11.05.2019 Views

A study of the solution photophysics of the isomeric aminobenzoic acids, N,N-dimethylaminobenzoic acids and their methyl esters

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

L<br />

A STUDY OF THE SOLUTION PHOTOPHYSICS OF THE ISOMERIC<br />

AM]INOBENZOIC ACIDS, N,N-DIMETHYLAMINOBENZOIC ACIDS AND<br />

THEIR METHYL ESTERS<br />

John Austin Timothy Revill<br />

being a Thesis submitted in partial fulfilment <strong>of</strong> <strong>the</strong> requirements for <strong>the</strong> degree<br />

<strong>of</strong><br />

DOCTOR OF PHILOSOPHY<br />

awarded by <strong>the</strong><br />

UNIVERSITY OF CENTRAL LANCASHIRE<br />

Sponsoring Establishment:<br />

Collaborating Establishment:<br />

University <strong>of</strong> Central Lancashire<br />

Department <strong>of</strong> Applied Physics,<br />

University <strong>of</strong> Strathclyde,<br />

Glasgow<br />

Department <strong>of</strong> Chemistry,<br />

University <strong>of</strong> Central Lancashire,<br />

Preston.<br />

February 1993


Acknowledgements<br />

I would like to thank Pr<strong>of</strong>essor J.D.Hepworth <strong>and</strong> especially Pr<strong>of</strong>essor<br />

KG.Brown, my supervisors, for <strong>the</strong>ir always available help <strong>and</strong> guidance during<br />

my years at <strong>the</strong> Polytechnic <strong>and</strong> later <strong>the</strong> University. I would also like to thank<br />

Dr. D.Birch <strong>of</strong> Strathclyde University <strong>and</strong> Pr<strong>of</strong>. W.Rettig <strong>of</strong> <strong>the</strong> Technical<br />

University <strong>of</strong> Berlin for invaluable discussions <strong>and</strong> advice. Also I would like to<br />

thank The Science <strong>and</strong> Engineering Research Council, without whose financial<br />

support this Ph.D. would have been impossible.<br />

I would like to say a special thankyou to my parents for moral support<br />

<strong>and</strong> helping to make life a little more pleasant.


Chapter 1<br />

Chapter 2<br />

Introduction<br />

1.1 The nature <strong>of</strong> light 1<br />

1.2 Molecular orbital <strong>the</strong>ory 2<br />

1.3 Absorption <strong>of</strong> light 3<br />

1.4 Transition moments 6<br />

1.4.1 Vibrational overlap integral 8<br />

1.4.2 Spin overlap integral 12<br />

1.4.3 Transition moment 14<br />

1.5 Extinction coefficients 15<br />

1.6 Dissipative pathways 19<br />

1.6.1 Vibrational relaxation 20<br />

1.6.2 Radiative transitions 20<br />

1.6.3 Non-radiative transitions 22<br />

1.6.4 Chemical reactions 23<br />

1.7 Quenching 23<br />

1.7.1 Photophysical quenching 23<br />

1.8 Kinetics <strong>of</strong> fluorescence 27<br />

1.8.1 Stern-Volmer equation 29<br />

1.9 Acid Base properties 30<br />

1.10 Literature background to work 31<br />

1.10.1 4-Di<strong>methyl</strong>aminobenzonitrile 35<br />

1.10.2 4-Di<strong>methyl</strong>amino <strong>and</strong> 4-Diethylamino 41<br />

benzoic acid <strong>and</strong> <strong>the</strong>ft <strong>methyl</strong> or ethyl ester<br />

1.10.3 Aims <strong>of</strong> project 42<br />

Experimental<br />

2.1 Materials 43<br />

2.1.1 Solvents 43<br />

2.1.2 Chemicals 43<br />

2.2 Instrumentation (techniques) 44<br />

2.2.1 Absorption spectra 44<br />

2.2.2 Ground <strong>and</strong> excited state pK a <strong>and</strong> pKb values 46<br />

2.2.3 Fluorescence spectra 49


2.2.4 Fluorescence decay pr<strong>of</strong>iles 51<br />

Chapter 3<br />

Solution Photophysics <strong>of</strong> 3-<strong>aminobenzoic</strong> acid <strong>and</strong><br />

3-N,N-di<strong>methyl</strong><strong>aminobenzoic</strong> acid <strong>and</strong> <strong>the</strong>ir <strong>methyl</strong><br />

<strong>esters</strong><br />

3.1 Ground state absorption spectra as a function <strong>of</strong> 55<br />

solvent<br />

3.2 Ground state absorption <strong>and</strong> fluorescence emission 60<br />

data as a function <strong>of</strong> pH<br />

3.3 Fluorescence emission properties as a function <strong>of</strong> 66<br />

solvent<br />

3.4 Fluorescence lifetime data as a function <strong>of</strong> solvent <strong>and</strong> 69<br />

pH<br />

3.5 Fluorescence emission properties as a function <strong>of</strong> 73<br />

temperature<br />

3.6 Summary <strong>of</strong> conclusions 76<br />

Chapter 4<br />

Solution Photophysics <strong>of</strong> 2-aminbbenzoic acid <strong>and</strong><br />

2-N,N-di<strong>methyl</strong><strong>aminobenzoic</strong> acid <strong>and</strong> <strong>the</strong>ir <strong>methyl</strong><br />

<strong>esters</strong><br />

4.1 Ground state absorption spectra as a function <strong>of</strong> 77<br />

solvent<br />

4.2 Ground state absorption <strong>and</strong> fluorescence emission 83<br />

properties as a function <strong>of</strong> pH<br />

4.3 Fluorescence emission properties as a function <strong>of</strong> 88<br />

solvent<br />

4.4 Fluorescence lifetime data as a function <strong>of</strong> solvent <strong>and</strong> 92<br />

pH<br />

4.5 Fluorescence emission properties as a function <strong>of</strong> 97<br />

temperature<br />

4.6 Summary <strong>of</strong> conclusions 97<br />

E


Chapter 5<br />

Solution Photophysics <strong>of</strong> 4-<strong>aminobenzoic</strong> acid <strong>and</strong><br />

4-N,N-di<strong>methyl</strong><strong>aminobenzoic</strong> acid <strong>and</strong> <strong>the</strong>ir <strong>methyl</strong><br />

<strong>esters</strong><br />

5.1 Ground state absorption spectra as a function <strong>of</strong> 99<br />

solvent<br />

5.2 Ground state absorption <strong>and</strong> fluorescence emission 102<br />

properties as a function <strong>of</strong> pH<br />

5.3 Fluorescence emission properties as a function <strong>of</strong> 106<br />

solvent<br />

5.4 Effect <strong>of</strong> excitation wavelength on <strong>the</strong> fluorescence 109<br />

properties <strong>of</strong> 4DMABA <strong>and</strong> M4DMAB<br />

5.5 Variation <strong>of</strong> fluorescence intensity <strong>of</strong> 4DMABA <strong>and</strong> 111<br />

M4DMAB as a function <strong>of</strong> concentration<br />

5.6 Fluorescence lifetime data as a function <strong>of</strong> solvent <strong>and</strong> 117<br />

pH<br />

5.7 Fluorescence properties <strong>of</strong> 4DMABA <strong>and</strong> M4DMAB 124<br />

in mixed solvents<br />

5.8 Fluorescence emission properties as a function <strong>of</strong> 133<br />

temperature<br />

5.9 Summary <strong>of</strong> conclusions 150<br />

Chapter 6 References 151<br />

Appendix A 159<br />

Appendix B 167


Abbreviations<br />

Solvents<br />

BCL I ip Chlorobutane / isopentane (4:1)<br />

BCN / iBCN Butyronitrile I isobutyronitrile (9:1)<br />

MCH I ip Methylcyclohexane / isopentane (4:1)<br />

Compounds<br />

DMABN 4-Di<strong>methyl</strong>aminobenzonitrile<br />

XABA<br />

(x=2, 3 or 4)-Aminobenzoic acid<br />

MxAB<br />

Methyl (x=2, 3 or 4)-aminobenzoate<br />

xDMABA (x=2, 3 or 4)-N,N-di<strong>methyl</strong><strong>aminobenzoic</strong> acid<br />

MXDMAB Methyl (x=2, 3 or 4)-N,N-di<strong>methyl</strong>aminobenzoate<br />

E4DEAB Ethyl 4-diethylaminobenzoate<br />

Miscellaneous<br />

TICT<br />

Twisted intra-molecular charge transfer


ABSTRACT<br />

The <strong>solution</strong> <strong>photophysics</strong> <strong>of</strong> <strong>the</strong> <strong>isomeric</strong> <strong>aminobenzoic</strong> <strong>acids</strong> <strong>and</strong> N,Ndi<strong>methyl</strong><strong>aminobenzoic</strong><br />

<strong>acids</strong> <strong>and</strong> <strong>the</strong>ir <strong>methyl</strong> <strong>esters</strong> were studied under a<br />

variety <strong>of</strong> conditions <strong>of</strong> solvent, pH, concentration <strong>and</strong> temperature.<br />

The most straightforward behaviour was obtained when <strong>the</strong> amino or<br />

N,N-di<strong>methyl</strong>amino substituent was in <strong>the</strong> meta- position relative to <strong>the</strong><br />

carboxylic add or ester function. The amino compounds absorb with a peak<br />

between 300 <strong>and</strong> 325 nm depending on solvent, <strong>the</strong>ir fluorescence peaks lie<br />

between 350 <strong>and</strong> 410 rim. The corresponding N,N-di<strong>methyl</strong>amino compounds<br />

usually absorb <strong>and</strong> fluoresce some 20 - 30 rim to <strong>the</strong> red <strong>of</strong> <strong>the</strong>se wavelengths.<br />

Both protonation <strong>and</strong> deprotonation <strong>of</strong> <strong>the</strong> ground state molecules takes place<br />

under weakly acidic conditions with pKi, values <strong>of</strong> approximately 3.0 <strong>and</strong> pK a<br />

values <strong>of</strong> approximately 5.0 being obtained. The observed variation in <strong>the</strong><br />

fluorescence <strong>of</strong> <strong>the</strong> meta isomers with pH is largely caused by ground state<br />

equilibria. Förster cycle calculations suggest that <strong>the</strong> excited state pK a is little<br />

changed from <strong>the</strong> ground state whereas pKC is much smaller. This is in good<br />

agreement with <strong>the</strong> observed fluorescence data <strong>and</strong> <strong>the</strong> nature <strong>of</strong> <strong>the</strong> excited<br />

state. Measurements <strong>of</strong> fluorescence decay pr<strong>of</strong>iles reveal a longer lived excited<br />

singlet state (usually some 10 -20 ns) than is seen for <strong>the</strong> o<strong>the</strong>r isomers. The<br />

fluorescence intensity <strong>of</strong> <strong>the</strong> meta- isomers was observed to rise with decreasing<br />

temperature, although this could possibly be attributed to absorbance changes.<br />

The <strong>photophysics</strong> <strong>of</strong> <strong>the</strong> corresponding ortho compounds is complicated<br />

by steric <strong>and</strong> hydrogen-bonding interactions especially in <strong>the</strong> compounds<br />

containing an N,N-di<strong>methyl</strong>amino group. The increased interaction between<br />

<strong>the</strong> two substituent groups leads to an increase in extinction coefficient <strong>and</strong><br />

fluorescence quantum yield <strong>and</strong> a decrease in fluorescence lifetime for 2-<br />

<strong>aminobenzoic</strong> acid <strong>and</strong> its <strong>methyl</strong> ester compared to <strong>the</strong> meta derivatives.<br />

Spectral b<strong>and</strong> positions, pK values <strong>and</strong> <strong>the</strong> effect <strong>of</strong> temperature on <strong>the</strong><br />

fluorescence are similar in <strong>the</strong> two sets <strong>of</strong> compounds. The 2-N,Ndi<strong>methyl</strong>amino<br />

compounds do not parallel <strong>the</strong> corresponding meta- derivatives<br />

due to steric hindrance <strong>and</strong> intra-molecular hydrogen bonding. These cause<br />

weak, blue shifted absorptions <strong>and</strong> low fluorescence quantum yields. There<br />

was no evidence found in this work for <strong>the</strong> presence <strong>of</strong> a zwitterionic species in<br />

<strong>the</strong> absorption properties <strong>of</strong> 2-N,N-di<strong>methyl</strong><strong>aminobenzoic</strong> acid.<br />

The photophysical properties <strong>of</strong> 4-<strong>aminobenzoic</strong> acid <strong>and</strong> its <strong>methyl</strong><br />

ester are in many ways very similar to those <strong>of</strong> <strong>the</strong> corresponding orthosubstituted<br />

compounds. However <strong>the</strong> same certainly cannot be said for 4-N,Ndi<strong>methyl</strong><strong>aminobenzoic</strong><br />

add <strong>and</strong> its <strong>methyl</strong> ester. These two compounds exhibit<br />

anomalous fluorescence in <strong>the</strong> form <strong>of</strong> two fluorescence b<strong>and</strong>s in some<br />

solvents. The appearance <strong>of</strong> this anomalous fluorescence is both concentration<br />

<strong>and</strong> solvent composition dependent. Various mechanisms to explain <strong>the</strong>se<br />

observations were tested; excimer, exciplex, TICT state <strong>and</strong> ground state dimer<br />

formation. The latter was found to be in best accord with <strong>the</strong> properties <strong>of</strong> <strong>the</strong>se<br />

molecules, but it was clear that in binary solvent mixtures such as hexane /<br />

acetonitrile <strong>and</strong> acetonitrile / water o<strong>the</strong>r mechanisms could also be involved.


Chapter 1


Chapter 1 Introduction<br />

1.1 The nature <strong>of</strong> light<br />

Photochemistry is concerned with <strong>the</strong> chemical <strong>and</strong> physical effects <strong>of</strong><br />

electronic excitation produced by <strong>the</strong> interaction <strong>of</strong> electromagnetic radiation<br />

with matter. There are two physical models which may be used to describe <strong>the</strong><br />

properties <strong>of</strong> electromagnetic radiation (emr). The wave model takes into<br />

account <strong>the</strong> electric <strong>and</strong> magnetic components <strong>of</strong> <strong>the</strong> radiation <strong>and</strong> envisages<br />

electromagnetic radiation in terms <strong>of</strong> oscillating electric <strong>and</strong> magnetic fields<br />

in planes which are perpendicular to each o<strong>the</strong>r <strong>and</strong> to <strong>the</strong> direction <strong>of</strong><br />

propagation 1 (figure 1.1). The z co-ordinate is in <strong>the</strong> direction <strong>of</strong> propagation,<br />

so that <strong>the</strong> electric vector is contained in <strong>the</strong> zy plane <strong>and</strong> <strong>the</strong> magnetic vector<br />

in <strong>the</strong> xz plane.<br />

I<br />

Figure 1.1 The instantaneous electric <strong>and</strong> magnetic field strength vectors as<br />

a function <strong>of</strong> position along axis <strong>of</strong> propagation <strong>of</strong> emr<br />

The quantum model considers light to have a particle nature (Planck<br />

1901) <strong>and</strong> to exist as discrete quanta called photons, whose energy E is given by<br />

E=hv= 1.1<br />

x<br />

where h = Plancks constant, c is <strong>the</strong> speed <strong>of</strong> light, V is <strong>the</strong> frequency <strong>of</strong> <strong>the</strong><br />

radiation <strong>and</strong> Xis its wavelength. Table 1.1 gives details <strong>of</strong> selected values <strong>of</strong><br />

<strong>the</strong>se parameters.<br />

1


Radiation Wavelength Frequency Energy<br />

/ rim v/s-1x10 15 E / kJ mo!-'<br />

Far UV 100 3.00 1200<br />

Near UV 250 1.20 480<br />

U V 350 0.86 343<br />

Blue-green 500 0.60 240<br />

Red 700 0.41 171<br />

Infrared 1000 0.30 120<br />

Table 1.1 Colour, frequency, wavelength <strong>and</strong> energy <strong>of</strong> light<br />

1.2 Molecular orbital <strong>the</strong>ory<br />

To underst<strong>and</strong> <strong>the</strong> interaction <strong>of</strong> matter with electromagnetic<br />

radiation, a knowledge <strong>of</strong> molecular orbital <strong>the</strong>ory is required. Molecular<br />

orbitals are formulated as linear combinations <strong>of</strong> <strong>the</strong> valence shell atomic<br />

orbitals. It is assumed that <strong>the</strong> inner electrons remain in <strong>the</strong>ir original atomic<br />

orbitals. For example <strong>the</strong> interaction <strong>of</strong> two identical atomic orbitals 23a <strong>and</strong><br />

Ob give rise to two molecular orbitals <strong>of</strong> <strong>the</strong> form:<br />

411= 0a+ 13b 1.2<br />

2=t a -lb 1.3<br />

One molecular orbital ljf I is bonding (more stable than <strong>the</strong> original<br />

atomic orbitals) <strong>and</strong> <strong>the</strong> o<strong>the</strong>r W2 is antibonding (less stable) 2. Those which<br />

are completely symmetrical about <strong>the</strong> internuclear axis are designated sigma<br />

orbitals <strong>and</strong> arise when <strong>the</strong> interacting atomic orbitals 13a <strong>and</strong> ôb are s<br />

orbitals or p orbitals. Molecular orbitals derived from <strong>the</strong> interaction <strong>of</strong> two<br />

parallel p orbitals are called a orbitals <strong>and</strong> again bonding <strong>and</strong> antibonding<br />

molecular orbitals are formed. If, prior to formation <strong>of</strong> <strong>the</strong> molecular orbital,<br />

<strong>the</strong> atomic orbitals are each occupied by a single electron or if one is occupied<br />

by two electrons <strong>and</strong> <strong>the</strong> o<strong>the</strong>r is vacant <strong>the</strong>n <strong>the</strong> electrons in <strong>the</strong> molecular<br />

system both occupy <strong>the</strong> lower energy bonding molecular orbital. This leads to<br />

a gain in stability over <strong>the</strong> isolated atoms <strong>and</strong> is <strong>the</strong> basis <strong>of</strong> <strong>the</strong> molecular<br />

orbital description <strong>of</strong> electron pair covalent bonding. Certain compounds


contain non-bonding valence shell electrons (n). These electrons are not<br />

involved in bonding relationships <strong>and</strong> can be regarded as being localised on<br />

<strong>the</strong>ir atomic nuclei. In terms <strong>of</strong> <strong>the</strong>ir energy, <strong>the</strong> molecular orbitals generally<br />

take <strong>the</strong> following order (figure 1.2).<br />

Antibonding *<br />

Antibonding .Jt *<br />

Energy<br />

Non-bonding<br />

Bonding<br />

It<br />

- Si<br />

Figure 1.2 Relative energies <strong>of</strong> molecular orbitals<br />

1.3 Absorption <strong>of</strong> light<br />

In a normal molecule, two electrons are assigned to each molecular<br />

orbital where <strong>the</strong> electrons fill <strong>the</strong> orbitals from <strong>the</strong> lowest energy orbital<br />

upwards in such a manner that <strong>the</strong>ir spins are paired, in order to have <strong>the</strong><br />

most stable energy configuration. In order for electronic excitation to occur<br />

<strong>the</strong> energy <strong>of</strong> <strong>the</strong> absorbed photon (E) must equal <strong>the</strong> energy difference<br />

between an occupied molecular orbital <strong>and</strong> an unoccupied molecular orbital.<br />

This electronic excitation <strong>of</strong> a molecule results in <strong>the</strong> promotion <strong>of</strong> an<br />

electron from one molecular orbital to ano<strong>the</strong>r <strong>of</strong> higher energy e.g. a-as,<br />

n_a* , it-it * or n-i? <strong>and</strong> retains <strong>the</strong> same electron spin. Figure 1.3 gives an<br />

example <strong>of</strong> a n-71* transition for methanal (formaldehyde).<br />

3


7t*<br />

HQ<br />

Op<br />

H'6I<br />

Ill<br />

hu -<br />

II<br />

n(p)<br />

n(sp)<br />

Figure 1.3 Electronic excitation <strong>of</strong> methanal<br />

11<br />

II<br />

In general terms <strong>the</strong> acY* transition requires <strong>the</strong> greatest energy <strong>and</strong><br />

<strong>the</strong> wavelength <strong>of</strong> absorption associated with <strong>the</strong> transition will generally be<br />

in <strong>the</strong> far UV. Both <strong>the</strong> cy-a<br />

<strong>and</strong> <strong>the</strong> nG* transitions will be <strong>of</strong> little<br />

significance when discussing <strong>the</strong> photochemistry <strong>of</strong> compounds containing Jr<br />

electrons since <strong>the</strong>se transitions involving it electrons are considerably more<br />

intense <strong>and</strong> <strong>of</strong> lower energy. In e<strong>the</strong>ne, <strong>the</strong> lowest energy transition is <strong>the</strong><br />

singlet ic-ic * which gives rise to an intense absorption in <strong>the</strong> ultraviolet 3. The<br />

triplet ic-ice transition (spin inversion <strong>of</strong> promoted electron) is so weak that a<br />

path length <strong>of</strong> 14m <strong>of</strong> liquid e<strong>the</strong>ne would be required for its detection 4 .<br />

Conjugation shifts it-ice transitions to longer wavelengths. This is because <strong>the</strong><br />

energy gap between <strong>the</strong> highest occupied molecular orbital (HOMO) <strong>and</strong> <strong>the</strong><br />

lowest unoccupied molecular orbital (LUMO) becomes progressively smaller<br />

as <strong>the</strong> number <strong>of</strong> conjugated double bonds increases. Consequently <strong>the</strong><br />

wavelength <strong>of</strong> maximum absorption increases. An example <strong>of</strong> this is found<br />

in butadiene which involves <strong>the</strong> interaction <strong>of</strong> two C=C double bonds,<br />

4


leading to a red shift in <strong>the</strong> absorption maximum relative to that found for<br />

e<strong>the</strong>ne.<br />

The n_it* transition is characteristic <strong>of</strong> molecules possessing<br />

chromophores with multiple bonded hetero atoms. It is <strong>of</strong> low intensity<br />

because it is symmetry <strong>and</strong>/or overlap forbidden <strong>and</strong> is normally <strong>the</strong> b<strong>and</strong><br />

occurring at longest wavelength. Acetone exhibits an example <strong>of</strong> a n ir*<br />

transition. In <strong>the</strong> ground state <strong>the</strong> non-bonding <strong>and</strong> it-bonding orbitals are<br />

doubly occupied, leaving <strong>the</strong> ir orbital empty, figure 1.4. In <strong>the</strong> lowest excited<br />

singlet state <strong>the</strong> non-bonding <strong>and</strong> ic orbitals are both singly occupied, leaving<br />

<strong>the</strong>ir-bonding orbital doubly occupied. This nir* transition has a wavelength<br />

<strong>of</strong> maximum absorption <strong>of</strong> 280 nm (E=18 dm- 3 mol-1 cm -1). However in <strong>the</strong><br />

region <strong>of</strong> 190 nm <strong>the</strong>re is an intense b<strong>and</strong> which may be due to an nG* or G-<br />

transition5.<br />

Plan<br />

Elevation<br />

Me\ Me.e Me.Q<br />

/CO<br />

Me Me / 0 Me43<br />

n-orbital n-orbital n *_orbita l<br />

Figure 1.4 Molecular orbitals <strong>of</strong> acetone<br />

Mixtures <strong>of</strong> electron donors <strong>and</strong> electron acceptors in <strong>solution</strong> <strong>of</strong>ten<br />

exhibit a new absorption b<strong>and</strong> which is shown by nei<strong>the</strong>r component<br />

separately. It is attributed to <strong>the</strong> presence in such mixtures <strong>of</strong> a donor-acceptor<br />

complex 6. Typical acceptors are picric acid <strong>and</strong> o<strong>the</strong>r polynitro aromatic<br />

compounds; typical donors include aromatic hydrocarbons. The transition is<br />

referred to as a charge-transfer transition <strong>and</strong> is generally broad <strong>and</strong><br />

structureless. The spectroscopic transition corresponds approximately to <strong>the</strong><br />

light induced transfer <strong>of</strong> an electron from <strong>the</strong> donor to <strong>the</strong> acceptor. An<br />

5


example <strong>of</strong> such a complex is between <strong>the</strong> acceptor 4-benzoquinone <strong>and</strong><br />

donor 2,3-di<strong>methyl</strong>butadiene. In <strong>methyl</strong>cyclohexane <strong>solution</strong> <strong>the</strong> mixture<br />

produces two b<strong>and</strong>s at 450 <strong>and</strong> 300 nm, but in pure 2,3-di<strong>methyl</strong>butadiene<br />

only a single peak is observed at 350 nm 7. A single molecule can include both<br />

donor, <strong>and</strong> acceptor groups such that a charge transfer (CT) transition can be<br />

observed in <strong>solution</strong>s <strong>of</strong> pure compounds.<br />

1.4 The transition moment <strong>and</strong> probability <strong>of</strong> transitions<br />

The probability <strong>of</strong> an optical transition between two states is given by<br />

<strong>the</strong> value <strong>of</strong> <strong>the</strong> transition moment (TM) integral 8<br />

TM =J tI'.4L'P 1 d'r 1.4<br />

where 'F. is <strong>the</strong> total wavefunction <strong>of</strong> <strong>the</strong> initial state, T f<br />

is <strong>the</strong> total<br />

wavefunction <strong>of</strong> <strong>the</strong> final state <strong>and</strong> j.t is <strong>the</strong> dipole moment operator. The<br />

transition moment in principle can be calculated from quantum mechanics<br />

but it can also be obtained from absorption spectra through its relation to <strong>the</strong><br />

oscillator strength f (see later). Although this integral cannot be precisely<br />

evaluated because <strong>the</strong> exact forms <strong>of</strong> <strong>the</strong> wave functions are not known,<br />

approximations can be made.<br />

Nuclear motion is considered to be very slow in comparison to that <strong>of</strong><br />

electrons because <strong>of</strong> <strong>the</strong> large mass differences. The motion <strong>of</strong> <strong>the</strong> electrons<br />

can <strong>the</strong>refore be considered as being independent <strong>of</strong> that <strong>of</strong> <strong>the</strong> nucleus. This<br />

approximation is known as <strong>the</strong> Born-Oppenheimer approximation, which<br />

bears up extremely well until degenerate orbitals are found where <strong>the</strong>ir<br />

potential energy surfaces cross. Thus <strong>the</strong> total wavefunction can be factorised<br />

into a nuclear (vibrational) wavefunction 0 <strong>and</strong> an electronic wavefunction<br />

w, ('PrOw,). The transition moment is <strong>the</strong>n given by


TM = I OI.WLJLOLWf di . 1.5<br />

Since i' operates only on <strong>the</strong> electrons, equation 1.5 can be written as<br />

TM =5 Oj.O f dt I W.J.LW1 d; 1.6<br />

where n <strong>and</strong> e refer to nuclei <strong>and</strong> electrons respectively. Fur<strong>the</strong>r<br />

approximations have to be made. It is assumed<br />

i) that w is <strong>the</strong> product <strong>of</strong> one electron wavefunctions (orbitals which are<br />

<strong>the</strong>mselves linear combinations <strong>of</strong> atomic orbitals) <strong>and</strong> that <strong>the</strong> same orbitals<br />

may be used to describe both ground <strong>and</strong> excited state <strong>and</strong><br />

ii)<br />

that only one electron is promoted during a transition. The transition<br />

moment can <strong>the</strong>n be written<br />

TM = JOi O f d'Tn J4aI.L4idte 1.7<br />

where O i <strong>and</strong> <strong>of</strong> are <strong>the</strong> wavefunctions <strong>of</strong> <strong>the</strong> initial <strong>and</strong> final orbitals <strong>of</strong> <strong>the</strong><br />

electron.<br />

The final approximation is that <strong>the</strong> orbitals can be factorised into a<br />

product <strong>of</strong> space <strong>and</strong> spin orbitals (Q <strong>and</strong> S respectively)<br />

1.8<br />

TM =fO i.Of dCn JSS1dt JQiIiQfdt e 1.9<br />

It should be noted that R only operates on <strong>the</strong> space co-ordinate.<br />

The first term in <strong>the</strong> transition moment expression (equation 1.9) is <strong>the</strong><br />

overlap integral <strong>of</strong> <strong>the</strong> wavefunctions for nuclear vibrations. The second<br />

7


term is a spin overlap integral <strong>and</strong> its value depends on <strong>the</strong> initial <strong>and</strong> final<br />

spin states <strong>of</strong> <strong>the</strong> promoted electron. The third term is called <strong>the</strong> electronic<br />

transition moment <strong>and</strong> its value depends on <strong>the</strong> symmetries <strong>and</strong> <strong>the</strong> amount<br />

<strong>of</strong> spatial overlap <strong>of</strong> <strong>the</strong> initial <strong>and</strong> final orbitals.<br />

This application <strong>of</strong> quantum mechanical <strong>the</strong>ory to electronic processes<br />

has led to a set <strong>of</strong> selection rules 9 which allow transitions to be classified as<br />

allowed or forbidden. The probability <strong>of</strong> occurrence <strong>of</strong> an electronic transition<br />

<strong>and</strong> hence <strong>the</strong> intensity <strong>of</strong> <strong>the</strong> absorption b<strong>and</strong> is dependent upon various<br />

factors. These factors are included in <strong>the</strong> selection rules which govern<br />

whe<strong>the</strong>r or not a transition will be allowed or forbidden. Transitions which<br />

obey <strong>the</strong> selection rules can give rise to very intense absorption b<strong>and</strong>s.<br />

Conversely, transitions which do not conform to <strong>the</strong> selection rules ei<strong>the</strong>r do<br />

not occur or else <strong>the</strong> probability <strong>of</strong> occurrence is so low that only very weak<br />

b<strong>and</strong>s are observed in <strong>the</strong> spectrum.<br />

The transition moment in equation 1.9 is a product <strong>of</strong> three separate<br />

integrals <strong>and</strong> its value will be zero if any one <strong>of</strong> <strong>the</strong> integrals is zero. When<br />

this happens <strong>the</strong> transition has a zero probability <strong>of</strong> occurrence <strong>and</strong> is said to<br />

be forbidden. When <strong>the</strong> transition moment is non-zero, <strong>the</strong> transition is said<br />

to be allowed.<br />

1.4.1 The vibrational overlap integral 10 JO1 O f dTn may be evaluated from a<br />

knowledge <strong>of</strong> <strong>the</strong> vibrational wavefunctions <strong>of</strong> <strong>the</strong> two states involved in <strong>the</strong><br />

transition (obtained by solving <strong>the</strong> appropriate Schrodinger equation) <strong>and</strong> <strong>the</strong><br />

application <strong>of</strong> <strong>the</strong> Franck-Condon principle 11 . The principle states that an<br />

electronic transition occurs so rapidly in comparison with vibrational<br />

frequencies that no change in internuclear separation occurs during <strong>the</strong><br />

course <strong>of</strong> <strong>the</strong> transition. This implies that <strong>the</strong> transition may be represented<br />

by a vertical line connecting <strong>the</strong> two potential energy surfaces. The most<br />

probable transition from this vibrational level at <strong>the</strong> same internuclear<br />

F1


distance will be <strong>the</strong> turning point <strong>of</strong> <strong>the</strong> oscillation. In order to evaluate <strong>the</strong><br />

f0 10 1d; integral, it is necessary to know <strong>the</strong> vibrational wavefunctions, which<br />

for simplicity will be treated as simple harmonic oscillators.<br />

-- awawaw<br />

R WV<br />

El an.<br />

111 way<br />

In<br />

I- -<br />

e<br />

a)<br />

Internuclear seporQtion Cr)<br />

Figure 1.5 Ground state vibrational wavefunctions <strong>and</strong> energy levels for a<br />

harmonic oscillator<br />

For such systems, <strong>the</strong> potential energy varies with internuclear separation as<br />

described by <strong>the</strong> parabolic curve <strong>of</strong> figure 1.5. The equilibrium separation re is<br />

given by <strong>the</strong> point a, where <strong>the</strong> potential energy is a minimum. Solution <strong>of</strong><br />

