-local potential’ or

otherwise known as B3LYP was employed in the study.44 The hybrid method uses a

combination of Hartree-Fock (HF) and Density Functional Theory (DFT) theories.36, 44 In

the HF method, wave function Ψ0 is written as a product of one electron wave function or

spin orbitals (Eq-17). The spin orbitals Ψi are expanded as a linear combination of basis

functions (Φ).

Ψi=Σ Ci uΦu (17) HF solves the wave function assuming that each electron is moving in an

average electron distribution produced by other electrons. But it does not take into

account the electron correlation effect, i.e., how electrons in a molecular system respond

to each others motion. It is also inefficient for accurate modelling of energetics of

reaction and bond dissociation energies, which is essential for examining ion dissociation

in this study.36, 45

In recent years, DFT methods have gained popularity, as it takes into

account the electron correlation effect. DFT achieves greater accuracy than HF theory at

a modest increase in computing time.36 DFT theories are based on the fact that the

ground-state electronic energy is completely determined by the electron density (ρ). DFT

methods compute electron correlation via general functionals of the electron density. The

Page 38: Experimental and computational studies of the unimolecular

24

approximate functionals used by DFT methods, partition the electronic energy as sum of

three terms, kinetic energy, attraction between nuclei and electron, Coulomb repulsion,

and an exchange-correlation term. The exchange-correlation term accounts for the

remainder of the electron-electron interaction, which itself are divided into separate

exchange and correlation components. The difference between the various DFT methods

lies in the way the exchange and correlation components are treated by the functionals.36,

46 A variety of functionals have been defined such as, local exchange and correlation

functionals and gradient-corrected functionals (non-local). The local exchange and

correlation involve only the values of electron spin densities whereas the gradient-

corrected functionals involve both the electron spin densities and their gradients.36 A

popular gradient-corrected exchange functional is one proposed by Becke in 1988.47 A

widely used gradient-corrected correlation functional is the LYP functional of Lee, Yang

and Parr.48 The combination of Becke’s gradient-corrected exchange functional and the

LYP functional forms the B-LYP method.

Several functionals have been defined, which treat exchange functionals as

linear combination of HF, local and gradient-corrected exchange terms. Such functionals

are referred to as the hybrid functionals. The most popular one of this type is B3LYP

method.44 B3LYP method has become increasing popular because of its ability to

generate accurate results with only a modest increase in computing resource.

Page 39: Experimental and computational studies of the unimolecular

25

References for chapter one

1. Goldstein, E., Berl. Ber. 1886, 39, 691. 2. Thomson, J. J., Rays of Positive Electricity and Application to Chemical Analyses.

2nd ed.; Longmans, Green & Co.: London, 1922. 3. Aston, F. W., Philosophical Magazine (1798-1977) 1920, 40, 628. 4. Smythe, W. R.; Rumbaugh, L. H.; West, S. S., Phys. Rev. 1934, 45, 724. 5. Downard, K., Mass spectrometry: A foundation course. Royal Society of

Chemistry: Cambridge, 2004. 6. Munson, M. S. B.; Field, F. H., J. Am. Chem. Soc. 1966, 88, (12), 2621. 7. Harrison, A. G., Chemical ionization mass spectrometry. 2nd ed.; CRC Press:

Boca Raton, 1992. 8. Smith, A. L. C.; Field, F. H., J. Am. Chem. Soc. 1977, 99, (20), 6471. 9. Harrison, A. G.; Jennings, K. R., J. Chem. Soc., Faradays, Trans. 1 1976, 72,

1601. 10. Chapman, J. R., Practical Organic Mass Spectrometry: A Guide for Chemical

and Biochemical Analysis. J. Wiley: Chichester; New York, 1993. 11. Mann, M.; Meng, C. K.; Fenn, J. B., Proceeding of the 36th Annual

Conference on Mass Spectrometry and Applied topics 1988, 1207-08. 12. Cech, N. B.; Enke, C. G., Mass Spectrom. Rev. 2001, 20, 362. 13. Fenn, J. B.; Mann, M.; Meng, C. K.; Wong, S. F.; Whitehouse, C. M., Science

1989, 246, (8), 64. 14. Taflin, D. C.; Ward, T. L.; Davies, J. D., Langmuir 1989, 5, 376. 15. Van Berkel, G. J.; Quirke, J. M E.; Tigani, R. A.; Dilley, A. S., Anal. Chem. 1998,

70, 1544. 16. Cole, R. B.; Zhu, J., Rapid Commun. Mass Spectrom. 1999, 13, 607. 17. Saf, R.; Mirtl, C; Hummel, K., Tetrahedron Lett. 1994, 35, 6653.

Page 40: Experimental and computational studies of the unimolecular

26

18. Busch, K. L.; Glish, G. L.; McLuckey, S. A., Mass Spectrometry/Mass

Spectrometry: Technques and Applications of Tandem Mass Spectrometry. VCH: NewYork: 1988.

19. Sleno, L.; Volmer, D. A., J. Mass Spectrom. 2004, 39, 1091. 20. McLafferty, F. W.; Todd, P. J.; McGilvery, D. C.; Baldwin, M. A., J. Am. Chem.

Soc. 1979, 102, (10), 3360. 21. Shukla, A. K.; Futrell, J. H., J. Mass Spectrom. 2000, 35, 1069. 22. Yost, Y. A.; Enke, C. G., J. Am. Chem. Soc. 1978, 100, 2274. 23. Glish, G. L.; Goeringer, D. E., Anal. Chem. 1984, 56, 2291. 24. Hoffman, E. De; Stroobant, V., Mass spectrometry: principles and applications.

Chichester; New York: Wiley: 2001. 25. McLafferty, F. W., Org. Mass Spectrom. 1980, 15, (3), 114. 26. McLafferty, F. W., Interpretation of mass spectra. Third Edition ed.; University

Science Books: Mill Valley, California, 1980. 27. Bowie, J. H., Mass Spectrom. Rev. 1990, 9, 349. 28. Budzikiewicz, H.; Poppe, A.; Stockyl, D., Int. J. Mass Spectrom. Ion Processes

1983, 47, 217. 29. Hayes, R. N.; Sheldon, J. C.; Bowie, J. H.; Lewis, J. Chem. Soc. Chem. Commun.

1984, 21, 1431. 30. Eichinger, P. C. H.; Dua, S.; Bowie, J. H., Int. J. Mass Spectrom. Ion Processes

1994, 133, 1. 31. Dua, S.; Bowie, J. H.; Taylor, M. S.; Buntine, M. A., Int. J. Mass Spectom. Ion

Processes 1997, 165/166, 139. 32. Payne, G. B., J. Org. Chem. 1962, 27, 3819. 33. Behrens, C. H.; Ko, S. Y.; Sharpless, B.; Walker, F. J., J. Org. Chem. 1985, 50,

5687. 34. Bonini, C.; Guikiano, C.; Righi, L.; Rossi, L., Tetrahedron Lett. 1992, 7429. 35. Alcami, M.; Mo, O.; Yanez, M., Mass Spectrom. Rev. 2001, 20, (4), 195.

Page 41: Experimental and computational studies of the unimolecular

27

36. Foresman, J. B.; Frish, E., Exploring chemistry with electronic structure method. Gaussian, Inc: Pittsburg, PA, 1996.

37. Hehre, W. J. Random, L. Schleyer, P. v. R. Pople, J. A., Ab Initio Molecular

Orbital Theory. Wiley: New York, 1986. 38. Gaussian 03 (Revision C.02). Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.;

Scuseria, G. E.; Robb, M. A.; Cheeseman, J. R.; Montgomery, Jr., J. A.; Vreven, T.; Kudin, K. N.; Burant, J. C.; Millam, J. M.; Iyengar, S. S.; Tomasi, J.; Barone, V.; Mennucci, B.; Cossi, M.; Scalmani, G.; Rega, N.; Petersson, G. A.; Nakatsuji, H.; Hada, M.; Ehara, M.; Toyota, K.; Fukuda, R.; Hasegawa, J.; Ishida, M.; Nakajima, T.; Honda, Y.; Kitao, O.; Nakai, H.; Klene, M.; Li, X.; Knox, J. E.; Hratchian, H. P.; Cross, J. B.; Bakken, V.; Adamo, C.; Jaramillo, J.; Gomperts, R.; Stratmann, R. E.; Yazyev, O.; Austin, A. J.; Cammi, R.; Pomelli, C.; Ochterski, J. W.; Ayala, P. Y.; Morokuma, K.; Voth, G. A.; Salvador, P.; Dannenberg, J. J.; Zakrzewski, V. G.; Dapprich, S.; Daniels, A. D.; Strain, M. C.; Farkas, O.; Malick, D. K.; Rabuck, A. D.; Raghavachari, K.; Foresman, J. B.; Ortiz, J. V.; Cui, Q.; Baboul, A. G.; Clifford, S.; Cioslowski, J.; Stefanov, B. B.; Liu, G.; Liashenko, A.; Piskorz, P.; Komaromi, I.; Martin, R. L.; Fox, D. J.; Keith, T.; Al-Laham, M. A.; Peng, C. Y.; Nanayakkara, A.; Challacombe, M.; Gill, P. M. W.; Johnson, B.; Chen, W.; Wong, M. W.; Gonzalez, C.; and Pople, J. A.; Gaussian 03, Revision

C.02, Gaussian, Inc.: Wallingford CT, 2004. 39. Hehre, W. J.; Stewart, R. F.; Pople, J. A., J. Chem. Phys. 1969, 51, 2657. 40. Hehre, W. J., Ab Initio Molecular Orbital Theory. Acc. Chem. Res. 1976, 9, 399. 41. Hehre, W. J.; Ditchfied, R.; Pople, J. A., J. Chem. Phys. 1972, 56, 2257. 42. Krishnan, R.; Binkley, J. S.; Seeger, R.; Pople, A. J., J. Chem. Phys. 1980, 72,

650. 43. Frisch, M. J.; Pople, J. A.; Binkley, J. S. J., J. Chem. Phys. 1984, 80, 3265. 44. Becke, A. D., J. Chem. Phys. 1993, 98, 5648. 45. Szabo, A.; Ostlund, N. S., Modern quantum chemistry. McGraw-Hill: New york,

1986. 46. Dreizler, R. M.; Gross, E. K. U., Density functional theory: An approach to the

quantum many-body problem. Springer-Verlag: Berlin; New York, 1990. 47. Becke, A. D., Phys. Rev. A 1988, 38, 3098. 48. Lee, C.; Yang, W.; Parr, R.G., Phys. Rev. B 1988, 37, 785.

Page 42: Experimental and computational studies of the unimolecular

28

CHAPTER TWO: EVIDENCE OF AN INTRAMOLECULAR

NUCLEOPHILIC AROMATIC SUBSTITUTION: AN

EXPERIMENTAL AND THEORETICAL STUDY OF THE

GAS-PHASE REARRANGEMENT OF AZO DYES

Abstract

NH2

SO3

NN

SO3

NHN

HN

SO3

HN NHN

SO3

NHN N

H

SO3

HN-N2

H

SO3

HN

SO3

NHN

NH

SO3

HNN2

SO3

NHN

NH

A surprising rearrangement was identified in the gas-phase by electrospray

ionization tandem mass spectrometry of anions of azo dye, and the rearrangement was

found to effect a loss of the azo moiety bridging aromatic rings as nitrogen. The collision-

induced dissociation (CID) of substituted analogues of the azo dyes, provide evidence for

the involvement of a nucleophlic aromatic substitution reaction mechanism in the

rearrangement process. Isotopic labeling of the azo dye anion, reveals the involvement of

the tautomer during the fragmentation process. The participation of tautomerization in

this intramolecular nucleophilic aromatic substitution mechanism was further supported

Page 43: Experimental and computational studies of the unimolecular

29

by electronic structure calculations carried out at B3LYP/6-31+G(d) level of theory.

Somewhat surprisingly, however, the computational data indicate that the rate

determining step is not via a Meisenheimer transition state. The novel rearrangement

was also found to occur in dianions of the azo dye, where Coulombic repulsions would be

expected to dominate the fragmentation process in the gas-phase.

Page 44: Experimental and computational studies of the unimolecular

30

2.1 Introduction

Azo dyes are characterized by the presence of one or more azo moieties

(−N=N−) bridging aromatic rings. The resulting extended conjugation gives rise to strong

absorption maxima in the visible spectrum and thus a range of desirable colours used in

diverse applications including textile dyes, paint pigments, printing inks, and food

colourings.1 As a result of their extensive use, millions of tones of azo dyes are produced

by dyeing industries around the world. To facilitate the dyeing process, azo dyes are often

modified to include sulfonate moieties that improve their water solubility.1

During the manufacture and use of azo dyes, large quantities of azo dyes,

together with their degraded products are discarded in wastewater and as solid residue.

The effluents from these industries often find their way into the aquatic environments.

Their presence in the aquatic environments, not only bring aesthetic objections but also

create hazardous degradation products, such as aromatic amines that are known

mammalian carcinogens.2 As a result, their analysis and detection in wastewater has

become very important, and led to increasing numbers of published methods.3

The most popular method for detecting the presence of azo dyes in

environmental samples makes use of their ultraviolet-visible (UV-Vis) spectral

properties4 coupled to High Pressure Liquid Chromatography (HPLC) or Thin Layer

Chromatography (TLC) for separation of components from each other and the matrix.4, 5

However, with the introduction of mass spectrometry, dyes have been analyzed by

coupling mass spectrometry with HPLC. The most popular ionization technique that has

been employed in conjunction with mass spectrometric techniques is electrospray

ionization, as the sulphonic acid moieties in the azo dyes are readily ionizable to produce

Page 45: Experimental and computational studies of the unimolecular

31

[M-H]- anions.6-8 However, azo dyes have also been analyzed by other ionization

methods namely Field Desorption (FD)5, Plasma Desorption (PD)9, Fast Atom

Bombardment (FAB)10, thermospray11, and Matrix Assisted Laser Desorption Ionization

(MALDI).8 Analysis of azo dyes in waste water has been simplified by using tandem

mass spectrometry. Tandem mass spectrometry or MS/MS offers selective detection of

analytes of interest from complex mixtures and hence the detection of azo dyes from the

waste water requires no purification steps.7, 8 Much of the previous work on detection of

azo dyes from environmental samples using MS/MS has focused on detection of

structurally diagnostic fragment ions arising from the azodye anions. Azo dyes show

characteristic fragments corresponding to the loss of SO3 and SO2 and these fragments

were routinely used to rapidly screen the presence of azo dyes in environmental samples.

11,12,13,14

Richardson and co-workers, in an attempt to obtain further structural

information to better characterize sulfonated azo dyes, have reported fragments

corresponding to the so-called ‘azo cleavage’, i.e., fragments arising from homolytic

bond cleavage from either side of the azo moiety.7 Interestingly, in their study a

characteristic fragment corresponding to the loss of 28 Da was also observed. Gaskell and

co-workers in their study have also noted the loss of 28 Da. The loss was proposed to be

the loss of N2 from the azo dyes but neither group has proposed a mechanism to account

for this surprising observation.7, 12 Bowie in his study with aromatic azo benzenes under

Electron-Ionization (EI) condition also reported this interesting loss of nitrogen. However

the mechanism effecting the reaction in radical cations, under EI conditions is likely to be

different.13 What is so interesting about this fragmentation is that the azo moiety bridging

Page 46: Experimental and computational studies of the unimolecular

32

the two aromatic rings is lost as nitrogen, leaving the two aromatic rings intact. This

suggests that new bonds are formed between the remaining two aromatic rings.

In the present study we have generated a range of sulfonated azodye

anions in the gas-phase by electrospray ionization of basic solutions of the azo dyes (Fig-

2.1). These ions were activated by collisions with neutral atoms such as Helium or Argon

to induce unimolecular fragmentation in order to study and elucidate the mechanism for

the loss of nitrogen.

Fig-2.1: Range of azo compounds

In our study, we have proposed two intramolecular mechanisms that could

conceivably account for the loss of N2, one representing a simple dissociation-

recombination reaction and the other representing an intramolecular nucleophilic

aromatic substitution reaction. The dissociation recombination mechanism involves a

simple bond homolysis of the azo moiety (Scheme-2.1), followed by the liberation of

nitrogen and formation of an ion-dipole complex. Bowie in his pioneering work with

fragmentation of negative ions has highlighted the importance of the formation of ion-

dipole complex, prior to dissociation to the fragment ions. The formation of fragment

X

SO3H

NN

Y

(a) X = NH2, Y = H

(b) X = NH2, Y = OMe

(c) X = NH2, Y = NO2

(d) X = NH2, Y = Ph

(e) X = NH2, Y = SO3H

(f) X = OH, Y = H

4-amino-3-(phenyldiazenyl)naphthalene-1-sulfonic acid

4-amino-3-((4-methoxyphenyl)diazenyl)naphthalene-1-sulfonic acid

4-amino-3-((4-nitrophenyl)diazenyl)naphthalene-1-sulfonic acid

4-amino-3-(biphenyl-4-yldiazenyl)naphthalene-1-sulfonic acid

4-amino-3-((4-sulfophenyl)diazenyl)naphthalene-1-sulfonic acid

4-hydroxy-3-(phenyldiazenyl)naphthalene-1-sulfonic acid

Page 47: Experimental and computational studies of the unimolecular

33

ions from the ion-neutral complex can be effected by reactions such as proton transfer

and nucleophilic substitution within the complex.14 Here we propose the naphthyl radical

anion and the phenyl radical would undergo recombination within the ion-dipole complex

to form the product.

SO3

NH2

NN

SO3

NH2

N2 recombination

SO3

NH2

Scheme-2.1: Proposed dissociation-recombination reaction mechanism

The nucleophilic aromatic substitution proposal involves the attack of

naphthyl amine on to the benzene ring at the ipso position (Scheme-2.2) to form a

resonance stabilized Meisenheimer intermediate (b). The Meisenheimer intermediate

would then undergo an elimination process to eject nitrogen as a neutral (Scheme-2.2)

leaving a charge separated intermediate (c). The resulting intermediate would then

undergo a facile 1,3-proton transfer to form the amine bridged product (d).

NH2

SO3

N

SO3

H2N

SO3

HN

SO3

N H2N NN

-N2 1,3-H+ Transferipso attack

(c) (d)(a) (b)

Scheme-2.2: Proposed intramolecular nucleophilic aromatic substitution reaction.

The mechanism as written in Scheme-2.2 could be considered as charge

remote reaction. A charge remote reaction is defined as one which is uninfluenced by the

charged centre and occurs remote from the centre. Interestingly, the mechanism in

Page 48: Experimental and computational studies of the unimolecular

34

Scheme-2.2 is not driven by the negative charge and indeed occurs remote from it.

However, the Hammett σp value for the SO3- is reported as +0.35 and thus is expected to

influence the nucleophilicity for the para-amine moiety. As a consequence the

mechanism might still be considered as influenced by the charge and thus not charge

remote by definition.f

Riveros has recently demonstrated bimolecular nucleophilic aromatic

substitution in the gas phase, in which the fluoride anions displace the nitro substituent

from the benzene ring to from fluoro-benzene (Scheme-2.3).15 It seems possible that the

azo dyes might present a possible unimolecular example of such a process in the gas-

phase.

NO2

+ F

NO2F F

+ NO2

Scheme-2.3: Nucleophilic aromatic substitution reaction in the gas-phase15

In the present study we elucidate the reaction mechanism for the loss of

nitrogen via (i) comparison of homologues substituted on the benzene ring (Fig-2.1(a-f)),

(ii) change of nucleophile, (iii) deuterium exchange and MS3 experiments and

(iv) quantum chemical calculations

2.2 Results and discussion

2.2.1 Collision induced dissociation (CID) of deprotonated azo dyes

Fig-2.2 shows the negative ion CID mass spectrum of the [M-H]- anions

from 4-amino-3-(phenyldiazenyl)naphthalene-1-sulfonic acid (Fig-2.1(a)). The spectrum

f Adams, J., Mass Spectrom. Rev., 1990, 9, 141.

Page 49: Experimental and computational studies of the unimolecular

35

was obtained using an electrospray ionization linear ion-trap mass spectrometer. The [M-

H]- molecular ion is observed at m/z 326.

Fig-2.2: CID Spectrum of the azodye 4-amino-3-(phenyldiazenyl)naphthalene-1-sulfonic acid as shown in Fig-2.1(a)

The prominent fragment ions in the spectrum are observed at m/z 262 (-64

Da) corresponding to the loss of SO2 and at m/z 246 which corresponds to the neutral loss

of SO3 (Scheme-2.4). The driving force for the loss of SO2, is the formation of stable SO2

molecule and a very stable phenoxide anion. Binkley and co-workers have reported such

fragmentation from aromatic sulfonate anions.16 The driving force for the loss of SO3, is

the formation of a stable SO3 molecule and a relatively stable phenide anion.

The product ion at m/z 221 corresponds to the loss of 105 Da resulting

from the concomitant loss of N2 and the phenyl radical and results from the so-called

azo-cleavage (Scheme-2.4).7 The other prominent fragment ion observed at m/z 234 (-92

Da) occurs from either the loss of anilinyl radical or from consecutive losses of SO2 and

220 240 260 280 300 320m/z

5

10

15

20

25

30

35

40

45

50

55

60

65

70

75

80

85

90

95

100

Re

lativ

e A

bund

ance

325.9

246.0

233.9

220.9 262.0

297.9

295.9308.9

NH2

SO3-

NN

CID

Page 50: Experimental and computational studies of the unimolecular

36

N2. The minor fragment ions at m/z 296 (-30 Da) and m/z 309 (-17 Da) could correspond

to the loss of hydrazine (N2H2) and ammonia (NH3) respectively. The fragment ions that

represent the facile loss of nitrogen, are observed at m/z 298 (-28 Da), with an abundance

of 15% of the molecular ion

SO3

NH2

NN

O

NH2

NN

m/z 262

m/z 326

NH2

NN

m/z 246

SO3

NH2m/z 221

SO3

NH

N

m/z 234

SO3

HN

m/z 298

SO2

SO3N2

NH

+ N2

Scheme-2.4: Fragmentation pathways for major ions in the CID spectrum of the azodye 4-amino-3-(phenyldiazenyl)naphthalene-1-sulfonate (Fig-2.1(a))

The loss of anilinyl radical, resulting in the formation of fragment ion at

m/z 234 is proposed to occur via an initial azo-hydrazone tautomerization followed by the

cleavage of the N-N bond from the hydrazone tautomer (Scheme-2.5). Fragmentation of

this type has been previously reported and rationalized by Gaskill and co-workers.12

SO3-

HN

N

NH

SO3-

NH

N

NH

SO3-

NH

N

NH

Scheme-2.5: Azo-hydrazone type fragmentation

Page 51: Experimental and computational studies of the unimolecular

37

The consecutive losses of N2 and SO2 from the parent anion resulting in

the formation of ions at m/z 234 were identified through MS3 experiments. One such

experiment involves the initial selection of the product ion at m/z 262 from the CID

spectrum of the molecular ion. The selected ion at m/z 262 when subjected to CID,

fragments to give the ion at m/z 234 (Fig-2.3). A plausible mechanism for this

fragmentation would involve an initial loss of SO2 from the parent azo anion to give a

phenoxide type intermediate. The intermediate could then undergo the loss of azo moiety

as nitrogen, by analogy to the parent ion.