<strong>the</strong> appropriate Schrodinger equation leads to <strong>the</strong> result that <strong>the</strong> vibrational<br />

energy (Ev) is quantized with values<br />

Ev=hU(V+ >) 1.10<br />

where 1) is <strong>the</strong> fundamental vibrational frequency <strong>and</strong> V <strong>the</strong> vibrational<br />

quantum number, which is an integer 0,1,2,.. .n. These values are represented<br />

by <strong>the</strong> equally spaced horizontal lines in <strong>the</strong> figure. The oscillator curves<br />

represent <strong>the</strong> wavefunctions associated with each level. Since <strong>the</strong> square <strong>of</strong><br />

<strong>the</strong> amplitude (0) <strong>of</strong> <strong>the</strong> function gives <strong>the</strong> probability <strong>of</strong> a particular nuclear<br />

configuration, we can see from <strong>the</strong> bottom vibrational level (V"=O) that <strong>the</strong>


nuclei are most likely to be found at <strong>the</strong> equilibrium point re <strong>and</strong> as V"<br />

increases it becomes more probable that <strong>the</strong> molecule will come close to <strong>the</strong><br />

configuration corresponding to <strong>the</strong> intersection <strong>of</strong> <strong>the</strong> horizontal lines <strong>and</strong><br />

parabolic curve, (<strong>the</strong> turning point <strong>of</strong> <strong>the</strong> vibration, where <strong>the</strong> total energy<br />

equals <strong>the</strong> potential energy, <strong>the</strong> kinetic energy is zero, <strong>and</strong> <strong>the</strong> nuclei are<br />

static). If upon excitation <strong>the</strong>re is no change in geometry, <strong>the</strong>n figure 1.6A is a<br />

schematic representation <strong>of</strong> <strong>the</strong> most likely transition. From <strong>the</strong> figure it can<br />

be seen that <strong>the</strong> most probable transition is V"=O to V'=O i.e. <strong>the</strong> O—*O<br />

transition.<br />

Excited<br />

state<br />

0'<br />

a;<br />

C<br />

Li<br />

:1<br />

C)<br />

=0<br />

Le<br />

Ground<br />

state<br />

Internuclear<br />

Internuclear separation (r separation (-)<br />

Figure 1.6 Ground state <strong>and</strong> excited state potential energy curves. (A) with<br />

no change in geometry upon excitation <strong>and</strong> (B) with change in<br />

geometry upon excitation, <strong>the</strong> most common scenario<br />

Figure 1.6B represents <strong>the</strong> more usual case where stretching <strong>of</strong> bonds<br />

upon excitation leads to a change in geometry (from exciting <strong>the</strong> electron into<br />

an antibonding orbital thus weakening <strong>the</strong> bond). In figure 1.68, <strong>the</strong> V"=O to<br />

V'=2 transition will be <strong>the</strong> most likely vibrational transition. Very similar<br />

considerations apply to emission spectra, with <strong>the</strong> proviso that <strong>the</strong> excited<br />

state will normally become <strong>the</strong>rmally equilibrated before emission, which<br />

i-El]


<strong>the</strong>refore takes place almost exdusively from <strong>the</strong> bottom vibrational level <strong>of</strong><br />

<strong>the</strong> excited state (V'=O); see figure 1.7.<br />

a.<br />

I-<br />

a)<br />

C<br />

Ui<br />

p<br />

Figure 1.7 Representation <strong>of</strong> <strong>the</strong> most likely emission <strong>and</strong> absorption<br />

transitions<br />

The Franck Condon principle states that electronic transitions occur in<br />

an exceedingly brief interval <strong>of</strong> time so that no change in nuclear position or<br />

nuclear kinetic energy occurs during <strong>the</strong> transition. This implies that <strong>the</strong><br />

transition may be represented by a vertical line connecting <strong>the</strong> two potential<br />

energy surfaces, <strong>and</strong> <strong>the</strong> most probable transition will be to that vibrational<br />

level with <strong>the</strong> same internuclear distance at <strong>the</strong> turning point <strong>of</strong> <strong>the</strong><br />

oscillation, eg lines AY <strong>and</strong> ZB in Figure 1.7. A transition represented by line<br />

AX is extremely improbable because <strong>the</strong> molecule, in arriving at point X,<br />

would have suddenly acquired an excess kinetic energy given by XI.<br />

There are however various assumptions inherent in this treatment, which<br />

are questionable given that<br />

11


i) real molecules are anharmonic oscillators. For a diatomic molecule <strong>the</strong><br />

potential energy diagram is approximated by a Morse curve 12, <strong>the</strong> higher<br />

vibrational levels <strong>of</strong> which become progressively closer toge<strong>the</strong>r until <strong>the</strong>y<br />

merge into a continuum at <strong>the</strong> dissociation limit.<br />

ii) For any molecule containing more than two atoms <strong>the</strong> potential<br />

energy curve becomes a surface in many dimensions, due to <strong>the</strong> large number<br />

<strong>of</strong> possible vibrational modes.<br />

iii) For a molecule which undergoes a change in geometry upon excitation,<br />

<strong>the</strong> vibrational <strong>and</strong> rotational transitions are so closely spaced that <strong>the</strong>y<br />

overlap giving rise to a smooth broad absorption. Hence in <strong>the</strong> majority <strong>of</strong><br />

cases little vibrational structure may be seen in visible or ultraviolet<br />

absorption b<strong>and</strong>s.<br />

1.4.2 The spin overlap integral<br />

The effect <strong>of</strong> electron spin upon transition intensities is given by <strong>the</strong><br />

spin overlap integral JSS1d The electrons involved in <strong>the</strong> transition can only<br />

have two spin states (designated a <strong>and</strong> 3) <strong>and</strong> <strong>the</strong>re are three common<br />

situations<br />

12


(a) Singlet (--* Singlet Transitions<br />

In singlet e-* singlet transitions no change occurs in <strong>the</strong> spin state <strong>of</strong> <strong>the</strong><br />

promoted electron <strong>and</strong> <strong>the</strong> spin overlap integral is unity<br />

Jaad=JJ3j3 d5= 1<br />

Lii<br />

because <strong>the</strong> spin wave functions are normalised. There are no spin<br />

restrictions on such transitions which are <strong>the</strong>refore fully allowed.<br />

(b) Triplet H Triplet Transitions<br />

Since <strong>the</strong> transition occurs with no change in multiplicity, <strong>the</strong> spin<br />

overlap integral is again unity <strong>and</strong> <strong>the</strong> transition is fully allowed.<br />

(c) Singlet *-* Triplet Transition<br />

Since <strong>the</strong> promoted electron changes its spin state <strong>the</strong> spin overlap<br />

integral is<br />

1j3j3d5=Jctccd5=O 1.12<br />

because <strong>the</strong> cx <strong>and</strong> ji spin wavefunctions are orthogonal. The transition<br />

moment is thus zero <strong>and</strong> <strong>the</strong> transition is strongly forbidden.<br />

Although singlet 4-> triplet transitions are strongly forbidden, <strong>the</strong>y do<br />

occur through spin-orbit coupling. Since <strong>the</strong> electron is charged <strong>and</strong> spinning,<br />

it is expected to have not only spin angular momentum but also a magnetic<br />

moment. In a singlet


that although total angular momentum is conserved, nei<strong>the</strong>r orbital nor spin<br />

angular momentum are individually conserved. This means that <strong>the</strong> concept<br />

<strong>of</strong> a pure spin state is no longer tenable but that intermediate situations must<br />

be contemplated in which a normal singlet state has a certain degree <strong>of</strong> triplet<br />

character <strong>and</strong> vice versa. In conclusion, although singlet e-* triplet transitions<br />

are strictly forbidden, weak b<strong>and</strong>s may be observed 4 .<br />

1.4.3 The electronic transition moment<br />

The electronic transition moment is intimately related to <strong>the</strong><br />

symmetries <strong>of</strong> <strong>the</strong> orbitals. Transitions may be space <strong>and</strong>/or overlap<br />

forbidden. For example, n1c* transitions involve two orbitals located in<br />

different regions <strong>of</strong> space (i.e. no overlap occurs) <strong>and</strong> <strong>the</strong>refore <strong>the</strong>se<br />

transitions are overlap forbidden. However, <strong>the</strong> two orbitals involved in <strong>the</strong><br />

ic-ir transition have large amplitudes <strong>of</strong> <strong>the</strong>ir wavefunctions in <strong>the</strong> same<br />

region <strong>of</strong> space <strong>and</strong> are <strong>the</strong>refore not overlap forbidden.<br />

Symmetry forbidden transitions should have zero intensity <strong>and</strong> be<br />

unobservable, although <strong>the</strong>y are found to have small but finite intensities.<br />

This feature arises because <strong>the</strong> vibrational (nuclear) <strong>and</strong> electronic motions<br />

are not completely independent <strong>of</strong> each o<strong>the</strong>r but are weakly coupled. The<br />

coupling, called vibronic coupling, is responsible for <strong>the</strong> non zero intensity <strong>of</strong><br />

symmetry forbidden transitions such as <strong>the</strong> benzene absorption at 254 nm.<br />

This transition is forbidden if <strong>the</strong> molecule is in <strong>the</strong> form <strong>of</strong> a regular<br />

hexagon but <strong>the</strong>re are vibrations which distort <strong>the</strong> hexagon <strong>and</strong> reduce its<br />

symmetry, so that <strong>the</strong> transition becomes weakly allowed. A fur<strong>the</strong>r type <strong>of</strong><br />

transition which involves symmetry is <strong>the</strong> parity forbidden transition. When<br />

<strong>the</strong> wave function <strong>of</strong> a molecule changes sign on inversion through a centre<br />

<strong>of</strong> symmetry it is called ungerade, while those not changing sign on reflection<br />

are called gerade. Selection rules state that ungerade -* gerade transitions will<br />

14


e allowed but gerade t4 gerade <strong>and</strong> ungerade #4 ungerade transitions are<br />

forbidden.<br />

The selection rules can be summarised as follows. If a completely<br />

allowed transition has an oscillator strength fa <strong>of</strong> <strong>the</strong> order <strong>of</strong> unity, <strong>the</strong>n<br />

o<strong>the</strong>r transitions have oscillator strengths f given by<br />

f=fo *fs *fsym *fa 1.13<br />

where f0 is <strong>the</strong> overlap factor <strong>and</strong> fs is <strong>the</strong> electron spin factor <strong>and</strong> fsym is <strong>the</strong><br />

symmetry or parity factor 14.<br />

1.5 Extinction coefficients<br />

Experimental measurements <strong>of</strong> absorption transitions usually use<br />

extinction coefficients (molar absorptivities) to quantify <strong>the</strong> transition<br />

intensity. Under normal circumstances absorption properties can be<br />

represented by <strong>the</strong> Beer-Lambert Law:<br />

I<br />

1.14<br />

= Ecl = Absorbance<br />

(L. )<br />

where Jo is <strong>the</strong> intensity <strong>of</strong> <strong>the</strong> incident monochromatic radiation, I t is <strong>the</strong><br />

intensity <strong>of</strong> <strong>the</strong> transmitted radiation, c is <strong>the</strong> concentration <strong>of</strong> <strong>the</strong> sample, 1 is<br />

<strong>the</strong> path length <strong>and</strong> S is <strong>the</strong> molar extinction coefficient. The molar extinction<br />

coefficient, or more correctly <strong>the</strong> decadic molar extinction coefficient, is<br />

characteristic <strong>of</strong> a particular compound at a particular wavelength <strong>of</strong><br />

radiation. This empirical law is valid except when very high intensities <strong>of</strong><br />

radiation are employed <strong>and</strong> a significant proportion <strong>of</strong> <strong>the</strong> molecules in a<br />

given region are in <strong>the</strong> excited state ra<strong>the</strong>r than <strong>the</strong> ground state at an one<br />

time.<br />

15


An alternative measure <strong>of</strong> absorption intensity, which can be related<br />

more readily to <strong>the</strong>oretical principles is <strong>the</strong> oscillator strength f given by <strong>the</strong><br />

equation 1.15<br />

= 12303mc 2<br />

,tNe2 ]Fjcdv = 4.315x1crFJ cdv ..........1.15<br />

[<br />

where <strong>the</strong> integration extends over <strong>the</strong> entire b<strong>and</strong> related to <strong>the</strong> transition<br />

from state S-'S; N is Avogadros's number, <strong>and</strong> F constitutes a correction<br />

factor related to <strong>the</strong> refractive index <strong>of</strong> <strong>the</strong> medium in which <strong>the</strong> absorbing<br />

molecule is dissolved. In most cases this value is taken as 1. The major<br />

difference between <strong>the</strong> oscillator strength <strong>and</strong> <strong>the</strong> extinction coefficient is that<br />

<strong>the</strong> former is a measure <strong>of</strong> <strong>the</strong> integrated intensity <strong>of</strong> absorption over a whole<br />

b<strong>and</strong>, whereas £ is a measure <strong>of</strong> <strong>the</strong> intensity <strong>of</strong> <strong>the</strong> absorption at a single<br />

wavelength.<br />

The lowest energy absorption transition in benzene, a 'Mg 3 1 132 u<br />

transition ( 1 A -* 'Lb in Platt notation), is a symmetry forbidden g -3 u<br />

transition which becomes partially allowed through vibronic coupling 15. The<br />

wavelength <strong>of</strong> maximum absorption lies around 255 nm with an extinction<br />

coefficient <strong>of</strong> 220 dm3 mol-1 cm-1 . The S0 -* S2 absorption is also forbidden,<br />

being a 1 Ai g -* 1 13 1 transition ('A -3 'L a), but exhibits a much higher<br />

extinction coefficient (see Table 1.2). The first allowed transition, 1 A1g -+ 'Ei<br />

('A -4 'B), is much more intense that ei<strong>the</strong>r <strong>of</strong> <strong>the</strong> first two transitions but<br />

lies in <strong>the</strong> vacuum ultraviolet.<br />

On substitution, <strong>the</strong> high symmetry <strong>of</strong> <strong>the</strong> benzene molecule (D6h) is<br />

removed <strong>and</strong> <strong>the</strong> So - transitions become more allowed. The addition <strong>of</strong><br />

both electron donating <strong>and</strong> electron withdrawing groups tends to cause a<br />

bathochromic shift in <strong>the</strong> lowest energy absorption b<strong>and</strong> 16. On adding a<br />

carboxylic acid group, a new absorption b<strong>and</strong> is observed around 280 nm (e<br />

16


1,000 dm3 mo!-1 cm-I)16 <strong>and</strong> <strong>the</strong>re is an increase in <strong>the</strong> absorption intensity at<br />

230 run, giving a peak with an extinction coefficient <strong>of</strong> 11,600 dm 3 cm-1 moF'<br />

16 On <strong>the</strong> addition <strong>of</strong> an amino group to <strong>the</strong> benzene ring, to form ani!ine,<br />

<strong>the</strong> new absorption peak is also around 280 nm 16, but <strong>the</strong> extinction<br />

coefficient is approximately twice that <strong>of</strong> benzoic acid (E 1,850 dm3 cm-1mo! -1 );<br />

Tab!e 1.2. When a N,N-di<strong>methyl</strong>amino group is substituted into <strong>the</strong> ring <strong>the</strong><br />

absorption spectrum is fur<strong>the</strong>r red shifted to 300 nm <strong>and</strong> <strong>the</strong> extinction<br />

coefficient becomes even larger (8 2,400 mo! -1 dm3 cm-1). One point <strong>of</strong> interest<br />

for aniline is that in cyclohexane <strong>the</strong> absorption spectrum is structured whilst<br />

in ethanol a c!ean structureless absorption b<strong>and</strong> is seen 16<br />

Molecule / Species Solvent Transition ?max<br />

chromophore<br />

(nm)<br />

Benzene neutra! none 256 220<br />

200 6,300<br />

180 100,000<br />

Benzoic acid neutral Inone 280 1,000<br />

230 11,600<br />

Ani!ine neutra! water —.g* 280 1,850<br />

230 8,600<br />

cation acid (aq) t—nr* 254 160<br />

203 7,500<br />

8<br />

dm 3 moicnr1<br />

Table 1.2<br />

Examples <strong>of</strong> transitions for various organic molecules <strong>and</strong> <strong>the</strong>ir<br />

extinction coefficients<br />

Wepster 17 showed that in <strong>the</strong> case <strong>of</strong> both aniline <strong>and</strong> N,Ndimethy!ani!ine<br />

that <strong>the</strong> !one pair <strong>of</strong> e!ectrons were conjugated with <strong>the</strong><br />

benzene it-system, but that <strong>the</strong> H or Me groups were "tilled', at an angle <strong>of</strong><br />

approximately 30 0 re!ative to <strong>the</strong> plane <strong>of</strong> <strong>the</strong> ring, Figure 1.8.<br />

17


O ;R<br />

Figure 1.8<br />

Diagram representing <strong>the</strong> "tilt" angle between substituents on<br />

<strong>the</strong> aniline nitrogen atom <strong>and</strong> <strong>the</strong> plane <strong>of</strong> <strong>the</strong> n-system<br />

The presence <strong>of</strong> two substituents in <strong>the</strong> benzene ring raises <strong>the</strong> twin<br />

possibilities <strong>of</strong> electronic <strong>and</strong> steric interactions between <strong>the</strong> two substituents.<br />

Fur<strong>the</strong>r work by Wepster 18 in <strong>the</strong> late 50's dealt with <strong>the</strong> steric effects<br />

observed in di- <strong>and</strong> higher- substituted benzene derivatives. Although much<br />

<strong>of</strong> his work was involved with ortho- substituted derivatives he also worked<br />

on meta- <strong>and</strong> para- disubstituted benzenes. The potential steric effect <strong>of</strong><br />

substituents in <strong>the</strong> 2- or 2,6- positions <strong>of</strong> molecules such as aniline is to cause<br />

<strong>the</strong> -NR2 group (Figure 1.8) to rotate about its bond to <strong>the</strong> ring <strong>and</strong> <strong>the</strong>reby<br />

twist <strong>the</strong> nitrogen lone pair out <strong>of</strong> conjugation with <strong>the</strong> n-system <strong>of</strong> <strong>the</strong> ring.<br />

Wepster 19 found that even for aniline, with bulky t-butyl groups attached to<br />

<strong>the</strong> 2,6- positions, <strong>the</strong> NH2 group was only twisted slightly out <strong>of</strong> conjugation<br />

with <strong>the</strong> ring n-system (approximately 30 0 as measured by a 50% decrease in<br />

<strong>the</strong> extinction coefficient <strong>of</strong> <strong>the</strong> 1 A - 1Lb transition). By contrast, <strong>the</strong><br />

extinction coefficient <strong>of</strong> <strong>the</strong> corresponding N,N-di<strong>methyl</strong> compound was<br />

reduced to approximately 1/20th <strong>of</strong> that <strong>of</strong> <strong>the</strong> N,N-di<strong>methyl</strong>aniline<br />

indicating, on <strong>the</strong> basis <strong>of</strong> equation 1.15a (where c, is <strong>the</strong> extinction coefficient<br />

<strong>of</strong> <strong>the</strong> unsubstituted aniline <strong>and</strong> 8 <strong>the</strong> extinction coefficient <strong>of</strong> <strong>the</strong> ring<br />

substituted compound), a twist angle (p) <strong>of</strong>> 750<br />

2/60 = cos 2 ( l.15a<br />

The presence <strong>of</strong> an electron withdrawing group in <strong>the</strong> 4-position<br />

relative to <strong>the</strong> amine helps to reduce <strong>the</strong> twist angle caused by <strong>the</strong> presence <strong>of</strong><br />

a particular 2- or 2,6- substituent 20, so that p for amine > (p for 4-ester > (p for<br />

iLJ


4-NO2. The absorption properties <strong>of</strong> <strong>the</strong> <strong>aminobenzoic</strong> <strong>acids</strong> are discussed in<br />

more detail in section 1.10.<br />

1.6 Dissipative pathways<br />

Having excited an electron into an orbital <strong>of</strong> higher energy, producing<br />

an excited state, <strong>the</strong> molecule has a number <strong>of</strong> possibilities open to it to<br />

dissipate <strong>the</strong> extra energy. These include vibrational relaxation, radiative<br />

transitions (fluorescence, delayed fluorescence <strong>and</strong> phosphorescence),<br />

radiationless transitions (internal conversion <strong>and</strong> intersystem crossing),<br />

energy transfer <strong>and</strong> chemical reaction.<br />

These processes are not mutually exclusive <strong>and</strong> <strong>the</strong>y compete with<br />

each o<strong>the</strong>r for <strong>the</strong> deactivation <strong>of</strong> an excited state. Vibrational relaxation <strong>of</strong>ten<br />

precedes <strong>the</strong> o<strong>the</strong>rs. They are portrayed in <strong>the</strong> Jablonski diagram shown in<br />

Figure 1.8. The radiative transitions are portrayed by straight lines <strong>and</strong> <strong>the</strong><br />

non-radiative transitions by jagged lines. S n <strong>and</strong> Tn refer to <strong>the</strong> nth excited<br />

singlet <strong>and</strong> triplet states, where n is an integer, except for % which is <strong>the</strong><br />

ground state.<br />

S3<br />

H<br />

Internal<br />

conversion<br />

Internal<br />

conversion<br />

SI<br />

Fluorescenc'<br />

T 1<br />

Absorption<br />

Ii<br />

coi<br />

S o<br />

Figure 1.8 A simplified Jablouski diagram.<br />

19


1.6.1 Vibrational Relaxation<br />

The energy <strong>of</strong> <strong>the</strong> absorbed photon will not <strong>of</strong>ten match exactly <strong>the</strong><br />

energy difference between <strong>the</strong> lowest vibrational level <strong>of</strong> <strong>the</strong> ground <strong>and</strong><br />

excited states <strong>of</strong> <strong>the</strong> absorbing molecule. Therefore <strong>the</strong> state initially produced<br />

is an upper vibrational / rotational state <strong>of</strong> <strong>the</strong> electronically excited<br />

molecule. The molecule loses excess vibrational energy through vibrational<br />

relaxation until <strong>the</strong> zeroth vibrational level <strong>of</strong> <strong>the</strong> excited state is reached.<br />

The loss <strong>of</strong> this excess vibrational (<strong>and</strong>/or rotational) energy is largely<br />

dependent on collisions, as a result <strong>of</strong> which <strong>the</strong> vibrational energy is<br />

converted into kinetic energy distributed between <strong>the</strong> partners in <strong>the</strong><br />

collision.<br />

1.6.2 Radiative transitions<br />

In <strong>the</strong> radiative transitions represented by straight lines on <strong>the</strong><br />

Jablonski diagram, an excited species passes from a higher energy state to one<br />

<strong>of</strong> lower energy with emission <strong>of</strong> a photon.<br />

a) Fluorescence is caused by a radiative transition between states <strong>of</strong> <strong>the</strong><br />

same multiplicity, although for polyatomic molecules <strong>the</strong> transition is<br />

usually S1—+S0. S2—*S0 fluorescence is occasionally observed but it is usually<br />

hidden by <strong>the</strong> far more intense Si—*So transition. Azulene fluoresces from 52<br />

<strong>and</strong> only negligibly from S 1 21 . In many cases <strong>the</strong> absorption <strong>and</strong> fluorescence<br />

spectra are mirror images when plotted on a frequency scale. The existence <strong>of</strong><br />

this mirror image relationship implies that only minor changes occur in <strong>the</strong><br />

molecular structure <strong>of</strong> <strong>the</strong> compound during excitation. However <strong>the</strong> solvent<br />

<strong>and</strong> substituent groups affect <strong>the</strong> position <strong>and</strong> structure <strong>of</strong> <strong>the</strong> spectrum.<br />

20


) Phosphorescence is <strong>the</strong> emission <strong>of</strong> radiation accompanying <strong>the</strong><br />

deactivation <strong>of</strong> an excited species to a lower state <strong>of</strong> different multiplicity e.g.<br />

T1—*S0. Rate constants are always smaller than those <strong>of</strong> <strong>the</strong> corresponding<br />

fluorescence transitions because <strong>the</strong> transition is spin forbidden.<br />

c) Delayed fluorescence, in which <strong>the</strong> luminescence decays more slowly<br />

than expected, can arise by several mechanisms <strong>of</strong> which triplet-triplet<br />

annihilation22 <strong>and</strong> <strong>the</strong>rmal activation 23 have been studied most closely.<br />

A measure <strong>of</strong> <strong>the</strong> efficiency <strong>of</strong> a photochemical process such as<br />

fluorescence is <strong>of</strong> fundamental value. The concept <strong>of</strong> a quantum yield (4))<br />

where<br />

- Number <strong>of</strong> molecules undergoing a process<br />

- Number <strong>of</strong> photons absorbed<br />

1.16<br />

can give valuable information on <strong>the</strong> nature <strong>of</strong> <strong>the</strong> process. Information can<br />

be ga<strong>the</strong>red from <strong>the</strong> magnitude <strong>of</strong> this quantity <strong>and</strong> <strong>the</strong> effects that<br />

temperature, pressure, concentration, substituent groups <strong>and</strong> solvents have<br />

on it. The quantum yield for formation <strong>of</strong> a product may not be <strong>the</strong> same as<br />

<strong>the</strong> quantum yield for <strong>the</strong> decomposition <strong>of</strong> starting material, depending on<br />

<strong>the</strong> complexity <strong>of</strong> <strong>the</strong> reaction.<br />

The fluorescence intensity F per unit volume is proportional to <strong>the</strong><br />

fractional light absorption Taper unit volume per unit time<br />

F41 1a 1.17<br />

where Ia=I0(1-e 23 ')<br />

1.18<br />

21


For dilute <strong>solution</strong>s <strong>the</strong> term inside <strong>the</strong> brackets can be exp<strong>and</strong>ed <strong>and</strong> if<br />

identical experimental conditions are maintained <strong>the</strong>n fluorescence<br />

intensities <strong>and</strong> concentrations are linearly related according to <strong>the</strong> expression<br />

1.19<br />

F=4f 2.30310 Ed 1.19<br />

On <strong>the</strong> o<strong>the</strong>r h<strong>and</strong>, at high concentrations where all <strong>the</strong> incident light<br />

is completely absorbed, <strong>the</strong> exponential term in equation 1.18 becomes<br />

negligible <strong>and</strong> consequently:<br />

F=4 1 10 1.20<br />

The fluorescence intensity should remain constant for any fur<strong>the</strong>r increase in<br />

concentration, but for many compounds <strong>the</strong> intensity starts decreasing . This<br />

decrease in fluorescence is due to <strong>the</strong> quenching <strong>of</strong> fluorescence by molecules<br />

<strong>of</strong> <strong>the</strong> same kind (concentration quenching) 14 .<br />

1.6.3 Non-radiative transitions<br />

Non-radiative transitions occur between degenerate vibrationalrotational<br />

levels <strong>of</strong> different electronic states. There is no change in <strong>the</strong><br />

energy <strong>of</strong> <strong>the</strong> system <strong>and</strong> consequently no photon is emitted.<br />

a) Internal conversion is a non-radiative transition between degenerate<br />

states <strong>of</strong> <strong>the</strong> same multiplicity. These transitions are <strong>of</strong>ten extremely fast.<br />

b) Intersystem crossing is a non-radiative transition between two states<br />

<strong>of</strong> different multiplicity which are not necessarily degenerate. The<br />

intersystem crossing S1—,T1, which competes with fluorescence, is <strong>the</strong> process<br />

by which triplet states are normally populated. One mechanism <strong>of</strong> delayed<br />

22


fluorescence is when <strong>the</strong> electron undergoes a T1—*S1 transition with <strong>the</strong> help<br />

<strong>of</strong> <strong>the</strong>rmal activation.<br />

1.6.4 Chemical reaction.<br />

The chemistry <strong>of</strong> an excited species usually differs from that <strong>of</strong> <strong>the</strong><br />

ground state species as a result <strong>of</strong> <strong>the</strong> excess energy carried by <strong>the</strong> excited<br />

species <strong>and</strong> also as a result <strong>of</strong> <strong>the</strong> different electronic arrangement <strong>of</strong> <strong>the</strong><br />

excited state. Chemical reactions can occur to yield a product or products <strong>of</strong><br />

lower energy.<br />

1.7 Quenching<br />

A substance which accelerates <strong>the</strong> decay <strong>of</strong> an electronically excited state<br />

is described as a quencher <strong>and</strong> is said to quench <strong>the</strong> state. Some possible<br />

quenching routes are shown in figure 1.10.<br />

M<br />

.<br />

Photochemical<br />

quenching<br />

Photophysical<br />

quenching<br />

Products<br />

Figure 1.9 Routes for possible quenching mechanisms<br />

M<br />

Self quenching<br />

Q Impurity ____<br />

guenching<br />

Electron<br />

transfer<br />

quenching<br />

quenching<br />

Energy<br />

transfer<br />

1.7.1 Photophysical quenching<br />

Unlike photochemical quenching, photophysical quenching does not<br />

lead to new ground state products. A number <strong>of</strong> different cases can be<br />

distinguished.<br />

i) Excimer or Exciplex formation. (Excited Dimer), (Excited complex)<br />

23


On increasing <strong>the</strong> concentration <strong>of</strong> a solute, <strong>the</strong> fluorescence intensity is<br />

decreased. This is caused by concentration quenching which also usually<br />

results in a new emission at a longer wavelength <strong>and</strong> can be explained in<br />

terms <strong>of</strong> excimer or exciplex formation. Such complexes exist only in <strong>the</strong><br />

excited state, being fully dissociated, <strong>and</strong> <strong>the</strong>refore undetectable, in <strong>the</strong> ground<br />

state. Excimers are formed between identical molecules one <strong>of</strong> which is in <strong>the</strong><br />

excited state <strong>and</strong> one in <strong>the</strong> ground state. Excimers are non-polar, while<br />

exciplexes, which are formed between an excited species <strong>and</strong> one or more<br />

different ground state molecules, are polar entities. They have large dipole<br />

moments as a result <strong>of</strong> some degree <strong>of</strong> electron transfer between <strong>the</strong> two<br />

components <strong>of</strong> <strong>the</strong> exciplex. In <strong>the</strong> case <strong>of</strong> organic molecules, excimer<br />

formation was first identified by Forster <strong>and</strong> Kasper 24 for pyrene in <strong>solution</strong>.<br />

Excimer <strong>and</strong> exciplex formation are <strong>of</strong>ten rapid diffusion controlled reactions<br />

<strong>and</strong> as a consequence <strong>of</strong> being diffusion controlled are concentration<br />

dependent. Care must be taken not to confuse <strong>the</strong>se two species with excited<br />

dimers or excited complexes which are formed in <strong>the</strong> ground state <strong>and</strong> <strong>the</strong>n<br />

absorb a photon <strong>of</strong> light.<br />

ii) Iwisted Intra-molecular Charge Iransfer (TJCT) states.<br />

Stable excited states can occur in molecules in which two<br />

chromophores, one a donor <strong>and</strong> one an acceptor, are separated by one<br />

(twisted) single bond. These have been used to explain anomalous<br />

fluorescence observed in various compounds. The basic <strong>the</strong>ory was<br />

developed by Lippert 25 <strong>and</strong> later refined by Grabowski <strong>and</strong> Rettig <strong>and</strong> coworkers26'<br />

27 In compounds with a near linear planar ground state<br />

conformation, <strong>the</strong> Franck-Condon excited state already possesses considerable<br />

charge transfer characteristics because <strong>of</strong> <strong>the</strong> mesomeric interaction <strong>of</strong> donor<br />

<strong>and</strong> acceptor groups. Rettig 28 called this state <strong>the</strong> delocalised excited state (DE).<br />

In <strong>the</strong> perpendicular conformation, this mesomeric interaction can no longer<br />

24


take place <strong>and</strong> <strong>the</strong> available excited states are ei<strong>the</strong>r different TICT states with<br />

electron transfer from <strong>the</strong> various donor orbitals to <strong>the</strong> various 1t*orbitals <strong>of</strong><br />

<strong>the</strong> acceptor, or locally excited (LE) states within <strong>the</strong> donor <strong>and</strong> acceptor subunits.<br />

According to Rettig27, <strong>the</strong> TICT state can be populated spontaneously if<br />

<strong>the</strong> transition DE -, TICT is exo<strong>the</strong>rmic (or only slightly endo<strong>the</strong>rmic) or <strong>the</strong><br />

activation barrier is low enough such that this transition can occur within<br />

<strong>the</strong> lifetime <strong>of</strong> <strong>the</strong> excited state. Freezing <strong>the</strong> <strong>solution</strong> increases <strong>the</strong> effective<br />

activation barrier such that in many cases TICT state formation is stopped. In<br />

<strong>the</strong> case <strong>of</strong> DMABN (<strong>the</strong> first molecule to be found to exhibit TICT behaviour)<br />

Rettig 27 assumed that <strong>the</strong> molecule was largely rigid, with only <strong>the</strong><br />

di<strong>methyl</strong>amino group twisting out <strong>of</strong> plane, <strong>and</strong> that <strong>the</strong> local viscosity in <strong>the</strong><br />

solvent shell surrounding <strong>the</strong> solute is not drastically altered by dipolar<br />

interactions. Rettig drew his conclusions for this from experimental<br />

evidence 29 where it was found that polar molecules exhibit <strong>the</strong> same<br />

reorientation relaxation time in polar or non-polar solvents. Work is<br />

currently underway to try <strong>and</strong> quantify <strong>the</strong> effects <strong>of</strong> excitation on molecular<br />

dipole moments30 .<br />

iii) Heavy atom quenching<br />

Molecular fluorescence is quenched by <strong>the</strong> presence <strong>of</strong> species<br />

containing heavy atoms. The phenomenon <strong>of</strong>ten arises through <strong>the</strong><br />

formation <strong>of</strong> a singlet exciplex which, because <strong>of</strong> <strong>the</strong> heavy atom effect,<br />

undergoes erthanced intersystem crossing to <strong>the</strong> triplet exciplex, which <strong>the</strong>n<br />

dissociates into its component parts. Thus <strong>solution</strong>s <strong>of</strong> anthracene <strong>and</strong> some<br />

<strong>of</strong> its derivatives become less fluorescent on addition <strong>of</strong> bromobenzene, while<br />

<strong>the</strong> triplet-triplet absorption intensity increases as a result <strong>of</strong> enhanced singlet<br />

to triplet inter-system crossing 31 .<br />

25


iv) Electronic energy transfer<br />

In this process, an excited donor molecule D* collapses to its ground<br />

state with <strong>the</strong> simultaneous transfer <strong>of</strong> its electronic excitation energy to an<br />

acceptor molecule A which is promoted to an excited state.<br />

D * +A—*A * +D 1.21<br />

The emission <strong>of</strong> D* is <strong>the</strong>refore quenched <strong>and</strong> replaced by <strong>the</strong><br />

characteristics <strong>of</strong> A*. The process may involve <strong>the</strong> emission <strong>of</strong> a photon by D*<br />

which is re-absorbed by A (radiative transfer) or a non-radiative process.<br />

Radiative transitions require <strong>the</strong> acceptor to absorb <strong>the</strong> photons<br />

emitted by <strong>the</strong> donor. The probability (rate) <strong>of</strong> transfer will depend upon <strong>the</strong><br />

quantum efficiency <strong>of</strong> emission <strong>of</strong> <strong>the</strong> donor, <strong>the</strong> number <strong>of</strong> acceptor<br />

molecules in <strong>the</strong> path <strong>of</strong> <strong>the</strong> emitted photon <strong>and</strong> <strong>the</strong> extent <strong>of</strong> overlap<br />

between <strong>the</strong> emission spectrum <strong>of</strong> <strong>the</strong> donor <strong>and</strong> <strong>the</strong> absorption spectrum <strong>of</strong><br />

<strong>the</strong> acceptor. Radiative energy transfer can occur over extremely long ranges<br />

as for instance photochemical changes occurring on Earth under <strong>the</strong><br />

influence <strong>of</strong> sunlight.<br />

Non-radiative energy transfer occurs through <strong>the</strong> interaction between<br />

<strong>the</strong> donor <strong>and</strong> acceptor molecules (no matter how small). Two <strong>the</strong>ories have<br />

been put forward. The coulombic 32 interaction can be expressed as ahumber<br />

<strong>of</strong> terms, dipole-dipole, dipole-quadrupole, multipole-multipole <strong>of</strong> decreasing<br />

significance. The contribution <strong>of</strong> <strong>the</strong> normally dominant dipole-dipole<br />

interaction to energy transfer has been studied in depth by F6rster 24. The<br />

distance between <strong>the</strong> excited donor molecule <strong>and</strong> <strong>the</strong> ground state acceptor<br />

molecule at which energy transfer to A <strong>and</strong> internal deactivation <strong>of</strong> D* are<br />

equally probable is known as <strong>the</strong> critical transfer distance. Forster has<br />

calculated that energy transfer by dipole-dipole interaction can be significant<br />

even at distances <strong>of</strong> <strong>the</strong> order <strong>of</strong> 10 nm. The only transfer processes allowed by<br />

<strong>the</strong> coulombic interaction are those in which <strong>the</strong>re is no change in spin. The<br />

26


ate constants for multipole-multipole <strong>and</strong> dipole-quadrupole interactions<br />

fall <strong>of</strong>f with distance much more rapidly than that for dipole-dipole transfer<br />

<strong>and</strong> <strong>the</strong>y are <strong>the</strong>refore only likely to be important at distances less than 4 run.<br />

The electron exchange 33 mechanism is an extremely short range<br />

process, as it involves <strong>the</strong> overlap <strong>of</strong> <strong>the</strong> orbitals to enable electrons to<br />

transfer. The energy transfer can be formulated thus:<br />

122<br />

Due to <strong>the</strong> intervention <strong>of</strong> a bimolecular intermediate, which could be<br />

an exciplex or a collision complex, <strong>the</strong> energy transfer will be subject to<br />

conservation <strong>of</strong> electron spin. Under Wigner's spin rules 34, both <strong>the</strong><br />

following transitions are occasionally allowed.<br />

3D* + 1A_* 1D + 3A* 1.23<br />

1D* + IA.....> 1D + 1A* 1.24<br />

The former (triplet-triplet energy transfer), though allowed under <strong>the</strong><br />

exchange interaction, is doubly forbidden under <strong>the</strong> coulombic dipole-dipole<br />

interaction, whilst singlet-singlet energy transfer is allowed under both<br />

interactions.<br />

1.8 Kinetics <strong>of</strong> fluorescence<br />

Consider <strong>the</strong> following scheme (Figure 1.10) where I is <strong>the</strong> rate <strong>of</strong><br />

absorption <strong>of</strong> photons, K 1 <strong>and</strong> K are <strong>the</strong> rate constants for fluorescence <strong>and</strong><br />

phosphorescence respectively <strong>and</strong> K ic <strong>and</strong> K1 are <strong>the</strong> rate constants for<br />

internal conversion <strong>and</strong> inter-system crossing.<br />

27


so<br />

hij(I)<br />

rS i<br />

____ lKiw<br />

Kic<br />

4<br />

S0 + hty'<br />

Figure 1.10 Diagram representing <strong>the</strong> processes involved in populating <strong>and</strong><br />

depopulating excited states<br />

Under continuous irradiation, <strong>the</strong> steady state approximation is<br />

applicable to excited states <strong>and</strong> it follows that <strong>the</strong> rates <strong>of</strong> formation <strong>and</strong><br />

destruction <strong>of</strong> Si are equal, i.e<br />

where<br />

1= [Si] E1K<br />

= Kk + Kf + isc<br />

1.25<br />

1.26<br />

The quantum yield <strong>of</strong> fluorescence is given by<br />

Of<br />

- Rate <strong>of</strong> emission by S 1 - K 1 [S 1] - K 1<br />

- Rate <strong>of</strong> absorption <strong>of</strong> photons by S - I -<br />

1.27<br />

If tf is <strong>the</strong> actual measured lifetime (t1 = 1/ 1K) <strong>the</strong>n<br />

4f = K't 1.28<br />

Given <strong>the</strong> definition <strong>of</strong> radiative lifetime; T o = 1/K1 where 'C 0 is <strong>the</strong> time<br />

taken for <strong>the</strong> population to diminish to lie <strong>of</strong> its initial concentration<br />

assuming that no o<strong>the</strong>r non-radiative processes are occurring, <strong>the</strong><br />

relationship between radiative lifetime <strong>and</strong> <strong>the</strong> actual measured lifetime <strong>of</strong> Si<br />

are given by <strong>the</strong> expression<br />

1f t4 1.29<br />

13


1.8.1 The Stern-Volmer equation<br />

The kinetics <strong>of</strong> fluorescence quenching may be treated by <strong>the</strong> use <strong>of</strong> <strong>the</strong><br />

Stern-Volmer equation. Consider a fluorescing excited species which also<br />

decays through a non-radiative route or quenching with a rate constant K q.<br />

Applying <strong>the</strong> steady state approximation to <strong>the</strong> concentration <strong>of</strong> S<br />

gives<br />

I = [Si].(K1 + IC4 + K q[QI) 1.30<br />

where Kd = K1 + 1Kisc <strong>and</strong> [Q] is <strong>the</strong> concentration <strong>of</strong> <strong>the</strong> quencher.<br />

Therefore<br />

Of = K1[Si]/I = K f / (Kf+Kd+Kq[Q]) 1.31<br />

The quantum yield <strong>of</strong> fluorescence in <strong>the</strong> absence <strong>of</strong> quencher (4) is given by<br />

Hence<br />

= K1 /(K1+K4) 1.32<br />

lKf +Kd +Kq [Q] Kq[Q] =1+K q +t[Q] .......... 1 33<br />

K1+K4<br />

Of<br />

This is <strong>the</strong> Stern-Volmer relationship, where tis <strong>the</strong> fluorescence lifetime<br />

in <strong>the</strong> absence <strong>of</strong> quencher.<br />

Under ideal conditions, <strong>the</strong> plot <strong>of</strong> $7 / o f against<br />

quencher<br />

concentration gives a straight line <strong>of</strong> gradient K 'tv. i t is known, <strong>the</strong><br />

quenching rate constant may be obtained.<br />

For many systems, Kq in mobile solvents is <strong>of</strong> <strong>the</strong> order <strong>of</strong> 10 9_10 1 dm3<br />

mol -1 sec-1 which is close to <strong>the</strong> diffusion controlled rate constant Kdiff. This<br />

suggests that in <strong>the</strong>se cases quenching is so rapid that <strong>the</strong> rate determining<br />

step is <strong>the</strong> diffusion <strong>of</strong> <strong>the</strong> quencher to within <strong>the</strong> active sphere <strong>of</strong> S1.


1.9 Acid-base properties<br />

Both ground state absorption spectra <strong>and</strong> excited state emission spectra<br />

are very much pH dependent. For <strong>the</strong> aromatic amino<strong>acids</strong>, <strong>the</strong>re is a<br />

possibility <strong>of</strong> absorption I emission from <strong>the</strong> anion, neutral <strong>and</strong> cation<br />

species. The ground state pK a <strong>and</strong> pKi, values can be calculated easily by<br />

varying <strong>the</strong> pH <strong>of</strong> <strong>the</strong> <strong>solution</strong> <strong>and</strong> measuring <strong>the</strong> absorption spectra.<br />

Measurement <strong>of</strong> <strong>the</strong> excited state pK a values is somewhat more complicated.<br />

Excited state pK values are usually very different from those <strong>of</strong> <strong>the</strong> ground<br />

state 35. This phenomenon was explained by Förster who postulated that<br />

proton exchanges are so rapid that an acid-base equilibrium is established<br />

between <strong>the</strong> excited molecule <strong>and</strong> its conjugate base during <strong>the</strong> lifetime <strong>of</strong><br />

<strong>the</strong> excited state species.<br />

HA<br />

ho) HA (_)<br />

1-<br />

HA+hi<br />

A+htf'<br />

Figure 1.11 Protonation <strong>and</strong> deprotonation reaction scheme<br />

In general, excited state molecules are stronger <strong>acids</strong> than <strong>the</strong><br />

corresponding ground state molecules. Two methods have been employed to<br />

measure <strong>the</strong> excited state pKa values <strong>of</strong> excited singlet states. Both methods<br />

assume proton equilibrium during <strong>the</strong> lifetime <strong>of</strong> <strong>the</strong> excited state.<br />

1) The p1Ca* is that pH at which <strong>the</strong> fluorescence <strong>of</strong> HA* drops to one half<br />

<strong>of</strong> <strong>the</strong> intensity observed in <strong>solution</strong>s so strongly acid that only HA* emits. It<br />

is also <strong>the</strong> pH at which <strong>the</strong> intensity <strong>of</strong> <strong>the</strong> fluorescence <strong>of</strong> A* drops to one<br />

half <strong>of</strong> <strong>the</strong> value observed in <strong>solution</strong>s so alkaline that only A* emits.<br />

GTXI


2) This method depends on <strong>the</strong> Forster-Weller cycle.<br />

*<br />

-I<br />

(S1 )i-<br />

(SD)<br />

S0 )<br />

Figure 1.12 Forster-Weller cycle for deriving pK values for excited states<br />

From <strong>the</strong> diagram it can be seen that<br />

AEHA-AEA.. = AH - = AG - AG* 1.34<br />

assuming that <strong>the</strong> entropy <strong>of</strong> dissociation is <strong>the</strong> same in both ground <strong>and</strong><br />

excited state. It follows that<br />

- AE A.. =<br />

LnI.-') =<br />

LK) RT kT<br />

1.35<br />

where K* <strong>and</strong> K are <strong>the</strong> dissociation constants for IIA* <strong>and</strong> HA, <strong>and</strong> bsv is <strong>the</strong><br />

difference in <strong>the</strong> frequency <strong>of</strong> absorption or emission <strong>of</strong> HA <strong>and</strong> A - .<br />

Knowledge <strong>of</strong> Av <strong>and</strong> K <strong>the</strong>refore allows calculation <strong>of</strong> K*<br />

1.10. Literature background to <strong>the</strong> project<br />

There have been several sporadic studies <strong>of</strong> <strong>the</strong> absorption <strong>and</strong><br />

fluorescence properties <strong>of</strong> <strong>the</strong> <strong>isomeric</strong> <strong>aminobenzoic</strong> <strong>acids</strong>, with <strong>the</strong> earliest<br />

reported experiments those <strong>of</strong> Doub <strong>and</strong> V<strong>and</strong>enbelt 36 in 1947, who studied<br />

<strong>the</strong> ultraviolet absorption characteristics <strong>of</strong> monosubstituted <strong>and</strong><br />

p-disubstituted benzene derivatives. Included in <strong>the</strong>ir studies were benzene,<br />

31


enzoic add, aniline <strong>and</strong> <strong>the</strong> protic species <strong>of</strong> 2-<strong>aminobenzoic</strong> acid. Doub <strong>and</strong><br />

V<strong>and</strong>enbelt36 were initially interested in <strong>the</strong> effect <strong>of</strong> substituent groups on<br />

<strong>the</strong> wavelength <strong>of</strong> maximum absorption <strong>and</strong> <strong>the</strong> appearance <strong>of</strong> a secondary,<br />

<strong>and</strong> in some cases, a tertiary absorption b<strong>and</strong>. They found that in <strong>the</strong> case <strong>of</strong><br />

ortho- or pa Ta- directing substituent groups that <strong>the</strong> absorption peak was red<br />

shifted with <strong>the</strong> increasing ability <strong>of</strong> <strong>the</strong> substituent group to donate electrons.<br />

In 1949, <strong>the</strong>y extended <strong>the</strong>ir studies to include a more extensive examination<br />

<strong>of</strong> <strong>the</strong> ortho- <strong>and</strong> meta- disubstituted benzenes 37 <strong>and</strong> studied <strong>the</strong> ground state<br />

absorption properties <strong>of</strong> <strong>the</strong>se molecules as a function <strong>of</strong> pH <strong>and</strong> solvent as<br />

well as considering substituent effects. In 1955, Doub <strong>and</strong> V<strong>and</strong>enbelt 38<br />

extended <strong>the</strong> range <strong>of</strong> solvents <strong>and</strong> substituent groups used, as well as<br />

investigating <strong>the</strong> effects <strong>of</strong> a third substituent group on <strong>the</strong> ground state<br />

absorption spectra. The first reported fluorescence spectra for <strong>the</strong>se<br />

compounds came from Melhuish 39, who reported <strong>the</strong> fluorescence quantum<br />

yields <strong>of</strong> 2-<strong>aminobenzoic</strong> acid <strong>and</strong> <strong>methyl</strong> 2-aminobenzoate in ethanol <strong>and</strong><br />

investigated oxygen quenching effects. The quantum yield <strong>of</strong> 2-<strong>aminobenzoic</strong><br />

acid was reported as 0.56 I 0.59 <strong>and</strong> for <strong>methyl</strong> 2-aminobenzoate 0.66 I 0.69 in<br />

<strong>the</strong> presence <strong>and</strong> absence <strong>of</strong> oxygen respectively. However <strong>the</strong> concentrations<br />

used for this <strong>study</strong> were in <strong>the</strong> order <strong>of</strong> 10.2 mol dm-3, leading to problems<br />

with inner filter effects <strong>and</strong> dimerisation. In 1963, Mataga 40 extended <strong>the</strong><br />

studies <strong>of</strong> Doub <strong>and</strong> V<strong>and</strong>enbelt 36 by measuring both <strong>the</strong> absorption <strong>and</strong><br />

emission properties <strong>of</strong> <strong>the</strong> <strong>isomeric</strong> <strong>aminobenzoic</strong> <strong>acids</strong> as a function <strong>of</strong><br />

solvent, finding that <strong>the</strong> solvent effects upon <strong>the</strong> absorption spectra differed<br />

markedly from those on <strong>the</strong> fluorescence spectra. He plotted Stokes shifts<br />

against <strong>the</strong> solvent parameter F(D,n) (Equation 1.36), concluding that<br />

deviations from this relationship were due to strong intra- or inter-molecular<br />

hydrogen bonding which was dependent on solvent.<br />

( D+l n 2 +1<br />

F(D,n)=i + I 1.36<br />

2D+1 2n2 +1)<br />

where n is <strong>the</strong> refractive index <strong>and</strong> D is <strong>the</strong> dielectric constant <strong>of</strong> <strong>the</strong> solvent.<br />

32


COOH<br />

NH2<br />

COOH<br />

I<br />

NH3<br />

Kz<br />

NK,<br />

/Ka<br />

6~1~l r<br />

z<br />

Figure 1.13 Reaction scheme proposed by Leggate <strong>and</strong> Dunn 41 utilising<br />

nomenclature proposed by Tramer37<br />

In 1964, Leggate <strong>and</strong> Dunn 41 reported an extensive <strong>study</strong> <strong>of</strong> ortho<strong>aminobenzoic</strong><br />

<strong>acids</strong>, <strong>and</strong> used <strong>the</strong> Hammett equation to calculate <strong>the</strong><br />

equilibrium constant between <strong>the</strong> neutral form <strong>of</strong> 2-<strong>aminobenzoic</strong> acid <strong>and</strong><br />

its zwitterion structure, utilising <strong>the</strong> above scheme.<br />

Leggate <strong>and</strong> Dunn 41 concluded that <strong>the</strong> ionisation constants <strong>of</strong> 2-<br />

<strong>aminobenzoic</strong> acid gave an excellent fit to <strong>the</strong> two parameter Hammett<br />

equation, suggesting that this acid at its isoelectric point is better represented<br />

as a mixture <strong>of</strong> zwitterion <strong>and</strong> neutral species, than as a single hybrid species.<br />

Kuhn <strong>and</strong> Geider 43 calculated K z for 2-<strong>aminobenzoic</strong> acid (2ABA), 2-N<strong>methyl</strong><strong>aminobenzoic</strong><br />

acid (2MABA) <strong>and</strong> 2-N,N-di<strong>methyl</strong><strong>aminobenzoic</strong> acid<br />

(2DMABA). Complementing this earlier work, Tramer 42 utilised both ground<br />

state absorption <strong>and</strong> JR studies to deduce more information about <strong>the</strong><br />

zwitterion / neutral equilibrium, finding startling differences between 2ABA<br />

33


<strong>and</strong> 2DMABA. He showed that 2MABA dissolved in inert solvents behaves<br />

in an analogous manner to 2ABA <strong>and</strong> exists in <strong>the</strong> neutral molecular form<br />

(M), <strong>and</strong> that <strong>the</strong>re is no evidence for an intra-molecular hydrogen bond. The<br />

reverse is <strong>the</strong> case in polar solvents, in that <strong>the</strong> predominant form is <strong>the</strong><br />

zwitterion structure, assumed to be stabilised by solute/ solvent interactions.<br />

When <strong>the</strong> strong intra-molecular hydrogen bond is present in 2DMABA, <strong>the</strong><br />

conjugation <strong>of</strong> <strong>the</strong> nitrogen lone pair electrons with <strong>the</strong> it electron system <strong>of</strong><br />

<strong>the</strong> benzene ring is completely removed as shown by <strong>the</strong> extremely weak<br />

fluorescence spectrum observed in hexane. The difference between 2MABA<br />

<strong>and</strong> 2DMABA in terms <strong>of</strong> ability to form a hydrogen bonded complex was<br />

thought to be due to <strong>the</strong> enhancement <strong>of</strong> <strong>the</strong> basic properties <strong>of</strong> <strong>the</strong> amino<br />

group by <strong>the</strong> inductive release <strong>of</strong> <strong>the</strong> two <strong>methyl</strong> groups thus increasing <strong>the</strong><br />

strength <strong>of</strong> <strong>the</strong> hydrogen bond. Tramer appears not to have looked into <strong>the</strong><br />

major difference between 2DMABA <strong>and</strong> 2MABA, namely that 2DMABA will<br />

involve considerably more steric hindrance between <strong>the</strong> COOH <strong>and</strong> N(CH3)2<br />

substituents because <strong>of</strong> <strong>the</strong> bulky nature <strong>of</strong> <strong>the</strong> N,N-di<strong>methyl</strong>amino groups.<br />

Tramer 44 extended his studies <strong>of</strong> <strong>the</strong>se three compounds by measuring <strong>the</strong>ir<br />

fluorescence <strong>and</strong> phosphorescence emission spectra as a function <strong>of</strong> solvent<br />

<strong>and</strong> pH. Although <strong>the</strong>re is little new information in this 44 report, dimer<br />

formation was observed in more concentrated <strong>solution</strong>s <strong>of</strong> 2MABA in polar<br />

<strong>solution</strong>s. Tramer's calculated pKa values are in accordance with those<br />

reported by o<strong>the</strong>r authors. Tramer 42 <strong>and</strong> Kuhn <strong>and</strong> Geider 43 have calculated<br />

K Z constants in reasonably concentrated <strong>solution</strong>s (approximately 10.2 mol<br />

dm -3) a concentration which in most solvents will lead to dimerisation. All<br />

<strong>the</strong>ir calculations are based on <strong>the</strong> appearance <strong>of</strong> a peak at around 320 nm<br />

which has an extinction coefficient <strong>of</strong> 10 dm 3 mol -' cm', so it is not<br />

surprising that a similar <strong>study</strong> by Jian et al. 45 in 1986 made no mention <strong>of</strong><br />

this Z-M equilibrium. In a similar manner to Tramer, Jian et al.45 measured<br />

<strong>the</strong> absorption <strong>and</strong> emission spectra <strong>of</strong> <strong>the</strong> <strong>isomeric</strong> <strong>aminobenzoic</strong> <strong>acids</strong> in


different solvents <strong>and</strong> as a function <strong>of</strong> pH. This is <strong>the</strong> first reported <strong>study</strong> <strong>of</strong> 3-<br />

<strong>aminobenzoic</strong> acid. They also studied <strong>the</strong> acid base equlibria in <strong>the</strong> ground<br />

<strong>and</strong> excited states. Jian et al.45 appear not to have taken into account <strong>the</strong><br />

potential for hydrogen bonding. They do however include <strong>the</strong> results <strong>of</strong><br />

CNDO calculations, calculating <strong>the</strong> wavelength <strong>of</strong> maximum absorption <strong>and</strong><br />

<strong>the</strong> oscillator strength <strong>of</strong> <strong>the</strong> various types <strong>of</strong> transitions for all three <strong>isomeric</strong><br />

<strong>aminobenzoic</strong> <strong>acids</strong>, which appear to correspond well with <strong>the</strong> experimental<br />

values. In conclusion <strong>the</strong>y reported that in all three <strong>isomeric</strong> <strong>aminobenzoic</strong><br />

<strong>acids</strong> <strong>the</strong> lone pair on <strong>the</strong> nitrogen is in conjugation with <strong>the</strong> it electrons <strong>of</strong><br />

<strong>the</strong> aromatic ring system <strong>and</strong> this may lead to <strong>the</strong> energy <strong>of</strong> <strong>the</strong> ic—its<br />

transition being nearly equal to or slightly lower than <strong>the</strong> n1t* transition<br />

energy. That <strong>the</strong> observed absorption b<strong>and</strong> is a ic-it" transition is borne out by<br />

<strong>the</strong> high extinction coefficients <strong>and</strong> <strong>the</strong> observed solvent effect, namely that<br />

increasing solvent polarity <strong>and</strong> / or hydrogen bonding ability leads to a red<br />

shift in both <strong>the</strong> absorption <strong>and</strong> emission spectra.<br />

There have only been three publications regarding <strong>the</strong> fluorescence <strong>of</strong><br />

4DMABA, two by Cowley et al. 46,47 <strong>and</strong> <strong>the</strong> third by Brown <strong>and</strong> Revill 48 with<br />

only a h<strong>and</strong>ful <strong>of</strong> papers on <strong>the</strong> <strong>methyl</strong> <strong>esters</strong> <strong>of</strong> 4DMABA (M4DMAB <strong>and</strong><br />

E4DEAB) 47' The fluorescence properties <strong>of</strong> <strong>the</strong>se three compounds are<br />

anomalous in that <strong>the</strong>y show dual fluorescence, <strong>the</strong> source <strong>of</strong> which is still a<br />

matter <strong>of</strong> debate.<br />

1.10.1 4-Di<strong>methyl</strong>aminobenzonitrile<br />

(During this discussion <strong>and</strong> throughout <strong>the</strong> Results we will utilise <strong>the</strong><br />

accepted terminology for <strong>the</strong>se systems where <strong>the</strong> "normal" blue fluorescence<br />

is attributed to an excited state labelled b* <strong>and</strong> <strong>the</strong> red "anomalous"<br />

fluorescence to a state labelled a".)<br />

Since <strong>the</strong> discovery <strong>of</strong> <strong>the</strong> dual fluorescence <strong>of</strong> 4-(N, N -<br />

di<strong>methyl</strong>amino)benzonitrile (DMABN) in 1962 by Lippert et al. 25, several<br />

"lii


<strong>the</strong>ories have been proposed to explain <strong>the</strong> source <strong>of</strong> <strong>the</strong> anomalous<br />

fluorescence. Kosower <strong>and</strong> Dodiuk 55 proposed that <strong>the</strong> dual fluorescence was<br />

due to a species in which <strong>the</strong> nitrogen on <strong>the</strong> amino group was protonated,<br />

<strong>the</strong> proton being donated by <strong>the</strong> solvent. However this <strong>the</strong>ory can be almost<br />

immediately ruled out since in this latter paper it is shown that dual<br />

fluorescence occurs in both acetonitrile <strong>and</strong> butyronitrile, both <strong>of</strong> which are<br />

normally incapable <strong>of</strong> proton donation. Although <strong>the</strong>re has been no more<br />

evidence to support this <strong>the</strong>ory, Kosower <strong>and</strong> Dodiuk also identified o<strong>the</strong>r<br />

emissions as being due to a dimer <strong>and</strong> perpendicular <strong>and</strong> planar monomers.<br />

The idea <strong>of</strong> a perpendicular monomer (conformational isomer) was also<br />

proposed by Lippert 56 <strong>and</strong> Grabowski et al. 26 . Grabowski et al. 26 . used <strong>the</strong><br />

range <strong>of</strong> compounds shown in Figure 1.14 to argue for <strong>the</strong> acceptance <strong>of</strong> <strong>the</strong><br />

idea <strong>of</strong> conformational isomers being responsible for <strong>the</strong> anomalous<br />

fluorescence. This was <strong>the</strong> initial proposal <strong>of</strong> <strong>the</strong> idea <strong>of</strong> a Twisted Intramolecular<br />

Charge Transfer or TICT state.<br />

Me\ /Me<br />

N<br />

Me<br />

/<br />

N<br />

- -<br />

Me I-<br />

-fl<br />

• I<br />

• I<br />

• I<br />

I<br />

I<br />

I<br />

I<br />

C<br />

It'<br />

N<br />

C<br />

I''<br />

N<br />

C<br />

It'<br />

N<br />

•<br />

C<br />

• •<br />

N<br />

. I<br />

I<br />

ED<br />

Wi<br />

Iv<br />

Figure 1.14 Compounds utilised by Grabowski et a126 to argue <strong>the</strong> existence<br />

<strong>of</strong> <strong>the</strong> TICT phenomenon


Structure IV is a representation <strong>of</strong> <strong>the</strong> TICT state <strong>of</strong> DMABN<br />

(compound I) which shows two fluorescence b<strong>and</strong>s in polar solvents, while II<br />

shows only a single fluorescence b<strong>and</strong>, which closely resembles a typical b*<br />

emission. Compound III, which one might expect to show a similar<br />

fluorescence spectrum, actually shows a fluorescence b<strong>and</strong> representative <strong>of</strong><br />

<strong>the</strong> a* emission. There is however one major difference between compounds<br />

II <strong>and</strong> Ill, <strong>the</strong> orientation <strong>of</strong> <strong>the</strong> lone pair <strong>of</strong> <strong>the</strong> nitrogen. In II <strong>the</strong> orientation<br />

<strong>of</strong> <strong>the</strong> lone pair is parallel to <strong>the</strong> plane <strong>of</strong> <strong>the</strong> ring system <strong>and</strong> in III orthogonal<br />

to <strong>the</strong> plane <strong>of</strong> <strong>the</strong> ring system. Thus <strong>the</strong> structural evidence seems to suggest<br />

that <strong>the</strong> anomalous emitting state in I corresponds to a twisted rotamer lv.<br />

Although this appears to be evidence <strong>of</strong> TICT formation, it is only based upon<br />

data acquired by static experiments <strong>and</strong> dynamic evidence was absent27' 7.<br />

In 1985, Heisel et al. 58 reported that using fast laser spectroscopy, <strong>the</strong>y<br />

had found <strong>the</strong> kinetic relationship between a* <strong>and</strong> b* which was first<br />

predicted by Grabowski et al. 54. Their work agrees with <strong>the</strong> findings <strong>of</strong><br />