Fig-2.3: CID mass spectrum of the mass selected product ion at m/z 262

The other experiment involves the mass selection of the product ion at m/z

298, from the CID spectrum of the molecular ion. The selected ion at m/z 298 when

subjected to CID, fragments to give ions at m/z 234 (Fig-2.4). A plausible mechanism for

this process could be the reverse of the previous mechanism, where the loss of SO2

NH2

SO3-

NN

140 160 180 200 220 240 260m/z

5

10

15

20

25

30

35

40

45

50

55

60

65

70

75

80

85

90

95

100

Rel

ativ

e A

bund

ance

233.9

183.9

261.9154.8

234.9142.9 232.9

127.9 192.9155.9 231.9168.8

CID

m/z 262

-SO2

CID

Page 52: Experimental and computational studies of the unimolecular

38

occurs from the m/z 298 that is formed from the loss of nitrogen from the parent azo

anion.

220 230 240 250 260 270 280 290 300m/z

5

10

15

20

25

30

35

40

45

50

55

60

65

70

75

80

85

90

95

100

Rel

ativ

e A

bund

ance

233.9

297.9

Fig-2.4: CID mass spectrum of the mass selected product ion at m/z 298

2.2.2 Solution phase labelling

In an effort to establish the structure of the ions at m/z 309, m/z 296 and

m/z 234, the sample 4-amino-3-(phenyldiazenyl)naphthalene-1-sulfonic acid (Fig-1(a))

was dissolved in a solvent mixture of D2O/acetonitrile (80% v/v). The most exchangeable

hydrogens in the molecule would be those on the amine functional group. The resulting

solution was then analysed by ESI-MS (Fig-2.5). The spectrum shows ions m/z 327, m/z

328 apart from the unlabeled molecular ion at m/z 326. The two ions at m/z 327 and m/z

328 represent one deuterium and two deuteria exchanged molecular ions.

NH2

SO3-

NN

m/z 298

CID -N2

CID

Page 53: Experimental and computational studies of the unimolecular

39

318 320 322 324 326 328 330 332 334 336m/z

5

10

15

20

25

30

35

40

45

50

55

60

65

70

75

80

85

90

95

100

Rel

ativ

e A

bund

ance

328.4

327.5

329.4

326.5 330.3

Fig-2.5: MS spectrum of deuterium labeled sample

Upon mass selection and CID of the ions at m/z 328, the fragment ion

representing the loss of NH3 in the original spectrum, is shifted by a mass unit of 2 Da

(m/z 309) (Fig-2.6). This indicates that the exchanged deuterium at the amine site of the

molecule is lost as a neutral NHD2. The result hence suggests that the amine moiety is

lost as ammonia, with one extra hydrogen extracted from the ring. However, the fragment

ion representing the loss of N2H2 in the original mass spectrum, showed no mass shift in

labeled ESI MS/MS spectrum (Fig-2.6). This suggests that the loss of N2H2 involves

hydrogens from the aromatic rings. However other isobaric losses such as CH2O or NO

can not be ruled out.

Interestingly, ESI MS/MS of the labeled sample also isolated the product

ion that was proposed to arise from the loss of anilinyl radical, which in the original

Page 54: Experimental and computational studies of the unimolecular

40

spectrum was masked with fragment ions from consecutive losses of SO2 and N2. The

ESI-MS/MS spectrum of the m/z 328 ion showed two major product ions at m/z 235 and

m/z 236 (Fig-2.6). The fragment ion at m/z 235 would correspond solely to the loss of

deuterated anilinyl radical (-93 Da), whereas the fragment ion at m/z 236 (-92 Da) would

correspond to ions arising from consecutive losses of N2 and SO2. Surprisingly, the

abundance of the ion at m/z 236 is very small, indicating that the contribution from the

loss of SO2 and N2 is neglible as compared to that from the anilinyl radical loss (Fig-2.6).

The result suggest that the tautomer is actively involved in the fragmentation process of

the 4-amino-3-(phenyldiazenyl)naphthalene-1-sulfonate anion.

Fig-2.6: CID mass spectrum of completely deuterium exchanged sample

220 240 260 280 300 320m/z

5

10

15

20

25

30

35

40

45

50

55

60

65

70

75

80

85

90

95

100

Rel

ativ

e A

bund

ance

248.1

328.1

235.0

264.1

223.0

300.1236.1

298.1 309.1234.0

ND2

SO3-

NN

CID

Page 55: Experimental and computational studies of the unimolecular

41

2.2.3 Evidence towards the mechanism for N2 loss

2.2.3.1 Positive ion experiment

In order to form positive ions of the azodye 4-amino-3-

(phenyldiazenyl)naphthalene-1-sulfonic acid was dissolved in aqueous acetonitrile (50%

v/v) solution containing approximately 2% formic acid solution, before subjecting them

to electrospray ionization. ESI MS/MS spectrum of the [M+H]+ ions of the 4-amino-3-

(phenyldiazenyl)naphthalene-1-sulfonic acid (Fig-2.1(a)) is shown in Fig-2.7. The

[M+H]+ molecular ion is observed at m/z 328.

Fig-2.7: CID mass spectrum of the positive ions of the azodye 4-amino-3-(phenyldiazenyl)naphthalene-1-sulfonic acid as in (Fig-2.1(a))

The prominent fragment ions in the spectrum are observed at m/z 312 (-16

Da) corresponding to the loss of NH2 or O and at m/z 247 which corresponds to the loss

of HSO3. The fragment ion corresponding to the loss of N2 and phenyl radical i.e from

140 160 180 200 220 240 260 280 300 320m/z

0

10

20

30

40

50

60

70

80

90

100

Rel

ativ

e A

bund

ance

246.9 327.9

247.9

311.9

310.9

221.8217.9

141.8 248.9230.9157.8 169.8 216.9 263.9

298.9

287.9189.8

x50

NH3

SO3H

NN

CID

Page 56: Experimental and computational studies of the unimolecular

42

the so-called azo cleavage is also observed at m/z 222. The fragment ions corresponding

to the loss of SO3 and OH were observed at m/z 248 and m/z 311.

Interestingly, the spectrum shows only a small fragment ion abundance at

m/z 299 corresponding to the loss of nitrogen. In Fig-2.7 it has been amplified 50 times

its original abundance. This result is in contrast to the Bowie’s observation with

positively charged radical ion, where the loss of nitrogen was more prevalent.13 In most

instances, reactions of radical ions differ from that of the closed-shell ions and hence

such a result is rather not surprising.

In regards to our proposal, in the positive ion mode, we would expect the

nucleophilic aromatic substitution mechanism to be switched off as the ammonium ion, is

too poor a nucleophile to effect nucleophilic substitution. Hence, the result suggests that

the dissociation-recombination mechanism might be occurring but only in a minor extent.

2.2.3.2 Dianion Experiment

The azo dye 4-amino-3-((4-sulfophenyl)diazenyl)naphthalene-1-sulfonic

acid (Fig-2.1(e)) would be expected to form dianions in the gas-phase, since it has two

ionizable sulphonic acid moieties. ESI-MS/MS spectrum of the [M-2H]2- ions of the

azodye was obtained and is presented in (Fig-2.8). The [M-2H]2- ions is observed at m/z

202.5 in the spectrum.

Page 57: Experimental and computational studies of the unimolecular

43

60 80 100 120 140 160 180 200 220 240 260 280m/z

0

5

10

15

20

25

30

35

40

45

50

55

60

65

70

75

80

85

90

95

100

Rel

ativ

e A

bund

ance

171.0

234.0

202.5152.0 188.5 221.0156.0

79.9 172.0

Fig-2.8: CID mass spectrum of [M-2H]2- ions from the azo dye 4-amino-3-((4sulfophenyl)diazenyl) naphthalene-1-sulfonic acid as in Fig-2.1(e)

The CID mass spectrum shows a pair of fragment ions at m/z 221 and m/z

156, resulting from the loss of N2 via the azo cleavage. The product ion at m/z 221

corresponds to the sulphonated naphthyl radical anion, while the ion at m/z 156

corresponds to the sulphonated phenyl radical anion. The CID mass spectrum also shows

a pair of fragment ions at m/z 234 and m/z 171, resulting from the cleavage of the N-N

bond from the hydrazone tautomer via an intial azo-hydrazone tautomerization. The

product ion at m/z 234 corresponds to the sulphonated diimino radical anion while the

product ion at m/z 171 corresponds to the sulphonated anilinyl radical anion. The

fragment ion at m/z 188.5 corresponds to the loss of 14 Da. Since the fragment ion

formed after the loss of N2 would also be doubly charged, the product ion at m/z 188.5

corresponds to the loss of N2.

NH2

SO3

NN

SO3

CID

Page 58: Experimental and computational studies of the unimolecular

44

The presence of the fragment ion at m/z 188.5 is surprising as it suggests a

rearrangement of a dianion, prior to fragmentation. The result also suggests that the loss

of nitrogen is unlikely to occur via the proposed dissociation-recombination mechanism

(Scheme-2.1). This is because, the initial homolytic cleavage of the parent ion, would

result in the formation of two anions from either side of the azo moiety. The resulting

anions would experience Coulombic repulsion and would not exist as the stable complex

(Scheme-2.1). Hence, the dianion experiment suggests that dissociation-recombination

reaction is unlikely to effect the facile loss of azo moiety as nitrogen.

2.2.3.3 Authentic product experiment

In order to verify the likelihood of nucleophilic aromatic substitution

reaction effecting the loss of N2, secondary amine product (Scheme-2.2(d)) from the

aromatic substitution reaction proposal was synthesized and was then subjected to

tandem mass spectrometric analysis (MS/MS). Figure-9 shows the comparison of the

MS/MS spectrum (Fig-2.9(a)) of the proposed amine product with that of the

MS/MS/MS spectrum (Fig-2.9(b)) from the negative ion of 4-amino-3-

(phenyldiazenyl)naphthalene-1-sulfonic acid (Fig(2.1(a)). It can be seen that both the

spectra have the same fragment ions at m/z 234 and m/z 218, corresponding to the loss of

SO2 and SO3 respectively. The abundance of the fragment ions compared with the

molecular ion, however are noticeably different, even with the same collisional

conditions. This could be, due to either (a) the presence of various positional isomers in

the sample of the synthesized amine. The authentic amine was synthesized by

sulphonating the secondary amine precursor in the presence of sulphuric acid. The

resonance contributing structures of the amine (Scheme-2.6), dictate the course of the

Page 59: Experimental and computational studies of the unimolecular

45

reaction and thus sulphonation can occur at a number of positions. The para positions

(Scheme-2.6(a, h)) are generally favoured over the others and thus the resulting mixture

that was analysed would have consisted of large proportions of the two para positional

isomers (Scheme-2,6(c), Scheme-2.6(h)).

NH NH NH NH

NHNHNHNH

(a) (b) (c) (d)

(e)(f)(g)(h)

Scheme-2.6: Resonance contributing structures of the secondary amine

(b) The differing ion abundances could also be due to the difference in internal energy of

the molecules in the two samples, as different energies could be imparted during

successive collisional experiments.

Page 60: Experimental and computational studies of the unimolecular

46

220 240 260 280 300m/z

10

20

30

40

50

60

70

80

90

100

10

20

30

40

50

60

70

80

90

100R

elat

ive

Abu

ndan

ce

234.1298.1

218.1234.1

298.0

218.1

MS/MS

MS/MS/MS

Fig-2.9: Comparision of MS/MS spectra of the secondary amine product with that of the MS3 spectrum of the azo dye anion, 4-amino-3-(phenyldiazenyl)naphthalene-1-sulfonate (Fig-1(a))

In general, complete sulphonation of aromatic compounds leads to

additions of sulfonate groups at multiple positions on the aromatic ring. In this case, the

favoured product after complete sulphonation would be a disubstituted product (4-(4-

sulfonatophenylamino)naphthalene-1-sulfonic acid) with substitution at the para position

of both rings. Hence dianions of this compound would be the product from the

nucleophilic aromatic substitution proposal (Scheme-2.2) for the azodye 4-amino-3-

((4sulfophenyl)diazenyl) naphthalene-1-sulfonic acid (Fig-2.1(e)) .

To overcome the ambiguity in the monosulfonated case, a similar

experiment was conducted by obtaining MS/MS spectrum of the corresponding

disulfonated amine product. Figure-2.10, shows the comparision between the MS/MS

NH2

SO3

NN

HN

SO3

CID (a)

CID

m/z 298

CID (b)

Page 61: Experimental and computational studies of the unimolecular

47

spectrum of the amine (Fig-2.10(a)) with the MS/MS/MS spectrum of the dianions

formed from 4-amino-3-((4-sulfophenyl)diazenyl)naphthalene-1-sulfonic acid (Fig-

2.10(b)). Interestingly, the major fragment ions at m/z 156.5 (-32 Da), m/z 297.1 and m/z

313.1 corresponding to the loss of SO2, SO3 and SO2- respectively all have the very

similar ion abundance. The result, in this case is a better match than that for the

monsulfonated azodye, possibly because of the absence of positional isomers in the

sample of synthesized amine product. The result hence provides a strong support for the

nucleophilic aromatic substitution proposal for the loss of N2.

MS/MS

MS/MS/MS

Fig-2.10: Comparision of MS/MS spectra of the amine product with that of the MS3

spectrum of the dianions from 4-amino-3-((4-sulfophenyl)diazenyl)naphthalene-1-sulfonic acid (Fig-2.1(e))

A similar experiment was conducted for the corresponding monoanions by

obtaining MS/MS spectrum of the monoanion that was formed in competition from the

HN

SO3

SO3

NH2

SO3

NN

SO3

150 200 250 300m/z

10

20

30

40

50

60

70

80

90

100

10

20

30

40

50

60

70

80

90

100

Rel

ativ

e A

bund

ance

188.5

156.5

297.1 313.1188.5

156.5

297.1 313.0

296.0221.0

CID (a)

CID (b)

m/z 188.5

CID

Page 62: Experimental and computational studies of the unimolecular

48

disulfonated amine product during electrospray ionization (Fig-2.11). The resulting

MS/MS spectrum (Fig-2.11(a)) was compared with the corresponding MS/MS/MS

spectrum of the monoanions from 4-amino-3-((4-sulfophenyl)diazenyl)naphthalene-1-

sulfonate (Fig-2.11(b)). The two spectra are significantly different with one having a

fragment ion at m/z 320 of significant abundance.

This is rather not surprising and arguably, this experiment could explain

and give evidence to the proposed nucleophilic aromatic substitution (Scheme-2.2). In

both cases, monoanions would have formed as a result of deprotonation of any one of the

sulphonic acid residues, resulting in a mixture of two separate monoanions. In the case of

the authentic amine product, the MS/MS spectrum (Fig-2.11(a)) would be the

representative of fragment ions arising from the mixture of two isomeric monoanions.

Whereas, its ambiguous for the monoanions from 4-amino-3-((4-

sulfophenyl)diazenyl)naphthalene-1-sulfonate and the representative MS/MS/MS

spectrum (Figure-2.11(b)) could actually depict only those fragment ions arising from

the monoanions carrying deprotonated sulphonic acid moiety on the naphthyl ring. This

is because the initial CID of the monoanion would have resulted in ions corresponding to

the loss of N2 at m/z 378, only from those ions, containing deprotonated sulphonic acid

moiety on the naphthyl ring. The other isomeric mono anion, with deprotonated

sulphonic acid on the phenyl ring would probably have shown no fragmentation

corresponding to the loss of N2. This is possibly because, electron withdrawing effect of

the sulphonic acid residue in this case, would have deactivated the amine group, and

hence from a possible nucleophilic substitution attack on to the phenyl ring (Scheme-2.7).

Page 63: Experimental and computational studies of the unimolecular

49

SO3

NH2

NN

SO3H

CID

-N2

SO3

HN

SO3H

vs

CIDProducts

EWG

SO3H

NH2

NN

SO3

EDG

CID

-N2

SO3H

HN

SO3

CIDProducts

EWG

Scheme-2.7: Substitution nucleophilic mechanism in the monoanions of 4-amino-3-((4-sulfophenyl)diazenyl)naphthalene-1-sulfonate

Page 64: Experimental and computational studies of the unimolecular

50

SO3

NH2

NN

SO3H

SO3H

NH2

NN

SO3

+

Fig-2.11: Comparision of MS/MS spectra of the monoanions from disubstituted amine product with that of the MS3 spectrum of the monoanions from 4-amino-3-((4-sulfophenyl)diazenyl) naphthalene-1-sulfonic acid (Fig-2.1(e))

2.2.3.4 Change of nucleophile

The rate of the nucleophilic substitution reactions will be dependent on the

characteristics of the attacking nucleophile and hence the activation barrier for the

reaction, will increase or decrease depending on the strength of the nucleophile. In

general, the order of the nucleophilicity is RNH2 > ROR > ROH, particularly when the

structures of the nucleophiles being compared are otherwise similar.17 Hence, we tested

our proposal by conducting MS/MS (Fig-2.12) on the [M-H]- anion from 4-hydroxy-3-

(phenyldiazenyl)naphthalene-1-sulfonic acid as shown in Fig-2.1(f) , which has OH

moiety replacing the amine at the same position of the naphthalene ring. The spectrum

shows a prominent ion at m/z 327 which corresponds to the [M-H]- molecular ion and a

300 320 340 360 380m/z

20

40

60

80

100

20

40

60

80

100R

elat

ive

Abu

ndan

ce297.8

377.8

319.8

311.8 362.9 373.5338.3 357.3297.8

377.8

375.8

MS/MS

MS/MS/MS

CID (b)

CID (a)

SO3

HN

SO3H

+

SO3

HN

SO3H

CID

m/z 378

Page 65: Experimental and computational studies of the unimolecular

51

fragment ion at m/z 247 (-80 Da) corresponding to the loss of SO3. The fragment ion

corresponding to the azo-cleavage was observed at m/z 222 (-105 Da) with low

abundances and hence has been magnified to 100 times it’s original abundance. The

fragment ion at m/z 299 is also only observed at a vanishingly low abundance

corresponding to only 0.22% of the base peak, and that has been amplified to 1000 times

it’s original abundance.

The fragment ion corresponding to the loss of nitrogen is negligible (Fig-

2.12). This result could be attributed to the fact, that the hydroxyl group is a weaker

nucleophile than the amine in the azo dye 4-amino-3-(phenyldiazenyl)naphthalene-1-

sulfonate and thus supports the proposed nucleophilic substitution mechanism in the

latter compound.

180 200 220 240 260 280 300 320 340m/z

10

20

30

40

50

60

70

80

90

100

Rel

ativ

e A

bund

ance

x1000x100 246.9

326.9

170.8

221.8

298.8

Fig-2.12: CID mass spectrum of negative ions from 4-hydroxy-(phenyldiazenyl) naphthalene-1-sulfonic acid (Fig-1(f))

OH

SO3

NN

CID

Page 66: Experimental and computational studies of the unimolecular

52

2.2.3.5 CID of Substituted analogues

Nucleophilic aromatic substitution reactions are generally favoured in

aromatic compounds bearing electron withdrawing groups due to stabilization of negative

charge in the Meisenheimer intermediate and consequent lowering of the activation

barrier.18, 19 To further examine the generality and the mechanism of loss of N2, via the

aromatic substitution mechanism, a para- nitro substituted analogue was subjected to ESI

MS/MS. Figure-2.13 shows the CID mass spectrum of the [M-H]- anion from 4-amino-3-

((4-nitrophenyl)diazenyl)naphthalene-1-sulfonic acid (Fig-2.1(c)) and was recorded at the

same collision energy as the spectrum of the unsubstituted homologue. The molecular

ion is observed at m/z 371 with prominent fragment ions at m/z 307(-64 Da) and m/z

291(-80 Da), corresponding to the basic fragments arising from SO2 and SO3 losses

respectively. The base peak in the spectrum at m/z 263 (-108 Da) corresponds to a

secondary fragment, following consecutive losses of SO3 and N2. Such consecutive loss

of SO3 and N2 were not previously seen in the MS/MS spectrum of the unsubstituted azo

dye.

Interestingly, the spectrum also separated the fragment ion arising from

the loss of anilinyl radical (m/z 234 (-137 Da)) from that arising from consecutive losses

of SO2 and N2 (m/z 279 (-80 Da)). This observation further provides evidence for the

anilinyl radical loss and is consistent with the labeling experiment. Other basic fragment

ions corresponding to the loss of NH3 and N2H2 were observed at m/z 355 and m/z 341.

The azo cleavage fragment ion was observed at m/z 221. Fragment ions corresponding to

the loss of the substituent as NO2 and HNO2 were observed at m/z 325 and m/z 324.

Page 67: Experimental and computational studies of the unimolecular

53

In accordance to the substitution prediction, the nitro substitution at the

para position would decrease the barrier for the nucleophilic substitution, thereby

increasing the abundance of product ion corresponding to the loss of N2. Surprisingly, the

fragment ion corresponding to the loss of N2 has similar ion abundance (12%), as

compared with the unsubstituted azodye (15%). Quantitative comparison of fragment ion

abundances between two different fragment ions can be ambiguous because (a)

introduction of new functional group provides competitive pathways for fragmentation

(e.g. –NO2, -HNO2), and (b) under the same collision conditions, the energy imparted to a

molecule depends on the mass of the precursor ion, which in turn affects the fragment ion

abundance.20

Fig-2.13: CID mass spectrum of negative ions from 4-amino-3-((4-nitrophenyl)diazenyl)naphthalene-1-sulfonic acid (Fig-1(c))

NH2

SO3

NN

NO2

150 200 250 300 350m/z

10

20

30

40

50

60

70

80

90

100

Rel

ativ

e A

bund

ance

262.9

290.9 306.9

370.8

324.8

342.8233.8

278.9247.8151.8 220.8 354.8

CID

Page 68: Experimental and computational studies of the unimolecular

54

Further qualitative comparison was obtained by subjecting negative ions

from a para substituted methoxy analogue, 4-amino-3-((4-

methoxyphenyl)diazenyl)naphthalene-1-sulfonic acid (Fig-2.1(b)). The CID spectrum is

presented as Fig-2.14 and shows no loss of N2, even when obtained under the same

conditions as other homologues. This observation is consistent with the destabilization of

the Meisenheimer intermediate by the electron withdrawing methoxy moiety.