Huppert et al.60 who found pulse limited rise times in <strong>the</strong> emission <strong>of</strong> <strong>the</strong> a*<br />

state which feeds directly <strong>the</strong> emission for <strong>the</strong> b* state. The fluorescence<br />

kinetics <strong>of</strong> <strong>the</strong> a* <strong>and</strong> b* states were found to vary with solvent <strong>and</strong><br />

temperature, with <strong>the</strong> decay <strong>of</strong> <strong>the</strong> a* showing single exponential kinetics <strong>and</strong><br />

b* dual exponential decay kinetics. These findings were interpreted as<br />

suggesting ei<strong>the</strong>r <strong>the</strong>rmally assisted intersystem crossing to a solvated triplet<br />

or <strong>the</strong>rmally activated internal conversion to <strong>the</strong> ground state.<br />

Heisel et al.59 . also found a rise time <strong>of</strong> approximately 40 ps for <strong>the</strong> a*<br />

emission <strong>and</strong> <strong>the</strong> kinetics <strong>of</strong> <strong>the</strong> a* peak were elucidated by analysing <strong>the</strong><br />

kinetics <strong>of</strong> <strong>the</strong> b* emission. They found that one <strong>of</strong> <strong>the</strong> components <strong>of</strong> <strong>the</strong><br />

decay <strong>of</strong> <strong>the</strong> b* fluorescence was <strong>the</strong> same as <strong>the</strong> rise time <strong>of</strong> <strong>the</strong> a* peak,<br />

lending more pro<strong>of</strong> to <strong>the</strong> <strong>the</strong>ory that a* <strong>and</strong> b* are formed sequentially. They<br />

also' found that <strong>the</strong> b* fluorescing state was formed within 5 ps followed by<br />

<strong>the</strong> a* state within 40 ps. Their data agree with <strong>the</strong> proposal <strong>of</strong> Rotkiewicz et<br />

37


al.61 that <strong>the</strong> b state is <strong>the</strong> kinetic precursor <strong>of</strong> <strong>the</strong> a* state with kinetics<br />

which are found to be very viscosity <strong>and</strong> temperature dependent. Later in<br />

1985 Heisel et al. 59 calculated <strong>the</strong> rates <strong>of</strong> TICT formation. Their experimental<br />

evidence shows that <strong>the</strong> apparent activation energy for TICT formation is<br />

very solvent dependent, which is in agreement with <strong>the</strong> ideas <strong>of</strong> Rettig <strong>and</strong><br />

co-workers 2527. Kobayashi et al.62 <strong>and</strong> August et al.63 have investigated <strong>the</strong><br />

<strong>solution</strong> <strong>photophysics</strong> <strong>of</strong> DMABN using laser induced fluorescence (LIF) <strong>and</strong><br />

proposed that <strong>the</strong> a* fluorescence was produced by a Van der Waals complex<br />

with <strong>the</strong> resulting spectra being attributed to <strong>the</strong> internal rotation <strong>of</strong> <strong>the</strong> N,Ndi<strong>methyl</strong>amino<br />

group. They found no evidence for dimer or higher aggregate<br />

formation. August et al.&i also investigated <strong>the</strong> jet cooled LW fluorescence <strong>of</strong><br />

DMABN in <strong>the</strong> gas phase. They interpreted <strong>the</strong>ir results in a similar manner<br />

to Kobayashi et al.62' finding excellent agreement with <strong>the</strong> expected dynamics<br />

<strong>of</strong> TICT state formation.<br />

However <strong>the</strong>re still appears to be some confusion in <strong>the</strong> literature,<br />

with Mataga <strong>and</strong> Nakashima 64 in 1973 publishing results which indicated<br />

that <strong>the</strong> anomalous fluorescence was concentration dependent. They found<br />

that in toluene <strong>and</strong> diethyl e<strong>the</strong>r <strong>solution</strong>s, <strong>the</strong> ratio <strong>of</strong> <strong>the</strong> a* / b*<br />

fluorescence intensities for DMABN increased over a concentration range <strong>of</strong><br />

1x1O -S to 0.1 mol dm-3 <strong>and</strong> that <strong>the</strong> a* b<strong>and</strong> showed a red shift at very high<br />

concentrations. Mataga <strong>and</strong> Nakashima 64 do not propose any mechanism for<br />

this apparent concentration dependence. However in polar solvents like<br />

methanol <strong>and</strong> acetonitrile, <strong>the</strong> intensity ratio decreased with <strong>the</strong> a* b<strong>and</strong><br />

showing a slight blue shift at higher concentrations( i.e. dimer / excimer<br />

formation). This shift in <strong>the</strong> wavelength <strong>of</strong> maximum emission was<br />

concluded to result from a change in <strong>the</strong> dipolar interaction between <strong>the</strong><br />

excited state <strong>and</strong> <strong>the</strong> ground state, which is facilitated by higher<br />

concentrations. In all solvents studied no observable change could be seen in<br />

<strong>the</strong> ground state absorption spectra. The only solvent in which <strong>the</strong>re was no<br />

RN


change in <strong>the</strong> observed fluorescence intensity was propylene glycol, from<br />

concentrations <strong>of</strong> 1x10 -5 up to 5x10 -3 mol dm-3' which is in complete<br />

contradiction to <strong>the</strong> findings <strong>of</strong> Khalil et al. 65 However, Khalil's findings<br />

agree with <strong>the</strong> rest <strong>of</strong> Mataga's results. An additional peak observed by Khalil<br />

et al. at 340 nm was ruled out by Mataga who found this was caused by an<br />

impurity which disappeared on repeated recrystallisation from n-hexane.<br />

In 1981 Lippert et al. 56 reported that <strong>the</strong> dual fluorescence <strong>of</strong> DMABN<br />

at sufficiently low (but unstated) concentrations was not concentration<br />

dependent, but in 1-chlorobutane was found to be temperature dependent.<br />

Ch<strong>and</strong>ros 66 has proposed that <strong>the</strong> anomalous fluorescence is due to a 1:1<br />

complex comprising an excited state monomer with a solvent molecule (i.e.<br />

an exciplex), being stabilised by more conventional specific solvent - solute<br />

interactions. Ch<strong>and</strong>ros 66 observed that <strong>the</strong> "normal" fluorescence b<strong>and</strong><br />

observed in <strong>methyl</strong>cyclohexane followed a linear Stern-Volmer plot on being<br />

quenched by additions <strong>of</strong> triethylamine, <strong>and</strong> that, on addition <strong>of</strong> propionitrile<br />

to <strong>the</strong> <strong>methyl</strong>cyclohexane <strong>solution</strong>, <strong>the</strong> b* fluorescence diminished <strong>and</strong> <strong>the</strong> a*<br />

fluorescence appeared. Utilising extremely similar experiments to Ch<strong>and</strong>ros,<br />

Varma et al. 50 proposed that <strong>the</strong> source <strong>of</strong> <strong>the</strong> exciplex must involve a a-type<br />

bond between <strong>the</strong> lone pair electrons <strong>of</strong> <strong>the</strong> amino nitrogen atom on <strong>the</strong> one<br />

h<strong>and</strong> <strong>and</strong> lone pair electrons <strong>of</strong> <strong>the</strong> polar solvent on <strong>the</strong> o<strong>the</strong>r. This has been<br />

challenged by Suppan 67, who has reported experimental evidence that <strong>the</strong><br />

dual fluorescence observed for DMABN comes from a dielectric-stabilised<br />

highly polar twisted intra-molecular charge transfer state <strong>of</strong> DMABN <strong>and</strong> not<br />

from solute-solvent exciplexes. He also found that DMABN in<br />

perfluorohexane (which has 14 lone pair donor F atoms) shows only single<br />

fluorescence whilst in 1-fluoropentane, which has one F atom, it shows dual<br />

fluorescence, concluding that <strong>the</strong> solvent lone pair <strong>of</strong> electrons are not<br />

involved in <strong>the</strong> a* emission.<br />

39


In 1988 Varma et al.68 69 extended <strong>the</strong>ir work to inyolve time resolved<br />

measurements <strong>and</strong> found that <strong>the</strong> normal fluorescence required <strong>the</strong> sum <strong>of</strong><br />

three exponentials to adequately fit <strong>the</strong> data. The three species were attributed<br />

to excited solute - solvent complexes, excited bare solutes <strong>and</strong> ground state<br />

solute-solvent complexes. Previous to this work, only dual exponential<br />

decays had been observed. Nag et al.51 studied <strong>the</strong> <strong>solution</strong> chemistry <strong>of</strong><br />

DMABN <strong>and</strong> in a similar experiment to Varmaet al.70 <strong>and</strong> Ch<strong>and</strong>ros 66 added<br />

propionitrile to a <strong>solution</strong> <strong>of</strong> DMABN in an inert solvent. They observed that<br />

as <strong>the</strong> percentage <strong>of</strong> added propionitrile was increased <strong>the</strong> b* emission<br />

decreased with increasing a* emission (as observed by Ch<strong>and</strong>ros 66 <strong>and</strong> Varma<br />

et al.70). This was attributed to a reduction in <strong>the</strong> activation energy <strong>of</strong> TICT<br />

state formation by <strong>the</strong> added propionitrile. This trend was only found to be<br />

true up to 15% v/v, when <strong>the</strong> addition <strong>of</strong> more propionitrile caused <strong>the</strong> a*<br />

fluorescence to decrease in intensity. This was explained in terms <strong>of</strong> nonradiative<br />

processes becoming more favourable <strong>and</strong> it was concluded that <strong>the</strong><br />

ratio <strong>of</strong> b*/a* emission vs % propionitrile is basically linear. Weisenborn et<br />

al.68 69 state that both <strong>the</strong> quenching <strong>of</strong> <strong>the</strong> a* <strong>and</strong> <strong>the</strong> growth <strong>of</strong> <strong>the</strong> b* exhibit<br />

linear Stern-Volmer behaviour as a function <strong>of</strong> <strong>the</strong> concentration <strong>of</strong> <strong>the</strong> polar<br />

compound P in <strong>the</strong> solvent mixture with cyclohexane <strong>and</strong> consider this to be<br />

strong evidence for exciplex formation between <strong>the</strong> excited DMABN<br />

molecule <strong>and</strong> P.<br />

As can be seen from <strong>the</strong> above discussion, <strong>the</strong>re is still much confusion<br />

about <strong>the</strong> source <strong>of</strong> <strong>the</strong> anomalous fluorescence shown by <strong>the</strong> widely studied<br />

DMABN system. There are also several closely related molecules which<br />

exhibit this dual fluorescence. In particular, <strong>the</strong> <strong>methyl</strong> <strong>and</strong> ethyl <strong>esters</strong> <strong>of</strong> 4-<br />

N,N-di<strong>methyl</strong><strong>aminobenzoic</strong> acid have received a certain amount <strong>of</strong><br />

attention46' 47, 70<br />

40


1.10.2 4-Di<strong>methyl</strong>amino <strong>and</strong> 4-Diethyl<strong>aminobenzoic</strong> acid <strong>and</strong> <strong>the</strong>ir<br />

<strong>methyl</strong> or ethyl <strong>esters</strong><br />

There is one important difference between <strong>the</strong> energy levels <strong>of</strong><br />

4DMABA, <strong>and</strong> M4DMAB (or <strong>the</strong>ir ethyl counterparts) as apposed to DMABN,<br />

in that <strong>the</strong> 'La <strong>and</strong> 'Lb states are inverted in energy 49. In similar experiments<br />

to those <strong>of</strong> Varma 50 <strong>and</strong> Nag et aL 51 , Cowley et a147 added small amounts <strong>of</strong><br />

polar solvent to an inert <strong>solution</strong> <strong>of</strong> M4DMAB in cyclohexane <strong>and</strong> noted <strong>the</strong><br />

appearance <strong>of</strong> dual fluorescence. Varma 50 attributed this to exciplex<br />

formation through <strong>the</strong> interaction <strong>of</strong> M4DMAB with a polar molecule.<br />

Cowley 47 however attributes this to solvent relaxation allowing <strong>the</strong><br />

formation <strong>of</strong> <strong>the</strong> TICT state i.e. lowering <strong>of</strong> <strong>the</strong> energy <strong>of</strong> activation for TICT<br />

state formation. Cowley 47 concluded that ground state dimer formation<br />

seemed unlikely, since <strong>the</strong> ratio <strong>of</strong> <strong>the</strong> two b<strong>and</strong>s was independent <strong>of</strong><br />

concentration <strong>and</strong> that excimer formation could be ruled out because <strong>of</strong> <strong>the</strong><br />

low concentration used (10 mol dm -3). However our results disagree with<br />

this statement as we find that <strong>the</strong> ratio <strong>of</strong> <strong>the</strong> two b<strong>and</strong>s are concentration<br />

dependent48' 52• Quantum chemical calculations by Rettig, taking into account<br />

a-electrons <strong>and</strong> solvent effects, suggest that for <strong>the</strong> related molecule ethyl 4-<br />

diethylaminobenzoate (E4DEAB) dual fluorescence will be observed in both<br />

polar <strong>and</strong> non-polar solvents due to formation <strong>of</strong> a TICT state. Varma <strong>and</strong> coworkers<br />

53 believe that 1:1 exciplex formation is responsible for <strong>the</strong> dual<br />

fluorescence, whilst Phillips' group 49 disagree with Rettig but are in partial<br />

agreement with Varma 53 <strong>and</strong> ourselves in that a ground state species or <strong>and</strong><br />

excimer / exciplex is <strong>the</strong> cause <strong>of</strong> <strong>the</strong> anomalous fluorescence.<br />

In conclusion, <strong>the</strong> recent studies on both 4-di<strong>methyl</strong>aminobenzonitrile<br />

<strong>and</strong> 4-di<strong>methyl</strong><strong>aminobenzoic</strong> acid <strong>and</strong> its <strong>esters</strong> suggest that <strong>the</strong> origin <strong>of</strong> <strong>the</strong><br />

observed dual fluorescence is not fully understood <strong>and</strong> may be due to some<br />

form <strong>of</strong> exciplex formation 66 or to ground sate dimer / aggregate formation 71<br />

or to a TICT state72 .<br />

41


1.10.3 Aims <strong>of</strong> this project<br />

By <strong>study</strong>ing both <strong>the</strong> ground <strong>and</strong> excited states <strong>of</strong> <strong>the</strong> 2, 3 <strong>and</strong> 4-<br />

<strong>aminobenzoic</strong> <strong>acids</strong>, N,N-di<strong>methyl</strong><strong>aminobenzoic</strong> <strong>acids</strong> <strong>and</strong> <strong>the</strong>ir <strong>methyl</strong><br />

<strong>esters</strong>, it was hoped to underst<strong>and</strong> how solvent, steric <strong>and</strong> substituent effects<br />

alter <strong>the</strong> <strong>photophysics</strong> <strong>of</strong> <strong>the</strong> molecules. In <strong>the</strong> case <strong>of</strong> <strong>the</strong> 4-isomers it was<br />

confirmed that when <strong>the</strong> substituent was a N,N-di<strong>methyl</strong>amino group, dual<br />

fluorescence was observed, so through various experiments utilising both<br />

steady state <strong>and</strong> time-resolved spectroscopy it was hoped to elucidate <strong>the</strong><br />

origin <strong>of</strong> this anomaly.<br />

42


Chapter 2


2.1 Materials<br />

2.1.1 Solvents.<br />

The pure solvents (purchased from <strong>the</strong> Aldrich Chemical Company)<br />

<strong>and</strong> binary mixtures used in this <strong>study</strong>, which are given in Table 2.1 below,<br />

were ei<strong>the</strong>r spectrophotometric or spectr<strong>of</strong>luorometric grade <strong>and</strong> were used<br />

as received except for acetonitrile, butyronitrile <strong>and</strong> water. The acetonitrile<br />

<strong>and</strong> butyronitrile were distilled over phosphorous pentoxide, collected<br />

under an inert atmosphere <strong>and</strong> <strong>the</strong>n distilled again over potassium<br />

carbonate. Even <strong>the</strong>n some fluorescence from <strong>the</strong> butyronitrile remained.<br />

Water was doubly distilled <strong>and</strong> <strong>the</strong>n passed through an ion exchange<br />

column. Hexanes refers to a mixture <strong>of</strong> non-cyclic six carbon compounds.<br />

Relative<br />

dielectric<br />

constant<br />

Dipole<br />

moment<br />

Viscosity<br />

coefficient<br />

Refractive<br />

Index<br />

Solvent or Binary mixture eie, 11<br />

acetonitrile 37.5 3.37 0.375 1.3441<br />

butyronitrile / isobutyronitrile, (9:1) 20.3 3.57 0.624 1.3860<br />

chlorobutane / isopentane (9:1) 7.09 2.11 1 0.469 1.4021<br />

ethanol 24.3 1.68 1 1.078 1.3613<br />

n-hexane 1.89 0.08 1 0.324 1.3748<br />

hexanes<br />

assumed similar to n-hexane<br />

<strong>methyl</strong>cyclohexane / isopentane (4:1) 1.95 1 0.00 0.324 1.3748<br />

water 81.7 1.85 0.891 1 1.3404<br />

Table 2.1<br />

Solvent properties<br />

The four pure solvents used in this work were chosen for <strong>the</strong>ir range<br />

<strong>of</strong> properties, varying from hydrogen bonding to inert <strong>and</strong> non-polar<br />

solvents. The three binary mixtures were chosen for <strong>the</strong> same reason but<br />

also for <strong>the</strong>ir ability to form good glasses at low temperature (77K). Some<br />

useful solvent properties are reported in table 2.1.<br />

2.1.2 Chemicals<br />

Quinine sulphate monohydrate <strong>and</strong> 9,10-diphenylanthracene which<br />

were used as quantum yield st<strong>and</strong>ards were purchased from <strong>the</strong> Aldrich<br />

43


Chemical company (Gold star) <strong>and</strong> were used as received. The<br />

<strong>aminobenzoic</strong> <strong>acids</strong> were purchased from Sigma I Aldrich or from Lancaster<br />

Syn<strong>the</strong>sis <strong>and</strong> were recrystallised before use. Details <strong>of</strong> <strong>the</strong>ir properties are<br />

given in Table 2.2. The <strong>methyl</strong> <strong>esters</strong> were prepared by one <strong>of</strong> two methods:<br />

Method a:73 A <strong>solution</strong> containing <strong>methyl</strong> iodide (0.03 mol) in a small<br />

quantity <strong>of</strong> benzene was added to a <strong>solution</strong> <strong>of</strong> 2-<strong>aminobenzoic</strong> acid (0.01<br />

mol) in benzene <strong>and</strong> 1,8-diazobicyclo[5, 4,0]undec -7-ene (0.03 mol). The<br />

reaction mixture was stirred at room temperature for four hours <strong>and</strong> <strong>the</strong>n<br />

washed with sodium hydrogen carbonate <strong>solution</strong>. Extraction with ethyl<br />

acetate <strong>and</strong> removal <strong>of</strong> <strong>the</strong> dried solvent gave <strong>methyl</strong> 2-aminobenzoate<br />

which was distilled twice.<br />

Method b: 74 A <strong>solution</strong> <strong>of</strong> 4-<strong>aminobenzoic</strong> acid (0.01 mol) in methanol (ca<br />

50 cm3) was added to boron trifluoride-methanol complex (0.03 mol) <strong>and</strong> <strong>the</strong><br />

mixture was boiled under reflux for four hours. After cooling, <strong>the</strong> mixture<br />

was poured into saturated sodium hydrogen carbonate <strong>solution</strong>. The ester<br />

was extracted using ethyl acetate <strong>and</strong> <strong>the</strong> <strong>solution</strong> dried with magnesium<br />

sulphate. The ethyl acetate was removed under reduced pressure <strong>and</strong> <strong>the</strong><br />

resulting <strong>methyl</strong> 4-aminobenzoate was twice recrystallised from ethanol.<br />

2.2 Instrumentation <strong>and</strong> Instrumental techniques<br />

22.1 Absorption spectra<br />

Absorption spectra were measured in <strong>the</strong> seven different solvent or<br />

binary mixtures on a Perkin-Elmer Lambda 3 spectrophotometer or a<br />

Hewlett-Packard HP8451A diode array spectrometer using matched quartz<br />

cells <strong>of</strong> variable pathlengths (1mm to 10mm). The accuracy <strong>of</strong> <strong>the</strong> peak<br />

wavelengths is estimated at +1- 2nrn <strong>and</strong> <strong>the</strong> extinction coefficients at +1-<br />

10%. Unless o<strong>the</strong>rwise stated <strong>the</strong> concentration <strong>of</strong> <strong>the</strong> <strong>solution</strong>s used was<br />

around 1*10.4 mol dm 3 .<br />

44


— — — —<br />

-<br />

VP<br />

'0<br />

ret<br />

Cl<br />

LI)<br />

'0<br />

— — — — — — —<br />

cek<br />

'.o<br />

£<br />

-Scnt ...<br />

'c -'<br />

as<br />

en reS cu<br />

0 .CIJ Cl 'a<br />

trt EEeut €E E<br />

—'--<br />

?-t O '<br />

.c-jt' en '0 tn in Cl<br />

rZ€'a j'at' "'.dt rC<br />

-toc't<br />

Gs<br />

2-S.2-2-a!S.S..2.aS.a<br />

-<br />

'a<br />

C<br />

IZ<br />

V<br />

a)<br />

-t<br />

I-.<br />

a)<br />

-a<br />

a)<br />

V<br />

. xx<br />

a<br />

U,<br />

5<br />

000<br />

II<br />

-<br />

II<br />

II<br />

— -<br />

—----<br />

00 00 00<br />

I I I II<br />

ZZZ .0<br />

II<br />

u U II II II<br />

0 00<br />

- N-<br />

CICI)C)<br />

enenen r ' --<br />

xxz z xx xx 0 00<br />

66 6 ±± . 6 66 6 66<br />

'00<br />

cQN<br />

Ot-'<br />

(I)(1) - —-<br />

I-.<br />

a)<br />

-E<br />

C<br />

ft<br />

m<br />

-a<br />

-ø<br />

ft<br />

U<br />

0<br />

a)<br />

0<br />

E<br />

cc<br />

C,<br />

'C<br />

a)<br />

S<br />

0<br />

a)<br />

V<br />

I 1 CI)U) C'. 1 1<br />

0 U)t<br />

tC t--<br />

Nb'.<br />

qa- -<br />

a- ento<br />

a — N aCflC'1 Cl<br />

u - I - I - - '- - Cl O.CN 0C1<br />

cq<br />

LrI Cl Cl -<br />

a-.<br />

e t' cno<br />

oc.-r'? rtl<br />

!7b?<br />

cn'd'<br />

0'<br />

ON<br />

a) cc' tact<br />

O cctct' .s<


2.2.2 Ground <strong>and</strong> excited state pK values<br />

a) The pK values for ground state protonation <strong>and</strong> deprotonation<br />

were calculated by measuring <strong>the</strong> ground state absorption spectra as a<br />

function <strong>of</strong> pH75. The method depends upon <strong>the</strong> direct determination <strong>of</strong> <strong>the</strong><br />

ratio <strong>of</strong> molecular species (neutral molecule) to ionised species in a series <strong>of</strong><br />

non- absorbing buffer <strong>solution</strong>s with a span <strong>of</strong> at least half a pH unit. For<br />

this purpose, <strong>the</strong> spectrum <strong>of</strong> <strong>the</strong> molecular species was first obtained in a<br />

buffer <strong>solution</strong> whose pH was chosen such that <strong>the</strong> compound studied was<br />

present wholly as this species. This spectrum was <strong>the</strong>n compared with that<br />

<strong>of</strong> <strong>the</strong> pure ionised species similarly isolated at ano<strong>the</strong>r suitable pH. A<br />

wavelength was <strong>the</strong>n chosen at which <strong>the</strong> greatest difference between <strong>the</strong><br />

absorbances <strong>of</strong> <strong>the</strong> species was observed. If it is assumed that Beer's law is<br />

obeyed for both species, <strong>the</strong>n <strong>the</strong> observed absorbance (A) at <strong>the</strong> chosen<br />

wavelength is due to <strong>the</strong> sum <strong>of</strong> <strong>the</strong> optical densities <strong>of</strong> <strong>the</strong> ionised species,<br />

A 1, <strong>and</strong> <strong>the</strong> molecular species Am. The absorbance is related to <strong>the</strong><br />

concentration <strong>of</strong> <strong>the</strong> relevant species by A=Ecl. The concentration <strong>of</strong> <strong>the</strong><br />

ionised species in <strong>the</strong> mixture is ac where a is <strong>the</strong> fraction <strong>of</strong> molecule<br />

ionised <strong>and</strong> hence its contribution to <strong>the</strong> observed absorbance is c 1acl <strong>and</strong><br />

similarly <strong>the</strong> contribution from <strong>the</strong> molecule is Em (1-(X)cl where a is given<br />

by<br />

i-a = Ka /([H+]+Ka); a =[H.i-]/([H+]-i.Iç) for <strong>acids</strong> ..........2.1<br />

i-a= [H+]/([H+]+K1,); a= Kb /([H+]+Kb) for bases ..........2.2<br />

If <strong>the</strong> same cell pathlength is used throughout <strong>the</strong>n<br />

A=(ej(X+em (1-cx))c<br />

<strong>and</strong> <strong>the</strong>refore, since E =A/c <strong>and</strong> a is defined above.<br />

46


EjK.<br />

= + €m[H] for<strong>acids</strong> . 2.3<br />

[H]+K0 [W]+K0<br />

£ = s_K0 + for bases 2.4<br />

[H]+K0 [H]+K0<br />

Provided that <strong>the</strong> same total concentration is used for all measurements,<br />

equations 2.3 <strong>and</strong> 2.4 may be written with <strong>the</strong> appropriate absorbances<br />

replacing <strong>the</strong> extinction coefficients. Each equation can be rearranged into<br />

two forms to suit different needs. When <strong>the</strong> functional group being<br />

determined is an acid, equation 2.5a is used if A 1 is greater than A. <strong>and</strong> 2.5b<br />

if <strong>the</strong> reverse is <strong>the</strong> case.<br />

pKa=pH+Iog A 2.5a<br />

..........2.Sb<br />

When <strong>the</strong> group being determined is a base, equation 2.6a is used if A is<br />

greater than Am, <strong>and</strong> 2.6b if <strong>the</strong> reverse is <strong>the</strong> case.<br />

pKb = pH + log t AJ 2.6a<br />

pK=pH+logAmA 2.6b<br />

A—A 1<br />

The pK a or p1


pH A ArnA<br />

Log(m) pKa<br />

1.50 0.1480 0.5948 0.1028 0.7624 2.26<br />

1.67 0.2002 0.5426 0.1550 0.5442 2.21<br />

1.82 0.2529 0.4899 0.2077 0.3727 2.19<br />

2.10 0.3491 0.3937 0.3039 0.1124 2.21<br />

2.26 0.4137 0.3291 0.3685 -0.0491 2.21<br />

2.54 0.5009 0.2419 0.4557 -0.2750 2.26<br />

2.73 0.5659 0.1769 0.5137 -0.4630 2.26<br />

Table 2.3 Determination <strong>of</strong> <strong>the</strong> ionisation constant for protonation <strong>of</strong> 2-<br />

<strong>aminobenzoic</strong> acid (1.92*104 mol dm-3). Temperature 20 0C,<br />

analytical wavelength 326 nm<br />

Once <strong>the</strong> pKa is known, it is possible to calculate <strong>the</strong> percentage <strong>of</strong> ionisation<br />

given <strong>the</strong> pH value using equations 2.7 <strong>and</strong> 2.8.<br />

100<br />

% Ionised = for <strong>acids</strong> 2.7<br />

1 + anti log(pK - pH)<br />

%Ionised =<br />

100<br />

1 + anti log(pH - pK)<br />

for bases 2.8<br />

b) Excited state pK* values may be determined in a similar<br />

manner to <strong>the</strong> ground state pK values detailed above 35 Variation <strong>of</strong><br />

fluorescence intensity with pH yields a values for <strong>the</strong> excited states<br />

involved in <strong>the</strong> photochemistry <strong>and</strong> <strong>the</strong>se data in turn may be used to<br />

determine pK* values via <strong>the</strong> Henderson-Hasselbalch relationship 80 . In<br />

experiments <strong>of</strong> this type, it is essential to excite <strong>the</strong> fluorescence at <strong>the</strong><br />

isobestic point for any ground state equilibrium which is taking place in <strong>the</strong><br />

pH range <strong>of</strong> interest. This ensures that changes in <strong>the</strong> fluorescence spectra<br />

due to (de)protonation processes are isolated from changes in <strong>the</strong> spectra<br />

caused by variations in <strong>the</strong> absorbed light intensity.<br />

B1


The effect <strong>of</strong> pH on <strong>the</strong> fluorescence spectra <strong>of</strong> all twelve compounds<br />

in aqueous buffers was studied in <strong>the</strong> pH range 11.0 - 2.00 <strong>and</strong> <strong>the</strong> results are<br />

reported in Chapters 3 - 5.<br />

c) The excited state pK values were calculated using equation 2.9<br />

based on <strong>the</strong> FOrster cycle (see figure 1.12)<br />

pK* - pK= 2.1 x 1ft 3(v -VA) 2.9<br />

where V<br />

base respectively.<br />

<strong>and</strong> VA are <strong>the</strong> b<strong>and</strong> maxima (in cm 4) <strong>of</strong> <strong>the</strong> acid <strong>and</strong> conjugate<br />

2.2.3 fluorescence spectra<br />

Fluorescence spectra were measured using ei<strong>the</strong>r a Perkin-Elmer LS5<br />

fluorescence spectrometer in <strong>the</strong> fully corrected mode or a Perkin-Elmer<br />

LS50. Unless o<strong>the</strong>rwise stated a concentration <strong>of</strong> around 1*106 mol dm 3<br />

was used. The low temperature fluorescence spectra were recorded using<br />

equipment at <strong>the</strong> SERC Daresbury Laboratory (see section 2.2.4), utilising a<br />

xenon light source in place <strong>of</strong> <strong>the</strong> synchrotron radiation source, <strong>and</strong> an<br />

Oxford Instruments CF204 cryostat. The <strong>solution</strong> was placed into <strong>the</strong> sample<br />

chamber <strong>and</strong> <strong>the</strong> sample jacket was evacuated. Liquid nitrogen was<br />

transferred to <strong>the</strong> cryostat. The temperature was modulated by ei<strong>the</strong>r<br />

restricting or increasing <strong>the</strong> flow <strong>of</strong> <strong>the</strong> liquid nitrogen. Measurements were<br />

taken at decreasing temperature from ambient ra<strong>the</strong>r than on warming up<br />

<strong>of</strong> <strong>the</strong> <strong>solution</strong> from 77K in order to reduce adsorption <strong>of</strong> water into <strong>the</strong><br />

<strong>solution</strong>. Samples were not used more than once for this same reason.<br />

Spectra were only corrected for blank solvent fluorescence, which in <strong>the</strong> case<br />

<strong>of</strong> <strong>the</strong> butyronitrile / iso-butyronitrile mixture at temperatures <strong>of</strong> 133K <strong>and</strong><br />

less was quite intense.<br />

49


Quantum yields were measured on optically dilute samples<br />

(absorbance .c 0.05) to eliminate <strong>the</strong> effect <strong>of</strong> self quenching <strong>and</strong> inner filter<br />

effects <strong>and</strong> so that <strong>the</strong> fluorescence intensity was proportional to<br />

concentration / absorbance (see introduction). The samples were degassed by<br />

bubbling with oxygen-free nitrogen for twenty minutes.<br />

Quinine sulphate (1.28 x 10-6 mol dm -3) dissolved in perchloric acid<br />

(0.01 mol dm4) was used as <strong>the</strong> quantum yield st<strong>and</strong>ard76' 77. A stock<br />

<strong>solution</strong> (0.00128 mol dm- 3) <strong>of</strong> quinine sulphate (0.0979g) in perchloric acid<br />

(1 dm3, 0.01 mol dm -3) was prepared. Dilutions were freshly made up on a<br />

daily basis. The lifetime <strong>of</strong> <strong>the</strong> stock <strong>solution</strong> was around one month.<br />

A BBC microcomputer (<strong>and</strong> later an IBM 386 personal computer) was<br />

interfaced through <strong>the</strong> chart output <strong>of</strong> <strong>the</strong> Perkin-Elmer LS5 fluorescence<br />

spectrometer <strong>and</strong> spectra were recorded on <strong>the</strong> computer. This enabled <strong>the</strong><br />

area under <strong>the</strong> curve to be calculated. By using <strong>the</strong> following equation,<br />

unknown quantum yields were calculated (see Appendix A for full listing <strong>of</strong><br />

program):<br />

• 1 (unknown) = Fluor. int. (unknown) x Abs(st<strong>and</strong>ard)<br />

4 1 (st<strong>and</strong>ard) Fluor. int. (s tan dard) x Abs(unknown)<br />

2.10<br />

where o f = fluorescence quantum yield, Fluor.int. = integrated area under<br />

<strong>the</strong> corrected fluorescence spectrum <strong>and</strong> Abs is <strong>the</strong> absorbance <strong>of</strong> <strong>the</strong> sample<br />

at <strong>the</strong> excitation wavelength.<br />

Phosphorescence spectra (77K) were measured on <strong>the</strong> Perkin-Elmer<br />

LSS spectr<strong>of</strong>luorimeter in phosphorescence mode with background<br />

correction using a cryostat attachment. An attempt was made to determine<br />

<strong>the</strong> phosphorescence lifetime by keeping <strong>the</strong> time gate (tg) constant <strong>and</strong><br />

measuring <strong>the</strong> intensity at different delay times ('Cd). This was unsuccessful.<br />

50


2.2.4 fluorescence decay pr<strong>of</strong>iles<br />

Time-resolved fluorescence spectroscopy was performed using <strong>the</strong><br />

single photon counting technique 78, using a variety <strong>of</strong> excitation sources;<br />

<strong>the</strong> Synchrotron Radiation Sources (SRS) at Daresbury, UK <strong>and</strong> BESSY,<br />

Berlin, Germany, a hydrogen flash lamp at <strong>the</strong> University <strong>of</strong> Strathclyde,<br />

Glasgow <strong>and</strong> a frequency doubled dye laser at <strong>the</strong> Ru<strong>the</strong>rford Appleton<br />

Laboratory (R.A.L). The main work was undertaken at <strong>the</strong> SRS at Daresbury.<br />

Operating in single bunch mode <strong>the</strong> SRS provides pulses <strong>of</strong> light :5 200 ps<br />

duration <strong>of</strong> frequency <strong>of</strong> = 3.0 Ml-Iz. Wavelength selection is achieved with a<br />

Spex 15005P Czerny-Turner monochromator with continuously variable<br />

slits. Sample fluorescence was detected at right angles to <strong>the</strong> excitation beam<br />

with a Peltier-cooled Mullard PM2254 photomultiplier. Wavelength<br />

selection on <strong>the</strong> emission side was achieved with Barr <strong>and</strong> Stroud<br />

interference filters (b<strong>and</strong>width 10 nm f.w.h.m). Pulses from <strong>the</strong><br />

photomultiplier were fed into an Ortec 583 constant fraction discriminator<br />

<strong>and</strong> <strong>the</strong>n into <strong>the</strong> START channel <strong>of</strong> an Ortec 457 time-to-amplitude<br />

converter (TAC).<br />

Stop pulses for <strong>the</strong> TAC were generated ei<strong>the</strong>r from a strip line,<br />

essentially an aerial close to <strong>the</strong> synchrotron ring which picks up a signal as<br />

<strong>the</strong> electrons pass by <strong>and</strong> an Ortec 473A constant-fraction discriminator, or<br />

from <strong>the</strong> 500 MHz signal used to drive <strong>the</strong> SRS klystron, divided down by a<br />

factor <strong>of</strong> 160 <strong>and</strong> passed through an Ortec 583 constant fraction<br />

discriminator. The TAC output was accumulated in 1024 channels <strong>of</strong> an mo-<br />

Tech 5400 multichannel analyser <strong>and</strong> <strong>the</strong>n stored on a winchester hard disk<br />

on a POP-il mini-computer before being passed to a CONVEX mainframe<br />

for processing <strong>and</strong> permanent storage. The fluorescence decay pr<strong>of</strong>iles were<br />

evaluated by computer deconvolution using a kinetic model <strong>of</strong> <strong>the</strong> form in<br />

equation 2.11<br />

51


Fluorescence intensity(t) = I i a1e'1 . 2.11<br />

where a i is <strong>the</strong> amplitude <strong>of</strong> a component lifetime, t1 (where i = 1, 2 or 3).<br />

The data were analysed using an iterative convolution method using<br />

weighted, non-linear least-squares curve fitting routines. The computed fits<br />

to <strong>the</strong> experimental data were evaluated on <strong>the</strong> basis <strong>of</strong> <strong>the</strong> X2 values <strong>and</strong><br />

<strong>the</strong> distribution <strong>of</strong> residuals. Care had to be taken because although x2 <strong>the</strong><br />

values were a good indication <strong>of</strong> <strong>the</strong> suitability <strong>of</strong> a fit, <strong>the</strong>y could not be<br />

completely relied upon. Consequently <strong>the</strong> distribution <strong>of</strong> <strong>the</strong> residuals, <strong>the</strong><br />

percentage <strong>of</strong> each component present <strong>and</strong> <strong>the</strong> percentage error on <strong>the</strong><br />

lifetimes were also used to choose <strong>the</strong> complexity <strong>of</strong> fit which most<br />

adequately described <strong>the</strong> decay kinetics <strong>of</strong> a particular species. Figure 2.1a<br />

shows a poor single exponential fit to <strong>the</strong> decay pr<strong>of</strong>ile <strong>of</strong> 2ABA in water<br />

<strong>and</strong> figure 2.1b shows <strong>the</strong> good double exponential fit to <strong>the</strong> same data.<br />

a


2—<strong>aminobenzoic</strong> acid (H20)<br />

to<br />

4-I<br />

io2<br />

I.<br />

\4..<br />

' I. ..<br />

lo t<br />

io o<br />

0<br />

2 4 6 0 10 12<br />

Fit I<br />

12<br />

Time/iO sec xso- 3.49<br />

C<br />

It<br />

a-,<br />

U) -12<br />

Figure 2.1a Single exponential fit to <strong>the</strong> fluorescence decay pr<strong>of</strong>ile <strong>of</strong> 2ABA<br />

in water<br />

53


2-amirrnbenzoic acid (H20)<br />

In<br />

=<br />

0 to<br />

10 1<br />

I,, :. .. •<br />

10 0 - .aI<br />

Mb<br />

.,<br />

Ga 2<br />

Cm<br />

0<br />

0. 2 4 B 8 10 12<br />

Fit 2 Time/10 ° sec X59 0.95<br />

Figure 2.lb Double exponential fit to <strong>the</strong> fluorescence decay pr<strong>of</strong>ile <strong>of</strong><br />

2ABA in water


Chapter 3


.. flu..nsnm<br />

3.1 Ground state absorption spectra as a function <strong>of</strong> solvent.<br />

For all <strong>the</strong> compounds investigated during <strong>the</strong> course <strong>of</strong> this <strong>study</strong>, <strong>the</strong><br />

meta- isomers would be expected to form <strong>the</strong> simplest case due to <strong>the</strong>ir<br />

inability to form a resonance stabilised charge transfer state. There is also no<br />

possibility <strong>of</strong> steric interaction between <strong>the</strong> carboxylic acid or ester <strong>and</strong> <strong>the</strong><br />

amino or N,N-di<strong>methyl</strong>amino substituent groups, although both substituent<br />

groups are capable <strong>of</strong> forming inter-molecular hydrogen bonds with suitable<br />

solvents. The importance <strong>of</strong> this steric interaction will be apparent in chapter<br />

four where <strong>the</strong> ortho- isomers are discussed. In all <strong>of</strong> <strong>the</strong> seven solvents <strong>and</strong><br />

binary mixtures studied, 3-<strong>aminobenzoic</strong> acid (3ABA), <strong>methyl</strong> 3-<br />

aminobenzoate (M3AB), 3-N,N-di<strong>methyl</strong><strong>aminobenzoic</strong> acid (3DMABA) <strong>and</strong><br />

<strong>methyl</strong> 3-N,N-di<strong>methyl</strong>amino-benzoate (M3DMAB) all exhibit generally<br />

weak, UV absorption (extinction coefficients <strong>of</strong> <strong>the</strong> order <strong>of</strong> 1-3,000 dm 3 mol-'<br />

cm 1). 3ABA (Figure 3.1) <strong>and</strong> M3AB exhibit a single absorption b<strong>and</strong> above 250<br />

nm with a wavelength <strong>of</strong> maximum absorption varying between 300 - 325<br />

nm (Table 3.1).<br />

0.16<br />

0.14<br />

0.12<br />

v 0.10<br />

C<br />

C<br />

0.08<br />

0<br />

0.06<br />

0.04<br />

0.02<br />

0.00<br />

200 220 240 260 280 300 320 340 360 380 400<br />

Wavelength (nm)<br />

Figure 3.1 Absorption spectrum <strong>of</strong> 3ABA in hexane<br />

55


The N,N-di<strong>methyl</strong>amino compounds have a wavelength <strong>of</strong><br />

maximum absorption varying between 320 - 346 nm (Figure 3.2 <strong>and</strong> Table 3.1).<br />

Two o<strong>the</strong>r absorption b<strong>and</strong>s are observed around 230 <strong>and</strong> 200 nm. The<br />

exceptions to <strong>the</strong>se general comments are 3ABA in hexane <strong>and</strong><br />

<strong>methyl</strong>cyclohexane / isopentane (4:1), where a shoulder is observed on <strong>the</strong><br />

main peak at 285 rim, some 20 rim blue shifted from <strong>the</strong> main peak (Figure<br />

3.1).<br />

The absorption spectra <strong>of</strong> 3ABA <strong>and</strong> 3DMABA (Figure 3.2) in<br />

unbuffered water shows unexpectedly low wavelengths <strong>of</strong> maximum<br />

absorption. However, given <strong>the</strong> pK a values reported later in this <strong>the</strong>sis (Table<br />

3.2), it is clear that this apparent anomaly is due to ionisation <strong>of</strong> <strong>the</strong> carboxylic<br />

acid group to form <strong>the</strong> anion which, in both cases, absorbs to <strong>the</strong> "blue" <strong>of</strong> <strong>the</strong><br />

neutral molecule. In ethanol <strong>the</strong>re is a similar though less noticeable effect<br />

with <strong>the</strong> wavelengths <strong>of</strong> maximum absorption being around 317 <strong>and</strong> 336 nm<br />

respectively (Figure 3.2 <strong>and</strong> Table 3.1).<br />

C,<br />

1<br />

0.9<br />

Acetonitrile<br />

0.8 Ethanol<br />

0.7 Water<br />

0.6<br />

' 0.5<br />

C<br />

. 0.4<br />

IC<br />

0.3<br />

0.2<br />

0.1<br />

0<br />

250 270 290 310 330 350 370 390<br />

Wavelength (nm)<br />

Figure 3.2 Absorption spectra <strong>of</strong> 3DMABA in water, ethanol <strong>and</strong><br />

acetonitrile<br />

As expected, replacement <strong>of</strong> <strong>the</strong> amino substituent with a N,Ndi<strong>methyl</strong>amino<br />

group results in a red shift <strong>of</strong> <strong>the</strong> absorption b<strong>and</strong> <strong>of</strong> some 20<br />

56


nm. It might be expected that <strong>the</strong>re would be a simultaneous increase in<br />

extinction coefficient, but <strong>the</strong>se seem to remain constant.<br />

Since <strong>the</strong> meta- isomer is incapable <strong>of</strong> forming a resonance stabilised<br />

charge transfer state, <strong>the</strong> only interaction between <strong>the</strong> amino <strong>and</strong> carboxyl<br />

groups in <strong>the</strong> excited state can be through <strong>the</strong> inductive effect. As can be seen<br />

from <strong>the</strong> data in Table 3.1, <strong>the</strong> substituent group has an important effect on<br />

both <strong>the</strong> intensity <strong>and</strong> wavelength <strong>of</strong> maximum absorption.<br />

Solvent Measured 3ABA M3AB MABA M3DMAB<br />

quantity<br />

______ _<br />

Acetonitrile 323 323 343<br />

___ S 7dm3 cnr1 mot1 2,400 2,300 0 2,200<br />

BuCN/IBuCN 322 323 341<br />

S /dm3cm4mo1' 2,700 3,300 0 2,100<br />

BCl/iP 320 321 342<br />

S 7dm3 cur 1 mol1 n/s 2,344 5 2,421<br />

Ethanol 317 324<br />

344<br />

S /dm3cnr1mol4 2,000 2,100 0 2,100<br />

Hexane 306 312 336<br />

S 7dm3 cm4 mol4 n/s 2,100 0 2,500<br />

MCH/iP 316 319<br />

£ /dm3cnr1mol1 n/s 2,400 0 2,600<br />

Water 302 317 321<br />

S /dm3cnr1mol4 1,100 2,500 0 1,400<br />

Neutral 310 312 E326322 322<br />

£ /dm3cm4mot1 1,000 2,000 1,000 1,400<br />

Anion 300 - 308 -<br />

C /dm3 cm 1 mo11 11,700 1 - 1,400 -<br />

Cation ?,max Inm 1 272 273 272 274<br />

C /dr&cm 1 mol4 1 1,000 1,000 900 850<br />

Table 3.1 Primary ground state absorption characteristics <strong>of</strong> 3-<br />

<strong>aminobenzoic</strong> <strong>and</strong> 3-N,N-di<strong>methyl</strong><strong>aminobenzoic</strong> <strong>acids</strong> <strong>and</strong> <strong>the</strong>ir<br />

<strong>methyl</strong> <strong>esters</strong> as a function <strong>of</strong> pH <strong>and</strong> solvent<br />

Alkyl substituents on <strong>the</strong> aromatic ring generally produce a red shift,<br />

with a <strong>methyl</strong> group having <strong>the</strong> largest effect. Both COOR <strong>and</strong> NR2 groups<br />

have <strong>the</strong> ability to interact with <strong>the</strong> electrons <strong>of</strong> <strong>the</strong> aromatic ring system,<br />

<strong>the</strong>reby increasing <strong>the</strong> wavelength <strong>of</strong> maximum absorption. The N,Ndi<strong>methyl</strong>amino<br />

group is a better electron donor than an amino group, whilst<br />

57


acid groups are better electron acceptors than <strong>the</strong>ir corresponding <strong>esters</strong>.<br />

However this latter difference is not as significant as <strong>the</strong> former. Figure 3.3<br />

shows <strong>the</strong> possible resonance structures available for aniline <strong>and</strong> benzoic add.<br />

HO 0<br />

cs<br />

O /0<br />

HO /0 HO<br />

C<br />

/O)<br />

C<br />

+af)LJ<br />

I<br />

HO 0<br />

o<br />

It<br />

S+<br />

'II<br />

I I<br />

II<br />

II<br />

I I<br />

I. .7<br />

INn2<br />

+<br />

NH2<br />

-0<br />

00<br />

+ +<br />

NH2<br />

NH2<br />

NH2<br />

-<br />

Figure 3.3 Mesomeric structures <strong>of</strong> Benzoic acid <strong>and</strong> Aniline<br />

Since <strong>the</strong> two substituent groups are meta - disposed <strong>the</strong>y cannot<br />

interact mesomerically through <strong>the</strong> ic-system. Any interaction must be<br />

through o-bonds, an effect which dies <strong>of</strong>f rapidly with distance. This <strong>the</strong>refore<br />

gives rise to <strong>the</strong> following trend in <strong>the</strong> amount <strong>of</strong> charge transfer ability <strong>of</strong> <strong>the</strong><br />

meta - compounds.<br />

3DMABA ~! M3DMAB> 3DABA 2: M3AB.<br />

By increasing <strong>the</strong> amount <strong>of</strong> delocalisation <strong>of</strong> <strong>the</strong> system <strong>the</strong> energy <strong>of</strong><br />

<strong>the</strong> 7t_it* transition will be lower, which will shift <strong>the</strong> wavelength <strong>of</strong><br />

maximum absorption to higher wavelengths. It is <strong>the</strong>refore predicted that <strong>the</strong><br />

absorption maxima in a given solvent would follow a similar trend <strong>and</strong> this<br />

is generally observed (Table 3.1).


There have been few reports <strong>of</strong> <strong>the</strong> absorption properties <strong>of</strong> <strong>the</strong>se<br />

compounds in <strong>the</strong> literature. In early literature (pre 1985), a passing <strong>study</strong> <strong>of</strong><br />

3DMABA, by Doub <strong>and</strong> Leanord37 can be found. They observed that at a pH <strong>of</strong><br />

3.75, three absorption b<strong>and</strong>s were observed at 218.5 am (C 14,000 dm 3 mol-'<br />

cm-1), 250 nm (8 4,400 dm3 mo!- 1 cm-') <strong>and</strong> 310 nm (S 650 dm3 mol-' cm-1 )37<br />

However at a pH <strong>of</strong> 3.75 <strong>the</strong> species being investigated will be a mixture <strong>of</strong><br />

cation <strong>and</strong> neutral species (30 / 70). Our results reproduce <strong>the</strong> tertiary<br />

absorption b<strong>and</strong> observed around 313 am, which has been attributed to a 1t-t<br />

transition.<br />

In a more recent <strong>study</strong>, Jian measured <strong>the</strong> absorption spectra <strong>of</strong> 3ABA<br />

as a function <strong>of</strong> solvent <strong>and</strong> pH <strong>and</strong> our results agree with his data within<br />

experimental error (see Table 3.1a).<br />

So!vent (absorption maxima in nm <strong>and</strong> respective oscifiator strength)<br />

MCH Acetonitri!e Methanol iO M HC1 IM HC! Alkaline<br />

Xmax Amt Xrnax f Amax f Xrnax I Xmax I<br />

340 320 1 314 0.080 312 0.013 271 0.014 301 0.078<br />

248 246 1 245 0.166 271 0.008 227 0.199 233 0.456<br />

208 228 6 216 0.675 221 intense 208 0.032 212 0.636<br />

Table 3.1a<br />

Absorption maxima <strong>of</strong> 3-<strong>aminobenzoic</strong> acid as a function <strong>of</strong><br />

solvent <strong>and</strong> pH, reported by Jian et al<br />

For both <strong>acids</strong> it was found that <strong>the</strong> cation absorbed at 272 am (C 1,000<br />

dm3 mol-' cm-'), with <strong>the</strong> neutral species absorbing at 326 nm (pH


3.2 Ground state absorption data as a function <strong>of</strong> pH.<br />

In water all <strong>the</strong> meta- substituted compounds are capable <strong>of</strong><br />

protonation <strong>and</strong> <strong>the</strong> two <strong>acids</strong> can also be deprotonated to form <strong>the</strong> anion.<br />

The latter species absorbs more strongly than <strong>the</strong> corresponding neutral<br />

molecule <strong>and</strong> its absorption maximum is blue shifted by some 15 - 30 nm<br />

(Table 3.1).<br />

.1<br />

0.35 .4.\<br />

0.304. \<br />

I<br />

'<br />

• ' I<br />

0.25 j\ % •u,<br />

020 '. •)<br />

. oistN...S'<br />

0.10 1<br />

INZ<br />

pH 9.23<br />

pH 5.93<br />

pHt9S<br />

pH 4.37<br />

----pH 3.94<br />

pH 3.67<br />

pH 3.39<br />

0.05<br />

0.00<br />

250 270<br />

290 310 330 350<br />

Wavelength (nm)<br />

370 390<br />

Figure 3.4 Absorption spectra <strong>of</strong> 3DMABA as a function <strong>of</strong> pH<br />

Figure 3.4 shows <strong>the</strong> ground state absorption properties <strong>of</strong> 3DMABA as<br />

a function <strong>of</strong> pH. It can be clearly seen that as <strong>the</strong> pH is reduced from a value<br />

<strong>of</strong> nine to a pH value <strong>of</strong> four, <strong>the</strong>re is a two fold reduction in <strong>the</strong> absorption<br />

intensity. At pH 4 <strong>the</strong> distribution <strong>of</strong> species is approximately 60% neutral<br />

molecule <strong>and</strong> 40% anion. The absorption maximum has changed from 313<br />

nm at high pH to 326 nm. As <strong>the</strong> pH is reduced to lower values <strong>the</strong> peak at<br />

326 run continues to decrease in intensity <strong>and</strong> <strong>the</strong> intense peak observed at<br />

270 nm slowly levels out until at a pH <strong>of</strong> less than two no fur<strong>the</strong>r change is<br />

observed in intensity until an acidity is reached which will produce <strong>the</strong> bi-


cation. This however requires <strong>the</strong> very strongly acidic conditions <strong>of</strong> a<br />

concentrated acid 45. These data are presented in Table 3.2. Also included in<br />

this table are <strong>the</strong> excited state pK values (pK*) which have been ei<strong>the</strong>r<br />

determined from fluorescence spectral measurements or calculated from <strong>the</strong><br />

Forster-Weller relationship 24. The latter have been calculated using both<br />

absorption <strong>and</strong> fluorescence maxima <strong>and</strong> both sets <strong>of</strong> results are presented.<br />

Ground<br />

Excited state S1<br />

state S0<br />

Compound pKb pKa pKb* plc*<br />

Experimental Calculated Experimental Calculated<br />

Abs Fluor Abs Fluor<br />

3ABA 2.75 4.49 3.03 -6.75 - 4.76 6.75 5.48<br />

(3.12) (4.74)<br />

(-7.1) ___ _____________ (7.2) (4.4)<br />

M3AB 3.52 - 5.00 -6.38 - - - -<br />

3DMABA 1 3.35 5.23 3.38 -9.44 - 5.24 9.00 5.78<br />

M3DMAB 3.79 - 4.88 -8.23 - - - -<br />

Table 3.2<br />

Ground <strong>and</strong> excited pK values for both protonation <strong>and</strong><br />

deprotonation. Data in bold are values reported by Jian et al<br />

For <strong>the</strong> <strong>esters</strong> at a pH > 5, <strong>the</strong> ground state consists <strong>of</strong> <strong>the</strong> pure neutral<br />

species only. On <strong>the</strong> basis <strong>of</strong> <strong>the</strong> calculated pK b*, one would not expect any<br />

protonation <strong>of</strong> <strong>the</strong> excited neutral species so this should lead to a constant<br />

fluorescence spectrum <strong>and</strong> intensity. Within <strong>the</strong> limits <strong>of</strong> experimental error<br />

this is observed. The converse is also probably true in that at pHs c 1.5 <strong>the</strong><br />

ground state will be composed <strong>of</strong> pure cation. Since <strong>the</strong> cation is nonfluorescent<br />

<strong>the</strong> amount <strong>of</strong> fluorescence observed will depend on <strong>the</strong> fraction<br />

<strong>of</strong> excited cation which loses a proton to form <strong>the</strong> excited neutral molecule. In<br />

<strong>the</strong> absence <strong>of</strong> protonation <strong>of</strong> <strong>the</strong> excited neutral species, it would again be<br />

expected that <strong>the</strong> fluorescence intensity would remain constant <strong>and</strong> this is<br />

also observed. The ratio <strong>of</strong> <strong>the</strong>se two constant intensities (i.e <strong>the</strong> fluorescence<br />

intensity at a pH c 1.5 against <strong>the</strong> fluorescence intensity <strong>of</strong> a pH> 5) gives <strong>the</strong><br />

proportion <strong>of</strong> excited cation which loses a proton relative to <strong>the</strong> population<br />

which decays non-radiatively. This ratio is 1/2.97 <strong>and</strong> 1/5.39 for M3AB <strong>and</strong><br />

M3DMAB respectively.<br />

ril


At intermediate pHs i.e. 1.5 c pH .c 5.0, <strong>the</strong> fluorescence intensity is<br />

principally determined by <strong>the</strong> ground state equilibrium between <strong>the</strong> cation<br />

<strong>and</strong> neutral ester. The fluorescence intensity <strong>of</strong> both M3AB <strong>and</strong> M3DMAB is<br />

independent <strong>of</strong> pH in alkaline <strong>and</strong> neutral <strong>solution</strong>s but begins to decrease as<br />

pH is decreased, from approximately 6.0 until at a pH <strong>of</strong> approximately 3.0, a<br />

constant (relatively low) level <strong>of</strong> fluorescence is observed. The pKb* values<br />

determined from <strong>the</strong>se spectral changes are 4.27 <strong>and</strong> 4.04 for M3AB <strong>and</strong><br />

M3DMAB respectively if <strong>the</strong> above feedback step from <strong>the</strong> excited cation is<br />

neglected, <strong>and</strong> 5.00 <strong>and</strong> 4.88 respectively if it is included.<br />

It is noticeable that <strong>the</strong>se values are vastly different from <strong>the</strong> pK b *<br />

values calculated via a Forster-Weller cycle which are also included in <strong>the</strong><br />

Table 3.2. The latter suggest that <strong>the</strong> amino- substituents are much weaker<br />

bases in <strong>the</strong> excited state than in <strong>the</strong> ground state, which would be in accord<br />

with a first excited singlet state where <strong>the</strong>re is a degree <strong>of</strong> charge transfer from<br />

<strong>the</strong> nitrogen atom to <strong>the</strong> carbonyl functional. Over <strong>the</strong> pH range used here<br />

<strong>the</strong> excited cation would be expected to loose a proton to form <strong>the</strong> excited<br />

neutral ester <strong>and</strong> thus <strong>the</strong> observed fluorescence intensity should remain<br />

constant when excited at <strong>the</strong> isobestic point. That this is not observed<br />

indicates that <strong>the</strong> lifetime <strong>of</strong> <strong>the</strong> excited cation is extremely short such that<br />

loss <strong>of</strong> a proton can only partially compete with <strong>the</strong> o<strong>the</strong>r decay processes. The<br />

fluorescence properties are <strong>the</strong>refore largely decided by <strong>the</strong> ground state cation<br />

4-4 neutral equilibrium <strong>and</strong> <strong>the</strong> pKb* determined from <strong>the</strong> fluorescence<br />

spectra should be approximately <strong>the</strong> same as <strong>the</strong> ground state pKb. The<br />

discrepancy between <strong>the</strong> pKj, values calculated from <strong>the</strong> absorption <strong>and</strong><br />

fluorescence spectra is not insignificant: 3.52 versus 5.00 for M3AB <strong>and</strong> 3.79<br />

versus 4.88 for M3DMAB. The reason for this is not yet known.<br />

On decreasing <strong>the</strong> pH for <strong>the</strong> two <strong>acids</strong>, it is found that <strong>the</strong> fluorescence<br />

intensity is reduced in a similar manner for that <strong>of</strong> <strong>the</strong> two benzoates (Figure<br />

3.5). However, <strong>the</strong> situation is more complex here because <strong>the</strong> two <strong>acids</strong> can<br />

62


also deprotonate in addition to <strong>the</strong> protonation observed for <strong>the</strong>ir<br />

corresponding <strong>esters</strong>. As may be seen from Table 3.2 <strong>the</strong> two protic equilibria<br />

overlap in <strong>the</strong> ground state with pK a <strong>and</strong> pKb values which differ by (just)<br />

less than two pH units. Application <strong>of</strong> <strong>the</strong> Forster-Weller cycle to <strong>the</strong> excited<br />

state equilibria once again suggests that he two <strong>acids</strong>, like <strong>the</strong> two <strong>esters</strong>, are<br />

much weaker bases in <strong>the</strong> excited state, but that <strong>the</strong>ir acidity is only slightly<br />

less than in <strong>the</strong> ground state. The comments made previously with respect to<br />

<strong>the</strong> excited cation <strong>the</strong>refore also apply here i.e. that it should loose a proton to<br />

form <strong>the</strong> excited neutral species if this process is able to complete with <strong>the</strong><br />

o<strong>the</strong>r excited state decay processes open to <strong>the</strong> cation. The consequence <strong>of</strong> <strong>the</strong><br />

weakening <strong>of</strong> acidity in <strong>the</strong> excited state should bias <strong>the</strong> excited state<br />

distribution in favour <strong>of</strong> <strong>the</strong> excited neutral species at <strong>the</strong> expense <strong>of</strong> <strong>the</strong><br />

excited anion. However, over a pH range <strong>of</strong> relevance / <strong>the</strong> [H+] available is<br />

insufficient to protonate <strong>the</strong> excited anion within its lifetime <strong>and</strong> it is no<br />

surprise to find that <strong>the</strong> pKa* values for 3ABA <strong>and</strong> 3DMABA determined<br />

from variations in fluorescence properties as. a function <strong>of</strong> pH (Table 3.2) are<br />

very similar to <strong>the</strong> ground state pK as. The same is also true <strong>of</strong> <strong>the</strong> pKb* values<br />

determined experimentally. There is some discrepancy between <strong>the</strong> pK a *<br />

values calculated using <strong>the</strong> Forster-Weller cycle <strong>and</strong> <strong>the</strong> absorption or<br />

fluorescence spectral maxima, but this is probably just a reflection <strong>of</strong> <strong>the</strong><br />

several assumptions used in deriving <strong>the</strong> cycle.<br />

[*1


9<br />

pH8.45<br />

g pH 639<br />

pH5.6.5<br />

f.<br />

6. - - - - -<br />

pH 536<br />

ñ<br />

U pH6<br />

2<br />

305 355 405 455 505<br />

Wavelength (nnt)<br />

555<br />

Figure 3.5 fluorescence spectra <strong>of</strong> 3DMABA as a function <strong>of</strong> pH<br />

Along with Jian et alA 5, we have been unable to observe any evidence<br />

for zwitterionic species in 3ABA or 3DMABA. Our ground state absorption<br />

data as a function <strong>of</strong> solvent <strong>and</strong> our data collected for <strong>the</strong> neutral species in<br />

water are in good agreement with those reported by <strong>the</strong> latter authors. There<br />

are two obvious reasons for <strong>the</strong> apparent disagreement with <strong>the</strong> prediction <strong>of</strong><br />

Kuhn 43 <strong>and</strong> Tramer42. Firstly, <strong>the</strong> concentrated <strong>solution</strong>s used in <strong>the</strong> initial<br />

studies 42 ' 43 will favour dimer formation, because <strong>of</strong> <strong>the</strong> very low extinction<br />

coefficient for <strong>the</strong> zwitterion (see introduction). Secondly, <strong>the</strong> complex<br />

protonation / deprotonation properties <strong>of</strong> <strong>the</strong> <strong>acids</strong> make it impossible to<br />

obtain a <strong>solution</strong> <strong>of</strong> just <strong>the</strong> pure neutral or pure zwitterion species.<br />

Whenever <strong>the</strong> neutral acid is present, some proportion is also protonated <strong>and</strong><br />

I or deprotonated to complicate observation <strong>of</strong> any zwitterion.<br />

Estimates <strong>of</strong> <strong>the</strong> pKa* <strong>and</strong> pKb* values for <strong>the</strong> excited state equlibria<br />

from <strong>the</strong> fluorescence emission <strong>and</strong> absorption spectra using <strong>the</strong> Forster<br />

equation24 vary considerably, especially for <strong>the</strong> 3DMABA. The differences<br />

between <strong>the</strong> pKa*.values calculated from <strong>the</strong> absorption <strong>and</strong> fluorescence<br />

ErA'


spectra can be attributed to <strong>the</strong> differences in energy resulting from solvent<br />

reorientation in <strong>the</strong> ground <strong>and</strong> excited state (Figure 3.6).<br />