Fig-2.14: CID mass spectrum of negative ions from 4-amino-3-((4-methoxyphenyl)diazenyl)naphthalene-1-sulfonic acid (Fig-1(b))

2.2.3.6 Electronic structure calculation

The experimental data discussed above suggest that nitrogen loss from the

azo anions proceed via the intramolecular nucleophilic aromatic substitution reaction

(Scheme-2.2) as opposed to the dissociation-recombination reaction mechanism

(Scheme-2.1). Preliminary electronic structure calculations using B3LYP/6-31+G(d)

level of theory failed to locate Meisenheimer structure (Scheme-2.2(b)) in the proposed

NH2

SO3

NN

OCH3

CID

220 240 260 280 300 320 340m/z

10

20

30

40

50

60

70

80

90

100

Rel

ativ

e A

bund

ance

276.9

356.9

341.8

Page 69: Experimental and computational studies of the unimolecular

55

intramolecular aromatic substitution reaction as either a stable intermediate or transition

state. Based on the experimental evidence of tautomerization (Scheme-2.5), a new

aromatic substitution mechanism involving the tautomeric structure was proposed

(Scheme-2.8). The new nucleophilic aromatic substitution mechanism involves initial

tautomerization of parent azo anion (a) to (b), then followed by the attack of the imine

onto the phenyl ring at the ipso position (Scheme-2.8) to form a resonance stabilized

Meisenheimer intermediate (c). The Meisenheimer intermediate would then undergo an

elimination process to form a stable intermediate neutral (d). The resulting intermediate

would then undergo a 1,3-H+ transfer to form the same amine bridged product (e),

consistent with the experimental findings.

NH2

SO3

N

N

SO3

NH

N

HN

SO3

HN NH

N

SO3

N

HN

N

HSO3

HN

(a) (b) (c)

(d)(e)

Scheme-2.8: Aromatic substitution reaction involving tautomeric structure

The calculations were conducted with a model system that employs a

phenyl ring instead of the larger naphthyl ring, to reduce the use of computing time.

Nevertheless, it is expected to be a representative of the larger system.

Page 70: Experimental and computational studies of the unimolecular

56

The calculated potential energy surface for the rearrangement of the model

azo anion 4-amino-3-(phenyldiazenyl)benzene-1-sulfonate (Fig-2.1(a)) and subsequent

loss of nitrogen from the anion by the newly proposed mechanism is shown in Fig-2.16,

with energies given relative to the parent anion. The electronic energies of the critical

stationary points are listed in Table-2.1 and the complete structural information of the

critical stationary points are given in Appendix-2. The full Cartesian information of the

stationary points are provided as Appendix-2. In the calculated mechanism the initial

tautomerization from the parent anion to the tautomer IM1 presents a barrier of 39 kJ

mol-1. The energy of the tautomer is 36 kJ mol-1 more than the parent anion. Interestingly,

even though during tautomerization, the aromaticity of the phenyl ring diminishes, the

calculated barrier of 39 kJ mol-1 is four fold lower than the aromatic stabilization energy

(155 kJ mol-1).17 A plausible explanation to support the finding is the presence and

influence of the resonance contributing structure for the hydrazone tautomer. Scheme-2.9

shows the resonance contributing structure of the tautomer.

SO3

NN

HNH

SO3

NN

NHH

Scheme-2.9: Resonance contributing structure of the tautomer

Upon investigation of the structure of the tautomer, it does not clearly

reveal the presence and the influence of the resonance structure as the C-C bond lengths

in the phenyl ring does not seem to have the same length (Fig-2.15 (a)). Nevertheless, the

resonance contributing structure with a strong intramolecular ion-dipole interaction could

Page 71: Experimental and computational studies of the unimolecular

57

have possibly reduced the activation barrier of tautomerization. Following the

tautomerization, the nitrogen of the imine, attacks the ipso carbon via the transition state

TS2. This presents a barrier of 197 kJ mol-1 to rearrangement to the stable intermediate,

IM2-Cn1, which lies some 68 kJ mol-1 above the azodye anion minimum. The structure

of TS2 is reminiscent of the proposed structure of the Meisenheimer intermediate, with

the nucleophile –NH completing a near tetrahedral geometry at the ipso position of the

benzene ring (Fig-2.15 (b)).15

(a) IM1 (b) TS1

Fig-2.15: Structures of the Tautomer (1M1) and Meisenheimer transition state (TS1) on the 4-amino-3-(phenyldiazenyl)benzene-1-sulfonate potential energy surface optimized at B3LYP/6-31+G(d) level of theory The intermediate IM2-Cn1 presents two other conformers IM2-Cn2 and

IM2-Cn3, with each conformer differing in the orientation of N=N—H. A transition state,

TS3 was located for the transfer of proton from N=N—H group on to the ipso position

from IM2-Cn2, and thus represents a 1,4- proton transfer. Intrinsic reaction coordinate

Page 72: Experimental and computational studies of the unimolecular

58

calculations on this transition state shows that it connects to the intermediate IM3. The

intermediate IM3 is also reminiscent of Meisenheimer intermediate. Intuitively, the

intermediate IM3 would lead to the ion dipole complex IM5 of the proposed product by

simple heterolytic cleavage of C-N bond. No transition state structure could be isolated

for the formation of IM5 from IM3, as isolation of transition state for simple bond

cleavages is often difficult. The ion-dipole complex IM5 presents energy of -155 kJ mol-1

relative to the parent diazo anion. The ion-dipole complex eventually dissociates to the

product Pro. The product Pro presents an energy of -151 kJ mol-1 relative to the parent

azo anion.

The reaction mechanism that could be summarized for 4-amino-3-

(phenyldiazenyl)naphthalene-1-sulfonate encompassing all the intermediates and

transition states from the calculation with the model system is given in Scheme-2.10.

NH2

SO3

NN

SO3

NHN

SO3

HN NHN

SO3

NHN

NH

SO3

NHN

NH

SO3

HN

N2

HN

Scheme-2.10: Calculated reaction mechanism-1 for 4-amino-3-(phenyldiazenyl) naphthalene-1-sulfonate

Page 73: Experimental and computational studies of the unimolecular

59

The other conformer IM2-Cn3 connects an alternative transition state TS4,

representating a 1,5-proton transfer on to the amine from the N=N−H. This presents a

barrier of 248 kJ mol-1. Intrinsic reaction coordinate calculations on the transition state

connect it to a charge separated intermediate minimum IM4, with an optimized energy of

189 kJ mol-1 relative to the parent anion. This charge separated intermediate IM4

undergoes a 1,3-proton transfer, through a transition state TS5 to the proposed product

Pro.

The other reaction mechanism that could be summarized for 4-amino-3-

(phenyldiazenyl)naphthalene-1-sulfonate encompassing all the intermediates and

transition states from the calculation with the model system is given in Scheme-2.11.

NH2

SO3

NN

SO3

NHN

HN

SO3

HN NHN

SO3

NHN N

H

SO3

HN-N2

H

SO3

HN

Scheme-2.11: Calculated reaction mechanism-2 for 4-amino-3-(phenyldiazenyl) naphthalene-1-sulfonate

Page 74: Experimental and computational studies of the unimolecular

60

196.8183.3

-151.2

78.7

35.8

295.9

349.9

365.0

285.6

67.7

0.0

278.7

39.5

68.9

247.9186.8

188.5

-155.8

Reaction Coordinate

kJ/m

ole

NH2

SO3-

NN

NH2

SO3-

+ +N2

SO3-

NH2

N

+N

SO3

HNNN

H

SO3-

NHN

NH

+ so3

NNHNH

so3

NN

HHN

so3

NHN

NH

so3

NNHN

H

SO3

NHN

NH

so3

NNHN

H

so3

HN H+ N2

SO3-

HN

N2

SO3-

HN

N2+

TS1

IM1

TS2

IM2-Cn1

IM2-Cn2

IM2-Cn3

TS3

IM3

TS4

IM4 TS5

Tau-pro

azo-pro2

azo-pro1

IM6

IM5

Pro

SO3

NHN

NH

so3

H2NN2

so3

H2N+ N2

SO3-

NNN

H H

Fig-2.16: Reaction coordinate diagram for the intramolecular rearrangement of the model diazo anion, 4-amino-3-(phenyldiazenyl)benzene-1-sulfonate model system calculated at B3LYP/6-31+G(d) level of theory. All energies are given in kJ mol-1. The uncorrected energies (not including (ZPE)) place TS4 higher than IM6 by 4.1 kJ mol-1.

Page 75: Experimental and computational studies of the unimolecular

61

Table-2.1: Optimized stationary points calculated for the loss of N2 from 4-amino-3-(phenyldiazenyl)benzenesulfonate anion (Electronic energy zero-point energy, relative energy and imaginary frequency calculated at B3LYP/6-31+G(d) level of theory)

Structure

Energy

Hartrees

Zero-point Energy

Hartrees

Relative energy

kJ mol-1

4-amino-3

(phenyldiazenyl)

benzenesulfonate

-1251.44306

0.20948

0.0

TS1 (-1453 cm-1

) -1251.42392

0.20539

39.5

IM1 -1251.43005

0.21009

35.8

TS2 (-370 cm-1

) -1251.36715

0.20852

196.8

IM2-Cn1 -1251.41741

0.20963

67.7

IM2-Cn2 -1251.41625

0.20893

68.9

IM2-Cn3 -1251.41271

0.20911

78.7

TS3 (-873 cm-1

) -1251.29701

0.20247

365.0

TS4 (-472 cm-1

) -1251.34418

0.20502

247.9

IM3 -1251.30546

0.20517

349.9

IM4 -1251.36796

0.20554

186.8

TS5 (-820 cm-1

) -1251.36501

0.20122

183.3

IM5 -1251.49873

0.20581

-155.8

IM6 -1251.36659

0.20481

188.5

Azo-pro1 -1251.32471

0.20382

295.9

Azo-pro2 -1251.32437

0.19956

285.6

Tau-pro -1251.33019

0.20276

278.7

Pro -1251.49676

0.20559

-151.2

Page 76: Experimental and computational studies of the unimolecular

62

Interestingly, from these calculations, the transition state barriers TS3 and

TS4, appear to dictate the energetics of the reaction and not the transition state that

involves the formation of Meisenheimer structure TS2, which was initially thought to be

the rate determining step. The calculation also suggests that the mechanism involving the

formation of charged separated intermediate IM4 (Scheme-2.11) could be the most

thermodynamically favoured pathway, as the transition TS4 in the mechanism presents a

lower barrier than the barrier TS3 in the other calculated mechanism (Scheme-2.10).

Calculations were also conducted for the tautomeric cleavage and azo

cleavage process. The products for tautomeric cleavage is represented by Tau-Pro with

its energy at 279 kJ mol-1, and that of the azo cleavage (Azo-Pro) at 296 kJ mol-1. It

could be seem that these pathways are endothermic while the rearrangement to the amine,

effecting the loss of nitrogen is exothermic (Fig-2.16).

The calculation could also suggest that the absence of fragment ion

corresponding to the loss of nitrogen for 4-hydroxy-3-(phenyldiazenyl)naphthalene-1-

sulfonate could be due to two important properties of the hydroxyl substituent. The

corresponding transition state TS2 could present a higher barrier as hydroxyl analogue is

less nucleophilic than the amine counterpart. Also, the transition state, corresponding to

the formation of the charged separated intermediate TS4, could present a higher barrier as

the basicity of the hydroxyl substituent is lower than the amine counterpart.

Initial calculations, were performed to investigate the effect of hydroxyl

moiety, by replacing the amine to a hydroxyl group in the model system (Fig-2.17). The

electronic energies of the critical stationary points are given in Appendix-2. The full

Cartesian information of the stationary points are provided as Appendix-2. Surprisingly,

Page 77: Experimental and computational studies of the unimolecular

63

the tautomerization occurs over a barrier of 9 kJ mol-1 (Hyd-TS1), which is four fold

lower than for the 4-amino-3-(phenyldiazenyl)benzene-1-sulfonate (Fig-2.17). This

explains why the abundance of the tautomeric products (m/z 222) is much higher for the

4-hydroxy-(phenyldiazenyl) naphthalene-1-sulfonate (Fig-2.12) as compared with 4-

amino-(phenyldiazenyl) naphthalene-1-sulfonate (Fig-2.2). The plausible explanation

could be that the hydroxyl substituent is better proton donor than the amine counterpart,

which effectively could decrease the activation barrier. Adding to the above result, the

energy of the tautomer also presents lower energy (2.4 kJ mol-1), which is significantly

lower than that found for the 4-amino-(phenyldiazenyl) naphthalene-1-sulfonate (Fig-

2.16).

The Meisenheimer transition state presents comparatively similar barrier

of 205 kJ mol-1 (Fig-2.17) as compared with the 4-amino-3-(phenyldiazenyl)benzene-1-

sulfonate (TS2, Fig-2.16). A rearrangement to the stable intermediate, Hyd-IM2-Cn1

occurs following the Meisenheimer transition state and that the intermediate lies some 89

kJ mol-1 above the parent anion. Here again the intermediate Hyd-IM2-Cn1 presents two

other conformers Hyd-IM2-Cn2 and Hyd-IM2-Cn3, with each conformer differing in

the orientation of N=N—H. A transition state, Hyd-TS3 was located for the transfer of

proton from N=N—H group on to the ipso position from Hyd-IM2-Cn2. The barrier to

this process, again is very similar to that of 4-amino-(phenyldiazenyl)benzene-1-sulfonate

(TS3, Fig-2.16) leading to the formation of the intermediate Hyd-IM3. The intermediate

Hyd-IM3 leads to the proposed product, which is -128 kJ mol-1 relative to the parent azo

anion. The result suggests that this pathway is highly unlikely to effect the loss of N2 due

to the high transition barrier, represented by Hyd-TS3.

Page 78: Experimental and computational studies of the unimolecular

64

Nevertheless, the reaction mechanism that could be summarized for 4-

hydroxy-3-(naphthyldiazenyl)naphthalene-1-sulfonate encompassing all the intermediates

and transition states from the calculation with the model system is given in Scheme-2.12.

OH

SO3

NN

SO3

O

N

SO3

O NH

N

SO3

N

O

NH

SO3

N

O

NH

SO3

O

N2

HN

Scheme-2.12: Calculated reaction mechanism-1 for 4-hydroxy-3(phenyldiazenyl) naphthalene-1-sulfonate

The transition state which was the rate determining transition state (TS4,

Fig-(2.16)) in 4-amino-3(phenyldiazenyl)benzene-1-sulfonate potential energy surface, in

the 4-hydroxy-3(phenyldiazenyl)benzene-1-sulfonate calculation Hyd-TS4, presents a

barrier of 335 kJ mol-1. Surprisingly, intrinsic reaction coordinate calculation on the

transition state connects it to a benzyne-like intermediate with a loss of phenol and

nitrogen as neutrals (Hyd-Pro2) as opposed to the loss of just the N2 in the 4-amino-

3(phenyldiazenyl)benzene-1-sulfonate. The process of the loss of N2 and phenol would

involve a cyclic cleavage reaction and the optimized energy of the products formed were

196 kJ mol-1 relative to the parent anion (Fig-2.17).

Page 79: Experimental and computational studies of the unimolecular

65

The other reaction mechanism that could be summarized for 4-hydroxy-3-

(naphthyldiazenyl)naphthalene-1-sulfonate encompassing all the intermediates and

transition states from the calculation with the model system is given in Scheme-2.13.

OH

SO3

NN

SO3

ON

HN

SO3

O NHN

SO3

NO N

H

-N2SO3 OH

+

Scheme-2.13: Calculated reaction mechanism-2 for 4-hydroxy-3(phenyldiazenyl) naphthalene-1-sulfonate

An alternative explanation for the loss of 28 Da could be the loss of CO arising

from the phenoxide isomer of the parent anion. Computational data (shown in chapter 3)

for dissociation of phenoxide indicate that CO loss occurs over a barrier of 439 kJ mol-1.

This is similar to the energetic requirement for the loss of N2 (381 kJ mol-1Fig-2.17,

TS3). Thus some participation of CO loss in the formation of the fragment ion at m/z

299 in 4-hydroxy-3(phenyldiazenyl) naphthalene-1-sulfonate cannot be excluded.

Page 80: Experimental and computational studies of the unimolecular

66

204.9

378.1

-128.4

195.7

119.2

-132.7

334.5

2.4

99.292.489.3

8.70.0

381.4

Reaction Coordinate

kJ

/mo

le

OH

SO3-

NN

SO3

ONN

H

SO3-

ONN

H

so3

NNHO

so3

NN

HO

so3

NO

NH

so3

NNO

H

SO3

NO

NH

SO3-

O

N2+

Hyd-TS1

Hyd-IM1

Hyd-TS2

Hyd-IM2-Cn1 Hyd-IM2-Cn2Hyd-IM2-Cn3

Hyd-IM3

Hyd-Pro

Hyd-TS4Hyd-TS3

SO3

NO

NH

Hyd-Pro-2

SO3

O

NN

H

SO3

+

OH

+ N2

SO3-

O

N2

Hyd-IM5

SO3OH

N2

Hyd-IM-New

Fig-2.17: Reaction coordinate diagram for the intramolecular rearrangement of the 4-hydroxy-(phenyldiazenyl)benzene-1-sulfonate model system calculated at B3LYP/6-31+G(d) level of theory. All energies are given in kJ mol-1.

Page 81: Experimental and computational studies of the unimolecular

67

In regards to the substituted analogues, rationalizing the result would be

rather difficult particularly because the Meisenheimer transition state, TS2 does not

represent the rate determining transition state. On consideration, in the methoxy analogue

the charge separated intermediate IM4 could be more stabilized due to the electron

donating nature of the methoxy substituent. This could present a counter effect on the

driving force to form the stabilized amine product and hence the transition state barrier

TS4 could present a higher energy barrier. Nevertheless, the Meisenheimer transition

state TS2, could present a lower barrier for the nitro substituted analogue and a higher

barrier for the methoxy analogue. Being, the first transition state, apart from the TS1,

which significantly requires a very high energy (200 kJ mol-1) it could largely affect the

proceeding of the rearrangement reaction. These facts could plausibly explain why the

methoxy analogue showed no rearrangement to effect the loss of nitrogen. More over on

close examination of the methoxy spectrum (Fig-2.14), we could see no fragment ions

corresponding to the loss of N2 and the phenyl radical and anilinyl radical. This suggests

that the competitive fragmentation pathway leading to the loss of methoxy substituent

(m/z 342) and sulfonate substituents (m/z 277) could have complicated other observed

fragmentation pathways (Fig-2.14). As well as for the nitro substituted analogue,

competitive fragmentation pathway with lower barriers could have largely reduced the

ion abundance that corresponds to the loss of nitrogen.

Generally in dianions, direct dissociation pathways are preferred and

rearrangement pathways are hindered due to Coulombic repulsions.21 Hence, we

conducted a prelimary calculation, to understand the influence of the Coulombic

Page 82: Experimental and computational studies of the unimolecular

68

repulsion on the identified rearrangement pathway. The various intermediates from the

calculated pathways of 4-amino-3-(phenyldiazenyl)benzene-1-sulfonate model were

optimized as stationary points with an additional sulphonate moiety on the other benzene

ring, under the same level of theory. The electronic energies of the critical stationary

points are given in Appendix-2. The full Cartesian information of the calculated

stationary points are provided as Appendix-2. The Coulombic repulsion energy for each

stationary points were then calculated by using the formula kq1q2/r2, where k was

assumed as permetivity of vaccum, the data for which is presented in Scheme-2.14.

Page 83: Experimental and computational studies of the unimolecular

69

Distance = 11.66 Å

Distance= 11.79 Å

Energy = 0 kJ mol-1

Energy = 34 kJ mol-1

NH2

SO3

NN

SO3

NHN

HN

SO3

HN NHN

SO3

NHN

NH

SO3

NHN

NH

SO3

HN

N2

SO3

NHN N

H

SO3

HN

N2

H

SO3

HN

Distance = 11.28 Å

Energy = 79 kJ mol-1

SO3 SO3

SO3

SO3 SO3SO3

SO3

SO3

SO3

Repulsion energy = 0 kJ mol-1

Repulsion energy = -2 kJ mol-1

Repulsion energy= 4 kJ mol-1Repulsion energy = 3 kJ mol-1

Distance = 11.33 Å

Distance = 11.21 Å

Energy = 84 kJ mol-1

Repulsion energy = 4 kJ mol-1

Energy = -142 kJ mol-1

Repulsion energy = 3 kJ mol-1

Distance = 11.33 Å

Energy = -142 kJ mol-1

DI-IM1

DI-IM2Cn2

DI-IM2Cn3

DI-Pro

DI-Pro

DI-TS1

Distance = 11.01 Å

Energy = 165 kJ mol-1

Repulsion energy = 6 kJ mol-1DI-IM6

Scheme-2.14: Preliminary calculation on the dianion system

Page 84: Experimental and computational studies of the unimolecular

70

The result suggest that the distance between the two sulfonate substituent

does not change much, from each stationary points in the two calculated pathways, thus

suggesting that Coulombic repulsion would play no major role in the rearrangement and

that the rearrangement is feasible in the dianions.

2.3 Conclusion

The combined experimental and theoretical study has revealed an

interesting rearrangement that effects the loss of azo moiety as nitrogen, upon CID of azo

dye anions. Initial substituent and nucleophile change studies supported a direct

intramolecular nucleophilic aromatic substitution effecting the loss of N2 and that the

energetics of the rearrangement is dictated by the transition state corresponding to the

nucleophilic substitution reaction. The complementary computational study however

suggested that the rearrangement proceeds via an initial tautomerization, followed by

nucleophilic aromatic substitution reaction and that the course of the reaction is more

complex with other transition states providing the rate determining step. Moreover, the

complementary computational studies on the nucleophile changed analogue reveals a

more complex reaction manifold and that does not follow the logical and straight forward

conclusions that could be derived from textbooks. The rearrangement was surprisingly

observed for the dianion, and that the preliminary calculation on the system suggests that

the reaction would proceed without any influence from Coulombic repulsion.

Page 85: Experimental and computational studies of the unimolecular

71

2.4 Experimental

2.4.1 Mass spectrometry

Standard solutions of azo dyes (10 µM) were prepared in aqueous

acetonitrile (50% v/v) solution. A few drops of aqueous ammonia (48% v/v) were added

to these solutions to aid deprotonation of the azo dyes. Mass spectra were obtained using

ThermoFinnigan LTQ linear ion-trap mass spectrometer (Thermo Electron Corporation,

San Jose, USA). Spectra were obtained by infusion of the standard solution (10µL/min),

typical settings were cone voltage (3.51 kV), capillary voltage (-5.89 V), source

temperature (270oC). Helium was used as the collision and buffer gas at a pressure of (7.6

x 10-6 torr).The dianion spectrum was obtained under the same settings except the source

temperature was set at 140oC to minimize decomposition in the source. The spectrum

presented results from the average of at least 50 scans and was baseline corrected to 5%

using Xcalibur software package (Thermo Electron Corporation, San Jose, USA).

2.4.2 Synthesis of azo compounds

All the azo dye compounds were prepared and characterised by

Christopher Gordon from the Department of Chemistry, University of Wollongong.g A

standardized procedure for the synthesis includes:

To a stirring solution of aniline (0.15 mL, 1.61 mmol) in HCl (4 M, 5.00

mL) at 0oC was added dropwise a solution of sodium nitrite (0.13 g, 1.93 mmol) in

distilled water (5 mL). The resulting mixture was stirred at 0oC for 15 minutes before the

g Gordon, C. J., Structure based design and synthesis of small molecules human immuno deficiency virus type 1 entegre inhibitors, in School of Chemistry. 2007, University of Wollongong, Wollongong.

Page 86: Experimental and computational studies of the unimolecular

72

corresponding naphthalene sulfonic acid was added (Fig-2.1). The mixture was then

vigorously stirred at room temperature for a further 30 minutes prior to the addition of

NaOH (2 M, 10 mL). The resulting coloured precipitate was collected and purified via

recrystalliation (10:10:1 acetone/EtOH/2M HCl) to effect the formation of brightly

coloured azo dyes as solids. These compounds were characterized by their melting point

(m.p.), 1H-NMR and 13C-NMR.

2.4.3 Synthesis of authentic amine compounds

A stirred suspension of N-phenyl-1-naphthleneamine (2.00 g, 9.13 mmol)

in H2O (5 mL) was cooled to 0 °C to which conc. H2SO4 (10 mL) was added dropwise.

The resulting mixture was stirred at 80 °C for 16 h after which the reaction was quenched

with NaHCO3(aq), extracted with EtOAc (4 × 25 mL), dried (MgSO4), and concentrated in

vacuo. The resulting brown oil was subjected to preparative layer chromatography (1:1,

EtOAc:MeOH) to afford both the monosulfonated and disulfonated products (0.15g, 4%)

as a dark brown viscous oil.

2.4.4 Calculations

Geometric optimizations were carried out with the B3LYP method22, 23

using 6-31+G(d) basis set within the Gaussian suite of programs.24 All stationary points

on the potential energy surface were characterized as either minima (no imaginary

frequencies) or transition state (one imaginary frequency) by calculation of frequency

using analytical gradient procedure. Frequency calculations provided zero-point energies,

which were added to the calculated energy. The minima connected by a given transition

state were confirmed by inspection of the animated imaginary frequency using the Gauss

Page 87: Experimental and computational studies of the unimolecular

73

view package 3.0 (Gaussian, Inc, Pittsburgh, USA) and by intrinsic reaction coordinate

calculation.25, 26

Page 88: Experimental and computational studies of the unimolecular

74

References for chapter two

1. Kulkarni, S. V.; Blackwell, C. D.; Blackard, A. I.; Stack-house, C. W.; Alexander,

M. W., Textile Dyeing Operations. Chap. G, Moyes Publications: Park Ridge, NJ, 1986.

2. Voyksner, K.; Straub, R.; Keever, J., Environ. Sci. Technol. 1993, 27, 1665-1672. 3. Cooks, J.; Thurnhear, T.; Kohler-Staub, D.; Galll, R.; Grossbacher, H., Swiss.

Biotech. 1988, 4, 14-17. 4. Venkataraman, K., The analytical chemistry of synthetic dyes. Wiley-Interscience:

1977. 5. Mattias, A.; Williams, A. E.; Games, D. E.; Jackson, A. H., Org. Mass Spectrom.

1976, 11, 266. 6. Straub, R.; Voyksner, R. D.; Keever, J. K., J. Chromatogr. 1992, 627, 173. 7. Richardson, S. D.; McGuire, J. M.; Thructon, A. D.; Baughman, G. L., Org. Mass

Spectrom. 1992, 27, 289. 8. Sullivan, A. G.; Gaskell, S. J., Rapid Commun. Mass Spectrom. 1997, 11, 803. 9. Panell, L. K.; Sokoloski, E. A.; Fales, H. M., Anal. Chem. 1985, 57, 1060. 10. Monaghan, J.J.; Baber, M.; Bordoli, R. S.; Sedgwick, R. D.; Tyler, A. N., Int. J.

Org. Mass Spectrom. Ion Phys. 1983, 46, 447. 11. Ballard, J. M.; Betowski, L. D., Anal. Chem. 1984, 56, 2604. 12. Sullivan, A. G., Garner, R., and Gaskell, S. J., Rapid Commun. Mass Spectrom.

1998, 12, 1207-1215. 13. Bowie, J. H.; Lewis, G. E., J. and Cooks, R. G., Chem. Soc. (B) 1967, 621. 14. Bowie, J. H., The fragmentations of even-electron organic negative ions. Mass

Spectrom. Rev. 1990, 9, 349-379. 15. Giroldo, T. A.; Xavier, L. A.; Riveros, J. A., Angew. Chem. Int. Ed 2004, 43,

3588-3590. 16. Binkley, R. W.; Flechtner, T. W.; Teversz, M. J. S.; Winnik, W.; Zhong, B., Org.

Mass Spectrom. 1993, 28, 769-772.

Page 89: Experimental and computational studies of the unimolecular

75

17. March, J., Advanced organic chemistry: reactions, mechanisms, and structure. 4 ed.; Wiley, c1992: New York, 1929; p 349 .

18. Meisenheimer, J., Liebigs Ann. Chem. 1902, 323, 205. 19. Artamkina, G. A.; Egorov, M. P.; Beletskaya, I. P., Chem. Rev. 1982, 82, 427. 20. Downard, K., Mass spectrometry: A foundation course. Royal society of

Chemistry: Cambridge, 2004. 21. Schroder, D., Angew. Chem. Int. Ed 2004, 43, 1329. 22. Becke, A. D., J. Chem. Phys. 1993, 98, 1372. 23. Lee, C. T.; Yang, W. T., Parr, R. G., Phys. Rev. B: Condens. Matter 1988, 37, 785. 24. Gaussian 03 (Revision C.02). Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.;

Scuseria, G. E.; Robb, M. A.; Cheeseman, J. R.; Montgomery, Jr., J. A.; Vreven, T.; Kudin, K. N.; Burant, J. C.; Millam, J. M.; Iyengar, S. S.; Tomasi, J.; Barone, V.; Mennucci, B.; Cossi, M.; Scalmani, G.; Rega, N.; Petersson, G. A.; Nakatsuji, H.; Hada, M.; Ehara, M.; Toyota, K.; Fukuda, R.; Hasegawa, J.; Ishida, M.; Nakajima, T.; Honda, Y.; Kitao, O.; Nakai, H.; Klene, M.; Li, X.; Knox, J. E.; Hratchian, H. P.; Cross, J. B.; Bakken, V.; Adamo, C.; Jaramillo, J.; Gomperts, R.; Stratmann, R. E.; Yazyev, O.; Austin, A. J.; Cammi, R.; Pomelli, C.; Ochterski, J. W.; Ayala, P. Y.; Morokuma, K.; Voth, G. A.; Salvador, P.; Dannenberg, J. J.; Zakrzewski, V. G.; Dapprich, S.; Daniels, A. D.; Strain, M. C.; Farkas, O.; Malick, D. K.; Rabuck, A. D.; Raghavachari, K.; Foresman, J. B.; Ortiz, J. V.; Cui, Q.; Baboul, A. G.; Clifford, S.; Cioslowski, J.; Stefanov, B. B.; Liu, G.; Liashenko, A.; Piskorz, P.; Komaromi, I.; Martin, R. L.; Fox, D. J.; Keith, T.; Al-Laham, M. A.; Peng, C. Y.; Nanayakkara, A.; Challacombe, M.; Gill, P. M. W.; Johnson, B.; Chen, W.; Wong, M. W.; Gonzalez, C.; and Pople, J. A.; Gaussian 03, Revision

C.02, Gaussian, Inc.: Wallingford CT, 2004. 25. Gonzalez, C.; Schlegel, H. B., J. Chem. Phys. 1989, 90, 2154. 26. Gonzalez, C.; Schlegel, H. B., J. Phys. Chem. 1990, 94, 5523.

Page 90: Experimental and computational studies of the unimolecular

76

CHAPTER THREE: COMPUTATIONAL INVESTIGATION

OF THE REARRANGEMENT AND FRAGMENTATION OF

PHENOXIDE ANION IN THE GAS-PHASE

Abstract

O

CO

CO

+ CO

H HCO

CO

CO

HCO

CO

CO

H

CO

O

O

O

CO

CH

O O O

OC

O

H OH C

O

The rearrangement and fragmentation of phenoxide anion in the gas-phase was

studied by electronic structure calculation carried out at B3LYP/6-311++G(d,p) level of

theory. The study investigated previously reported Collision-Induced Dissociation (CID)

studies of phenoxide anions, where product ions consistent with the loss of CO were

observed. The data from another study indicating the involvement of benzene-oxide and

oxepin intermediates during the fragmentation was also investigated. Primarily, the

Page 91: Experimental and computational studies of the unimolecular

77

electronic structure calculations reveal that the loss of CO occurs via two competing

reaction pathways involving ketene like intermediates and transition states. The

computational data also reveals the existence of rearrangement pathway of phenoxide

anion to benzene-oxide and oxepin structures eventually effecting the loss of CO.

Page 92: Experimental and computational studies of the unimolecular

78

3.1 Introduction

Phenols and related poly phenols have been used in the medical industry

for nearly over two centuries. The first use of phenol in the field of medicine was as an

antiseptic and it was Sir Joseph Lister (1827-1912) who showed that phenols can be used

as an antiseptic during surgery. Since then, phenolic compounds have been used in the

production of variety of useful commodities such as asprin, weed killer, wood

preservatives, disinfectants and bakelite.1

As a result of their extensive industrial and domestic use, large amounts of

synthetic phenols find their way into environmental waterways. Phenols, if present in

large quantities in environmental water-ways are known to be toxic for aquatic life.2

Phenols are also known to cause a number of health effects in humans. Some of the

reported health effects include cardiovascular effects, respiratory distress, metabolic

acidosis, renal failure, neurological effects, shock, coma and death. Gastrointestinal

effects such as nausea and vomiting have been reported as major health effects. Dermal

effects such as dermal inflammation and erythema are also very common upon exposure.

Moreover, exposures to some phenols are known to cause permanent genetic damage.h

Interestingly phenolic compounds are also biologically produced and

secreted by all living forms, including plants and humans. The phenols secreted by plants

are popularly known as polyphenols and they contain multiple chains of aromatic rings

with the hydroxyl moiety. Food sources that are rich in polyphenols include onion, apple,

tea, red wine, red grapes, grape juice, strawberries, raspberries, blueberries, cranberries,

and certain nuts. 3

h International Programme on Chemical Safety (IPCS) (1999). Phenol. Poisons Information Monograph. PIM 412.

Page 93: Experimental and computational studies of the unimolecular

79

For many years polyphenols as a dietary component was considered

antinutritive, and was thought to account only for the colour and flavour of certain food

stuffs. Currently polyphenols as dietary components are classified under phytonutrients

which refers to those bioactive components that promote significant positive health

effects in humans.3 Polyphenols basically protect plants and humans from oxidative

damage. These polyphenols are also known to block specific enzymes that cause

inflammation and allergies. i ,4,5 They also help the liver detoxify and inhibit specific

enzymes such as the angiotensin-converting enzyme (ACE) that raises blood pressure.

Polyphenols can be classified as non-flavonoids and flavonoids. j However, only the

flavonoids quercetin and phytoestrogens have been extensively studied. Quercetin was

found to possess anti-allergy, anti-inflammatory, immune modulating, anti-viral,

anticancer, lipid antioxidant and gastro-protective properties.k Phytoestrogens are those

flavonoid compounds that have the similar structure to mammalian estrogens. Food

sources that are rich in phytoestrogens are nuts, oil seeds, soy beans, cereals, breads,

legumes and flax seeds. Studies have shown that eating foods containing phytoestrogens

can help to reduce the symptoms of menopause and cardiovascular diseases.l,m,n However

phytoestrogens are also known to be endocrine-disrupting chemicals (EDC) with strong

anti-estrogenic activity.o

i Hertog, M.G.; Feskens. E. J.; Hollman. P. C.; Katan. M. B.; Kromhout. D., Lancet., 1993, 342, 8878. j Kinsell, J. E.; Frankel, E.; German, B.; Kanner, J., Food Technology, 1993, 47, 85. k Murray, R. G.; Granner, D. K.; Mayes, P. A.; Rodwell, V. W., Harper’s Biochemistry. 23 ed.; Appleton and Lange, New York, 1929; p 196. l Osaski, A. L.; Kennelly, E. J. Phytother. Res. 2003, 17, 845. m Linseisen, L.; Pillar, R.; Hermann, S.; Change-Claude, J., Int. J. Cancer 2004, 110, 284. n Bingham, S. A.; Atkinson, C.; Liggins, J.; Bluck, L.; Coward, A., Br. J. Nutr. 1998, 79, 393. o Barrett, J., Environ. Health Perspect. 1996, 39, 104.

Page 94: Experimental and computational studies of the unimolecular

80

Such huge interest in the phenolic compounds led to several studies on the

analysis and detection of phenols and polyphenols from biological and environmental

samples by using a combination of chromatographic and mass spectrometric techniques

or perhaps by using tandem mass spectrometry.p,6,q,r,s

The continuing interest in analysis and detection of phenolic compounds

from environmental and biological samples over the past two decades led to a numerous

studies, aimed at obtaining further structural information to better characterize phenolic

compounds. Numerous Collision-Induced Dissociation (CID) experiments on phenoxide

anions were conducted to enhance the understanding of basic fragmentation mechanism

of phenoxide anion, which as such has provided characteristic fragmentation finger

prints for identification and detections of phenolic compounds from biological and

environmental samples.7-11 Interestingly, in all the CID studies of phenoxide anion,

phenoxide anion was consistently found to fragment via a loss of 28 Da corresponding to

the loss of CO. Other prominent losses were the loss of H2O, C2H4 leading to the

formation of benzyne and cyclobutenone anions.7-11 This facile loss of CO was also

observed in the dissociation of polyphenolic compounds such as catechol, resorcinol and

hydroquinones.12-14

Binkley and co-workers proposed a mechanism for the facile loss of CO

involving the loss of the ipso carbon (Scheme-3.1).11 In the proposed mechanism, the

resonance structure (B) undergoes a ring closure reaction to form a bicyclic structure (C).

The ring closure reaction is perhaps a nucleophilic reaction, where the carbon at the ortho

p Globig, D.; Weickhardt, C., Anal. BioAnal. Chem. 2003, 377, 1124. q Kang, J.; Hick, L. A.; Price, W. E., Rapid Commun. Mass Spectrom. 2007, 21, 4065. r Kang, J.; Hick, L. A.; Price, W. E., Rapid Commun. Mass Spectrom. 2007, 21, 857. s Kang, J.; Price, W. E.; Hick, L. A, Rapid Commun. Mass Spectrom. 2006, 20, 2411.

Page 95: Experimental and computational studies of the unimolecular

81

position attacks an equivalent ortho carbon. The structure (C) eventually undergoes a ring

opening reaction to form the cyclopenta-2,4-dienyl(oxo)methanide anion (D). The

methanide anion subsequently forms the resonance stabilised cyclopentadienyl anion (E)

via a loss of CO. Even, though the transition states may demand high energy input, the

driving force for this process could be the formation of resonance stabilised

cyclopentadienyl anion and stable CO neutral (Scheme-3.1).11

O O O

+ CO

A B C D

C O

E

Scheme-3.1: Binkley and coworker’s mechanism of phenoxide fragmentation11

Interestingly, it was a CID study of phenoxy ethoxide anion by Bowie and

co-workers that provided the evidence for Binkley’s mechanism. In the study, it was

hypothesized that the phenoxy ethoxide anion undergoes a Smiles rearrangement in the

gas-phase i.e. an intramolecular nucleophilic aromatic substitution reaction at the ipso

position of the aromatic ring (Scheme-3.2).

O

OO

O

O

O

Scheme-3.2: Smiles rearrangement in the Gas-Phase

The CID of phenoxy ethoxide anion, resulted in the formation of a product

ion at m/z 93 corresponding the phenoxide anion (Scheme-3.3). The formation of

phenoxide anion during fragmentation was confirmed by comparing the CID mass

spectrum of the product ion at m/z 93 with that of

Page 96: Experimental and computational studies of the unimolecular

82

the authentic phenoxide anion.

O

OO

O

O

Smiles rearrangement

O CID_

Scheme-3.3: Fragmentation of phenoxy ethoxide anion

To probe the occurrence of Smiles rearrangement in the gas-phase, Bowie

and co-workers conducted labeling experiments involving, the isotopic labeling of the

terminal oxygen atom with 18O-oxygen in phenoxy ethoxide anion (Ph16O(CH2)218O-) and

then subjecting it to CID. The product ion spectrum obtained after CID, contained equal

proportions of the two possible phenoxide fragment ions , one at m/z 93 (Ph16O-), and the

other at m/z 95 (Ph18O-) (Fig-3.1), thus providing compelling evidence for the occurrence

of Smiles rearrangement in the gas-phase (Scheme-3.4).15

Page 97: Experimental and computational studies of the unimolecular

83

Fig-3.1: CID Mass Spectrum of the labeled ethoxide anion (Ph16O(CH2)218O-)15

O

O1818O

O

O18

O

CID

m/z 93

Smiles rearrangement

O CID

18O

m/z 95

Scheme-3.4: Fragmentation of the labeled phenoxy ethoxide anion (Ph16O(CH2)218O-)

Furthermore, to probe into the reaction specificity at the ipso position of

the aromatic ring, the ipso carbon of PhO(CH2)2O- was labeled with 13C-carbon and then

was subjected to CID. The product phenoxide ion obtained from collision-induced

dissociation of the labeled phenoxy ethoxide anion, upon further collisional activation,

Page 98: Experimental and computational studies of the unimolecular

84

fragmented exclusively by the loss of 13CO thus providing compelling evidence for the

reaction specificity (Scheme-3.5).15

Scheme-3.5: Fragmentation of the 13C labeled phenoxy ethoxide anion (PhO(CH2)2O-)

As such, both labelling studies provided evidence for the Smiles

rearrangement, but in addition the results provided evidence for the loss of ipso carbon as

CO during the fragmentation of the phenoxide anion. Infact, these results were used by

Binkley and co-workers in conjunction with their study, to explore and propose the

mechanism for the fragmentation of phenoxide anion involving the ipso carbon, leading

to formation of cyclopentadienyl anion via a loss of CO (Scheme-3.1).

In a recent study on the dissociation of perbenzoate anions by Blanksby

and co-workers, fragment ions corresponding to the formation of phenoxide anion were

observed following the loss of CO2.16 The spectrum obtained after further dissociation of

the fragment ion matched with that of the authentic phenoxide anion. Even though,

experimental evidence supported the formation of phenoxide anion in the rearrangement

13 O

O

13

O

O

O

13O

CID

m/z 94

Smiles rearrangement

O CID

13O

m/z 94

CID CID13CO13CO

C5H5- C5H5

-

Page 99: Experimental and computational studies of the unimolecular

85

process, the complementary computational study on the rearrangement surprisingly,

revealed that the facile loss of CO2 occurs via an intramolecular epoxidation of benzene

ring, leading to the formation of benzene-oxide and oxepin as intermediates (Scheme-

3.6).The result suggests that the benzene-oxide and oxepin intermediates may be linked

via facile rearrangement to phenoxide anion in the gas-phase.16

OOO

ortho

ipso

OO

O OO

O OCO2

OO

OOOO OC

O

O

CO2O

Scheme-3.6: Fragmentation of Perbenzoate anion16

Our study primarily aims at elucidating the mechanism for the loss of CO from

phenoxide anion by computational methods but in addition we aim to elucidate the

possible rearrangement of benzene-oxide and oxepin intermediates to phenoxide anion.

3.2 Materials and methods

Geometric optimizations were carried out with the B3LYP method17, 18

using 6-311++G(d,p) basis set within the Gaussian suite of programs.19 All stationary

points on the potential energy surface were characterized as either minima (no imaginary

frequencies) or transition state (one imaginary frequency) by calculation of frequency

using analytical gradient procedure. Frequency calculations provided zero-point energies,

which were added to the calculated energy. The minima connected by a given transition

state were confirmed by inspection of the animated imaginary frequency using the Gauss

Page 100: Experimental and computational studies of the unimolecular

86

view package 3.0 (Gaussian, Inc, Pittsburgh, USA) and by intrinsic reaction coordinate

calculation.20, 21

3.3 Results and discussion

In an effort to explore the validity of the mechanism proposed by Binkley

and co-workers, the structures in Scheme-3.1 were optimized at the B3LYP/6-311++G(d,

p) level of theory. The bicyclic structure in Scheme-3.1(c) however could not be

optimized either as a stable minima or transition state. Hence, a new mechanism was

proposed, which involves ketene as an intermediate (Scheme-3.7). In the newly proposed

mechanism, the phenoxide anion (A), ring opens to form a ketene intermediate (B). The

ketene subsequently undergoes a ring closure reaction to form cyclopenta-2,4-

dienyl(oxo)methanide anion (C). The anion subsequently undergoes a decarbonylation

reaction to give the observed cyclo pentadienyl anion (Scheme-3.7).

OC

O CO

+ CO

A B C D

Scheme-3.7: Proposed fragmentation pathway for phenoxide decomposition

Electronic structure calculations have been performed to investigate the

mechanism in detail. The calculated potential energy for the fragmentation of the

phenoxide anion is shown in Figure-3.2 with energies given relative to the phenoxide

anion. The electronic energies of the critical stationary points are listed in Table-3.1 with

their molecular structure provided in Figure-3.3. The full Cartesian coordinates of the

Page 101: Experimental and computational studies of the unimolecular

87

stationary points are provided in Appendix-3. The energies of the various intermediates

and transition states are given relative to the phenoxide anion

Page 102: Experimental and computational studies of the unimolecular

88

Fig-3.2: Reaction coordinate diagram for the fragmentation of the phenoxide anion in the gas-phase calculated at B3LYP/6-311++G(d,p) level of theory. All energies given in kJ mol-1.

0.0

438.7403.4

139.6

390.0

352.9371.6

396.7

359.6

166.4 151.5

TS1

IM1 TS2IM2

TS3

IM3

TS4 TS5

P1

P2

+

Page 103: Experimental and computational studies of the unimolecular

89

(a) Phenoxide (b) TS1

(c) IM1 (d) TS2

(e) IM2 (f) TS3

Page 104: Experimental and computational studies of the unimolecular

90

(g) TS4 (h) IM3

(i) TS5 (j) P2

Page 105: Experimental and computational studies of the unimolecular

91

(k) P1

Fig-3.3: Structures of the stationary points on the phenoxide potential energy surface optimized at B3LYP/6-311++G(d,p) level of theory. All lengths given in angstroms and angles in degree. All the dihedral angles (Ø) are defined with respect to the plane below, containing the three carbon atoms. Considering the phenoxide fragmentation via ketene like intermediate, the

phenoxide anion undergoes a ring opening reaction via the transition state TS1 (Fig-3.2).