Excited molecule<br />

in ground state<br />

solvent environment<br />

Excited molecules<br />

state<br />

solvent environment<br />

Solvent<br />

adjust to accommodate<br />

excited molecules<br />

Energy<br />

Absorption<br />

TFluorescence<br />

Ground siate molecule<br />

in ground state solvent<br />

environment<br />

Solvent molecules<br />

adjust to minimise<br />

energy Ground state molecule SE<br />

in excited state solvent<br />

environment<br />

Figure 3.6 Energy diagram representing change in solvent orientation<br />

around ground <strong>and</strong> excited state molecules<br />

The energy <strong>of</strong> a solute / solvent system is determined to a large extent<br />

by electrostatic forces <strong>and</strong> so <strong>the</strong> interaction is greatest when a polar solute is<br />

used in a polar solvent. This leads to <strong>the</strong> greatest difference between <strong>the</strong> most<br />

stable <strong>and</strong> least stable configurations. When a molecule undergoes electronic<br />

excitation <strong>the</strong>re is usually a significant change in its polarity. Consequently,<br />

<strong>the</strong> distribution <strong>of</strong> solvent molecules which gives <strong>the</strong> most stable<br />

configuration for <strong>the</strong> ground state does not necessarily minimise <strong>the</strong> energy<br />

for <strong>the</strong> same molecule in <strong>the</strong> excited state. The absorption process is virtually<br />

instantanepus compared with molecular motion <strong>and</strong> <strong>the</strong> solvent molecules<br />

subsequently adjust to achieve <strong>the</strong> most stable configuration for <strong>the</strong> excited<br />

65


state with a slight reduction in <strong>the</strong> energy. It is from this state that <strong>the</strong><br />

molecule returns to <strong>the</strong> ground state where again <strong>the</strong> energy <strong>of</strong> <strong>the</strong> system is<br />

above <strong>the</strong> minimum because <strong>the</strong> distribution <strong>of</strong> <strong>the</strong> solvent molecules<br />

corresponds to <strong>the</strong> excited state situation. Subsequent adjustment <strong>the</strong>n<br />

reduces <strong>the</strong> energy to <strong>the</strong> value from which <strong>the</strong> excitation originally occurred<br />

<strong>and</strong> <strong>the</strong> ground state environment is restored. Consequently <strong>the</strong> emitted<br />

photon is <strong>of</strong> lower energy than that <strong>of</strong> <strong>the</strong> corresponding exciting photon <strong>and</strong><br />

<strong>the</strong>refore <strong>the</strong> emission wavelength appears at longer wavelengths than that<br />

<strong>of</strong> <strong>the</strong> absorption spectrum.<br />

The differences between pK a <strong>and</strong> pKa* <strong>and</strong> pKb <strong>and</strong> pKb* are worthy <strong>of</strong><br />

comment. The large decrease in pKb upon excitation is underst<strong>and</strong>able given<br />

that <strong>the</strong> excited state involves charge transfer from <strong>the</strong> amino nitrogen to <strong>the</strong><br />

carbonyl function. The electron density on <strong>the</strong> nitrogen is <strong>the</strong>refore reduced<br />

<strong>and</strong> it will be less basic. One might anticipate that <strong>the</strong> acid function would be<br />

stronger in <strong>the</strong> excited state 81 i.e. pKa* < pKa, but <strong>the</strong> reverse appears to be <strong>the</strong><br />

case. Once again <strong>the</strong> charge transfer from <strong>the</strong> nitrogen atom increases <strong>the</strong><br />

electron density on <strong>the</strong> carbonyl carbon <strong>and</strong> <strong>the</strong>reby reduces <strong>the</strong> acidity <strong>of</strong> <strong>the</strong><br />

acidic OH.<br />

3.3 fluorescence emission properties as a function <strong>of</strong> solvent.<br />

The fluorescence spectra for <strong>the</strong>se compounds are generally<br />

structureless <strong>and</strong> red shifted from <strong>the</strong> wavelength <strong>of</strong> maximum absorption by<br />

approximately 4 - 5,000 cm -1. Unlike <strong>the</strong> ground state absorption spectra, <strong>the</strong><br />

fluorescence spectra show a distinct trend with solvent polarity <strong>and</strong> hydrogen<br />

bonding ability. In strong hydrogen bonding or polar solvents, <strong>the</strong><br />

fluorescence emission maximum is red shifted. In such solvents, 3ABA <strong>and</strong><br />

M3AB emit in <strong>the</strong> region <strong>of</strong> 400 nm (Table 3.3), whilst <strong>the</strong> di<strong>methyl</strong>amino<br />

derivatives emit some 30 nm to <strong>the</strong> red. All four compounds fluoresce quite<br />

za


strongly (4 ~! 0.1) with only moderate variation from compound to<br />

compound <strong>and</strong> from solvent to solvent. Examples <strong>of</strong> <strong>the</strong> spectra are given in<br />

Figures 3.7 <strong>and</strong> 3.8.<br />

Solvent<br />

3ABA M3AB 3DMABA M3DMAB<br />

Quantity<br />

Species<br />

Acetonitrile Fluor Xrnax /nm 400 403 438 434<br />

0.27 0.35 0.25 0.25<br />

Of<br />

BuCN / iBuCN Huor, 390 392 425 420<br />

Of<br />

0.28 0.31 0.24 0.23<br />

* * *<br />

PhosXmax/nm *<br />

BCI 'jP Fluorkmax/rim 380 377 407 406<br />

0.33 0.35 0.33 0.36<br />

Of<br />

PhosXmax/nm 485 495 * 521<br />

Ethanol FluorXmax/nni 431 435 436 459<br />

0.24 0.26 0.19 0.16<br />

Of<br />

Hexanes FluorXmax/nm 370 354 378 378<br />

0.16 0.10 0.21 0.20<br />

Of<br />

MCH / iP FluorXmax/nm 380 362 395 388<br />

0.29 0.12 0.28 0.26<br />

Of<br />

PhosXmax/nm 425 509 * *<br />

Water Fluorlmax/nm 407 414 438 435<br />

0.40 0.32 0.26 0.26<br />

Of<br />

Neutral FluorXrnax/nm 407 355 440 435<br />

Anion Fluor kmax /nrn 415 -<br />

0.42 -<br />

Of<br />

* No measurable phosphorescence<br />

0.31 0.41 0.21 0.30<br />

Table 3.3 Excited state emission characteristics <strong>of</strong> 3-<strong>aminobenzoic</strong> <strong>and</strong> 3-<br />

N,N-di<strong>methyl</strong><strong>aminobenzoic</strong> <strong>acids</strong> <strong>and</strong> <strong>the</strong>ir <strong>methyl</strong> <strong>esters</strong> as a<br />

function <strong>of</strong> solvent <strong>and</strong> pH<br />

435<br />

0.23<br />

-<br />

-<br />

The only reported emission <strong>study</strong> <strong>of</strong> any <strong>of</strong> <strong>the</strong> meta isomers is by Jian<br />

et al. 45 (Table 3.3a), who measured <strong>the</strong> emission spectra <strong>of</strong> 3ABA as a<br />

function <strong>of</strong> solvent <strong>and</strong> pH, but did not measure <strong>the</strong> quantum yields. In <strong>the</strong><br />

solvents acetonitrile, ethanol <strong>and</strong> <strong>methyl</strong>cyclohexane, <strong>the</strong> results are in<br />

67


agreement with those obtained in <strong>the</strong> present work within <strong>the</strong> limits <strong>of</strong><br />

experimental error.<br />

Solvent (Fluorescence maxima in nm)<br />

Cyclohexane MCH Acetonitrile Ethanol 10-3 M HCI 10.2 M KOH<br />

367 385 396 428 400 1 403<br />

Table 3.3a fluorescence maxima <strong>of</strong> 3ABA as a function <strong>of</strong> solvent <strong>and</strong> pH,<br />

data reported by Jian et al .45<br />

There is a discrepancy with <strong>the</strong> anion <strong>and</strong> neutral species; as Jian et al.<br />

observed that <strong>the</strong> anionic species <strong>of</strong> 3ABA has a wavelength <strong>of</strong> maximum<br />

emission <strong>of</strong> 403 nm against our value <strong>of</strong> 415 nm. However <strong>the</strong> ground state<br />

pKs are in close agreement with those reported by Jian et al.45 (see table 3.2).<br />

9.00<br />

8.00<br />

6.00 .<br />

r<br />

I<br />

/<br />

3ABA<br />

M3AB<br />

3DMABA<br />

0)<br />

U<br />

5.00.<br />

'I<br />

- -M3DMAB<br />

0)<br />

3.00 ..<br />

2.00 / '<br />

1.00 . . .<br />

if<br />

N<br />

330 380 430 480<br />

Wavelength (nm)<br />

530 580<br />

Figure 3.7 Fluorescence emission spectra <strong>of</strong> <strong>the</strong> four meta compounds in<br />

acetonitrile<br />

It can be seen from Figures 3.7 <strong>and</strong> 3.8 that <strong>the</strong> difference in<br />

fluorescence emission maxima between <strong>the</strong> four compounds is very much<br />

dependent upon solvent, <strong>and</strong> that <strong>the</strong> <strong>methyl</strong> isomers behave in a very<br />

similar manner to each o<strong>the</strong>r, as do <strong>the</strong> two di<strong>methyl</strong>amino isomers. The<br />

M.


effect <strong>of</strong> esterification <strong>of</strong> <strong>the</strong> acid upon <strong>the</strong> fluorescence is minimal, but<br />

<strong>methyl</strong>ation <strong>of</strong> <strong>the</strong> nitrogen causes a dramatic change.<br />

2<br />

a,<br />

C -<br />

C<br />

a,<br />

S.<br />

0<br />

300 350 400 450 500 550<br />

Wavelength (nm)<br />

Figure 3.8 Fluorescence emission spectra <strong>of</strong> <strong>the</strong> four meta compounds in<br />

<strong>methyl</strong>cyclohexane I isopentane (4:1)<br />

3.4 fluorescence lifetime data as a function <strong>of</strong> solvent <strong>and</strong> pH<br />

Fluorescence decay pr<strong>of</strong>iles were measured for all four compounds in<br />

all <strong>of</strong> <strong>the</strong> seven solvents <strong>and</strong> solvent mixtures. It was found that <strong>the</strong> decay<br />

pr<strong>of</strong>iles varied in complexity with compound, solvent <strong>and</strong> pH. These results<br />

are collated in Table 3.4. The vast majority <strong>of</strong> <strong>the</strong> measured decays are clean<br />

single exponentials, with <strong>the</strong> exception <strong>of</strong> 3ABA <strong>and</strong> 3DMABA in <strong>the</strong> nonpolar<br />

solvents, where double exponential decay pr<strong>of</strong>iles are found. Lifetime<br />

values are in <strong>the</strong> region <strong>of</strong> 10 ns, rising to between 15 - 20 ns upon degassing.<br />

However <strong>the</strong> value observed for M3AB in hexane <strong>of</strong> 1.85 ns is unexpectedly<br />

low in comparison to that observed in a <strong>methyl</strong>cyclohexane / isopentane (4:1)<br />

mixture.


CMD Solvent V<br />

ti(ns) a1<br />

T2(')<br />

a2 'r.<br />

me an<br />

3ABA Acetonitrile 1.25 14.21<br />

BuCN/IBuCN 1.11 18.36<br />

BC1 / iF 1.02 16.87<br />

Ethanol 1.51 14.02<br />

Hexane 1.31 1.75 0.7061 4.35 0.2939 2.51<br />

MCH I iF 1.09 3.72 0.7973 16.66 0.2027 6.34<br />

Water 1.67 21.59<br />

Anion 1.12 23.78<br />

____ Neutral 1.12 23.07<br />

M3AB Acetonitrile 1.08 9.97<br />

BuCN / IBuCN 0.98 9.11<br />

BCI / iF 1.13 9.71<br />

Ethanol 1.22 13.76<br />

Hexane 1.35 1.85<br />

MCH / iF 1.07 9.47<br />

_ Water 1.29 23.82<br />

3DMABA Acetonitrile 1.19 13.13<br />

BuCN / IBuCN 1.02 19.19<br />

BC1 / i-P 1.05 19.42<br />

Ethanol 1.35 14.07<br />

Hexane 0.88 2.84 0.2046 6.44 0.7954 5.70<br />

MCH / iF 1.16 5.40 0.2446 9.49 0.7554 8.49<br />

Water 1.12 22.20<br />

Anion 1.11 20.31<br />

Neutral 1.10 11.88<br />

M3DMAB Acetonitrile 1.08 13.22<br />

BuCN / iBuCN 1.83 14.54<br />

BCI / iP 1.08 9.40<br />

Ethanol 1.14 14.29<br />

Hexane 1.40 4.95<br />

MCH / iF 1.04 7.66<br />

Water 11.27 22.45<br />

Table 3.4<br />

Fluorescence decay pr<strong>of</strong>iles as a function <strong>of</strong> solvent <strong>and</strong> pH for<br />

<strong>the</strong> four ineta compounds in all solvents at room temperature<br />

The non-polar solvents clearly provide a less than ideal solvent<br />

environment for <strong>the</strong> two <strong>acids</strong> as <strong>the</strong>re are not only solubility problems but<br />

also <strong>the</strong>se are <strong>the</strong> only cases where a single exponential decay is not observed<br />

for <strong>the</strong>se compounds. Aggregation is well known in <strong>the</strong>se solvents 67 <strong>and</strong> this<br />

may explain not only <strong>the</strong> double exponential fits required for 3ABA <strong>and</strong><br />

3DMABA in hexane <strong>and</strong> <strong>methyl</strong>cyclohexane I isopentane, but also <strong>the</strong><br />

generally low lifetimes for M3AB <strong>and</strong> M3DMAB in <strong>the</strong>se solvents.<br />

70


"Water, a singularly effective ionising solvent on account <strong>of</strong> its high<br />

dielectric constant <strong>and</strong> its ion-solvating ability, is able to stabilise <strong>the</strong> anionic<br />

state <strong>of</strong> <strong>the</strong> aromatic <strong>aminobenzoic</strong> <strong>acids</strong> through hydrogen bonding to <strong>the</strong><br />

nitrogen groups as well as possibly through <strong>the</strong> carboxylic group" 82. In <strong>the</strong><br />

light <strong>of</strong> this statement it is perhaps surprising that <strong>the</strong> decay pr<strong>of</strong>iles observed<br />

in <strong>the</strong> polar <strong>and</strong> hydrogen bonding solvents are not also more complex. It is<br />

<strong>the</strong>refore probable that <strong>the</strong>re is one dominant solute / solvent arrangement<br />

<strong>and</strong> this is <strong>the</strong> one which is observed in <strong>the</strong> fluorescence. On <strong>the</strong> basis <strong>of</strong> <strong>the</strong><br />

pKa values calculated earlier <strong>and</strong> <strong>the</strong> lifetime data presented in Table 3.4, it is<br />

clear that both 3ABA <strong>and</strong> 3DMABA exist as <strong>the</strong> anion in unbuffered water.<br />

The absence <strong>of</strong> any complexity in <strong>the</strong> aqueous decay pr<strong>of</strong>iles backs up <strong>the</strong><br />

earlier conclusions about <strong>the</strong> pKa* values <strong>and</strong> <strong>the</strong> ability <strong>of</strong> <strong>the</strong> excited anion<br />

to gain a proton within its lifetime.<br />

The lifetime <strong>and</strong> quantum yield data has been combined to calculate<br />

rate constants for <strong>the</strong> radiative (fluorescence K1) <strong>and</strong> non-radiative (K nr) decay<br />

processes in <strong>the</strong> four molecules as a function <strong>of</strong> solvent (Table 3.5). The<br />

values observed for both rate constants are quite remarkably similar for all<br />

<strong>the</strong> solvents employed here if <strong>the</strong> K nr values in hexane <strong>and</strong><br />

<strong>methyl</strong>cyclohexane / isopentane (4:1) are omitted. This perhaps reflects <strong>the</strong><br />

relative lack <strong>of</strong> interaction between <strong>the</strong> two substituents on <strong>the</strong> benzene ring<br />

when <strong>the</strong>y are meta- orientated.<br />

71


— —<br />

—<br />

I-<br />

C) — l LI) N en 'oa'<br />

c4r-enen<br />

'r in<br />

0 N0'<br />

— LI) C'I ('I l Q en'o<br />

rCLtS2eliO<br />

x<br />

_ - -<br />

cq cn<br />

1!<br />

o C)'C 'ON<br />

r-J z<br />

a)<br />

-'-a<br />

Ct<br />

I- -----<br />

C)<br />

01C\LI)'0C'IN<br />

Ct NCflflNri<br />

.Ii,-LI)rLLi<br />

Lci<br />

ci)<br />

•<br />

CCI —<br />

—<br />

.Ct<br />

CtC5<br />

I U)<br />

—<br />

a)<br />

C-)<br />

C)<br />

0-.<br />

U)<br />

I-i<br />

C)<br />

C')<br />

U)<br />

C)<br />

C,<br />

C<br />

I-.<br />

C)<br />

'C<br />

I<br />

Ct<br />

I-i<br />

C % ON 0' N l<br />

ca<br />

Qreii et'6 4 vi ceá<br />

Ct<br />

t<br />

cqZ<br />

c)<br />

• eflC'1 ON LO'0 00 O<br />

a) M aN qcr LO t40' Lnen<br />

> enenNentI)<br />

Ii<br />

o<br />

0<br />

C)<br />

hi<br />

.• O ,t —('4 O. 0\N<br />

c Q'ri M LO ON 00 '<br />

Lei m'o V-itt)<br />

a.,<br />

1.)<br />

o -,<br />

en<br />

en<br />

•<br />

.<<br />

_<br />

I-)<br />

en<br />

I- — C<br />

-C<br />

Cc<br />

C)<br />

U) L.<br />

0_C<br />

U -s<br />

•d-aCt<br />

ci<br />

H<br />

—I<br />

Ct<br />

-a<br />

Ct<br />

I-<br />

C)<br />

'4-<br />

C<br />

C')<br />

C<br />

CC<br />

C-,<br />

0 C)<br />

'a. E c;<br />

C-C<br />

hi<br />

Ct —()<br />

-C<br />

tt_C a)<br />

Ct<br />

-:1<br />

Ct•_<br />

r'a<br />

0<br />

c'S<br />

•.<br />

U)<br />

•2<br />

a<br />

V–li<br />

U 2 gr<br />

hi<br />

Ct<br />

CC<br />

hi<br />

a)<br />

-c<br />

72


3.5 fluorescence emission properties as a function <strong>of</strong> temperature<br />

For variable temperature measurements only <strong>the</strong> three binary solvent<br />

mixtures were used, <strong>and</strong> <strong>the</strong>se were chosen for <strong>the</strong>ir ability to form good<br />

glasses at low temperature. The results are presented in Table 3.7. Although<br />

no new information was expected to be gained from this experiment, it was<br />

undertaken to enable a comparison to be made with <strong>the</strong> results obtained for<br />

<strong>the</strong> para- isomers where temperature effects on <strong>the</strong> normal <strong>and</strong> anomalous<br />

fluorescence were investigated.<br />

Cmd Solvent Quantumyield as a function <strong>of</strong> temperature<br />

293K 230K 193K 153K 133K 113K<br />

3ABA BuCN / iBuCN 0.28 0.50 0.55 0.58 0.58 0.58<br />

BC! / i-P 0.33 0.42 0.30 0.54 0.61 0.65<br />

MCH / iP 0.29 0.58 0.64 0.22 0.64 0.64<br />

M3AB BuCN / iBuCN 0.31 0.47 0.56 0.60 0.62 0.62<br />

BCI / iF 0.35 0.50 0.71 0.87 0.99 1.00<br />

MCH / iF 0.12 0.16 0.26 0.31 0.27 0.47<br />

3DMABA BuCN / IBuCN 0.24 0.41 0.43 0.42 0.39 0.39<br />

BCI / i-P 0.33 0.38 0.24 0.25 0.35 0.38<br />

MCH / iP 0.28 0.43 0.38 0.34 0.20 0.19<br />

M3DMAB BuCN I 1BuCN 0.23 0.47 0.56 0.53 0.54 0.59<br />

BCI / iF 0.36 0.69 0.82 0.84 0.86 0.86<br />

MCH / iF 0.26 0.31 0.39 0.43 0.43 0.44<br />

Table 3.7<br />

Excited state emission characteristics <strong>of</strong> <strong>the</strong> four meta<br />

compounds as a function <strong>of</strong> solvent <strong>and</strong> temperature<br />

In general, <strong>the</strong> fluorescence spectra <strong>of</strong> all <strong>the</strong> compounds broaden with<br />

lowering <strong>of</strong> temperature with an increase in <strong>the</strong> fluorescence quantum yield,<br />

which in some cases reaches nearly unity. These data have to be viewed with<br />

some caution however as <strong>the</strong>re has been no attempt to correct for absorption<br />

changes as a function <strong>of</strong> temperature. This increase in quantum yie!d imp!ies<br />

<strong>the</strong> presence <strong>of</strong> an activation barrier <strong>of</strong> some sort. Since fluorescence rate<br />

constants are usually temperature independent, <strong>the</strong> energy barrier is<br />

presumably associated with a competing radiationless process. It appears that<br />

this process is intersystem crossing 83 from Si to a higher triplet T n which is<br />

73


approximately degenerate with S1. The effect <strong>of</strong> temperature is to increase <strong>the</strong><br />

population <strong>of</strong> higher vibrational <strong>and</strong> rotational sub-levels <strong>of</strong> S1 from which<br />

faster rates <strong>of</strong> intersystem crossing may occur. Hence as <strong>the</strong> temperature rises<br />

<strong>the</strong> rate <strong>of</strong> intersystem crossing increases <strong>and</strong> consequently a smaller<br />

proportion <strong>of</strong> s1 molecules have an opportunity to fluoresce, <strong>and</strong> <strong>the</strong><br />

quantum yield <strong>of</strong> fluorescence will decrease accordingly. The wavelength <strong>of</strong><br />

maximum emission also red shifts. This shift is very dependent on solvent<br />

<strong>and</strong> substituent group. In chiorobutane <strong>and</strong> butyronitrile, <strong>the</strong> shift is <strong>of</strong> <strong>the</strong><br />

order <strong>of</strong> 5 rim, whilst in <strong>the</strong> inert solvent <strong>methyl</strong>cyclohexane <strong>the</strong> shift is as<br />

much as 15 nm. Unfortunately <strong>the</strong> accuracy <strong>of</strong> <strong>the</strong> wavelength measurements<br />

is +1- 5 nm which may somewhat disguise <strong>the</strong> real wavelength <strong>of</strong> maximum<br />

emission (see for example Figure 3.9).<br />

250000<br />

293K<br />

> 200000<br />

-a<br />

C<br />

4,<br />

U<br />

150000<br />

2<br />

0<br />

315 335 355 375 395 415 435 455 475 495 515<br />

Wavelength (nm)<br />

Figure 3.9 fluorescence spectra <strong>of</strong> M3AB in chiorobutane I isopentane (9:1)<br />

as a function <strong>of</strong> temperature<br />

1<br />

In <strong>the</strong> non-polar <strong>methyl</strong>cyclohexane, 3ABA <strong>and</strong> 3DMABA show <strong>the</strong><br />

appearance <strong>of</strong> a secondary peak at 400 rim which is not observed in <strong>the</strong> o<strong>the</strong>r<br />

solvents used. This second peak is observed at temperatures as high as 230K<br />

(Figure 3.10 <strong>and</strong> 3.11) <strong>and</strong> may indicate <strong>the</strong> onset <strong>of</strong> aggregation.<br />

n<br />

74


—<br />

7000.00<br />

6000.00<br />

• 5000.00<br />

C<br />

0?<br />

.E 4000.00<br />

C,<br />

o<br />

I<br />

+<br />

i<br />

+<br />

30000 4 .<br />

U,<br />

(I<br />

I- al/I<br />

2000.00<br />

1 000.00<br />

Nt"<br />

293K<br />

-----230K<br />

193K<br />

------153 K<br />

133K<br />

-------113K<br />

77K<br />

0.00<br />

280.00 330.00 380.00 430.00<br />

Wavelength (mu)<br />

480.00 530.00<br />

Figure 3.10 Fluorescence spectra <strong>of</strong> 3ABA in <strong>methyl</strong>cyclohexane /<br />

isopentane (4:1) as a function <strong>of</strong> temperature.<br />

700000<br />

293K<br />

230K<br />

C)<br />

400000<br />

193K<br />

153K<br />

—- 133K<br />

113K<br />

g200000<br />

77K<br />

W.<br />

100000<br />

0*<br />

330<br />

380 430 480 530<br />

Wavelength (urn)<br />

Figure 3.11 Fluorescence spectra <strong>of</strong> 3DMABA in <strong>methyl</strong>cyclohexane /<br />

isopentane (4:1) as a function <strong>of</strong> temperature.<br />

The largest change in wavelength <strong>of</strong> maximum emission is observed<br />

for M3AB in <strong>methyl</strong>cyclohexane. Although phosphorescence might be<br />

expected at 77K, none is seen. This feature fits well with <strong>the</strong> observed increase<br />

75


in o f as <strong>the</strong> temperature is decreased, particularly if <strong>the</strong> interpretation<br />

advanced a little earlier regarding decreased inter-system crossing is <strong>the</strong><br />

correct one.<br />

3.6 . Summary <strong>of</strong> conclusions<br />

From absorption spectroscopic studies <strong>the</strong> pK a <strong>and</strong> pKb values have<br />

been found for <strong>the</strong> four compounds, anlong with <strong>the</strong>ir ground state<br />

absorption properties as a function <strong>of</strong> solvent. Although <strong>the</strong>re is evidence<br />

from <strong>the</strong> spectra <strong>of</strong> 3ABA <strong>and</strong> 3DMABA that two species may be present in<br />

hexane <strong>and</strong> <strong>methyl</strong>cyclohexane / isopentane (4:1). This is a possible indication<br />

<strong>of</strong> ground state dimerisation but concentration measurements have not been<br />

undertaken to check this. If this is <strong>the</strong> case, <strong>the</strong>n some evidence for this might<br />

be expected in <strong>the</strong> steady state room temperature emission spectra. The<br />

fluorescence emission spectra <strong>of</strong> 3ABA <strong>and</strong> 3DMABA at lower temperatures<br />

do however appear to show <strong>the</strong> appearance <strong>of</strong> ano<strong>the</strong>r peak at 400 nm. The<br />

lowering <strong>of</strong> temperature may well favour formation <strong>of</strong> dimers or higher<br />

polymers 69. There are several o<strong>the</strong>r excited state possibilities such as excimer<br />

or exciplex formation which could account for <strong>the</strong> new peak, but <strong>the</strong>re is no<br />

emission evidence for <strong>the</strong>se except at lower temperatures. Fur<strong>the</strong>r data from a<br />

<strong>study</strong> <strong>of</strong> <strong>the</strong> concentration dependence <strong>of</strong> <strong>the</strong>se low temperature spectra (both<br />

absorption <strong>and</strong> fluorescence) coupled with low temperature fluorescence<br />

decay measurements will help to resolve <strong>the</strong> origin <strong>of</strong> <strong>the</strong> new b<strong>and</strong>s.<br />

76


Chapter 4


Chapter 4 Photophysical properties <strong>of</strong> 2-<strong>aminobenzoic</strong> acid, <strong>and</strong> 2-N,N-di<strong>methyl</strong><strong>aminobenzoic</strong><br />

acid <strong>and</strong> <strong>the</strong>ir <strong>methyl</strong> <strong>esters</strong><br />

4.1 Ground state absorption spectra as a function <strong>of</strong> solvent.<br />

Compared to <strong>the</strong> meta- isomers discussed in chapter 3, <strong>the</strong> absorption<br />

properties <strong>of</strong> <strong>the</strong> ortho- isomers are much more complex. However <strong>the</strong>y have<br />

been studied in greater detail by several authors notably Kuhn <strong>and</strong> Geider 43,<br />

Tramer 42' 44, Mataga <strong>and</strong> Ottolenghi 40, Doub <strong>and</strong> V<strong>and</strong>enbelt 38, Leggate <strong>and</strong><br />

Dunn 41 <strong>and</strong> Jian et al. 45 . The ortho- <strong>and</strong> para- <strong>isomeric</strong> amino <strong>and</strong> N,Ndi<strong>methyl</strong><strong>aminobenzoic</strong><br />

<strong>acids</strong> <strong>and</strong> <strong>the</strong>ir <strong>methyl</strong> <strong>esters</strong> differ from <strong>the</strong>ir metacounterparts<br />

in that <strong>the</strong> lone pair <strong>of</strong> electrons on <strong>the</strong> nitrogen are able to<br />

interact with <strong>the</strong> carbonyl group via <strong>the</strong> aromatic ring system <strong>and</strong><br />

consequently <strong>the</strong> excited states can involve some form <strong>of</strong> resonance<br />

interaction (stabilisation).<br />

As well as <strong>the</strong> more usual inter-molecular hydrogen bonds, <strong>the</strong> orthoisomers<br />

can also form intra-molecular hydrogen bonds. The close proximity<br />

<strong>of</strong> <strong>the</strong> two substituent groups leads to steric hindrance, so that each group will<br />

have an influence on <strong>the</strong> o<strong>the</strong>r <strong>and</strong> so may be twisted out <strong>of</strong> <strong>the</strong> plane <strong>of</strong> <strong>the</strong><br />

aromatic it system. This may well affect <strong>the</strong> amount <strong>of</strong> conjugation between<br />

<strong>the</strong> nitrogen lone pair <strong>and</strong> <strong>the</strong> carbonyl function <strong>and</strong> consequently affect <strong>the</strong><br />

amount <strong>of</strong> charge transfer between <strong>the</strong> two substituents.<br />

In all <strong>of</strong> <strong>the</strong> pure solvents <strong>and</strong> binary mixtures used in this <strong>study</strong>, 2-<br />

<strong>aminobenzoic</strong> acid (2ABA), <strong>methyl</strong> 2-aminobenzoate (M2AB), 2-N,Ndi<strong>methyl</strong><strong>aminobenzoic</strong><br />

acid (2DMABA) <strong>and</strong> <strong>methyl</strong> 2-N,N-di<strong>methyl</strong>aminobenzoate<br />

(M2DMAB), generally show structureless absorption spectra with a<br />

wavelength <strong>of</strong> maximum absorption ranging from 280 to 345 nm, depending<br />

upon <strong>the</strong> polarity <strong>of</strong> <strong>the</strong> solvent, its hydrogen bonding ability <strong>and</strong> on <strong>the</strong><br />

substituents (Figures 4.1 <strong>and</strong> 4.2 <strong>and</strong> Table 4.1).<br />

ViA


AVA<br />

0.6<br />

03<br />

L<br />

0<br />

0.4<br />

I<br />

o<br />

I •<br />

20.31-<br />

< I<br />

0.24-<br />

'-<br />

01<br />

<strong>of</strong><br />

Pt'<br />

I<br />

•7•- N.<br />

Acetonitrile<br />

-----Ethanol<br />

Hexane<br />

Water<br />

235 255 275 295 315 335 355 375 395<br />

Wavelength (nm)<br />

Figure 4.1 Absorption spectrum <strong>of</strong> 2ABA in all four pure solvents<br />