The transition state presents a barrier to 397 kJ mol-1 above the entrance channel (Fig-

3.2). From the molecular structure of TS1 (Fig-3.3(b)), a rather surprising dihedral angle

of 96° for the terminal C=C bond about the plane of the aromatic ring was observed. This

suggests that the ring opening reaction is just not a simple heterolytic cleavage, but also

encompasses rotation of the terminal C=C out of the plane of the ring. IRC calculation on

the transition state TS1 locates a ketene intermediate, IM1 (Fig-3.2), in which the

terminal C=C bond has rotated nearly 180° from its original position in phenoxide (Fig-

3.3(c)). Following the formation of the ketene like intermediate IM1, another rotomer,

represented by IM2 on the potential surface was identified (Fig-3.2). The two rotomers

are connected by a barrier of 372 kJ mol-1 and represented by TS2 (Fig-3.2). The

intermediate IM2 differ from IM1, in the orientation of the ketene moiety (Fig-3.3(d)),

and the transition state TS2 shows an intermediate dihedral angle of 80° (Fig-3.3(c)), thus

Page 106: Experimental and computational studies of the unimolecular

92

confirming that the process involves the formation of the rotomer. The intermediate IM2

may undergo a 1,5-proton transfer via the transition state TS3 over a barrier of 390 kJ

mol-1 to form a proton transferred ketene-like intermediate, IM3 (Fig-3.2). However, the

transition state TS3, again encompasses the process of rotation of the terminal C=C bond,

as could be seen from dihedral angle of 80° (Fig-3.3(e)). The proton transferred ketene

IM3, is 150 kJ mol-1 lower in energy than the ketene intermediate IM2 (Fig-3.2). The

stability of the intermediate IM3 is due to the delocalization of charge on to oxygen, as

indicated in Scheme-3.8. The structure of IM3 supports the presence of the resonance

contributing structure as in the structure the length of C=C bond of the ketene moiety is

1.25 Å, clearly providing evidence for the existence of a triple bond and hence the charge

delocalization on to the oxygen atom (Fig-3.3(h)).

Scheme-3.8: Resonance contributing structure of the intermediate IM3

The proton transferred ketene IM3 subsequently undergoes a complex

unimolecular reaction involving the coupling of a proton transfer together with ring

closing process to give the cyclopentadienyl anion via the loss of CO. The initial ion-

dipole complex P1 presents an energy of 139 kJ mol-1 (Fig-3.2) and the structure of P1

(Fig-3.3(k)) indicates a strong π electron interaction which stabilizes the complex. The

reaction occurs over a barrier of 439 kJ mol-1, and the representative transition state to

the process TS5 provides evidence for the complex reaction process (Fig-3.3(i)).

CO

CO

Page 107: Experimental and computational studies of the unimolecular

93

The generalized mechanism that could be summarized encompassing all

the important transition states and intermediates from the calculation is given in Scheme-

3.9.

O

C

O

CO

+CO

H HCO

CO

CO

Scheme-3.9: Calculated fragmentation mechanism for phenoxide anion

Surprisingly, a direct transition state TS4, leading to the proton transferred

ketene IM3 and was confirmed by IRC calculation (Fig-3.2). This is rather fascinating, as

both ring opening and proton transfer reactions are coupled together in the same

transition state and the representative transition state TS4 and the processes occur over a

comparable barrier of 403 kJ mol-1 (Fig-3.2). Investigation of the structure of TS4

identifies a intermediate dihedral angle of 78° for C=C of the carbonyl group, which in

this case provides the change in orientation required for the proton transfer to form IM3

(Fig-3.3(g)). The plausible explanation for the coupling of the proton transfer reaction

could be because of the high basicity of the terminal alkene moiety and that the open

ketene structure in the orientation of TS4 is unstable leading to the formation of IM3

immediately after ring opening of the phenoxide anion.

The generalized mechanism encompassing the complex ring cleavage

coupling a proton transfer can be summarized as in Scheme-3.10.

Page 108: Experimental and computational studies of the unimolecular

94

+CO

HC

O

CO

C

OOC

O

H

Scheme-3.10: Calculated pathway of phenoxide fragmentation

From these calculations, it is suggestive that the ketene intermediates

dictate the course of the reaction, that the two calculated reaction pathways are equally

competitive. Both the calculated pathways are consistent with the loss of ipso carbon as

CO.

To investigate the involvement of benzene-oxide and oxepin intermediates

in the fragmentation of phenoxide anion, electronic calculations were performed at

B3LYP/6-311++G(d,p) level of theory. The calculated potential energy for the

rearrangement to benzene-oxide and oxepin intermediates is shown in Figure-3.4 with

energies given relative to the phenoxide anion. The electronic energies of the critical

stationary points are listed in Table-3.1 with their molecular structure given in Figure-

3.5. The full Cartesian coordinates of the various stationary points are provided in

Appendix-3.

Page 109: Experimental and computational studies of the unimolecular

95

0.0

530.1

408.5

336.5

233.9

470.9

451.2

362.6

507.5

387.7 388.2

151.5

396.1

479.2

337.8

41.8

139.6

TS7

TS11

TS10

IM4IM6

TS13

IM7

TS6

IM8

P1P2

TS12

+

IM5

TS8

TS9

+

Fig-3.4: Potential energy diagram showing the involvement of benzene-oxide and oxepin structures in the fragmentation of phenoxide calculated at B3LYP/6-311++G(d,p) level of theory. All energies given in kJ mol-1.

Page 110: Experimental and computational studies of the unimolecular

96

(a) Phenoxide (b) TS6

(c) TS7 (d) IM4

(e) TS8 (f) IM5

Page 111: Experimental and computational studies of the unimolecular

97

(g) TS9 (h) IM6

(i) TS10 (j) IM7

(k) TS11 (l) IM8

Page 112: Experimental and computational studies of the unimolecular

98

(m) TS12 (n) TS13

(o) P1 (p) P2

Fig-3.5: Structures of the stationary points on the benzene-oxide and oxepin potential energy surface optimized at B3LYP/6-311++G(d,p) level of theory. All lengths, given in angstroms and angles in degrees. All the dihedral angles (Ø) are defined with respect to the plane below, containing the three carbon atoms.

Considering the rearrangement pathway, the phenoxide anion undergoes a

rearrangement to oxepin intermediate IM4 via two different transition states. One of the

transition states, TS7 (Fig-3.5(c)) represents a complex process involving an initial

heterolytic cleavage of C-C bond followed by C-O bond formation to form the seven

membered oxepin intermediate IM4 (Fig-3.4(d)).The transition state TS7 was confirmed by

Page 113: Experimental and computational studies of the unimolecular

99

an IRC calculation and it presents a barrier of 470.9 kJ mol-1 (Fig-3.4). The other transition

state TS6 represents an attack of oxygen on to ring at the ortho position and the reaction was

found to be a cyclization process as there was a alternation of double bonds from TS6 (Fig-

3.5(b)) to IM4 (Fig-3.5(d)).The barrier to the process TS6 was found to be 451.2 kJ mol-1

relative to the phenoxide anion (Fig-3.4). The transition state TS6 resembles that of a

benzene-oxide anion as the bond angle of that of a epoxy moiety (C−O−C) is 56° with

dihedral of -56° out of the plane of the benzene ring (Fig-3.5(b)) and is comparable with that

of benzene-oxide where the angle of epoxy moiety is 66° with the corresponding dihedral

angle of -52° (Fig-3.5(j)). The alternation of double bonds in TS6 also matches with that of

the benzene-oxide (IM7) (Fig-3.5(j)).

Following the formation of the oxepin intermediate, a rather surprising one

step ring closure reaction to effect the loss of CO was identified. The transition state TS13 for

this process occurs over a barrier of 508 kJ mol-1 (Fig-3.4), where the loss of CO occurs

perpendicular to the plane of the molecule (Fig-3.5(n)).

The generalised mechanism encompassing all the intermediates to this process

is given in Scheme-3.11.

OC

O

O

O

O

+ CO

Scheme-3.11: Calculated pathway for oxepin fragmentation

Page 114: Experimental and computational studies of the unimolecular

100

Now considering the oxepin to benzene-oxide rearrangement, the oxepin

structure IM4 rearranges to IM5, through a ring opening reaction (Fig-3.5). The transition

state TS8 was confirmed by an IRC calculation and it presents a barrier of 408 kJ mol-1 (Fig-

3.4). The new intermediate IM5, is 130 kJ mol-1 lower in energy than the oxepin intermediate

IM4 (Fig-3.4). The stability of the intermediate IM4 is due to the delocalization of the charge

on to the oxygen atom (Scheme-3.12). The structure of IM4 supports the presence of the

resonance contributing structure as in the structure, we could see the simultaneous presence

of triple bond character at the terminal C-C bond and double bond characteristics in the

carbonyl group, thus clearly providing evidence to charge delocalization on to the oxygen

atom (Fig-3.3(h)).

C

HO

C

HO

Scheme-3.12: Resonance contributing structure of the intermediate IM5

Following its formation, the intermediate IM5, undergoes a ring closure

reaction IM6 to form the other oxepin isomer IM6 (Fig-3.4). The transition state TS9 (Fig-

3.5(f)) represents a barrier of 337 kJ mol-1 (Fig-3.4).

The intermediate IM6 then rearranges to the benzene-oxide anion IM7 over a

barrier of 388 kJ mol-1. The transition state to this process is represented by TS10 (Fig-3.4)

and as seen from the transition state structure the process involves a ring closure reaction to

form an epoxy group, with shortening of both the C-C bond distance (1.74 Å) adjacent to the

oxygen atom and the C-O-C bond angle (75°) (Fig-3.5(i)) as compared to the oxepin

intermediate (C-C = 2.35 Å, C-O-C = 108°) (Fig-3.5(h)). The energy of the oxepin

intermediate IM5 is 334 kJ mol-1 and hence the rearrangement occurs over a small barrier of

Page 115: Experimental and computational studies of the unimolecular

101

55 kJ mol-1 relative to that of the oxepin intermediate (Fig-3.4). Such rearrangements were

studied for neutral benzene-oxide and oxepin. The two structures are known to exist in

equilibrium in solution, with equilibrium position favouring oxepin at lower temperatures and

benzene-oxide at ambient temperatures.22 Calculation on the neutral system under the same

level of theory found this barrier to rearrangement at 30 kJ mol-1. Calculations were also

conducted for the conversion of the oxepin IM4 to its corresponding benzene-oxide.

However, in this situation the benzene oxide could not be located as stable minima and the

transition state TS10 could be considered as the benzene-oxide structure (Fig-3.4).

The benzene-oxide then undergoes a complex rearrangement to form IM8

over a barrier of 480 kJ mol-1. The transition state to this process is represented as TS11 (Fig-

3.4) and was confirmed by an IRC calculation. On close observation of the structure and after

detailed investigation of the mechanism, the transition state process involves a cross ring

nucleophilic substitution of the negatively charged carbon, thus aiding in the formation of

CHO moiety (Fig-3.5(l)). The transition state structure in fact resembles the intermediate

proposed by Binkley and co-workers (Scheme-3.1(c)).11 The methanolate anion, IM7, is

surprisingly stable, only 42 kJ mol-1, more energetic than phenoxide anion (Scheme-3.13).

C OH

C OH

C OH

C OH

Scheme-3.13: Resonance structures of methanolate anion

Finally, intermediate IM8, then subsequently undergoes a simple 1,2-proton

transfer, via transition state TS12 (Fig-3.5(m)), over a barrier of 396 kJ mol-1 to effect the

formation of cyclopentadienyl anion and the loss of CO (Fig-3.4).

Page 116: Experimental and computational studies of the unimolecular

102

The generalized mechanism encompassing all the intermediates to the process

can be hence summarized as in Scheme-3.14.

O CO

O

O

HCH

O

OO

OC O

H

OH

+ CO

CO

Scheme-3.14: Rearrangement of the oxepin to benzene-oxide anion

Interestingly, the intermediate IM8 could also undergo a homolytic cleavage

to effect a loss of CHO.. Bowie and co-workers have observed such loss of CHO. from

phenoxide anions, but only to a minor extent as compared with the CO loss (CHO.:CO

≈1:5).15 The calculated energy of the cyclopentadienyl radical anion and formy radical is 530

kJ mol-1 above the phenoxide anion. This is 50 kJ mol-1 higher in energy than the barrier (s)

for the CO loss (Fig-3.4). The result from the calculation is hence consistent with the

experimental observation. Furthermore, Bowie and co-workers observed the loss of CDO.

Page 117: Experimental and computational studies of the unimolecular

103

from 2,4,6 D3 phenoxide anion. The experimental observation here again is consistent with

the mechanism arising from our calculation.

O

CO

O

O

D

CD

O

OOOC O

D

OD

+ DCO

D

D

D

D

D

D

D

D

D

DD

D

D

D

D

D D

D

DD

D

DD

D

D

D

Scheme-3.15: Fragmentation pathway for the loss of CHO.

These calculations have identified a pathway for the loss of CO via

rearrangement to the oxepin structure from phenoxide anion (Scheme-3.11). This calculation

suggest that the transition state TS13 in the direct decarbonylation pathway is a rate

determining step, and that the loss of CO via this pathway could be unfavourable. Our

calculations clearly suggest phenoxide anion rearranges to oxepin and benzene-oxide

intermediates and that the rearrangement pathway eventually effect the loss of CO (Scheme-

Page 118: Experimental and computational studies of the unimolecular

104

3.13). The rearrangement pathway is more energetically favourable than the direct

decarbonylation pathway (Fig-3.4).

The other intriguing question would be existence of rearrangement pathways

between ketene intermediates and the oxepin/benzene-oxide intermediates. We might expect

the existence of a transition state between the oxepin intermediate, IM4 (Fig-3.4) and ketene

intermediate, IM1 (Fig-3.2) (Scheme-3.16). Unfortunately, no transition state structure could

be identified between the two intermediates.

CO O

Scheme-3.16: Possible rearrangement between ketene and oxepin intermediates

As discussed earlier, a combined mass spectrometry and computational study

on the rearrangement of perbenzoate anion by Blanksby and co-workers, revealed that a facile

loss of CO2 occurs via an intramolecular epoxidation of the benzene ring, leading to the

formation of benzene-oxide and oxepin as intermediates (Scheme-3.6). Isotopic labelling

studies were also conducted, where the ipso carbon was labeled with 13C carbon and was

subjected to CID. The resulting ion, after the loss of CO2, upon further collisional activation

resulted in an exclusive loss of unlabeled CO.16 As such if these results when interpreted

with the calculated benzene-oxide and oxepin rearrangement from this study suggest that it

would result in the loss of unlabeled CO if benzene-oxide is formed from the perbenzoate

anion i.e. the loss of CO2 occurs via initial nucleophilic substitution at the ortho position of

the perbenzoate anion (Scheme-3.17).

Page 119: Experimental and computational studies of the unimolecular

105

13C

OOO

or tho 13C OO

O

13C

OO

O

13C

OCO2

13C

O

13CC

O

H

13C

OH

13C

+ CO

13C

CO

Scheme-3.17: Rearrangement of the labeled perbenzoate anion through initial nucleophilic attack at the ortho position Since, the rearrangement of oxepin to benzene-oxide is the more energetically

favoured pathway, the formation of oxepin from the perbenzoate anion, i.e. the loss of CO2

occurs via initial nucleophilic substitution at the ipso position of the perbenzoate anion would

eventually again lead to the loss of labeled CO (Scheme-3.18).

Page 120: Experimental and computational studies of the unimolecular

106

13C

OOO

13CO

OO

13C

OO

O13C

OC

O

O

CO2

13CO

ipso

13C O

H

13C

HO

13CO

C13CO

13C O

+

Scheme-3.18: Rearrangement of the labeled perbenzoate anion through initial nucleophilic attack at the ipso position The calculation does provide compelling evidence to the formation of

benzene-oxide anion i.e. the loss of CO2 occurs via initial nucleophilic substitution at the

ortho position of the perbenzoate anion, which more or less is in accordance with the study.16

Page 121: Experimental and computational studies of the unimolecular

107

Table-3.1: Optimized stationary points calculated for the loss of CO from phenoxide anion (Electronic energy zero-point energy, relative energy and imaginary frequency calculated at B3LYP/6-311++G (d, p) level of theory)

Structure

Energy

Hartrees

Zero-point

Energy

Hartrees

Relative energy

kJ mol-1

Phenoxide

-306.99406

0.08945

0

TS1 (-135 cm-1

) -306.83661

0.083076

396.7

IM1 -306.85145

0.08379

359.6

TS2 (-57 cm-1

) -306.84693

0.08386

371.6

IM2 -306.85418

0.08400

352.9

TS3 (-191 cm-1

) -306.83917

0.08311

390.0

IM3 -306.92784

0.08662

166.4

TS4(-133 cm-1

) -306.83413

0.08317

403.4

TS5(-453 cm-1

) -306.82017

0.08265

438.7

P1 -306.93597

0.08452

139.6

TS6(-824 cm-1

) -306.81671

0.08396

451.2

TS7(-225 cm-1

) -306.81033

0.08508

470.9

IM4 -306.85362

0.08712

362.6

IM6 -306.86299

0.08705

337.8

TS10(423 cm-1

) -306.84382

0.08705

388.2

IM7 -306.84514

0.08821

387.7

TS11(310 cm-1

) -306.80762

0.08551

479.2

IM8

-306.97735

0.08867

41.8

TS13(-476 cm-1

) -306.79677

0.08547

507.5

Page 122: Experimental and computational studies of the unimolecular

108

TS12-1741cm-1

) -306.83577

0.08204

396.1

Cyclopentadienyl

anion

-193.58082 0.077923

CO -113.34905 0.00504

TS8(-248cm-1

) -306.83085

0.08184

408.5

IM5 -306.90020

0.08467

233.9

TS9(-329cm-1

) -306.86238

0.08593

336.5

3.4 Conclusion

Computational investigations have provided an insight into the

fragmentation of phenoxide anion via the loss of CO and that the fragmentation mechanism

involves ketene like intermediates and transition states and that it truly complements

previously conducted isotopic labeling studies. The calculation also provides evidence for

direct decarbonylation of oxepin intermediate. Our calculations suggest the existence of

rearrangement pathway of phenoxide anion to benzene-oxide and oxepin structures

eventually effecting the loss of CO. The calculations some what provides complementary

and supporting evidence for the previously reported study involving benzene-oxide and

oxepin products from the decarboxylation of perbenzoate anions.

Page 123: Experimental and computational studies of the unimolecular

109

References for chapter three

1. Zweig, G.; Sherma, J.; Hanai, J., Phenols and organic acids. CRC Press: Boca Raton,

Fla, 1982. 2. Afghan, B. K.; Chau, S. Y., Analysis of Trace Organics in the Aquatic Environment.

CRC Press: Boca Raton, Fla, 1989. 3. Shahidi, F.; Ho, Chi-Tang., Phenolic compounds in foods and natural health products.

Oxford University Press: Oxford, 2005. 4. Bartsch, H.; Nair, J.; Owen, R. W., J. Biol. Chem. 2002, 383, 915. 5. Hashim, Y. Z.; Eng, M.; Gill, C. I.; McGlynn, H.; Rowland, I. R., Nutr. Rev. 2005, 63,

374. 6. Knust, U.; Erben, G.; Spiengelhalder, B.; Bartsch, H.; Owen, R. W., Rapid Commun.

Mass Spectrom. 2006, 20, 3119. 7. Anderson, G. B.; Gillis, R. G.; Johns, R. B.; Porter, Q. N.; Strachan, M. G., Org. Mass

Spectrom. 1984, 19, 99. 8. Anderson, G. B.; Gillis, R. G.; Johns, R. B.; Porter, Q. N.; Strachan, M. G., Org. Mass

Spectrom. 1984, 19, 583. 9. Bowie, J. H., Mass Spectrom. Rev. 1990, 9, 349. 10. Busch, K. L.; Norstom, A.; Nilsson, C. A.; Bursey, M. M.; Hass, J. R., Environ.

Health Perspec. 1980, 36, 125. 11. Binkley, R. W.; Fletcher, T. W.; Winnik, W., J. Org. Chem. 1992, 57, 5507. 12. Binkley, R. W.; Dillow, G. W.; Fletcher, T. W.; Winnik, W.; Tevesz, M. J. S., Org.

Mass Spectrom. 1994, 29, 491. 13. Flechtner, T. W.; Winnik, B.; Winnik, W.; Tevesz, M. J. S., J. Mass Spectrom. 1996,

31, 377. 14. Eichinger, P. C. H.; Dua, S.; Bowie, J. H., Rapid Commun. Mass Spectrom. 1997, 11,

1996. 15. Eichinger, P. C. H.; Bowie, J. H.; Hayes, R. N., J. Am. Chem. Soc. 1989, 111, 4224. 16. Harman, D. G.; Ramachandran, A.; Gracanin, M.; Blanksby, S. J., J. Org. Chem. 2006,

71, 7996.

Page 124: Experimental and computational studies of the unimolecular

110

17. Becke, A. D., J. Chem. Phys. 1993, 98, 1372. 18. Lee, C. T.; Yang, W. T., Parr, R. G., Phys. Rev. B: Condens. Matter 1988, 37, 785. 19. Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.;

Cheeseman, J. R.; Montgomery, Jr., J. A.; Vreven, T.; Kudin, K. N.; Burant, J. C.; Millam, J. M.; Iyengar, S. S.; Tomasi, J.; Barone, V.; Mennucci, B.; Cossi, M.; Scalmani, G.; Rega, N.; Petersson, G. A.; Nakatsuji, H.; Hada, M.; Ehara, M.; Toyota, K.; Fukuda, R.; Hasegawa, J.; Ishida, M.; Nakajima, T.; Honda, Y.; Kitao, O.; Nakai, H.; Klene, M.; Li, X.; Knox, J. E.; Hratchian, H. P.; Cross, J. B.; Bakken, V.; Adamo, C.; Jaramillo, J.; Gomperts, R.; Stratmann, R. E.; Yazyev, O.; Austin, A. J.; Cammi, R.; Pomelli, C.; Ochterski, J. W.; Ayala, P. Y.; Morokuma, K.; Voth, G. A.; Salvador, P.; Dannenberg, J. J.; Zakrzewski, V. G.; Dapprich, S.; Daniels, A. D.; Strain, M. C.; Farkas, O.; Malick, D. K.; Rabuck, A. D.; Raghavachari, K.; Foresman, J. B.; Ortiz, J. V.; Cui, Q.; Baboul, A. G.; Clifford, S.; Cioslowski, J.; Stefanov, B. B.; Liu, G.; Liashenko, A.; Piskorz, P.; Komaromi, I.; Martin, R. L.; Fox, D. J.; Keith, T.; Al-Laham, M. A.; Peng, C. Y.; Nanayakkara, A.; Challacombe, M.; Gill, P. M. W.; Johnson, B.; Chen, W.; Wong, M. W.; Gonzalez, C.; and Pople, J. A.; Gaussian 03

(Revision C.02), Gaussian, Inc: Wallingford CT, 2004. 20. Gonzalez, C.; Schlegel, H. B., J. Chem. Phys. 1989, 90, 2154. 21. Gonzalez, C.; Schlegel, H. B., J. Phys. Chem. 1990, 94, 5523. 22. Kassaee, M. Z.; Arshadi, S.; Taheri, N. A., J. Mol. Structure; THEOCHEM 2005, 715,

107.

Page 125: Experimental and computational studies of the unimolecular

111

Appendix-1 Table-A_1.1: Tandem mass spectra of [M-H+]- Ions from azo dyes anions para substituted phenyl and sulphonic acid analogues (where phenyl is represented as ph, and Anilinyl as An) Ortho- Ph (m/z) Abundance Loss or fragment

402 100 [M-H]-

385 1 -NH3

374 8 - N2

372 2 -N2H2

338 13 -SO2

322 58 -SO3

310 1 -SO2 -N2

294 4 -SO3 -N2

234 21 -An

221 6 -N2 -ph

Ortho- SO3Hm/z) Abundance Loss or fragment

406 100 [M-H]-

390 9 -NH2

378 5 - N2

376 4 -N2H2

342 ? -SO2

326 50 -SO3

310 4 ?

298 23 -SO3 -N2

296 4 -H2 SO3 - N2

234 6 -An

221 2 -N2 -ph

Page 126: Experimental and computational studies of the unimolecular

112

Appendix-2 Table-A_2.1: The Cartesian coordinates for all the stationary points for the fragmentation of 4-amino-3(phenyldiazenyl)benzenesulfonte calculated at B3LYP/6-31+G(d) level as illustrated in Figure-2.16.

Structure

Geometry

4-amino-3(phenyldiazenyl)

Benzenesulfonate

Atom Label X Y Z

C 2.749175 1.540861 0.027289 C 1.698733 2.44557 0.028348 C 2.513212 0.153352 -0.017103 C 0.356253 2.012522 -0.012218 C 1.210566 -0.296502 -0.052588 C 0.107484 0.600014 -0.045459 H 1.008607 -1.362408 -0.08282 S 3.931968 -0.995466 0.008552 O 3.323058 -2.339877 -0.194214 O 4.790477 -0.521124 -1.116073 O 4.535106 -0.782304 1.356817 N -0.666322 2.927536 -0.050497 H -0.462836 3.883088 0.207468 H -1.607353 2.566647 0.07553 N -1.130179 -0.019395 -0.047216 N -2.185176 0.697695 -0.023125 C -3.393792 -0.045469 -0.007159 C -4.581617 0.703357 0.008337 C -3.468405 -1.451568 -0.005012 C -4.711118 -2.080747 0.012682 C -5.894966 -1.330085 0.028388 C -5.823496 0.06627 0.025894 H -2.547554 -2.024192 -0.016897 H -4.759716 -3.167647 0.014808 H -6.860499 -1.830374 0.042521 H -6.734831 0.660167 0.037306 H -4.51108 1.788193 0.004551 H 1.899309 3.516286 0.051928 H 3.774911 1.897838 0.05401

NH2

SO3-

NN

Page 127: Experimental and computational studies of the unimolecular

113

IM1

Atom X Y Z

C -1.266711 -0.37664 -0.000192 C -0.10646 0.485148 -0.00014 C -0.29745 1.95695 -0.00001 C -1.67048 2.422849 0.000096 C -2.72312 1.558057 0.000073 C -2.52687 0.131133 -8.7E-05 N 1.066401 -0.1364 -0.00018 N 2.180606 0.574577 -0.00011 C 3.426874 -0.05073 -4.8E-05 C 4.577843 0.758786 -1.6E-05 C 5.843793 0.175346 0.000066 C 5.985295 -1.21679 0.000117 C 4.83668 -2.01761 0.000082 C 3.56339 -1.45042 0.000001 N 0.752463 2.75163 0.000008 S -4.00804 -0.92653 0.000028 O -3.48523 -2.31961 -0.00108 O -4.71764 -0.52073 -1.249 O -4.71638 -0.52225 1.250263 H -1.10842 -1.45042 -0.00033 H 0.461723 3.732985 0.000108 H 2.671666 -2.06653 -1.9E-05 H 4.47043 1.841111 -5.4E-05 H 4.931416 -3.10121 0.000124 H 6.724262 0.814037 0.000092 H 6.973316 -1.66995 0.000185 H -1.83234 3.500276 0.00019 H -3.74369 1.930433 0.00013 H 2.081391 1.609212 -5.8E-05

SO3-

NNN

H H

Page 128: Experimental and computational studies of the unimolecular

114

TS2

Atom X Y Z

C -1.41391 0.916089 0.035798 C -0.01961 0.752001 -0.27623 C 0.457102 -0.51705 -0.80167 C -0.47519 -1.59603 -0.89431 C -1.79521 -1.39652 -0.57374 C -2.2782 -0.12919 -0.11096 N 0.794462 1.788933 -0.08191 N 2.089289 1.663589 -0.3193 C 2.902378 0.413128 -0.16678 C 4.172813 0.483663 -0.84807 C 5.297224 -0.15202 -0.34932 C 5.283925 -0.80779 0.896324 C 4.100331 -0.7734 1.640055 C 2.951816 -0.1502 1.15498 N 1.73759 -0.55232 -1.16966 S -4.05408 0.017562 0.275648 O -4.22118 1.41725 0.752275 O -4.26496 -1.03097 1.316028 O -4.71469 -0.28631 -1.02711 H -1.76436 1.875118 0.403931 H 2.098581 -1.47447 -1.42091 H 2.052084 -0.11828 1.764599 H 4.2046 0.99693 -1.80641 H 4.068868 -1.22377 2.630965 H 6.218345 -0.11445 -0.92932 H 6.17762 -1.28846 1.284251 H -0.1236 -2.56269 -1.25037 H -2.50982 -2.20954 -0.6661 H 2.599324 2.481298 0.009087

so3

NNHNH

Page 129: Experimental and computational studies of the unimolecular

115

IM2-Cn1

Atom X Y Z

C -1.43986 -0.91528 -0.24549 C -0.07075 0.805951 -0.56898 C 0.507465 -0.48868 -0.71046 C -0.3417 -1.60067 -0.58436 C -1.69688 -1.46798 -0.29066 C -2.25317 -0.19984 -0.09752 N 1.857444 -0.73302 -1.05441 C 2.974388 -0.46579 -0.25517 N 1.761506 2.08538 -1.08258 N 0.533227 2.063494 -0.81057 C 2.890422 0.199486 0.979991 C 4.042552 0.426106 1.734261 C 5.296111 -0.00249 1.287427 C 5.381183 -0.66967 0.05979 C 4.23813 -0.89649 -0.7039 S -4.01554 -0.03864 0.368119 O -4.35014 1.377154 0.051172 O -4.02232 -0.36695 1.822644 O -4.69717 -1.05178 -0.48741 H -1.86769 1.908045 -0.14924 H 1.968608 -1.56843 -1.61642 H 1.924567 0.524791 1.351274 H 4.316963 -1.40055 -1.66593 H 3.952382 0.941616 2.687993 H 6.345073 -1.01357 -0.30999 H 6.187745 0.176924 1.882795 H 0.089785 -2.59453 -0.69932 H -2.33147 -2.34503 -0.20304 H 1.987208 3.079161 -1.25589

so3

NN

HHN

Page 130: Experimental and computational studies of the unimolecular

116

IM2-Cn2

Atom X Y Z

C -1.35739 -0.76935 0.402679 C 0.003946 0.554668 -0.68283 C 0.545468 -0.74509 -0.58745 C -0.30093 -1.79818 -0.2065 C -1.6506 -1.57697 0.057515 C -2.18314 -0.28535 -0.03294 N 1.898762 -1.01486 -0.90305 C 3.025522 -0.54388 -0.22278 N 0.656527 2.780193 -0.86846 N 0.856193 1.588048 -1.18227 C 2.94694 0.30448 0.896431 C 4.110475 0.73367 1.535607 C 5.372976 0.327354 1.093172 C 5.453751 -0.52463 -0.01392 C 4.299688 -0.95208 -0.66698 S -3.95668 -0.00063 0.321814 O -4.15104 1.447758 0.03042 O -4.08424 -0.36957 1.760174 O -4.65891 -0.92679 -0.61015 H -1.81514 1.747056 -0.52911 H 2.0406 -1.9017 -1.36915 H 1.978424 0.620514 1.268203 H 4.375507 -1.59862 -1.53979 H 4.022396 1.39145 2.397471 H 6.423891 -0.85435 -0.38016 H 6.273432 0.667395 1.597923 H 0.11795 -2.79835 -0.10844 H -2.29991 -2.40113 0.337887 H -0.13743 2.872911 -0.19078

so3

NHN

NH

Page 131: Experimental and computational studies of the unimolecular

117

IM2-Cn3

Atom X Y Z

C -1.41549 0.908048 0.282747 C -0.08258 1.044897 -0.12703 C 0.532341 0.01201 -0.8726 C -0.23442 -1.10412 -1.22377 C -1.58416 -1.19679 -0.87631 C -2.1798 -0.18789 -0.11757 N 1.909161 0.076662 -1.26998 C 2.976852 -0.22691 -0.40908 N 1.369063 2.849986 -0.431 N 0.561003 2.254943 0.311645 C 2.832299 -0.27593 0.98836 C 3.938467 -0.54379 1.796966 C 5.201729 -0.773 1.244876 C 5.346175 -0.72859 -0.14642 C 4.251555 -0.4572 -0.96435 S -3.95837 -0.28704 0.311339 O -4.61632 0.482183 -0.78286 O -4.04182 0.353468 1.65223 O -4.24506 -1.74928 0.293187 H -1.85864 1.680486 0.903945 H 2.060724 -0.25291 -2.21596 H 1.860648 -0.10654 1.439735 H 4.377546 -0.41609 -2.04546 H 3.802533 -0.5748 2.875559 H 6.318675 -0.90497 -0.60132 H 6.05604 -0.98153 1.883491 H 0.24719 -1.9173 -1.7651 H -2.17586 -2.06409 -1.153 H 1.498102 2.3438 -1.33349

so3

NNHN

H

Page 132: Experimental and computational studies of the unimolecular

118

TS3

Atom X Y Z

C -1.38058 -0.83745 -0.49749 C 0.005068 0.540779 -0.86816 C 0.607199 -0.70637 -0.70264 C -0.16674 -1.71701 -0.11548 C -1.50713 -1.49112 0.262957 C -2.1208 -0.27011 0.052978 N 1.968444 -0.92927 -1.05207 C 3.067891 -0.49931 -0.30039 N 0.259204 2.743667 -0.58452 N 0.667403 1.736708 -1.11917 C 2.940319 0.329968 0.828325 C 4.075781 0.741505 1.528105 C 5.355011 0.335966 1.137128 C 5.48345 -0.49617 0.019162 C 4.359194 -0.90581 -0.69436 S -3.91785 -0.07611 0.318833 O -4.39352 0.431641 -1.00059 O -4.00238 0.917499 1.425005 O -4.39349 -1.44177 0.673086 H -1.98674 1.415518 -1.20472 H 2.122349 -1.79658 -1.55085 H 1.957943 0.648922 1.158666 H 4.472442 -1.53837 -1.57342 H 3.95053 1.385539 2.395548 H 6.468001 -0.82555 -0.30665 H 6.232713 0.66089 1.689619 H 0.304079 -2.67423 0.098765 H -2.10354 -2.30089 0.673256 H -0.99434 1.927285 0.127951

SO3

NHN

NH

Page 133: Experimental and computational studies of the unimolecular

119

IM3

Atom X Y Z

C -1.39365 -0.82687 -0.31779 C 0.045275 0.662404 0.119711 C 0.498068 -0.77874 0.013794 C -0.42203 -1.69792 0.506292 C -1.77108 -1.33395 0.661407 C -2.2749 -0.10699 0.128133 N 1.738354 -1.18167 -0.47069 C 2.994468 -0.64664 -0.17198 N 1.466365 2.463537 -0.28747 N 0.84178 1.569738 -0.73965 C 3.203532 0.19284 0.938565 C 4.478265 0.696364 1.2004 C 5.566378 0.359585 0.3907 C 5.362066 -0.4905 -0.70411 C 4.091339 -0.98031 -0.992 S -4.08543 0.079214 -0.04484 O -4.25524 1.331393 -0.82995 O -4.55287 0.154273 1.369422 O -4.49946 -1.16793 -0.75134 H -1.69976 1.730593 -0.8343 H 1.749214 -2.08866 -0.92107 H 2.375988 0.443373 1.593638 H 3.932329 -1.6133 -1.8629 H 4.617861 1.354192 2.054761 H 6.195795 -0.76066 -1.34849 H 6.556661 0.752182 0.606032 H -0.11737 -2.72818 0.688987 H -2.47904 -2.04812 1.070587 H 0.103756 0.96999 1.2111

SO3

NHN

NH

Page 134: Experimental and computational studies of the unimolecular

120

IM4

Atom X Y Z

C -1.61498 1.158878 0.137483 C -0.22906 1.310851 -0.14367 C 0.303455 0.185675 -0.74865 C -0.32863 -1.01331 -1.10234 C -1.68488 -1.09411 -0.80046 C -2.32323 -0.00626 -0.17593 N 1.781839 0.382968 -1.03533 C 2.755714 -0.48307 -0.35424 N 3.116623 3.582561 -0.02066 N 3.034815 4.526604 0.549033 C 2.551536 -0.79411 0.990301 C 3.495375 -1.58472 1.647509 C 4.620533 -2.06414 0.967177 C 4.806871 -1.7482 -0.3805 C 3.871511 -0.94939 -1.04617 S -4.09701 -0.19696 0.242012 O -4.72343 -0.65911 -1.0345 O -4.54102 1.153104 0.692273 O -4.09985 -1.23533 1.318187 H -2.17475 1.961676 0.612523 H 1.978694 0.398453 -2.04216 H 1.660434 -0.43063 1.492564 H 4.01199 -0.69976 -2.09622 H 3.343163 -1.8347 2.693907 H 5.673684 -2.12315 -0.91839 H 5.344829 -2.68789 1.484705 H 0.190158 -1.84728 -1.57729 H -2.25906 -1.98292 -1.04319 H 1.819285 1.374631 -0.68081

so3

H2NN2

Page 135: Experimental and computational studies of the unimolecular

121

TS5

Atom X Y Z

C -1.46498 -0.77986 -1.11741 C -0.10218 -1.15247 -1.12304 C 0.554062 -0.91519 0.081887 C 0.03102 -0.37659 1.257901 C -1.31877 -0.0236 1.205876 C -2.05563 -0.22065 0.024644 N 1.957369 -1.38639 -0.12905 C 3.072776 -0.45638 -0.05878 C 2.888911 0.869401 -0.46118 C 3.983002 1.736459 -0.46138 C 5.247642 1.291236 -0.06199 C 5.419281 -0.03714 0.337648 C 4.332092 -0.9145 0.336891 S -3.80074 0.343209 -0.0003 O -4.31032 0.017602 1.366094 O -4.44012 -0.43514 -1.09928 O -3.69504 1.809605 -0.26863 H -2.09535 -0.91979 -1.99227 H 2.1726 -2.24465 0.387636 H 1.900844 1.205945 -0.75755 H 4.463624 -1.94928 0.648139 H 3.841341 2.768419 -0.77146 H 6.39637 -0.39292 0.65489 H 6.092063 1.975691 -0.05825 H 0.617698 -0.22792 2.163799 H -1.82013 0.393738 2.073658 H 1.581664 -1.65671 -1.19236

so3

HN H+ N2

Page 136: Experimental and computational studies of the unimolecular

122

so3

H2N

IM5

Atom X Y Z

C -1.61498 1.158878 0.137483 C -0.22906 1.310851 -0.14367 C 0.303455 0.185675 -0.74865 C -0.32863 -1.01331 -1.10234 C -1.68488 -1.09411 -0.80046 C -2.32323 -0.00626 -0.17593 N 1.781839 0.382968 -1.03533 C 2.755714 -0.48307 -0.35424 N 3.116623 3.582561 -0.02066 N 3.034815 4.526604 0.549033 C 2.551536 -0.79411 0.990301 C 3.495375 -1.58472 1.647509 C 4.620533 -2.06414 0.967177 C 4.806871 -1.7482 -0.3805 C 3.871511 -0.94939 -1.04617 S -4.09701 -0.19696 0.242012 O -4.72343 -0.65911 -1.0345 O -4.54102 1.153104 0.692273

IM6

Atom X Y Z

C -1.50204 -0.81894 -1.14742 C -0.15023 -1.24767 -1.23506 C 0.547014 -1.01579 -0.06169 C 0.09841 -0.44948 1.137413 C -1.23614 -0.05199 1.155615 C -2.03078 -0.23299 0.008852 N 1.984572 -1.48541 -0.20168 C 3.06386 -0.49222 -0.10823 C 2.901054 0.73585 -0.75045 C 3.939274 1.666505 -0.69303 C 5.117746 1.376631 0.004287 C 5.262469 0.144694 0.646079 C 4.231424 -0.79872 0.587487 S -3.75493 0.386481 0.063766 O -4.23052 0.024968 1.434343 O -4.4607 -0.32467 -1.04025 O -3.61016 1.859324 -0.14988 H -2.17684 -0.94237 -1.99184 H 2.190499 -2.28153 0.412077 H 1.969799 0.949181 -1.26597 H 4.338525 -1.75973 1.087065 H 3.820081 2.626716 -1.18735 H 6.170964 -0.08503 1.196659 H 5.916583 2.112173 0.05254 H 0.735645 -0.31167 2.011951 H -1.67816 0.387096 2.044552 H 1.858557 -1.85296 -1.18506

SO3-

HN

N2

Page 137: Experimental and computational studies of the unimolecular

123

Azo-pro1/Azo-pro-2

Atom X Y Z

C 1.934839 -1.20547 -0.02328 C 2.716391 0.061187 -0.00982 C 2.005069 -1.15581 -0.02633 C 0.608099 -1.16421 -0.04039 C -0.12256 0.027429 -0.04643 C 0.560635 1.256581 -0.04249 H -0.00833 2.181896 -0.0623 N 4.133409 0.098227 -0.04387 H 4.558978 -0.6773 0.45445 H 4.502764 0.97683 0.305301 S -1.95327 -0.00367 0.01175 O -2.35718 1.337537 -0.50085 O -2.26655 -0.22644 1.453961 O -2.32336 -1.14752 -0.87193 H 2.558259 -2.09633 -0.03554 H 0.068532 -2.1064 -0.05976

Azo-pro1

Atom X Y Z

N -2.88199 -0.32815 -0.000016 N -2.01717 -0.48214 -0.000007 C -0.58992 -0.19137 0.000023 C -0.09765 1.119118 0.000014 C 0.268718 -1.29078 0.000007 C 1.28147 1.320279 -0.000006 H -0.79173 1.954338 0.000026 C 1.649442 -1.07831 -0.000004 H -0.15204 -2.29208 0.000018 C 2.15511 0.224649 -0.000007 H 1.678265 2.332144 -0.000001 H 2.327266 -1.92748 -0.000017 H 3.229312 0.389446 -0.000027

NH2

SO3-

NN

Page 138: Experimental and computational studies of the unimolecular

124

NH

Azo-pro2

Atom X Y Z

C -0.000012 -1.40007 0.000005 C -1.228195 -0.77313 -0.00013 C 1.228203 -0.77311 0.000124 C -1.215812 0.63329 0.00011 H -2.164547 -1.32522 -8.20E-05 C 1.215825 0.633255 -0.00012 H 2.164525 -1.32526 0.000085 C -0.000009 1.326232 0.000004 H -2.156816 1.179302 0.000137 H 2.156795 1.179331 -0.00013 H 0.000042 2.413073 0.000002

Tau-pro

Atom X Y Z

N -2.35783 -0.13736 -0.00037 C -1.02058 -0.02988 0.000541 C -0.30317 1.214242 -0.00044 C 1.081736 1.236063 0.000194 C 1.811138 0.033046 -2.00E-06 C 1.135356 -1.20072 -0.00031 C -0.2488 -1.23958 0.000307 H -0.7927 -2.17938 0.001005 H -0.86957 2.143507 0.0002 H 1.704161 -2.12694 -0.00053 H 1.609443 2.186585 0.001076 H 2.897506 0.058742 -0.00083 H -2.77814 0.799915 -0.00011

Page 139: Experimental and computational studies of the unimolecular

125

Tau-Pro

Atom X Y Z

C -0.49202 -1.0748 -0.0002 C -1.963 0.971687 0.000087 C -2.59081 -0.41762 -7.90E-05 C -1.65596 -1.54373 0.00036 C -0.31567 -1.36586 0.000343 C 0.277026 -0.03436 0.000019 N -2.68288 2.020334 0.00052 N -3.87781 -0.52063 -0.00063 S 2.098429 0.043367 -5.90E-05 O 2.428298 1.491417 -0.00037 O 2.454362 -0.68909 -1.24975 O 2.454415 -0.6886 1.249915 H -0.04188 2.061569 -0.00038 H -4.13502 -1.51486 -0.00058 H -2.08904 -2.54347 0.000523 H 0.361836 -2.21437 0.00057

Pro

Atom X Y Z

S 3.883004 -0.30459 -0.07761 C 2.115594 -0.14733 0.052195 C 1.249392 0.637403 0.822068 C -0.10139 0.309059 0.929539 C -0.61411 -0.80329 0.241617 C 0.255032 -1.58618 -0.52912 C 1.61369 -1.26554 -0.61333 H 1.646371 1.496034 1.356081 H -0.75944 0.898951 1.562521 H -0.13674 -2.44735 -1.06953 O 4.487431 -0.80846 -0.86529 O 3.864654 1.620423 -0.78023 O 4.34022 0.381258 1.340143 H 2.297724 -1.8698 -1.20147 N -1.97711 -1.18793 0.371817 C -3.10264 -0.40972 0.122614 H -2.13905 -2.18582 0.346957 C -4.38006 -0.98539 0.30471 C -5.53826 -0.25032 0.06446 C -5.4649 1.082751 -0.35651 C -4.20268 1.656315 -0.54018 C -3.03333 0.930347 -0.31034 H -4.4522 -2.01786 0.643787 H -6.50688 -0.72367 0.212857 H -6.36882 1.658094 -0.5385 H -4.11913 2.687237 -0.87788 H -2.06665 1.391719 -0.47903

SO3-

NHN

SO3-

HN

Page 140: Experimental and computational studies of the unimolecular

126

N2

Atom X Y Z

N 0 0 -0.00256 N 0 0 1.102563

Page 141: Experimental and computational studies of the unimolecular

127

(a) 4-amino-3(phenyldiazenyl)benzenesulfonate (b) TS1

(c) IM1 (d) TS2

Page 142: Experimental and computational studies of the unimolecular

128

(e) IM2-Cn1 (f) IM2-Cn2

(f) IM2-Cn3 (g) TS3

Page 143: Experimental and computational studies of the unimolecular

129

(h) IM3 (i) TS4

K

(j) IM6 (k) Pro

Page 144: Experimental and computational studies of the unimolecular

130

(l) Tau-Pro (m) Azo-Pro1

(n) Azo-Pro2 (Benzene Radical) (o) nitrogen

Fig-A_2.1:Structures of the stationary points on the 4-amino-3(phenyldiazenyl)benzenesulfonate potential energy surface optimized at B3LYP/6-311++G(d,p) level of theory. All lengths given in angstroms and angles in degrees.