Solvent<br />

Measured 2ABA M2AB 2DMABA M2DMAB<br />

Species<br />

quantity<br />

Acetonitrile Am /run 336 337 281 336<br />

El dm3 moF 1 cm4 4,700 5,200 800 2,600<br />

BuCN / IBuCN 338 337 283 339<br />

8/ dm3 mol4 car1 4,100 6,000 1,200 3,200<br />

BCI / iF Xmax 337 335 289 335<br />

________________ 8/ dm3 mol-1 y 1 5,000 5,300 1,000 3,300<br />

Ethanol A.max /p'n 335 339 271 339<br />

6/ dm3 mo!4 y1 4,400 4,900 800 2,700<br />

Hexane kmax /nm 338 333 281 336<br />

8/ dm3mol'cnr' 4,900 5,000 800 3,100<br />

MCH / i1' Xmax /nrn 336 333 287 336<br />

________________ 8/ dm3 mol-1 3,500 6,600 1.000 2,800<br />

Water Xmax /rm 314 327 270 327<br />

1,900 3,600 700 1,400<br />

Neutral c/ dm3 mo!-1 car1 326 326 300 328<br />

Xmax /nrrt 2,100 3,900 700 1,400<br />

Anion Xmax 310 - 260 -<br />

8/ dm3 mo!4 cm-1 3,000 900 -<br />

Cation I Xmax /nm 272 272 272 272<br />

2/ dm3 mo!4 -1 940 840 500 960<br />

in


Table 4.1 Primary ground state absorption characteristics <strong>of</strong> 2-<br />

<strong>aminobenzoic</strong>, 2-N,N-di<strong>methyl</strong><strong>aminobenzoic</strong> acid <strong>and</strong> <strong>the</strong>ir<br />

<strong>methyl</strong> <strong>esters</strong> as a function <strong>of</strong> pH <strong>and</strong> solvent<br />

'Ii<br />

DIIQ<br />

0.07 AcetonUrile<br />

0.06 -----Ethanol<br />

C 0.05 Is<br />

/ Hexane<br />

1<br />

--<br />

Water<br />

'C<br />

0.04<br />

0.03<br />

V.<br />

0.02 S.<br />

I-. .<br />

0.01<br />

------.<br />

-. -. --<br />

0 I I I I I I<br />

250 260 270 280 290 300 310 320 330 340 350<br />

Wavelength (nm)<br />

Figure 4.2 Ground state absorption spectra <strong>of</strong> 2DMABA in all four pure<br />

solvents<br />

Compared to <strong>the</strong> corresponding meta isomers, 2ABA <strong>and</strong> M2AB<br />

absorption b<strong>and</strong>s are red shifted by some 15 - 30 nm <strong>and</strong> exhibit enhanced<br />

extinction coefficients usually by a factor <strong>of</strong> 2 - 3. However, <strong>the</strong> absorption<br />

b<strong>and</strong>s in <strong>the</strong> ortho- isomers are not as intense as those for <strong>the</strong> para-isomers,<br />

probably reflecting <strong>the</strong> steric crowding in <strong>the</strong> former. The 2-N,Ndi<strong>methyl</strong>amino<br />

compounds are clearly very different to both <strong>the</strong>ir rneta- <strong>and</strong><br />

para- counterparts. The usual red shift <strong>of</strong> some 20 nm when <strong>the</strong> amino group<br />

is <strong>methyl</strong>ated is absent in <strong>the</strong> ortho- compounds <strong>and</strong>, in <strong>the</strong> case <strong>of</strong> 2DMABA,<br />

is replaced by a 50 nm blue shift. Extinction coefficients in 2DMABA <strong>and</strong><br />

M2DMAB are much lower than might be expected, reflecting not only steric<br />

crowding but also apparently intra-molecular hydrogen bonding in <strong>the</strong> case <strong>of</strong><br />

2DMABA.<br />

79


In all <strong>of</strong> <strong>the</strong> solvents <strong>and</strong> binary mixtures studied, 2DMABA exhibits a<br />

more structured spectrum, with two peaks in <strong>the</strong> 265 - 280 nm region,<br />

generally split by 10 nm (see for example Figure 4.2). In water <strong>and</strong> ethanol<br />

ano<strong>the</strong>r peak is observed around 300 <strong>and</strong> 315 nm respectively. In hexane <strong>and</strong><br />

acetonitrile it is also possible to identify ano<strong>the</strong>r peak around 350 nm (c


secondary absorption b<strong>and</strong> in <strong>the</strong> 220-250 nm range, which varies<br />

considerably in position <strong>and</strong> strength, depending on compound <strong>and</strong> solvent.<br />

The absorption properties <strong>of</strong> M2AB are very similar to those <strong>of</strong> its parent add<br />

(Table 4.1). 2ABA <strong>and</strong> M2AB have fairly strong extinction coefficients <strong>of</strong><br />

around 4-5,000 dm 3 mol-' cm -', whilst 2DMABA <strong>and</strong> M2DMAB have much<br />

smaller extinction coefficients, <strong>of</strong> <strong>the</strong> order <strong>of</strong> 2-3,000 dm 3 mol-' cm -1 for <strong>the</strong><br />

benzoate <strong>and</strong> 700 - 1,200 dm 3 mol-1 cm -1 for <strong>the</strong> acid. This difference between<br />

<strong>the</strong> amino <strong>and</strong> N,N-di<strong>methyl</strong>amino substituent group can be explained in<br />

terms <strong>of</strong> <strong>the</strong> amount <strong>of</strong> steric hindrance between <strong>the</strong> substituent groups in<br />

each compound. Depending on <strong>the</strong> orientation <strong>of</strong> <strong>the</strong> carboxylic acid<br />

substituent, esterification will have little or no effect, but replacement <strong>of</strong> <strong>the</strong><br />

hydrogen atoms on <strong>the</strong> nitrogen by <strong>methyl</strong> groups will make <strong>the</strong> nitrogen<br />

lone pair more available for bonding by increasing <strong>the</strong> basicity <strong>of</strong> <strong>the</strong> amino<br />

group, but also will potentially result in <strong>the</strong> N,N-di<strong>methyl</strong>amino group being<br />

twisted out <strong>of</strong> plane, thus decreasing <strong>the</strong> extent <strong>of</strong> conjugation <strong>and</strong><br />

consequently <strong>the</strong> amount <strong>of</strong> charge transfer 17-20. For 2DMABA <strong>the</strong>re is <strong>the</strong><br />

added complication <strong>of</strong> <strong>the</strong> hydroxy group which creates a potential for two<br />

forms <strong>of</strong> hydrogen bonding (see Figure 44).<br />

10 o<br />

H<br />

0% /°H 0% /°H<br />

c<br />

I<br />

H<br />

Me<br />

C/ "H or \ H but only Me for ZDMABA<br />

Figure 4.4 Possible hydrogen-bonded structures for <strong>the</strong> ortho <strong>acids</strong> in inert<br />

solvents<br />

In terms <strong>of</strong> <strong>the</strong> variation <strong>of</strong> <strong>the</strong> wavelength <strong>of</strong> maximum absorption as<br />

a function <strong>of</strong> solvent, 2DMABA is a special case, being blue shifted some 40 -<br />

50 nm from 2ABA, M2AB <strong>and</strong> M2DMAB, which all show similar<br />

wavelengths <strong>of</strong> maximum absorption in a given solvent. M2DMAB has<br />

EjI


extinction coefficients approximately one half <strong>of</strong> those observed for 2ABA<br />

<strong>and</strong> M2AB. This as previously described can be attributed to steric hindrance<br />

between <strong>the</strong> carboxyl <strong>and</strong> N,N-di<strong>methyl</strong>amino groups which twists one or<br />

both <strong>of</strong> <strong>the</strong>m out <strong>of</strong> plane. Consequently a lower extinction coefficient would<br />

be expected as a result <strong>of</strong> loss <strong>of</strong> conjugation. 2DMABA however appears to<br />

exhibit anomalous properties having very low extinction coefficients <strong>of</strong> <strong>the</strong><br />

order <strong>of</strong> 1,000 dm3 mol-' cm-1 <strong>and</strong> a structured absorption spectrum. 2DMABA<br />

is behaving in a similar manner to benzene which has a primary absorption<br />

b<strong>and</strong> at 230 nm 16 <strong>and</strong> a secondary structured absorption b<strong>and</strong> around 280 nm<br />

(C = 1,045 dm3 mol- 1 cm-1). In 1969, Tramer42,44 investigated <strong>the</strong> absorption<br />

properties <strong>of</strong> 2-<strong>aminobenzoic</strong> acid <strong>and</strong> two <strong>of</strong> its derivatives namely, N-<br />

<strong>methyl</strong> 2-<strong>aminobenzoic</strong> add <strong>and</strong> N,N-di<strong>methyl</strong><strong>aminobenzoic</strong> acid by infrared<br />

<strong>and</strong> ultraviolet spectroscopy 42 <strong>and</strong> later by fluorescence emission<br />

spectroscopy 44. He showed that in <strong>the</strong> molecular form (Figure 1.13) none <strong>of</strong><br />

<strong>the</strong> three compounds form an intra-molecular hydrogen bond between <strong>the</strong><br />

carboxyl <strong>and</strong> amino groups. In <strong>the</strong> quasi-zwitterionic form (Figure 1.13) a<br />

strong intra-molecular hydrogen bond is formed such that it changes <strong>the</strong><br />

conformation <strong>of</strong> <strong>the</strong> amino group <strong>and</strong> removes <strong>the</strong> conjugation between <strong>the</strong><br />

nitrogen lone pair <strong>and</strong> <strong>the</strong> it-electron system.<br />

Literature data on 2ABA <strong>and</strong> 2DMABA absorption properties are<br />

reproduced in Table 4.1a. As can be seen by comparing <strong>the</strong> literature values<br />

with those in Table 4.1, <strong>the</strong> results are in excellent agreement within<br />

experimental error. However, Tramer 44 has reported a peak in <strong>the</strong> region <strong>of</strong><br />

345 nm with a quoted extinction coefficient <strong>of</strong> around 85 dm 3'mol- 1 cm 1 in<br />

acetonitrile. We were unable to observe this b<strong>and</strong>, but this is not surprising<br />

given its very weak intensity. At <strong>the</strong> concentrations which would be<br />

necessary to observe this b<strong>and</strong> (~ 1 mmol dm 3), <strong>the</strong> probability <strong>of</strong> aggregation<br />

<strong>and</strong> dimer formation must be high.<br />

IM


Solvent<br />

Measured<br />

quantity<br />

Compound<br />

2ABA<br />

2DMABA<br />

17/ 2nd B<strong>and</strong><br />

Acetonitrile )Ur.ax (rim) 334 346 / 280<br />

£ moF 1 cm 1 dm - 857 11000 "<br />

Cydohexane ?tjnax (nm) 349 / 281"<br />

8 mo1 1 cm 1 dm3 14 / 950<br />

Ethanol )m,< (rim) 336<br />

2 mo!-1 cm-1 dr& -<br />

Water Lnax flI<br />

327 322 / 270 IT<br />

£ mo1 1 cm4 dm3 1,900 40 15 / 1,000<br />

Table 4.1a Absorption maxima <strong>of</strong> 2ABA <strong>and</strong> 2DMABA as a function <strong>of</strong><br />

solvent <strong>and</strong> pH, as reported by various authors<br />

4.2 Ground state absorption <strong>and</strong> fluorescence emission properties as a<br />

function <strong>of</strong> pH.<br />

Measurement <strong>of</strong> absorption spectra in water as a function <strong>of</strong> pH for <strong>the</strong><br />

four ortho- compounds allows <strong>the</strong> calculation <strong>of</strong> pKb <strong>and</strong> pICa values for <strong>the</strong><br />

two <strong>acids</strong>. Sample spectra are shown in Figure 4.5 for 2ABA.<br />

0.8 p1-I 7.25<br />

0.7 — —pH 6.00<br />

0.6 pH 5.00<br />

0.5 — - — - — - pH400<br />

250 270 290 310 330 350 370 390<br />

Wavelength (rim)<br />

410<br />

Figure 4.5 Absorption spectra <strong>of</strong> 2ABA as a function <strong>of</strong> pH<br />

As can be seen from Figure 4.5, as <strong>the</strong> pH decreases from around 6.0 <strong>the</strong><br />

absorption begins to diminish <strong>and</strong> becomes red shifted, until at pH 4 <strong>the</strong><br />

XV


absorption maximises at approximately 326 nm <strong>and</strong> ano<strong>the</strong>r new b<strong>and</strong> begins<br />

to appear at 272 nm, which is associated with <strong>the</strong> cation. At this pH it would<br />

be expected to have 98.5% <strong>of</strong> <strong>the</strong> neutral species <strong>and</strong> 1.5% <strong>of</strong> <strong>the</strong> anion. On<br />

lowering <strong>the</strong> pH to a value <strong>of</strong>


ut it is clear that <strong>the</strong> value quoted by Jian et aL 45 for <strong>the</strong> pKb <strong>of</strong> 2.55 is clearly<br />

<strong>the</strong> "odd man out" There is apparently no report in <strong>the</strong> literature regarding<br />

<strong>the</strong> pKb values for M2AB <strong>and</strong> M2DMAB, but <strong>the</strong> values given in Table 4.2 for<br />

<strong>the</strong>se two quantities are similar to <strong>the</strong> pKb values for <strong>the</strong> parent <strong>acids</strong>.<br />

Comparison <strong>of</strong> our pK data for 2DMABA with <strong>the</strong> works <strong>of</strong> Tramer 44<br />

<strong>and</strong> Kuhn43 is difficult because <strong>of</strong> our inability to distinguish separate<br />

absorption features for <strong>the</strong> zwitterion <strong>and</strong> neutral species which <strong>the</strong>y believe<br />

to exist. However, Tramer's results 44 are shown in Table 4.2a<br />

equilibrium pK pK*<br />

pKBcPKBC*<br />

- 1.5<br />

PKCM 4 -11.5<br />

PKcz 1.4 2.7<br />

pKMA 6 10.0<br />

pKm 8.58 0.3<br />

pK1 -2.6 9.8<br />

Table 4.2a pK values obtained by Tramer 42' 44 calculated from Figure 1.13<br />

Tramer 44 states that <strong>the</strong> zwitterionic form <strong>of</strong> 2DMABA absorbs at 270<br />

run <strong>and</strong> that <strong>the</strong> neutral species absorbs at 322 nm. This argument is based on<br />

<strong>the</strong> fact that <strong>the</strong> ground state absorption spectrum <strong>of</strong> 2DMABA is almost<br />

identical with that <strong>of</strong> benzoic acid, which is considered to be a strong<br />

argument in favour <strong>of</strong> a zwitterionic structure, because <strong>the</strong> influence <strong>of</strong> <strong>the</strong><br />

quaternary ammonium group on <strong>the</strong> spectrum <strong>of</strong> benzene derivatives is<br />

insignificant, compared to that <strong>of</strong> an alkyl group 44. Tramer appears to have<br />

missed <strong>the</strong> fact that if <strong>the</strong> lone pair on <strong>the</strong> nitrogen were in fact to be removed<br />

from interaction with <strong>the</strong> plane <strong>of</strong> <strong>the</strong> aromatic ring system <strong>the</strong>n a similar<br />

effect would be observed. This could easily happen if <strong>the</strong>re was ei<strong>the</strong>r an<br />

intra- or inter-molecular hydrogen bond formed or as a result <strong>of</strong> steric<br />

interactions. For all <strong>the</strong> compounds studied in water as a function <strong>of</strong> pH <strong>the</strong><br />

peak observed in <strong>the</strong> 265 nm region decreases along with <strong>the</strong> 330 nm peak as<br />

<strong>the</strong> pH is lowered. Both 2ABA <strong>and</strong> M2AB show an isosbestic point at around<br />

E:l


290 nm whilst <strong>the</strong> N,N-di<strong>methyl</strong>amino derivatives do not. Since nei<strong>the</strong>r<br />

M2AB or M2DMAB can form a zwitterion, <strong>the</strong> 265 nm peak must be due to<br />

ano<strong>the</strong>r transition within <strong>the</strong> molecule to form <strong>the</strong> second excited singlet<br />

state. This may invalidate <strong>the</strong> conclusions <strong>of</strong> Tramer 44 <strong>and</strong> o<strong>the</strong>r authors<br />

who have used his results. Our spectra <strong>of</strong> 2DMABA closely resemble those<br />

reported by Tramer 44, although he does not appear to have studied <strong>the</strong> ester.<br />

It appears more likely that <strong>the</strong> two amino substituted compounds 2ABA <strong>and</strong><br />

M2AB, if <strong>the</strong>y are hydrogen bonded at all, form inter-molecular hydrogen<br />

bonds involving <strong>the</strong> carbonyl group <strong>and</strong> <strong>the</strong> N-H functions. The absorption<br />

spectra <strong>of</strong> 2ABA <strong>and</strong> M2AB suggest that <strong>the</strong>re is good conjugation between<br />

<strong>the</strong> NH2 <strong>and</strong> COOR groups <strong>and</strong> that <strong>the</strong>re are negligible differences between<br />

<strong>the</strong> acid (R=H) <strong>and</strong> <strong>the</strong> <strong>methyl</strong> ester (R=CH3). The latter would probably not<br />

be <strong>the</strong> case if <strong>the</strong> acid OH in 2ABA were hydrogen bonded to <strong>the</strong> nitrogen<br />

lone pair.<br />

The situation is not as straightforward for <strong>the</strong> N,N-di<strong>methyl</strong>amino<br />

compounds. The addition <strong>of</strong> <strong>the</strong> <strong>methyl</strong> groups to <strong>the</strong> nitrogen will enhance<br />

its basic properties <strong>and</strong> consequently <strong>the</strong> lone pair <strong>of</strong> electrons on <strong>the</strong> nitrogen<br />

will be more available for ei<strong>the</strong>r intra- or inter-molecular hydrogen bonding.<br />

With M2DMAB this hydrogen bond must be with <strong>the</strong> solvent because <strong>the</strong><br />

carboxylate ester group has no acidic proton. This could explain why<br />

M2DMAB shows larger extinction coefficients than those observed for<br />

2DMABA which may preferentially form an intra-molecular hydrogen bond<br />

involving <strong>the</strong> lone pair <strong>of</strong> electrons on <strong>the</strong> nitrogen, consequently losing<br />

conjugation with <strong>the</strong> aromatic ring system. Ano<strong>the</strong>r consequence <strong>of</strong> this<br />

intra-molecular hydrogen bond is that <strong>the</strong> pK a for 2DMABA is considerably<br />

greater than <strong>the</strong> pK a for 2ABA. The observed increase in pK a <strong>of</strong> nearly 3.8 pH<br />

units on di-N,N-<strong>methyl</strong>ation is much greater than that seen for <strong>the</strong> metaisomers<br />

(0.7 - 0.8 units) <strong>and</strong> <strong>the</strong> para- isomers (no change). The presence <strong>of</strong> <strong>the</strong><br />

intra-molecular hydrogen bond makes <strong>the</strong> COOH proton considerably more<br />

ET


difficult to remove from <strong>the</strong> 2DMABA molecule <strong>and</strong> a slightly basic pH<br />

ra<strong>the</strong>r than a weakly acidic one is now required.<br />

Since it is not possible to assign any particular absorption to a<br />

zwitterion structure, our results are calculated from what we believe to be<br />

neutral molecular form (m). For <strong>the</strong> ground state protonation, our pKj, value<br />

<strong>of</strong> 2.93 is between Tramer's 44 pK values <strong>of</strong> 4 <strong>and</strong> 1.4 for cation to molecule<br />

pKcm <strong>and</strong> cation to zwitterion pK cz respectively, whilst our deprotonation<br />

value <strong>of</strong> 8.46 corresponds well to Tramer's 44 value calculated for <strong>the</strong><br />

zwitterion to anion equilibrium pK za. Similar comments apply to <strong>the</strong> excited<br />

state deprotonation reaction. However our observations are in good<br />

agreement with those <strong>of</strong> Jian et al. 45 who were also unable to observe <strong>the</strong><br />

zwitterion.<br />

V<br />

C<br />

3.5<br />

pH 6.27<br />

3.. ,<br />

/ pH4.90<br />

2.5 /1<br />

2<br />

1.5<br />

o i. I,<br />

0.5<br />

0<br />

I<br />

•'<br />

-<br />

I<br />

'Is I<br />

/r<br />

•<br />

s.<br />

J<br />

/<br />

St.<br />

p1-13.97<br />

--—"pH2.99<br />

-----"pHl.14<br />

4 1<br />

ts — - . ----.--- -<br />

300 350 400 450 500<br />

Wavelength (nm)<br />

550<br />

Figure 4.6 Fluorescence spectra <strong>of</strong> 2ABA as a function <strong>of</strong> pH<br />

Figure 4.6 shows typical fluorescence emission spectra for <strong>the</strong> orthocompounds<br />

at different pH values. As can be seen from Figure 4.6, on<br />

lowering <strong>the</strong> pH <strong>the</strong> emission becomes red shifted <strong>and</strong> <strong>the</strong> intensity is<br />

reduced. This can be explained in a similar manner as that described for <strong>the</strong><br />

meta compounds in chapter. three.


The fluorescence intensity <strong>of</strong> <strong>the</strong> four ortho compounds varies in a<br />

similar manner to that <strong>of</strong> <strong>the</strong> ground state absorption spectra. 2ABA has <strong>the</strong><br />

largest <strong>of</strong> all <strong>the</strong> quantum yields (Table 4.3) with <strong>the</strong> pure anion having a<br />

quantum yield <strong>of</strong> around 0.43 compared to <strong>the</strong> neutral species which has a<br />

quantum yield <strong>of</strong> 0.30. Its <strong>methyl</strong> ester, M2AB shows a quantum yield <strong>of</strong> 0.13.<br />

2DMABA has <strong>the</strong> lowest <strong>of</strong> all <strong>the</strong> quantum yields with a value <strong>of</strong> around<br />

0.001 in water; this is fur<strong>the</strong>r strong evidence for <strong>the</strong> fact that <strong>the</strong> lone pair <strong>of</strong><br />

electrons on <strong>the</strong> nitrogen is no longer conjugated to <strong>the</strong> aromatic it-system.<br />

M2DMAB has a quantum yield which is greater by a factor <strong>of</strong> more than 10.<br />

However, it is noticeable that <strong>the</strong> quantum yield for <strong>the</strong> 2DMABA anion is<br />

also very small. Here <strong>the</strong> intra-molecular hydrogen bond no longer exists yet<br />

both <strong>the</strong> absorption intensity <strong>and</strong> fluorescence quantum yield for <strong>the</strong><br />

2DMABA anion are still low. The negative charge on <strong>the</strong> carboxylate anion<br />

does not appear to reduce <strong>the</strong>se properties for 2ABA. Steric effects may reduce<br />

<strong>the</strong> absorption intensity <strong>and</strong> quantum yield for M2DAB but not to <strong>the</strong> low<br />

levels observed for <strong>the</strong> anion. It is not clear at this juncture as to <strong>the</strong> reasons<br />

for <strong>the</strong>se anomalous anion properties.<br />

4.3 fluorescence <strong>and</strong> emission properties as a function <strong>of</strong> solvent.<br />

Fluorescence emission spectra <strong>and</strong> quantum yields were measured for<br />

<strong>the</strong> four ortho- compounds in <strong>the</strong> solvent range used in this <strong>study</strong>. 2ABA <strong>and</strong><br />

its <strong>methyl</strong> ester have high fluorescence quantum yields, ranging from 0.54 in<br />

acetonitrile to 0.20 in a mixture <strong>of</strong> hexanes (Table 4.3). However <strong>the</strong>re is a<br />

dramatic difference between <strong>the</strong>se two compounds <strong>and</strong> <strong>the</strong>ir N,Ndi<strong>methyl</strong>amino<br />

analogues. As mentioned earlier, <strong>the</strong>re is likely to be steric<br />

hindrance due to <strong>the</strong> close proximity <strong>of</strong> <strong>the</strong> carboxylic acid / carboxylate ester<br />

function <strong>and</strong> N,N-di<strong>methyl</strong>amino substituent leading to one or both <strong>of</strong> <strong>the</strong><br />

groups being twisted out <strong>of</strong> <strong>the</strong> plane <strong>of</strong> <strong>the</strong> aromatic ring system. In <strong>the</strong> case<br />

<strong>of</strong> 2DMABA <strong>the</strong> lone pair on <strong>the</strong> nitrogen may also be involved in a strong<br />

[$3


intra-molecular hydrogen bond. Both effects result in a loss <strong>of</strong> conjugation<br />

<strong>and</strong> consequently a loss <strong>of</strong> charge transfer efficiency in <strong>the</strong> system. This is<br />

supported by <strong>the</strong> low fluorescence quantum yields observed for M2DMAB,<br />

but even <strong>the</strong>se are an order <strong>of</strong> magnitude greater than those found for <strong>the</strong><br />

2DMABA. This again is fur<strong>the</strong>r evidence that <strong>the</strong> lone pair <strong>of</strong> electrons on <strong>the</strong><br />

nitrogen are used to form a strong hydrogen bond.<br />

Solvent Measured 2ABA M2AB 2DMABA M2DMAB<br />

Species<br />

quantity<br />

Acetonitrile FluorXmax/nm 394 399 420 429<br />

0.54 0.50 0.002 0.03<br />

BuCN / IBuCN Puor 390 392 436 412<br />

Of<br />

0.42 0.36 0.007 0.05<br />

PhosXmax/nm 451 453 415 480<br />

UC1/iP jquor ?.atnj /nan 389 388 425 415<br />

0.47 0.51 0.004 0.16<br />

PhosXmax/nm 454 452 480 500<br />

Ethanol Fluor Xinax /run 406 413 401 425<br />

0.32 0.42 0.002 0.04<br />

Hexane's FluorXrrtax/nm 380 377 400 395<br />

Of 0.20 0.30 0.008 0.05<br />

MCH / iF Fluor " /nm 386 378 400 398<br />

0.27 0.37 0.003 0.07<br />

PhosXmax/nm 453 451 478 474<br />

Water FluorXmax/nm 405 427 430 438<br />

0.38 0.10 0.003 0.03<br />

Neutral Fluor Xmax /nm 415 421 423 427<br />

0.30 0.13 0.001 0.04<br />

Anion Fluor 400 - 416 -<br />

0.43 - 0.001 -<br />

Table 4.3 Excited state emission characteristics <strong>of</strong> <strong>the</strong> 2-<strong>aminobenzoic</strong> acid,<br />

2-N,N-di<strong>methyl</strong><strong>aminobenzoic</strong> acid <strong>and</strong> <strong>the</strong>ir <strong>methyl</strong> <strong>esters</strong> as a<br />

function <strong>of</strong> solvent <strong>and</strong> pH<br />

In general <strong>the</strong> quantum yields <strong>of</strong> all <strong>the</strong> ortho- compounds increase<br />

with solvent polarity <strong>and</strong> hydrogen bonding ability. There is a similar trend<br />

for <strong>the</strong> wavelength <strong>of</strong> maximum emission in that in <strong>the</strong> non-polar solvents<br />

<strong>the</strong> wavelength <strong>of</strong> maximum emission is blue shifted in comparison to <strong>the</strong>


more polar <strong>solution</strong>s. However in all cases, <strong>the</strong> longest fluorescence<br />

wavelengths are seen in <strong>the</strong> hydrogen bonding solvents ethanol <strong>and</strong> water.<br />

Fur<strong>the</strong>rmore <strong>the</strong> fluorescence from <strong>the</strong> N,N-di<strong>methyl</strong>amino compound is<br />

red shifted some 20 nm from <strong>the</strong>ir corresponding amino compounds. There<br />

is one exception to this <strong>and</strong> that is 2DMABA in butyronitrile / isobutyronitrile<br />

(9:1) where <strong>the</strong> wavelength <strong>of</strong> maximum emission is red shifted<br />

a fur<strong>the</strong>r 15 nm (Figure 4.7). Typical spectra <strong>of</strong> all <strong>the</strong> compounds in hexane<br />

are shown in Figure 4.8.<br />

4500<br />

a. •<br />

2ABA<br />

3500<br />

3000<br />

2500<br />

'C<br />

)<br />

I<br />

I<br />

I<br />

I<br />

2DMABA<br />

-- M2DMAB<br />

2000<br />

I<br />

I<br />

C<br />

1500<br />

1000<br />

04-<br />

300 350 400 450 500 550 600<br />

Wavelength (nm)<br />

Figure 4.7 Fluorescence emission spectra <strong>of</strong> 2ABA, M2AB, 2DMABA <strong>and</strong><br />

M2DMAB in butyronitrile I isobutyronitrile (9:1)


160000 2ABA<br />

140000 a-'<br />

.C' \\<br />

------M2AB<br />

120000. 'I<br />

80000 .<br />

01 I<br />

II<br />

I<br />

Is<br />

I:<br />

I.<br />

A<br />

,<br />

/ /•. \<br />

40000 us S<br />

20000-<br />

- 2DMABA<br />

-----M2DMAB<br />

51 -<br />

0:" I I<br />

310 360 410 460 510<br />

Wavelength (nm)<br />

Figure 4.8 fluorescence emission spectra <strong>of</strong> 2ABA, M2AB, 2DMABA <strong>and</strong><br />

M2DMAB in hexane<br />

I .<br />

There have been a number <strong>of</strong> reports <strong>of</strong> <strong>the</strong> fluorescence properties <strong>of</strong><br />

<strong>the</strong> four ortho compounds. These are summarised in Table 4.3a. Our results<br />

are in excellent agreement with those found in <strong>the</strong> literature.<br />

Solvent 2ABA<br />

Species<br />

?unaxnrn<br />

Acetonitrile 395 -'<br />

400 40<br />

Ethanol 386<br />

413 40<br />

Anion 394 45<br />

39043<br />

2DMABA<br />

naxnm<br />

422<br />

410 40<br />

422 44<br />

Neutral 42043 4407T<br />

MCH 389<br />

387 40<br />

Water 40043<br />

408'"<br />

Table 4.3a fluorescence emission properties which have been previously<br />

reported in <strong>the</strong> literature<br />

91


4.4 fluorescence lifetime data as a function <strong>of</strong> solvent, pH <strong>and</strong><br />

temperature<br />

Fluorescence decay pr<strong>of</strong>iles were measured for <strong>the</strong> four compounds in<br />

all seven solvents <strong>and</strong> binary solvent mixtures. It was found that <strong>the</strong> decay<br />

pr<strong>of</strong>iles varied in complexity with compound <strong>and</strong> solvent / pH. The results<br />

are tabulated in Table 4.4. There have been no o<strong>the</strong>r reported lifetime<br />

measurements on <strong>the</strong> 2-aminoben.zoic <strong>acids</strong>. As mentioned in chapter 3,<br />

water is a singularly effective ionising solvent which is able to stabilise <strong>the</strong><br />

anionic state <strong>of</strong> <strong>the</strong> aromatic <strong>aminobenzoic</strong> <strong>acids</strong> through hydrogen bonding<br />

to <strong>the</strong> nitrogen groups as well as possibly through <strong>the</strong> carboxylic acid group.<br />

Thus for 2ABA <strong>and</strong> M2AB all <strong>the</strong> decays are mono exponential apart from<br />

those in water where a sum <strong>of</strong> two exponentials is required to adequately fit<br />

<strong>the</strong> data with values <strong>of</strong> 2.63 / 5.58 ns for 2ABA <strong>and</strong> 1.77 / 8.51 ns for M2AB. In<br />

<strong>the</strong> case <strong>of</strong> 2ABA, <strong>the</strong> neutral molecular species show a dual exponential<br />

decay with lifetimes <strong>of</strong> 3.01 <strong>and</strong> 5.70 ns while <strong>the</strong> anion shows a single<br />

exponential decay with a lifetime <strong>of</strong> 9.18 ns. In view <strong>of</strong> <strong>the</strong> range <strong>of</strong> different<br />

intra- <strong>and</strong> inter-molecular hydrogen bonds which could possibly exist for<br />

<strong>the</strong>se compounds in water , <strong>the</strong> observation <strong>of</strong> dual exponential fluorescence<br />

kinetics is underst<strong>and</strong>able although this was not seen for <strong>the</strong> meta- isomer.<br />