Page 145: Experimental and computational studies of the unimolecular

131

Table-A_2.2: Optimized stationary points calculated for the loss of N2 from 4-hydroxy-3-(phenyldiazenyl)benzenesulfonate Anion (Electronic energy zero-point energy, relative energy and imaginary frequency calculated at B3LYP/6-31+G(d) level of theory)

Structure

Energy

Hartrees

Zero-point Energy

Hartrees

Relative energy

kJ mol-1

4-hydroxy-3

(phenyldiazenyl)

benzenesulfonate

-1271.31579

0.19772

0.0

Hyd-TS1(-1337cm-1

) -1271.30811

0.19335

8.7

Hyd-IM1 -1271.31496

0.19782

2.4

Hyd-TS2(-287cm-1

) -1271.23567

0.19562

204.9

Hyd-IM2-Cn1 -1271.28081

0.19677

89.3

Hyd-IM2-Cn2 -1271.27927

0.19638

92.4

Hyd-IM2-Cn3 -1271.27686

0.19658

99.2

Hyd-TS3 (-655cm-1

) -1271.16294 0.19015

381.3

Hyd-IM3 -1271.16655

0.19250

378.1

Hyd-IM5 -1271.36196

0.19333

-132.7

Hyd-TS4(-509cm-1

) -1271.18053

0.18987

334.5

Hyd-Pro -1271.35995

0.19297

-128.4

Hyd-IM-New -1271.26244

0.18976

119.2

Hyd-Pro2 -1271.23115

0.18763

195.7

Page 146: Experimental and computational studies of the unimolecular

132

Table-A_2.3: Optimized stationary points calculated for the loss of N2 from (4-(4-sulfonatophenylamino) benzene-1-sulfonic acid) (Electronic energy zero-point energy, relative energy and imaginary frequency calculated at B3LYP/6-31+G(d) level of theory)

Structure

Energy

Hartrees

Zero-point Energy

Hartrees

Relative energy

kJ mol-1

(4-(4-sulfonato

phenyl amino)

benzene-1-sulfonic

acid)

-1874.69247

0.21196

0.0

DI-IM1 -1874.67982

0.21233

34.2

DI-IM2Cn2 -1874.66206

0.21192

79.7

DI-IM2Cn3 -1874.66017

0.21151

83.6

DI-IM6

-1765.095745

0.2022

-

DI-Pro -1765.21296

0.20242

-

Page 147: Experimental and computational studies of the unimolecular

133

Table-A_2.4: The Cartesian coordinates for all the stationary points for the fragmentation of 4-amino-3(phenyldiazenyl)benzenesulfonte calculated at B3LYP/6-31+G(d) level as illustrated in Figure-2.17.

Structure

Geometry

4-hydroxy-

3(phenyldiazenyl)

Benzenesulfonate

Atom X Y Z

C 2.705722 1.559882 -0.00495 C 1.620101 2.430033 0.000222 C 2.524348 0.163992 -0.01389 C 0.312921 1.928757 -0.00139 C 1.241411 -0.34594 -0.01738 C 0.112411 0.512938 -0.01121 H 1.085107 -1.42007 -0.02854 S 3.987846 -0.93131 0.004487 O 3.422991 -2.29879 -0.16807 O 4.805988 -0.4421 -1.14239 O 4.601816 -0.67437 1.339026 N -1.12198 -0.11463 -0.01297 N -2.15916 0.628609 -0.00371 C -3.3985 -0.05622 -0.00176 C -4.55142 0.744413 0.005648 C -3.52777 -1.4575 -0.00517 C -4.79561 -2.03413 -0.00107 C -5.94642 -1.23372 0.006354 C -5.81814 0.15839 0.009586 H -2.63139 -2.06778 -0.01056 H -4.88981 -3.11777 -0.00335 H -6.93194 -1.69323 0.009822 H -6.70405 0.789136 0.015447 H -4.43597 1.82519 0.00848 H 1.758235 3.507942 0.005046 H 3.71795 1.954226 -0.008 O -0.7218 2.797103 0.004333 H -1.55197 2.245355 0.003473

OH

SO3-

NN

Page 148: Experimental and computational studies of the unimolecular

134

Hyd-TS1

Atom X Y Z

C -2.63478 1.569959 0.000518 C -1.51927 2.385275 0.000719 C -2.53416 0.150744 -0.00042 C -0.22167 1.816546 0.000147 C -1.29623 -0.43818 -0.00102 C -0.11765 0.366658 -0.00067 H -1.2033 -1.51995 -0.00181 S -4.06971 -0.83489 0.000041 O -3.60269 -2.24877 -0.00615 O -4.75649 -0.40718 1.253338 O -4.76289 -0.39825 -1.24664 N 1.082917 -0.26722 -0.0009 N 2.125129 0.505772 -0.00063 C 3.401691 -0.07494 -0.00031 C 4.504774 0.795911 -0.0007 C 3.612923 -1.46569 0.00051 C 4.913704 -1.96466 0.000952 C 6.015117 -1.09881 0.000573 C 5.801813 0.28357 -0.00026 H 2.75517 -2.129 0.000853 H 5.070276 -3.04097 0.001637 H 7.026346 -1.4979 0.00095 H 6.647693 0.966988 -0.00057 H 4.331831 1.869107 -0.00134 H -1.60345 3.46861 0.001319 H -3.62834 2.01 0.00087 O 0.865922 2.543169 0.000464 H 1.754394 1.691361 -0.00018

SO3

ONN

H

Page 149: Experimental and computational studies of the unimolecular

135

Hyd-IM1

Atom X Y Z

C -1.37417 -0.89325 -0.18345 C -0.00174 0.649238 -0.4561 C 0.471214 -0.68793 -0.58989 C -0.43178 -1.7498 -0.45922 C -1.77492 -1.48394 -0.21397 C -2.25297 -0.16231 -0.07007 N 0.841359 1.717313 -0.59611 N 2.06368 1.403764 -0.8509 C 2.792939 -0.00042 -0.31174 C 4.067365 -0.17168 -0.9816 C 5.256025 -0.22594 -0.28151 C 5.298214 -0.1328 1.124975 C 4.084365 -0.00372 1.803673 C 2.866244 0.040163 1.128938 S -4.02987 0.110473 0.267331 O -4.15422 1.589658 0.381222 O -4.25987 -0.63836 1.535511 O -4.70326 -0.48616 -0.92107 H -1.72637 1.913758 -0.06891 H 1.938994 0.118129 1.690576 H 4.052647 -0.26316 -2.06489 H 4.072763 0.047553 2.891129 H 6.181624 -0.35364 -0.84092 H 6.241953 -0.17625 1.660626 H -0.06119 -2.76605 -0.55964 H -2.48195 -2.30346 -0.12149 H 2.694041 2.211064 -0.83569 O 1.770526 -0.9436 -0.86557

SO3-

ONN

H

Page 150: Experimental and computational studies of the unimolecular

136

Hyd-IM2Cn1

Atom X Y Z

C -1.504 0.986694 0.438193 C -0.16242 1.311497 0.173893 C 0.592235 0.449273 -0.64764 C 0.003386 -0.68369 -1.20787 C -1.34126 -0.9665 -0.96717 C -2.10085 -0.12753 -0.14474 C 2.924766 -0.02803 -0.33669 N 1.208465 3.198251 0.312002 N 0.339755 2.472412 0.843462 C 2.711026 -0.86195 0.765587 C 3.792266 -1.56295 1.308186 C 5.074456 -1.43831 0.767378 C 5.273189 -0.59719 -0.33357 C 4.204757 0.106457 -0.88842 S -3.87493 -0.49325 0.133188 O -4.53488 0.05379 -1.08593 O -4.19982 0.23019 1.392569 O -3.91156 -1.97974 0.233728 H -2.07981 1.618093 1.107781 H 1.719029 -0.96093 1.192951 H 4.340988 0.762532 -1.74335 H 3.62256 -2.20915 2.166043 H 6.265178 -0.48868 -0.76611 H 5.908083 -1.98739 1.197379 H 0.612096 -1.33862 -1.8264 H -1.80724 -1.84847 -1.39593 H 1.476911 2.848205 -0.63071 O 1.933544 0.730142 -0.92583

so3

NN

HO

Page 151: Experimental and computational studies of the unimolecular

137

Hyd-IM2-Cn2

Atom X Y Z

C -1.48561 1.043111 0.181789 C -0.1149 1.210167 -0.08576 C 0.563127 0.194891 -0.80239 C -0.13454 -0.9317 -1.23128 C -1.49011 -1.09063 -0.92881 C -2.16975 -0.10423 -0.21104 C 2.90486 -0.14777 -0.43776 N 1.666532 2.54918 0.424324 N 0.416421 2.444564 0.370939 C 2.716664 -0.80062 0.784835 C 3.830192 -1.27241 1.486762 C 5.122569 -1.10295 0.98415 C 5.296173 -0.44696 -0.24088 C 4.196038 0.029328 -0.95184 S -3.92859 -0.34148 0.241689 O -4.4762 1.041741 0.301386 O -3.85334 -1.03288 1.56036 O -4.46724 -1.18469 -0.86267 H -2.01393 1.836853 0.700686 H 1.716788 -0.93485 1.183674 H 4.313239 0.545316 -1.90038 H 3.677021 -1.77641 2.438347 H 6.295518 -0.30412 -0.64652 H 5.981829 -1.47399 1.537109 H 0.401471 -1.68039 -1.80931 H -2.03651 -1.96666 -1.26525 H 1.848888 3.507982 0.765531 O 1.882281 0.33485 -1.21463

so3

NO

NH

Page 152: Experimental and computational studies of the unimolecular

138

Hyd-IM2-Cn3

Atom X Y Z

C -1.35412 -0.94011 -0.15633 C 0.006722 0.93115 -0.50957 C 0.599332 -0.28156 -0.90276 C -0.14534 -1.46062 -0.91192 C -1.49052 -1.43864 -0.54236 C -2.09761 -0.235 -0.16436 C 2.9478 -0.34449 -0.43228 N 0.557188 3.079502 0.181792 N 0.8148 2.110161 -0.5619 C 2.771276 -0.51861 0.944757 C 3.893859 -0.55724 1.778083 C 5.182392 -0.42872 1.254909 C 5.34396 -0.25767 -0.12558 C 4.235495 -0.21479 -0.96906 S -3.86477 -0.21409 0.320757 O -4.15252 1.229741 0.550138 O -3.89036 -1.05762 1.549046 O -4.5639 -0.81496 -0.84867 H -1.87354 1.863768 0.086423 H 1.774367 -0.62192 1.36023 H 4.341782 -0.07436 -2.04067 H 3.750218 -0.68909 2.848006 H 6.340173 -0.15133 -0.54924 H 6.048373 -0.45783 1.911188 H 0.342296 -2.3837 -1.21381 H -2.08121 -2.34973 -0.55673 H -0.24859 2.873998 0.818937 O 1.916655 -0.32161 -1.34012

so3

NNO

H

Page 153: Experimental and computational studies of the unimolecular

139

Hyd-IM3

Atom X Y Z

C -1.73204 -1.13467 -0.21334 C -0.36708 1.500774 0.270544 C 0.65633 0.446223 -0.03268 C 0.26541 -0.83895 0.272406 C -1.11947 -1.09726 0.436244 C -2.1158 -0.14522 0.08214 C 3.011078 0.08243 -0.21909 N 0.278882 3.657262 0.61553 N 0.053701 2.846533 -0.21956 C 3.352498 -0.36543 1.060093 C 4.542745 -1.07374 1.236221 C 5.387259 -1.328 0.14957 C 5.035352 -0.86562 -1.12173 C 3.844808 -0.15893 -1.31174 S -3.84584 -0.71769 -0.10183 O -4.51415 0.385371 -0.84394 O -4.29853 -0.89023 1.307967 O -3.71557 -1.99685 -0.85732 H -2.39999 1.873386 -0.64274 H 2.695538 -0.15604 1.898658 H 3.551237 0.210408 -2.28962 H 4.813166 -1.42199 2.230071 H 5.685447 -1.05656 -1.97203 H 6.312286 -1.88005 0.294537 H 0.969371 -1.66675 0.273069 H -1.43986 -2.1058 0.679405 H -0.42235 1.526239 1.412327 O 1.887494 0.862708 -0.44147

SO3

NO

NH

Page 154: Experimental and computational studies of the unimolecular

140

Hyd-Pro

Atom X Y Z

S -3.76879 -0.33994 -0.01268 C -2.02758 -0.22384 0.022057 C -1.38924 -0.60602 -1.16087 C -0.03977 -0.97266 -1.14835 C 0.660186 -0.95254 0.056788 C 0.032625 -0.59066 1.248819 C -1.31564 -0.22374 1.225462 H -1.96221 -0.62503 -2.08306 H 0.470419 -1.27808 -2.05856 H 0.596523 -0.60376 2.178342 O -4.3011 -0.04076 1.326827 O -3.65478 1.811874 -0.22867 O -4.36887 -0.40116 -1.15856 H -1.83271 0.050137 2.140255 C 3.004815 -0.45226 0.020727 C 4.309262 -0.9598 0.119897 C 5.398941 -0.09261 0.063142 C 5.2046 1.28558 -0.09182 C 3.901862 1.780679 -0.19063 C 2.797491 0.924289 -0.1364 H 4.44283 -2.03126 0.239921 H 6.406191 -0.49627 0.141437 H 6.055511 1.960694 -0.13458 H 3.732648 2.848212 -0.31193 H 1.789758 1.317469 -0.21426 O 1.996781 -1.37989 0.076944

SO3-

O

Page 155: Experimental and computational studies of the unimolecular

141

Hyd-TS3

Atom X Y Z

C -1.726605 -1.1425 -0.201079 C -0.322947 1.480859 0.021144 C 0.654355 0.434858 -0.293828 C 0.255288 -0.88197 -0.125888 C -1.109013 -1.17471 0.038159 C -2.100534 -0.17233 -0.039408 C 3.020761 0.087467 -0.260778 N 0.308415 3.307268 0.909642 N 0.082673 2.867915 -0.204693 C 3.208685 -0.38582 1.041128 C 4.382846 -1.07508 1.349807 C 5.363997 -1.28396 0.37402 C 5.165202 -0.79494 -0.920302 C 3.992027 -0.10806 -1.243761 S -3.865822 -0.66648 -0.010989 O -4.583854 0.535999 -0.512765 O -4.104559 -0.98547 1.424705 O -3.910313 -1.8511 -0.914213 H -2.461589 1.924577 -0.361049 H 2.448065 -0.20964 1.795482 H 3.816952 0.281473 -2.242075 H 4.533059 -1.44263 2.361911 H 5.922299 -0.94977 -1.68518 H 6.276062 -1.82042 0.622556 H 0.977328 -1.68955 -0.208421 H -1.42634 -2.20848 0.134323 H -0.209086 1.566262 1.272603 O 1.91553 0.847134 -0.608259

SO3

NO

NH

Page 156: Experimental and computational studies of the unimolecular

142

Hyd-IM5

Atom X Y Z

C -0.90928 0.731762 0.061358 C 0.398906 1.110223 -0.25439 C 1.236045 0.196679 -0.89367 C 0.784879 -1.07699 -1.23709 C -0.52512 -1.44588 -0.91727 C -1.3716 -0.54698 -0.26226 C 3.572713 0.406477 -0.39884 N -5.42848 2.947255 0.354312 N -6.30709 2.303277 0.169751 C 3.445264 -0.17711 0.868289 C 4.579953 -0.31823 1.673739 C 5.834466 0.112578 1.235141 C 5.948949 0.694277 -0.03363 C 4.828265 0.842325 -0.84892 S -3.05783 -1.06482 0.224708 O -3.85091 0.197618 0.186396 O -2.87141 -1.62531 1.594474 O -3.44099 -2.07322 -0.80353 H -1.58684 1.427227 0.547457 H 2.475426 -0.51423 1.217463 H 4.900281 1.290998 -1.83571 H 4.47267 -0.77171 2.656513 H 6.91769 1.035757 -0.39214 H 6.70941 -0.00188 1.869958 H 1.45362 -1.76157 -1.75295 H -0.90749 -2.4266 -1.18389 H 0.771083 2.103603 -0.01627 O 2.529262 0.591789 -1.26845

SO3-

O

N2

Page 157: Experimental and computational studies of the unimolecular

143

Hyd-TS4

Atom X Y Z

C -1.49319 0.9831 0.485291 C -0.13221 1.252553 0.251238 C 0.44063 0.440017 -0.67902 C -0.12547 -0.59843 -1.40437 C -1.46956 -0.85205 -1.12012 C -2.14742 -0.06865 -0.17179 C 2.892492 -0.06993 -0.30445 N 1.907884 3.109701 -0.20508 N 0.948577 3.164001 0.423794 C 2.854007 -0.17327 1.083229 C 3.828251 -0.95197 1.712698 C 4.81322 -1.60279 0.962 C 4.827439 -1.47913 -0.43083 C 3.858343 -0.70645 -1.07708 S -3.88602 -0.48882 0.23214 O -4.43346 -1.03275 -1.04597 O -4.49209 0.800467 0.668678 O -3.75809 -1.5059 1.317614 H -2.07068 1.587763 1.181942 H 2.066285 0.327459 1.638052 H 3.841196 -0.59672 -2.15702 H 3.8117 -1.0511 2.794705 H 5.587982 -1.9869 -1.01826 H 5.565802 -2.20744 1.461395 H 0.431226 -1.17279 -2.14218 H -2.01068 -1.63744 -1.63825 H 2.064182 1.769989 -0.76269 O 1.935492 0.717475 -0.99579

SO3

O

NN

H

Page 158: Experimental and computational studies of the unimolecular

144

Hyd-IM-New

Atom X Y Z

C -3.56215 -0.91717 0.023938 C 3.992283 -2.08061 -0.58629 C 3.346984 -3.07081 -1.0076 C 1.95718 -3.16803 -0.92602 C 1.380684 -2.03285 -0.31877 C 2.154794 -0.94856 0.137538 C -2.87379 0.2434 -0.68922 N 2.393509 3.834396 -1.17836 N 1.712422 4.410718 -0.52679 C -4.12358 0.107325 -1.32054 C -5.21793 -0.3947 -0.61568 C -5.09012 -0.76971 0.727356 C -3.84441 -0.63244 1.350765 C -2.73774 -0.13187 0.66041 S 1.313318 0.488308 0.904295 O 0.542732 1.103578 -0.2321 O 2.410806 1.352189 1.400505 O 0.430678 -0.09907 1.946956 H 4.14111 -0.07685 0.385867 H -4.21194 0.402577 -2.36303 H -1.77509 -0.03526 1.158158 H -6.1779 -0.49272 -1.12024 H -3.72314 -0.91898 2.393695 H -5.94393 -1.16066 1.276259 H 1.350964 -3.99978 -1.27294 H 0.30247 -1.99365 -0.18944 H -1.01034 0.82115 -0.90559 O -1.85245 0.730994 -1.42921

SO3OH

N2

Page 159: Experimental and computational studies of the unimolecular

145

Hyd-pro2 (product-1)

Atom X Y Z

C 0.972065 -1.36174 -0.000091 C 2.346222 1.217444 0.00004 C 3.086356 0.201938 0.000043 C 2.554906 -1.092422 0.00002 C 1.144284 -1.101949 -0.000094 C 0.378856 0.080231 -0.000151 S -1.462762 -0.038818 0.000012 O -1.767627 -0.789978 -1.250182 O -1.91632 1.380347 -0.001523 O -1.767527 -0.787207 1.251895 H 0.370496 2.261981 -0.000174 H 3.130419 -2.014799 0.00006 H 0.618935 -2.05329 -0.000197

SO3

Page 160: Experimental and computational studies of the unimolecular

146

Hyd-pro2 (product-2)

Atom X Y Z

C -0.939798 -0.023554 0.000232 C -0.263735 1.201313 0.000117 C 1.134209 1.220458 0.000055 C 1.859905 0.027024 -0.000147 C 1.172265 -1.192406 -0.000049 C -0.222087 -1.224634 0.000194 H -0.8247 2.134767 -0.000044 H -0.76704 -2.163978 0.000371 H 1.652712 2.175981 -0.000073 H 1.725387 -2.128273 -0.00016 H 2.946061 0.045231 -0.000304 H -2.697608 0.777793 0.000421 O -2.30992 -0.111342 -0.000327

OH

Page 161: Experimental and computational studies of the unimolecular

147

Table-A_2.5: The Cartesian coordinates for all the stationary points for the fragmentation of (4-(4-sulfonatophenylamino) benzene-1-sulfonic acid) calculated at B3LYP/6-31+G(d) level as illustrated in Figure-2.17.

Structure Geometry

(4-(4-sulfonato phenyl amino) benzene-1-

sulfonic acid)

Atom X Y Z

C 4.669307 1.321165 0.063962 C 3.737058 2.348685 0.062299 C 4.257727 -0.01986 -0.00464 C 2.351925 2.09081 -0.00461 C 2.903104 -0.29916 -0.06137 C 1.924618 0.726609 -0.05624 H 2.566496 -1.33012 -0.10463 S 5.525384 -1.33306 0.011835 O 4.764296 -2.60213 -0.16629 O 6.425038 -0.98413 -1.1303 O 6.183174 -1.18886 1.346642 N 1.459083 3.144802 -0.06013 H 1.776516 4.030039 0.312438 H 0.483906 2.891166 0.081424 N 0.601821 0.266488 -0.07262 N -0.3297 1.130684 -0.04972 C -1.65251 0.621522 -0.04086 C -2.67373 1.584582 -0.02601 C -1.99942 -0.74467 -0.04606 C -3.33963 -1.1191 -0.03481 C -4.35651 -0.15046 -0.01761 C -4.01664 1.203576 -0.01421 H -1.20978 -1.48893 -0.06008 H -3.61418 -2.17018 -0.04476 H -4.81376 1.940496 -0.00859 H -2.39513 2.636479 -0.02763 H 4.06874 3.386719 0.099985 H 5.730974 1.54674 0.111888 S -6.11086 -0.66351 0.028284 O -6.27673 -1.27792 1.379146 O -6.2412 -1.63549 -1.09684 O -6.87664 0.604355 -0.16282

NH2

SO3

NN

SO3

Page 162: Experimental and computational studies of the unimolecular

148

DI-IM1

Atom X Y Z

C 4.646263 1.36658 0.004679 C 3.703247 2.35382 0.005436 C 4.279206 -0.02005 0.001567 C 2.284227 2.063159 0.003242 C 2.959679 -0.36399 -0.00078 C 1.918533 0.633446 -0.00021 H 2.666023 -1.40906 -0.00326 S 5.619268 -1.25118 0.001014 O 4.935857 -2.57471 -0.00323 O 6.386316 -0.9437 -1.24543 O 6.382048 -0.94951 1.251468 N 1.339695 2.988467 0.003903 H 1.754864 3.923957 0.006174 H -0.10293 2.012512 -0.00053 N 0.666258 0.158448 -0.00357 N -0.3375 0.998098 -0.00335 C -1.67539 0.569451 -0.00731 C -2.67205 1.558637 -0.00955 C -2.04141 -0.78693 -0.01206 C -3.39149 -1.133 -0.01811 C -4.38953 -0.14906 -0.01876 C -4.01978 1.197703 -0.01459 H -1.26724 -1.54676 -0.014 H -3.68226 -2.17951 -0.03052 H -4.80037 1.952074 -0.02393 H -2.3846 2.608683 -0.00967 H 3.997328 3.403678 0.007789 H 5.703948 1.616056 0.006384 S -6.15358 -0.62539 0.006727 O -6.36854 -1.17584 1.378305 O -6.27795 -1.64413 -1.07715 O -6.88574 0.64912 -0.25928

SO3

NHN

HN

SO3

Page 163: Experimental and computational studies of the unimolecular

149

DI-IM2Cn3

Atom X Y Z

C -3.00643 -0.68825 0.56789 C -1.68901 -0.26363 0.859575 C -1.22066 0.976866 0.331933 C -2.14982 1.752318 -0.39811 C -3.45572 1.338517 -0.62352 C -3.89588 0.095544 -0.14897 N 0.046232 1.537661 0.527231 C 1.32743 0.981197 0.315984 N 0.046507 -0.72678 2.269323 N -1.02508 -1.13849 1.751567 C 1.551677 -0.28065 -0.25013 C 2.859492 -0.73115 -0.46828 C 3.957839 0.061038 -0.13315 C 3.735698 1.321401 0.433803 C 2.438972 1.772966 0.660843 S -5.61422 -0.43431 -0.46984 O -5.75546 -1.72601 0.261752 O -5.69338 -0.55406 -1.95635 O -6.4412 0.68165 0.082209 H -3.33282 -1.64306 0.967254 H 0.042165 2.537687 0.368686 H 0.709716 -0.90401 -0.53517 H 2.273521 2.744578 1.12639 H 3.034904 -1.71202 -0.90017 H 4.588782 1.933839 0.711423 H -1.81025 2.701588 -0.81274 H -4.13915 1.967949 -1.18652 H 0.365912 -1.48565 2.893384 S 5.663916 -0.48871 -0.4798 O 5.56746 -1.974 -0.59921 O 6.028477 0.199578 -1.7563 O 6.453716 -0.01421 0.697604

so3

NNHN

SO3

H

Page 164: Experimental and computational studies of the unimolecular

150

DI-IM2Cn2

Atom X Y Z

C -2.96327 -0.66037 -0.52904 C -1.65585 0.265274 -0.87207 C -1.17565 -1.00729 -0.48599 C -2.03834 -1.83652 0.257006 C -3.33451 -1.44287 0.575758 C -3.80598 -0.18469 0.182355 N 0.084702 -1.49742 -0.8679 C 1.360323 -0.95333 -0.56066 N -0.92049 2.296431 -1.73002 N -0.8436 1.0501 -1.74446 C 1.559479 0.164997 0.258665 C 2.858258 0.593815 0.559777 C 3.972948 -0.0831 0.06409 C 3.775121 -1.19338 -0.76351 C 2.486919 -1.61735 -1.07993 S -5.53717 0.290531 0.530796 O -5.60675 1.73227 0.153212 O -5.70184 0.020812 1.989316 O -6.34599 -0.61226 -0.34142 H -3.36754 1.608329 -0.87353 H 0.106315 -2.50987 -0.87618 H 0.707812 0.694286 0.675483 H 2.340768 -2.4701 -1.7427 H 3.013875 1.462375 1.192797 H 4.63894 -1.71211 -1.16957 H -1.66814 -2.80504 0.590797 H -3.98822 -2.10334 1.138141 H -1.568 2.630987 -0.97592 S 5.668222 0.401875 0.538124 O 5.537835 1.801013 1.042192 O 6.461748 0.26952 -0.72161 O 6.055777 -0.58803 1.590275

SO3

NHN

NH

SO3

Page 165: Experimental and computational studies of the unimolecular

151

DI-IM6

Atom X Y Z

C 3.469271 -0.9552 -1018229 C 2.178332 1.543745 -1.096197 C 1.345813 1.11286 -0.07688 C 1.622305 0.217073 0.961449 C 2.909698 -0.31926 0.98335 C 3.835175 0.061293 -0.003517 N 0.014759 1.858438 -0.103928 C -1.28511 1.165657 -0.035815 C -1.53889 0.069977 -0.859321 C -2.80887 -0.5139 -0.833275 C -3.80728 -0.01354 0.009548 C -3.52805 1.067283 0.84863 C -2.26443 1.661244 0.825046 S 5.512755 -0.6771 0.05403 O 5.820928 -0.79457 1.515124 O 6.39244 0.286725 -0.673879 O 5.371446 -2.00021 -0.630429 H 4.232537 1.209047 -1.751501 H 0.026669 2.589339 0.614969 H -0.75212 -0.31819 -1.499815 H -2.04591 2.507802 1.475672 H -3.03355 -1.37213 -1.458958 H -4.30107 1.428883 1.519623 H 0.885349 -0.06667 1.713279 H 3.213688 -1.01246 1.761432 H 0.17032 2.348468 -1.014966 S -5.50185 -0.70937 -0.02915 O -5.98905 -0.55473 1.372207 O -6.21061 0.15798 -1.01575 O -5.31534 -2.11837 -0.474654

SO3

H2N

SO3

Page 166: Experimental and computational studies of the unimolecular

152

DI-Pro

Atom X Y Z

S -5.66495 -0.59167 -0.02587 C 3.942605 0.011465 -0.01467 C 2.885022 -0.79225 -0.43999 C 1.566174 -0.32066 -0.43468 C 1.280227 0.978931 0.015942 C 2.355479 1.78215 0.453065 C 3.663831 1.308947 0.429318 H 3.106324 -1.79381 -0.79691 H 0.769082 -0.95372 -0.81025 H 2.150382 2.790504 0.813731 O 6.410517 0.431299 -0.82212 O 6.064084 -0.62719 1.41437 O 5.589961 -1.93566 -0.67244 H 4.479897 1.945745 0.75967 N -6E-06 1.553657 0.000032 C -1.28022 0.978884 -0.01597 H -2.7E-05 2.564002 0.000111 C -2.35549 1.782121 -0.45304 C -3.66382 1.308865 -0.4294 C -3.94256 0.011307 0.014398 C -2.88497 -0.79242 0.439645 C -1.56613 -0.32079 0.434436 H -2.15041 2.79051 -0.81361 H -4.47986 1.94565 -0.75984 H -3.10625 -1.79409 0.796291 H -0.76902 -0.95389 0.809902 S -5.66497 -0.59164 0.025966 O -6.06583 -0.62289 -1.41389 O -6.40946 0.429079 0.826098 O -5.58937 -1.93751 0.66857

SO3

HN

SO3

Page 167: Experimental and computational studies of the unimolecular

153

Appendix-3 Table-A_3.1: The Cartesian coordinates for all the stationary points for the fragmentation of phenoxide anion calculated at B3LYP/6-31+G(d) level as illustrated in Figure-3.3. Structure

Geometry

Phenoxide

Atom X Y Z

C 0.287528 -1.21168 -5E-06 C -1.10006 -1.20007 0.000007 C -1.82771 0.000001 -1E-06 C -1.10016 1.199922 0.000007 C 0.287595 1.211748 -4E-06 C 1.077275 0.00015 -6.1E-05 O 2.347194 -8.2E-05 0.000031 H -1.63712 2.148344 0.000025 H -2.91315 -7.5E-05 0.000007 H -1.63712 -2.14843 0.000026 H 0.83173 -2.15297 0.000015 H 0.831296 2.153325 0.000015

TS1

Atom X Y Z

O 2.214017 0.577364 0.099588 C 1.314469 -0.17405 -0.0044 C 0.631164 -1.31688 -0.1086 C -0.82099 -1.45256 -0.09815 C -1.62293 -0.39783 0.121901 C -1.01337 0.930792 0.394179 C -0.8321 1.867998 -0.57647 H 1.264937 -2.19225 -0.23142 H -1.21777 -2.44336 -0.30874 H -2.70417 -0.51577 0.062882 H -0.7199 1.060732 1.45749 H -0.2727 2.726898 -0.14766

IM1

Atom X Y Z

C -2.69504 -1.06365 0.000023 C -1.45592 -0.49671 0.000005 C -1.21718 0.927171 -1.3E-05 C -0.012 1.55016 -8E-06 C 1.309946 0.915773 -2.6E-05 C 1.666603 -0.35721 0.000014 O 2.181175 -1.41725 -2E-06 H 2.184674 1.567446 0.000122 H 0.024927 2.634988 0.00002 H -2.11216 1.546131 0.000015 H -0.51847 -1.09357 -4.6E-05 H -2.60688 -2.17021 -6.2E-05

Page 168: Experimental and computational studies of the unimolecular

154

TS2

Atom X Y Z

O -2.57008 -1.12974 -0.38629 C -1.92274 -0.26739 0.075935 C -1.20108 0.696538 0.591052 C -0.07246 1.306721 -0.17862 C 1.164687 0.764255 -0.32339 C 1.712415 -0.47237 0.186594 C 3.01894 -0.81446 -0.04974 H -1.46898 1.015444 1.598346 H -0.28552 2.270829 -0.63639 H 1.886744 1.342704 -0.90129 H 1.003208 -1.08388 0.767384 H 3.226599 -1.78691 0.451271

IM2

Atom X Y Z

C -3.49348 -0.33691 0.000058 C -2.1274 -0.41221 -2.9E-05 C -1.29131 0.769857 0.000018 C 0.064155 0.876105 -2.9E-05 C 0.978115 -0.27215 -0.00012 C 2.289585 -0.242 -8.5E-05 O 3.468378 -0.23052 0.000118 H 0.550586 -1.27266 0.000051 H 0.524536 1.859204 0.000015 H -1.85741 1.699945 0.000064 H -1.57189 -1.37048 -5.4E-05 H -3.91081 -1.3681 0.0001

TS3

Atom X Y Z

O -3.29409 -0.37229 -0.0029 C -2.11998 -0.28205 0.01177 C -0.81472 -0.19839 0.029164 C -0.0108 1.019433 -0.12656 C 1.32732 1.005526 0.008821 C 2.152774 -0.17547 0.395579 C 2.763436 -0.98587 -0.50134 H -0.25556 -1.12905 0.136625 H -0.5351 1.933887 -0.39336 H 1.840423 1.952427 -0.17517 H 2.2137 -0.28983 1.499858 H 3.301041 -1.78813 0.050618

Page 169: Experimental and computational studies of the unimolecular

155

IM3

Atom X Y Z

C 1.682158 1.543474 0.000042 C 2.224029 0.30226 -8.7E-05 C 1.569396 -0.98344 0.000039 C 0.217389 -1.26067 0.000149 C -0.86755 -0.39461 -0.00014 C -2.04121 0.040756 -4.9E-05 O -3.1468 0.541927 0.000027 H 0.609745 1.691002 0.000315 H -0.01388 -2.33425 -5.4E-05 H 2.234199 -1.84509 -3.1E-05 H 3.31529 0.245989 -0.00019 H 2.323735 2.420319 0.000012

TS4

Atom X Y Z

C -1.37808 1.561554 0.405081 C -1.97378 0.471265 -0.16333 C -1.3824 -0.85468 -0.34203 C -0.18582 -1.40004 -0.01168 C 0.950712 -0.72218 0.670809 C 1.730162 0.121051 0.043866 O 2.444327 0.859081 -0.51833 H 1.097811 -0.79029 1.747142 H -0.03 -2.45247 -0.24907 H -2.05658 -1.56765 -0.82864 H -3.01292 0.483864 -0.55287 H -2.11767 2.392045 0.413829

TS5

Atom X Y Z

C -1.07301 1.421784 0.024071 C -2.07046 0.527979 -0.12553 C -1.68646 -0.87995 -0.16209 C -0.38757 -1.18579 0.0646 C 0.555917 -0.11103 0.398061 C 1.944314 -0.25552 0.131265 O 2.842175 0.391192 -0.34904 H 0.448249 0.360945 1.375955 H -0.0174 -2.19807 -0.04533 H -2.41235 -1.65773 -0.39336 H -3.13807 0.771207 -0.24628 H -1.31418 2.489264 0.119104

Page 170: Experimental and computational studies of the unimolecular

156

P1

Atom X Y Z

C 1.594189 0.046369 -0.87705 C 1.138544 1.157282 -0.13251 C 0.444393 0.67032 1.000234 C 0.472964 -0.7454 0.953619 C 1.184767 -1.12829 -0.20763 C -1.84869 -0.0006 -0.69031 O -2.84123 0.000191 -0.12506 H -0.02261 1.273995 1.771338 H 0.031855 -1.41682 1.682842 H 1.372625 -2.14639 -0.53238 H 2.146477 0.088274 -1.81009 H 1.284529 2.201325 -0.38929

TS6

Atom X Y Z

C 1.275008 -1.14587 0.459232 C 1.659987 0.054496 -0.05941 C 0.908248 1.279538 -0.14001 C -0.44123 1.494146 0.039336 C -1.52074 0.573955 0.277015 C -1.46285 -0.72722 -0.17634 O -0.66643 -1.62714 -0.26118 H -2.5081 0.991638 0.443968 H -0.76507 2.529755 -0.09197 H 1.484881 2.15854 -0.43168 H 2.690282 0.1655 -0.44293 H 1.918866 -2.00265 0.213083

TS7

Atom X Y Z

O -1.34593 -0.40941 0.90176 C -0.91404 -0.8376 -0.2778 C 0.463355 -1.34641 -0.38407 C 1.479862 -0.54447 0.131645 C 1.262196 0.848969 0.233358 C 0.040263 1.434174 -0.15562 C -1.08033 0.657454 -0.51508 H 0.685985 -2.18627 -1.03856 H 2.492244 -0.92607 0.298342 H 2.079183 1.494436 0.547165 H -0.01994 2.514233 -0.28423 H -1.97783 1.106191 -0.93144

Page 171: Experimental and computational studies of the unimolecular

157

IM4

Atom X Y Z

C 1.232104 -0.9242 -0.23717 C 0.091179 -1.68274 -0.25795 O -0.99981 -0.97617 0.553123 C -1.47536 0.090693 -0.1192 C -0.82056 1.27204 -0.23887 C 0.556873 1.460103 0.145994 C 1.470474 0.445931 0.164693 H 2.105058 -1.44158 -0.64321 H 2.507445 0.721324 0.369702 H 0.900065 2.476803 0.334056 H -1.35835 2.101612 -0.69635 H -2.484 -0.0197 -0.53417

IM6

Atom X Y Z

C -1.44107 -0.701 0.190013 C -0.40657 -1.54905 -0.2913 C 0.873114 -1.27161 -0.25096 C 1.034256 1.069165 -0.0516 C -0.2613 1.454317 -0.23589 C -1.42947 0.677043 0.078156 O 1.420466 -0.07222 0.563798 H 1.716322 -1.70515 -0.7754 H 1.849112 1.701954 -0.41938 H -0.40079 2.44801 -0.66267 H -2.3725 1.218654 0.176156 H -2.36965 -1.15896 0.540346

TS10

Atom X Y Z

O 1.374762 -0.03244 0.770744 C -0.29343 -1.53808 -0.31896 C 0.951511 -0.87806 -0.33049 C 0.954948 0.857322 -0.21952 C -0.35562 1.407551 -0.19215 C -1.44011 0.619297 0.147482 C -1.37458 -0.81138 0.15483 H 1.825015 -1.26234 -0.85638 H 1.770034 1.400022 -0.70136 H -0.49608 2.415574 -0.57778 H -2.42627 1.08566 0.190796 H -2.32713 -1.31936 0.331628

Page 172: Experimental and computational studies of the unimolecular

158

IM7

Atom X Y Z

C 0.945491 0.76993 -0.28419 C 0.944067 -0.79757 -0.34782 C -0.3067 -1.54705 -0.28178 C -1.3897 -0.81303 0.141334 C -1.44183 0.633812 0.141216 C -0.35817 1.402086 -0.17686 O 1.415109 -0.02926 0.799125 H -0.45636 2.456217 -0.42361 H -2.41974 1.113651 0.222896 H -2.35234 -1.30961 0.298184 H 1.799334 -1.23679 -0.86489 H 1.749273 1.321539 -0.77702

TS11

Atom X Y Z

C -0.84727 1.252495 -0.04578 C 0.432155 1.058663 -0.49635 C 1.228487 -0.34236 -0.32162 C 0.06271 -1.25927 -0.54259 C -1.13738 -1.1304 0.113437 C -1.67139 0.160562 0.297683 O 1.778053 -0.05416 0.848801 H -1.20635 2.269962 0.096895 H -2.7291 0.300333 0.511471 H -1.80419 -1.9888 0.203662 H 1.96017 -0.45927 -1.14629 H 1.151165 1.872829 -0.48483

IM8

Atom X Y Z

C -1.74498 -0.88349 0.000039 C -1.97245 0.532717 0.000051 C -0.73512 1.16628 -0.00007 C 1.683353 0.435869 0.000035 C 0.28803 0.158203 -6.2E-05 C -0.37338 -1.11189 -5.6E-05 O 2.637901 -0.35708 0.000041 H -2.94234 1.018002 0.000141 H -2.51825 -1.64473 0.000086 H 0.128329 -2.07056 -0.00016 H 1.911776 1.531905 0.000041 H -0.55547 2.235825 -6.8E-05

Page 173: Experimental and computational studies of the unimolecular

159

TS13

Atom X Y Z

C -0.33575 1.119144 0.098748 C -1.68201 0.920387 -0.02861 C -1.95051 -0.50803 -0.12209 C -0.76442 -1.17892 -0.02582 C 0.31736 -0.19426 0.12508 C 1.759875 -0.56169 0.049653 O 2.601652 0.32452 -0.20345 H -2.93519 -0.95155 -0.22245 H -2.44831 1.691389 -0.03104 H 0.194573 2.05726 0.181396 H 0.9051 -0.72512 1.157159 H -0.59663 -2.24793 -0.03921

TS12

Atom X Y Z

C -0.45815 0.994669 0.583741 C -1.50135 0.660658 -0.42478 O -1.35576 -0.6375 -0.51806 C -0.31686 -0.97909 0.544528 C 0.998775 -1.15355 0.063092 C 1.57255 0.068466 -0.35379 C 0.859069 1.208528 -0.02492 H -0.75409 1.772148 1.30645 H 1.302629 2.198613 -0.08693 H 2.564318 0.127021 -0.80012 H 1.586261 -2.05474 0.219862 H -0.77725 -1.74115 1.178071

IM5

Atom X Y Z

C -2.455714 -0.9735 -0.11956 C -1.73104 -0.00458 -0.01104 C -1.00696 1.197287 0.112774 C 0.349314 1.453207 0.022569 C 1.535443 0.695541 -0.12476 C 1.777467 -0.6986 -0.05351 O 1.006787 -1.6433 0.178146 H -3.08807 -1.81896 -0.21479 H 2.864086 -0.93679 -0.21811 H 2.441038 1.285113 -0.26066 H 0.555855 2.525596 0.06731 H -1.63828 2.075359 0.24223

Page 174: Experimental and computational studies of the unimolecular

160

TS8

Atom X Y Z

O -0.93331 -1.48953 0.395571 C 1.651907 -0.59444 -0.08602 C 1.386956 0.785563 -0.24476 C 0.198868 1.49878 0.041722 C -1.1132 1.115652 0.219013 C -1.67564 -0.17693 -0.03816 C -1.79894 -1.42824 -0.32687 H 2.688773 -0.87222 -0.412 H 2.249263 1.3833 -0.53501 H 0.341364 2.579331 0.124796 H -1.8286 1.893118 0.466455 H -2.81703 -0.26967 -0.19835

TS9

Atom X Y Z

C -1.54764 -0.50508 0.198279 C -1.33169 0.85199 0.07201 C -0.07403 1.471895 -0.24052 C 1.177906 0.93935 -0.05576 O 1.46869 -0.20003 0.555473 C 0.661342 -1.45864 -0.26557 C -0.62461 -1.48443 -0.252 H 1.469809 -1.96143 -0.76948 H 2.036402 1.503449 -0.45155 H -0.09507 2.476151 -0.66471 H -2.19378 1.51703 0.165419 H -2.53454 -0.82549 0.537841

Cyclopentadienyl anion

Atom X Y Z

C 1.202957 0.049914 -2.6E-05 C 0.32422 1.159454 -0.00015 C -1.00255 0.666643 -7.2E-05 C -0.94384 -0.74744 0.000168 C 0.419204 -1.12858 0.000075 H 0.61665 2.20513 -0.00014 H 2.287813 0.095008 0.000716 H 0.797289 -2.14641 -0.00063 H -1.79502 -1.42156 -0.00051 H -1.90668 1.267894 0.000606

CO

Atom X Y Z

C 0 0 -0.64439 O 0 0 0.483295

Page 175: Experimental and computational studies of the unimolecular

161

C6H5

.-

Atom X Y Z

C -1.16261 -0.45078 0.000007 C 0.000001 -1.23131 0.000243 C -1.16261 -0.45079 -0.00018 C -0.70606 0.901241 0.000068 C 0.706059 0.901241 -0.00022 H 2.191222 -0.78956 -0.00011 H 1.343803 1.780757 0.000232 H -1.34381 1.780755 0.000759 H -2.19122 -0.78956 -0.00038

CHO.

Atom X Y Z

C 0.06177 0.58372 0 O 0.06177 -0.59059 0 H -0.86478 1.222377 0