Whe<strong>the</strong>r <strong>the</strong> decays are truly dual exponential or actually represent a range <strong>of</strong><br />

fluorescent solute / solvent species which are best fitted by a dual exponential<br />

is impossible to judge.<br />

In <strong>the</strong> o<strong>the</strong>r solvents studied for 2ABA, <strong>the</strong> lifetimes vary from 4.63 ns<br />

in <strong>the</strong> non-polar hexane to 7.49 ns in <strong>the</strong> polar hydrogen bonding solvent<br />

ethanol. There is no evidence (i.e. multi-component decays) for multiple<br />

hydrogen bonded species in ethanol. M2AB shows very similar properties<br />

with <strong>the</strong> shortest lifetime <strong>of</strong> 3.77 ns in hexanes <strong>and</strong> <strong>the</strong> longest lifetime<br />

observed in ethanol <strong>of</strong> 7.99 ns. It was expected that <strong>the</strong> next simplest case<br />

would be M2DMAB, but this does not appear to be so as it appears to show


dual exponential decays in all solvents studied. There is a need for fur<strong>the</strong>r<br />

lifetime studies because <strong>of</strong> <strong>the</strong> conflicting data which have been obtained. The<br />

most recent <strong>study</strong> using laser excitation suggests that <strong>the</strong> decays may actually<br />

be single exponential. Unfortunately, attempts to obtain complementary data<br />

using, <strong>the</strong> synchrotron radiation source at BESSY have proved inconclusive<br />

because <strong>of</strong> instrumental artefacts. In a trend similar to that found for 2ABA<br />

<strong>and</strong> M2AB, <strong>the</strong> observed lifetime values are shortest in <strong>the</strong> non-polar<br />

solvents hexane <strong>and</strong> <strong>methyl</strong>cyclohexane / isopentane (4:1) <strong>and</strong> longest in <strong>the</strong><br />

more polar <strong>and</strong> hydrogen bonding solvents. It is difficult to envisage how<br />

<strong>the</strong>re might be two species <strong>of</strong> M2DMAB which might fluoresce o<strong>the</strong>r than<br />

through different solvation arrangements. In some solvents it is possible that<br />

<strong>the</strong> lone pair on <strong>the</strong> nitrogen is involved in some form <strong>of</strong> bonding with <strong>the</strong><br />

solvent. This could involve <strong>the</strong> formation <strong>of</strong> an exciplex with a C- type bond<br />

between <strong>the</strong> solvent <strong>and</strong> <strong>the</strong> molecule as proposed by Varma et al. for 4-<br />

di<strong>methyl</strong>aminobenzonitrile (DMABN) 50' 6869. This however does not explain<br />

why dual exponential decays are observed in <strong>the</strong> non-polar, non-hydrogen<br />

bonding solvents <strong>methyl</strong>cyclohexane / isopentane (4:1) <strong>and</strong> hexane unless<br />

<strong>the</strong>re is perhaps some form <strong>of</strong> dimer formation. Fur<strong>the</strong>r data are clearly<br />

needed.<br />

2DMABA shows <strong>the</strong> same complex behaviour, but this is perhaps<br />

more easily explained. In all but <strong>the</strong> non-polar, non-hydrogen bonding<br />

solvents 2DMABA shows dual exponential decays, with very short lifetimes<br />

<strong>of</strong> <strong>the</strong> order <strong>of</strong> 1 - 1.5 ns <strong>and</strong> a longer lifetime <strong>of</strong> <strong>the</strong> order <strong>of</strong> 6-9 ns. In hexane<br />

<strong>and</strong> <strong>methyl</strong>cyclohexane / isopentane (4:1), <strong>the</strong> single species observed has a<br />

lifetime <strong>of</strong> around 3.5 ns <strong>and</strong>, because <strong>of</strong> <strong>the</strong> inability <strong>of</strong> <strong>the</strong> solvent to interact<br />

with <strong>the</strong> molecule this is probably due to a molecular species with a strong<br />

intra-molecular hydrogen bond. However, in <strong>the</strong> more polar <strong>solution</strong>s <strong>the</strong>re<br />

must be an interaction or a series <strong>of</strong> different interactions between <strong>the</strong> solvent<br />

<strong>and</strong> <strong>the</strong> molecule. This leads to <strong>the</strong> possibility <strong>of</strong> competition between <strong>the</strong><br />

93


intra-molecular hydrogen bonded species <strong>and</strong> a solvated complex <strong>of</strong> a specific<br />

solute/solvent type. The percentage <strong>of</strong> <strong>the</strong> shorter lifetime species appears to<br />

be around 35 % <strong>and</strong> so is a substantial proportion.<br />

Compound Solvent<br />

a1 t2(ns) a2<br />

Species<br />

mean<br />

2ABA Acetonitrile 1.17 6.68<br />

BuCN / IBuCN 1.33 6.06<br />

BC! / iF 1.37 5.43<br />

Ethanol 1.14 7.49<br />

Hexane 1.21 4.63 1<br />

MCH / iF 1.37 5.21<br />

Water 0.86 2.63 0.4047 5.58 0.5953 4.39<br />

Anion 1.19 9.18 __<br />

Neutral 0.97 3.01 0.7398 5.70 0.2602 3.71<br />

M2AB Acetonitrile 0.99 6.57<br />

BuCN / iBuCN 1.45 7.54<br />

BC! / i-F 1.30 4.72<br />

Ethano! 0.94 7.99<br />

Hexane 0.87 3.77<br />

Water 0.64 1.77 0.8982 8.51 0.1018 2.45<br />

2DMABA Acetonitrile 1.14 0.83 0.8482 7.91 0.1518 3.25<br />

BuCN / iBuCN 0.99 1.51 0.8637 8.90 0.1362 2.51<br />

BC1 / iF 1.14 0.81 0.8482 7.91 0.1518 1.89<br />

Ethanol 1.71 1.73 0.8574 11.08 0.1426 3.06<br />

Hexane 0.83 4.10<br />

MCH / iF 1.46 2.90<br />

Water 1.00 0.93 0.8202 13.51 0.1798 3.19<br />

Anion 1.90 0.79 0.8585 6.50 0.1415 1.59<br />

Neutral 1.50 1.42<br />

M2DMAB Acetonitrile 1.31 5.75 0.6987 11.37 0.3013 7.44<br />

BuCN / IBuCN 1.90 5.51 0.1737 11.80 0.8263 10.71<br />

BC! / iF 1.25 5.32 0.8032 12.36 0.1968 6.71<br />

Ethanol 1.01 6.80 0.2209 13.50 1 0.7791 1 12.01<br />

Hexane 0.98 2.67 0.2894 4.33 0.7106<br />

1<br />

3.85<br />

MCH / iF 1.70 3.51 0.1465 5.38 0.8535 5.11<br />

Water 1.51 5.72 0.8952 13.62 0.1048 6.55<br />

Table 4.4<br />

Fluorescence decay pr<strong>of</strong>iles as a function <strong>of</strong> solvent pH for <strong>the</strong><br />

ortho compounds<br />

The lifetime <strong>and</strong> quantum yie!d data has been combined to calculate<br />

rate constants for <strong>the</strong> radiative (Kf) <strong>and</strong> non-radiative (K nr) decay processes in<br />

<strong>the</strong> four molecules as a function <strong>of</strong> solvent (Table 4.5). For <strong>the</strong> two amino<br />

94


substituted compounds K1 does not vary very much with solvent. However<br />

<strong>the</strong> value <strong>of</strong> K I, for 2ABA in hexane appears to rise steeply, this could again<br />

be due to aggregation. This steep rise is not observed for M2AB. In polar /<br />

hydrogen-bonding solvents both <strong>the</strong> N,N-di<strong>methyl</strong>amino substituted<br />

compounds show low K1 values, with 2DMABA being much less than<br />

M2DMAB, while <strong>the</strong> Knr values are similar to those found for <strong>the</strong> non<br />

<strong>methyl</strong>ated compounds.<br />

ii<br />

OY


a) * * *<br />

Lf) e"t<br />

'N o-tn<br />

'4<br />

Ct¼R,t LI). q o• .<br />

OC'4 tflr.4<br />

06- 'enomo-<br />

V<br />

m r-tf<br />

* *<br />

NC<br />

(<br />

CD<br />

t.<br />

4-,<br />

cc<br />

cz<br />

.t<br />

Cu<br />

I<br />

C<br />

a * * * *<br />

CNN<br />

Cue1oc.4c4ot%'tm<br />

t LI)OctJflrO0t%<br />

Cu<br />

c)<br />

* * *<br />

- -'qt'. o. q N en'R<br />

'ON<br />

d&<br />

C .J Ct) N 00 Q\ LI)<br />

V CLI) N''J ('IL<br />

crL)<br />

0)<br />

• o - -<br />

0)<br />

LI<br />

* *<br />

*<br />

N<br />

* *<br />

'o<br />

q oP<br />

'cq<br />

cO'ONN CRcno<br />

0 -<br />

C<br />

Ct<br />

U<br />

Cu<br />

U<br />

C<br />

N<br />

C<br />

.0<br />

C<br />

C<br />

-- as<br />

Ec<br />

(-'I-<br />

I-.<br />

'to<br />

a-a<br />

.45<br />

as<br />

cr c-I<br />

Cc<br />

U<br />

to<br />

I-ct<br />

-4 41<br />

— —<br />

Ctca<br />

41<br />

to<br />

cE 'c<br />

Ccl -<br />

CO<br />

p<br />

CU<br />

Cu<br />

Cl)<br />

.4-Jo<br />

a- -<br />

Iq<br />

V<br />

4-a<br />

U<br />

C)<br />

(I)<br />

a)<br />

'4<br />

Cl)<br />

C)<br />

Cl)<br />

Cl)<br />

C)<br />

U<br />

0<br />

04<br />

C)<br />

• -I<br />

as<br />

Cu<br />

• -4<br />

Cu<br />

C<br />

0<br />

C<br />

C<br />

Cu<br />

C)<br />

Cu<br />

Cu<br />

'-4<br />

C)<br />

ca<br />

aj<br />

cz<br />

t<br />

0<br />

u<br />

<<br />

'e Cu<br />

C) V<br />

— Eu<br />

F- *0<br />

I<br />

—<br />

- ('I r


4.5 Effect <strong>of</strong> temperature on <strong>the</strong> fluorescence emission properties <strong>of</strong> <strong>the</strong> 2-<br />

<strong>aminobenzoic</strong> acid, 2-N,N-di<strong>methyl</strong><strong>aminobenzoic</strong> acid <strong>and</strong> <strong>the</strong>ir<br />

<strong>methyl</strong> <strong>esters</strong>.<br />

Fluorescence spectra <strong>of</strong> <strong>the</strong> four ortho-isomers were measured in <strong>the</strong><br />

three binary solvent mixtures butyronitrile / isobutyronitrile (9:1),<br />

chiorobutane / isopentane (9:1) <strong>and</strong> <strong>methyl</strong>cyclohexane / isopentane (4:1) in<br />

<strong>the</strong> temperature range 77 - 298K. Unlike <strong>the</strong> corresponding meta- isomers,<br />

2ABA, M2AB <strong>and</strong> M2DMAB exhibit no obvious change in <strong>the</strong> shape <strong>of</strong> <strong>the</strong>ir<br />

spectra as <strong>the</strong> temperature is changed. However, in all three solvents<br />

2DMABA phosphoresces much more strongly than it fluoresces at low<br />

temperature. This observation may provide a clue to <strong>the</strong> generally<br />

anomalous properties <strong>of</strong> 2DMABA.<br />

As Table 4.6 shows, <strong>the</strong> general trend, although not as marked as for<br />

<strong>the</strong> meta- isomers, is for an increase in <strong>the</strong> fluorescence quantum yield as <strong>the</strong><br />

temperature decreases.<br />

Cmd Solvent Quantumyield as a function <strong>of</strong> temperature<br />

293K 230K 193K 153K 133K 113K<br />

2ABA BuCN I IBuCN 0.42 0.51 0.70 0.79 0.80 0.79<br />

BC1 uP 0.47 0.55 0.59 0.70 0.71 0.70<br />

MCH / iP 0.27 0.29 0.35 0.39 0.22 0.29<br />

M2AB BuCN / iBuCN 0.36 0.49 0.53 0.54 0.56 0.57<br />

BC1 / iP 0.51 0.67 0.73 0.77 0.80 0.83<br />

MCH / iP 0.37 0.41 0.48 0.51 0.53 0.54<br />

2DMABA BuCN / iBuCN 0.007 0.007 0.004 0.005 10.018 0.029<br />

BC! / iF 0.004 0.003 0.002 0.001 0.001 0.002<br />

MCI-I / iF 0.003 0.003 0.002 0.001 0.001 0.001<br />

M2DMAB BuCN / iBuCNJ 0.05 0.06 0.10 0.12 0.12 0.14<br />

____________ BC1 uP 0.16 0.20 0.22 0.19 0.16 0.16<br />

MCH / iF 0.07 0.09 0.13 0.15 0.16 0.14<br />

Table 4.6<br />

Excited state emission characteristics <strong>of</strong> <strong>the</strong> ortho -compounds as<br />

a function <strong>of</strong> solvent <strong>and</strong> temperature<br />

97


4.6 Summary <strong>of</strong> conclusions<br />

In conclusion, 2ABA <strong>and</strong> M2AB in all solvents except water exist in <strong>the</strong><br />

neutral molecular form. In water both compounds form hydrogen bonded<br />

complexes which are in equilibrium with <strong>the</strong> neutral molecular form.<br />

2DMABA in inert solvents exists as an intra-molecular hydrogen bonded<br />

complex, but in polar or hydrogen bonded solvents, in a similar manner to<br />

M2DMAB, it exists as at least two species, <strong>the</strong> composition <strong>and</strong> structure <strong>of</strong><br />

which can only be conjectured. The data could be explained by <strong>the</strong> formation<br />

<strong>of</strong> an exciplex which involves <strong>the</strong> formation <strong>of</strong> a bond utilising <strong>the</strong> lone pair<br />

<strong>of</strong> electrons on <strong>the</strong> nitrogen. The o<strong>the</strong>r possibility is <strong>of</strong> course a TICT state, but<br />

fur<strong>the</strong>r experiments are required along <strong>the</strong> lines <strong>of</strong> those described in more<br />

detail in chapter 5 to try to elucidate <strong>the</strong> structures, <strong>and</strong> indeed to verify that<br />

M2DMAB does show dual exponential decays in all <strong>the</strong> solvents studied,<br />

especially <strong>the</strong> inert solvents. As can be seen from Table 4.5, 2DMABA is<br />

clearly anomalous, but <strong>the</strong> ester behaviour suggests that <strong>the</strong>re are two<br />

different situations, corresponding to non-polar / medium polar <strong>and</strong> in<br />

highly polar solvents.<br />

02


Chapter 5


Chapter 5<br />

5.1 Ground state absorption spectra as a function <strong>of</strong> solvent.<br />

Unlike <strong>the</strong> ortho-isomers described in chapter four <strong>the</strong> 4-amino<br />

(4ABA) <strong>and</strong> 4-N,N-di<strong>methyl</strong><strong>aminobenzoic</strong> acid (4DMABA)<strong>and</strong> <strong>the</strong>ir <strong>methyl</strong><br />

<strong>esters</strong> (M4AB <strong>and</strong> M4DMAB) cannot form intra-molecular hydrogen bonds.<br />

In <strong>the</strong> case <strong>of</strong> <strong>the</strong> two <strong>acids</strong> in inert solvents <strong>and</strong> only slightly polar solvents,<br />

dimer formation is feasible 54. Both <strong>acids</strong> are partially soluble in <strong>the</strong> nonpolar<br />

solvents used (hexane <strong>and</strong> <strong>methyl</strong>cyclohexane I isopentane (4:1)) <strong>and</strong><br />

4DMABA is also only partially soluble in water. In all o<strong>the</strong>r solvents<br />

studied, <strong>the</strong> four compounds are completely soluble.<br />

These four compounds exhibit <strong>the</strong> strongest UV absorption b<strong>and</strong>s <strong>of</strong><br />

all <strong>the</strong> compounds studied <strong>and</strong> in all <strong>the</strong> solvents used (Table 5.1). The<br />

absorption properties <strong>of</strong> <strong>the</strong> <strong>methyl</strong> <strong>esters</strong> are very similar to those <strong>of</strong> <strong>the</strong><br />

parent <strong>acids</strong> in terms <strong>of</strong> <strong>the</strong>ir wavelength <strong>of</strong> maximum absorption, but <strong>the</strong><br />

extinction coefficients are more varied with no obvious overall pattern as a<br />

function <strong>of</strong> solvent. As would be expected, replacement <strong>of</strong> <strong>the</strong> amino<br />

substituent by an N,N-di<strong>methyl</strong>amino group results in a red shift <strong>of</strong> <strong>the</strong><br />

absorption b<strong>and</strong> <strong>of</strong> some 2 - 4,000 cm -1 <strong>and</strong> a general increase in <strong>the</strong> strength<br />

<strong>of</strong> absorption (Figure 5.1). The same effect was observed with both <strong>the</strong> ortho<strong>and</strong><br />

meta-isomers.<br />

4ABA <strong>and</strong> M4AB show a strong absorption b<strong>and</strong> in <strong>the</strong> 280 rim<br />

region, whilst 4DMABA <strong>and</strong> M4DMAB show generally stronger absorbances<br />

in <strong>the</strong> 310 nm region. The absorption spectra for all four compounds in<br />

hexane <strong>and</strong> <strong>methyl</strong>cyclohexane / isopentane (4:1) are more structured, <strong>and</strong><br />

4DMABA also exhibits this structural detail in acetonitrile <strong>and</strong> ethanol. The<br />

occurrence <strong>and</strong> position <strong>of</strong> this structured absorption b<strong>and</strong> varies from<br />

solvent to solvent <strong>and</strong> between compounds.


Solvent<br />

Measured 4ABA M4AB 4DMABA M4DMAB<br />

Species<br />

quantity<br />

Acetonitrile Am /nin 287 286 310 308<br />

c/ dm3 mol4 cm-1 11,200 17,500 27,100 19,700<br />

BuCN / iBuCN Xmax trim 286 288 307 309<br />

ci din3 mol4 15,300 25,300 24,400 29,700<br />

BC! / iF Xm /rim 278 275 310 305<br />

c/ din3 mol4 cnr1 17,300 16,000 19,400 20,600<br />

Ethanol Ax ,m 289 295 299 299<br />

8/ dm3 mo!-1 cm-1 16,200 17,900 24,400 19,800<br />

Hexane's Xmax /jjn 275 271 308 272<br />

________________ e/ din3 mo!-1 cm-1 Is 17,000 n/s 17,000<br />

MCH / iF )m /nm 276 271 307 304<br />

E/ dm3 mo!4 am1 n/s 22,400 n/s 25,700<br />

Water Amax /nm 278 285 311 316<br />

c/ dm3 mo!4 cm-1 12,800 14,200 n/s 16,600<br />

Neutral Xmax/nm 284 284 314 314<br />

E/ din3 mol -1 cm-i 11,500 16,000 n/s 16,000<br />

Anion Xmax/nm 266 - 284 -<br />

e/ dm3 moF1 cm-1 12,600 - n/s -<br />

Cation Xmax /nm 272 270 272 276<br />

c/ dm3 mo!1 cm-1 950 1,200 n/s 1,100<br />

Table 5.1<br />

Ground state absorption characteristics <strong>of</strong> 4-<strong>aminobenzoic</strong>,<br />

4-N,N-di<strong>methyl</strong><strong>aminobenzoic</strong> <strong>acids</strong> <strong>and</strong> <strong>the</strong>ir <strong>methyl</strong> <strong>esters</strong> as<br />

a function <strong>of</strong> pH <strong>and</strong> solvent<br />

030<br />

0.25<br />

0-20<br />

C<br />

I.<br />

- 0.15<br />

-C<br />

0.10<br />

0.05<br />

0.00<br />

220 240 260 280 300 320 340 360<br />

Wavelength (nn.)<br />

Figure 5.1 Absorption spectra <strong>of</strong> <strong>the</strong> para compounds in acetonitrile<br />

100


It is worth noting that 4ABA in hexane shows ano<strong>the</strong>r peak at 310<br />

nm, which could be an indication <strong>of</strong> <strong>the</strong> formation <strong>of</strong> ei<strong>the</strong>r a dimer or a<br />

zwitterionic species, although after consideration <strong>of</strong> <strong>the</strong> corresponding<br />

spectra <strong>of</strong> <strong>the</strong> ortho- compounds <strong>the</strong> latter does seem unlikely. There is no<br />

reason to doubt <strong>the</strong> conclusions <strong>of</strong> Jian et al. 45 that <strong>the</strong> lowest energy<br />

absorption b<strong>and</strong> in <strong>the</strong>se compounds is due to a it-iC * charge transfer<br />

transition. The slight red shifts in <strong>the</strong> absorption maximum <strong>of</strong> all four<br />

compounds in polar solvents <strong>and</strong> <strong>the</strong> high extinction coefficients which are<br />

observed provide fur<strong>the</strong>r confirmation.<br />

A second absorption b<strong>and</strong> which varies considerably in strength <strong>and</strong><br />

position is observed in <strong>the</strong> 220 - 250 nm region, which is in accordance with<br />

those observations made for both <strong>the</strong> ortho- <strong>and</strong> meta-isomers. The values<br />

presented in Table 5.1 are in excellent agreement with those reported in <strong>the</strong><br />

literature (Table 5.1a) 45,46,47,64'84•<br />

Solvent measured<br />

quantity<br />

4ABA<br />

Acetonitrile x.ax /nm 286<br />

MCH Amax /nxn 275<br />

Neutral Xrnax /rJn 286<br />

Cation A.max /nm 266<br />

Anion Xmax/nm 265<br />

Table 5.1a Ground state absorption characteristics <strong>of</strong> 4-<strong>aminobenzoic</strong> acid<br />

reported from Jiterature 447<br />

It was also found that <strong>the</strong> absorption spectra <strong>of</strong> 4DMABA <strong>and</strong><br />

M4DMAB in acetonitrile over a ten fold concentration range (10 - - 10 5 mol<br />

dm 3) did not vary ei<strong>the</strong>r in terms <strong>of</strong> shape or extinction coefficient within<br />

experimental error (Table 5.2)<br />

101


Concentration<br />

mol dm<br />

Pathlength<br />

111111<br />

4DMABA<br />

c<br />

dm3 mo!-1 cm-1<br />

M4DMAB<br />

c<br />

dm3 mo!-1 cm4<br />

1.09x10-4 1.00 24,659 20,329<br />

5.08x10-5 2.00 22,126 22,993<br />

2.32110-5 5.00 23,181 19,666<br />

1.02x10-5 10.0 1 25,643 1 19,275<br />

Table 5.2<br />

Ground state absorption data for 4DMABA <strong>and</strong> M4DMAB as a<br />

function <strong>of</strong> concentration<br />

These experiments were undertaken in view <strong>of</strong> <strong>the</strong> conclusion (see<br />

later) that <strong>the</strong> anomalous fluorescence in <strong>the</strong>se compounds is at least<br />

partially due to ground state dimer formation.<br />

5.2 Ground state absorption <strong>and</strong> fluorescence emission spectra as a<br />

function <strong>of</strong> pH.<br />

In water, <strong>the</strong> four compounds are capable <strong>of</strong> protonation <strong>and</strong> both<br />

<strong>acids</strong> are capable <strong>of</strong> deprotonation to form <strong>the</strong> anion. The latter species<br />

absorbs more strongly but to <strong>the</strong> blue <strong>of</strong> <strong>the</strong> neutral molecule. The cations<br />

also show similar properties to those seen for <strong>the</strong> ortho- <strong>and</strong> ineta-isomers<br />

<strong>and</strong> absorb much more weakly <strong>and</strong> are strongly blue shifted with respect to<br />

<strong>the</strong> neutral acid. Absorption <strong>and</strong> fluorescence spectral measurements as a<br />

function <strong>of</strong> pH a!low <strong>the</strong> calculation <strong>of</strong> pK values for ground <strong>and</strong> excited<br />

protonation <strong>and</strong> deprotonation <strong>and</strong> <strong>the</strong>se are presented in Tab!e 5.3. A<br />

typical example <strong>of</strong> <strong>the</strong> variation <strong>of</strong> <strong>the</strong> absorption spectra with pH is given<br />

in Figure 5.2. Comparison with <strong>the</strong> literature values for <strong>the</strong> only compound<br />

where <strong>the</strong>se are avai!able (4ABA) is very good.<br />

102


1.6<br />

1.4<br />

'••7C<br />

0.6<br />

- - I<br />

/ . F '•;<br />

I<br />

I' /<br />

/ I<br />

. I .•<br />

I /<br />

I I -<br />

0-4 .'\\\\\ / 1/ •'--<br />

••<br />

------.<br />

a2T<br />

01.<br />

220 240 260 280 320<br />

pH - 1.61<br />

pH-Z09<br />

pH-3.W<br />

pH-3.91<br />

pH-t77<br />

340<br />

Wavelength (nm)<br />

Figure 5.2 Absorption spectra <strong>of</strong> 4ABA as a function <strong>of</strong> pH<br />

Ground state So<br />

Excited state Si<br />

Compound plC1, plc p K1,* pK*<br />

Abs Fluor Abs Fluor<br />

4ABA 2.75 4.76 -6.72 - 9.76 6.17<br />

(2.14) (4.85) (-3.4) (10.8) (6.9)<br />

M4AB 2.33 - -7.23 - - -<br />

4DMABA 2.34 4.79 -7.99 - 11.85 7.02<br />

M4DMAB 2.41 1 - -7.98 - - -<br />

Table 5.3<br />

Ground <strong>and</strong> excited pK values for both protonation <strong>and</strong><br />

deprotonation (values in brackets are taken from literature) 45<br />

There is only a difference <strong>of</strong> 2.0 - 2.5 pK units between <strong>the</strong> PK a <strong>and</strong><br />

pKb values for <strong>the</strong> two <strong>acids</strong>, <strong>and</strong> it is <strong>the</strong>refore impossible to measure an<br />

absorption spectrum for <strong>the</strong> pure neutral acid in water, since it will always<br />

be contaminated by a few percent <strong>of</strong> <strong>the</strong> cation <strong>and</strong> / or anion. In unbuffered<br />

water, both 4ABA <strong>and</strong> 4DMABA will be partially ionised to <strong>the</strong>ir anion <strong>and</strong><br />

this ionisation is observed in <strong>the</strong> absorption spectra <strong>of</strong> <strong>the</strong> two <strong>acids</strong> in<br />

unbuffered water. The fluorescence spectra <strong>of</strong> <strong>the</strong> four pa Ta- compounds in<br />

water vary with pH. Naturally, <strong>the</strong> behaviour is more complex for <strong>the</strong> two<br />

<strong>acids</strong>, which involve two protic equilibria, than for <strong>the</strong> two <strong>esters</strong>, which<br />

have only one. Förster cycle calculations <strong>of</strong> pK a* <strong>and</strong> pKb* predict that <strong>the</strong><br />

ilulci


carboxylic acid groups will be weaker <strong>acids</strong> in <strong>the</strong> excited state than <strong>the</strong><br />

ground state <strong>and</strong> that <strong>the</strong> amino <strong>and</strong> di<strong>methyl</strong>amino substituents will be<br />

weaker bases in <strong>the</strong> excited state. These predictions are identical to those for<br />

<strong>the</strong> ortho- <strong>and</strong> ,neta-isomers<br />

For 4ABA <strong>and</strong> 4DMABA at pH values which are neutral or basic, <strong>the</strong><br />

observed fluorescence spectra are independent <strong>of</strong> pH. The principal species<br />

in absorption is <strong>the</strong> anion <strong>and</strong> <strong>the</strong> same is true in emission even though <strong>the</strong><br />

pKa* appears to lie in <strong>the</strong> neutral / weakly basic pH region (Table 5.3). The<br />

absence <strong>of</strong> any fluorescence from <strong>the</strong> neutral species reflects <strong>the</strong> inability <strong>of</strong><br />

<strong>the</strong> excited anion to gain a proton within its excited state lifetime, even<br />

though PK a* may apparently favour formation <strong>of</strong> <strong>the</strong> excited neutral<br />

molecule. At proton concentrations <strong>of</strong> 10.6 mol dm-3 <strong>and</strong> less, <strong>the</strong> second<br />

order reaction between a proton <strong>and</strong> an excited anion simply cannot<br />

compete with <strong>the</strong> radiative <strong>and</strong> non-radiative processes which give <strong>the</strong><br />

excited anion a sub-nanosecond lifetime (see later). As <strong>the</strong> pH is decreased<br />

from 7.0 to a value <strong>of</strong> approximately 3.5, <strong>the</strong> fluorescence spectra <strong>of</strong> both<br />

4ABA <strong>and</strong> 4DMABA in water red shift <strong>and</strong> decrease in intensity. The spectra<br />

consist <strong>of</strong> a mixture <strong>of</strong> fluorescence from <strong>the</strong> anion <strong>and</strong> <strong>the</strong> neutral species<br />

(Figure 5.3). Excitation at <strong>the</strong> isosbestic point <strong>and</strong> analysis <strong>of</strong> <strong>the</strong> resulting<br />

spectra as a function <strong>of</strong> pH yields pK a values <strong>of</strong> 4.73 <strong>and</strong> 4.93 for 4ABA <strong>and</strong><br />

4DMABA respectively. (The fluorescence spectra were analysed in a similar<br />

maimer to that adopted for <strong>the</strong> ground state absorption spectra; see chapter<br />

two).<br />

These values are in excellent agreement with <strong>the</strong> values obtained from<br />

<strong>the</strong> absorption spectra <strong>of</strong> <strong>the</strong> two <strong>acids</strong>. The behaviour <strong>of</strong> <strong>the</strong> fluorescence<br />

spectra in this pH region is <strong>the</strong>refore governed by <strong>the</strong> ground state equilibrium.<br />

Once again <strong>the</strong> proton concentrations are too low for protonation to compete<br />

with o<strong>the</strong>r excited state processes <strong>and</strong> <strong>the</strong> excited anion / neutral ratio produced,<br />

by absorption remains unchanged. In <strong>the</strong> pH range 3.5 - 10.0, 4DMABA also<br />

104


exhibits anomalous fluorescence in water (Figure 5.3), which will be described<br />

more fully in <strong>the</strong> following section. Fur<strong>the</strong>r reduction <strong>of</strong> <strong>the</strong> pH from<br />

approximately 3.5 to a proton concentration in <strong>the</strong> region <strong>of</strong> 1.0 mol dm 3<br />

decreases <strong>the</strong> intensity <strong>of</strong> fluorescence from <strong>the</strong> neutral species, but does not<br />

eradicate it. On <strong>the</strong> basis <strong>of</strong> <strong>the</strong> p1


position <strong>of</strong> <strong>the</strong> a* b<strong>and</strong> is very solvent dependent. The origin <strong>of</strong> this<br />

anomalous fluorescence is controversial, with TICT state formation 28 '<br />

exciplexes7° <strong>and</strong> ground or excited state dimers 484952 havhig been postulated.<br />

Solvent<br />

Species<br />

Measured<br />

4ABA M4AB 4DMABA<br />

___<br />

M4DMAB<br />

______ ________<br />

b* a* b*<br />

quantity<br />

Acetonitrile FluorXmax 331 333 350 480 350 480<br />

0.27 0.34 0.007* 0.009* 0.006* 0.011*<br />

BuCN / IBuCN Fluor Xmax ,'nm 338 333 340 460 342 460<br />

Of<br />

0.36 0.38 0.003* 0.031* 0.005* 0.022*<br />

Phos Xmax /nm 430 420 440 - 440 -<br />

BC1 / iP FluorAmax /nrn 325 325 350 425 352 415<br />

0.28 0.40 0.040* 0.12* 0.086* 0.17*<br />

Of<br />

PhosA.max/nm 448 430 450 - 450 -<br />

Ethanol FluorXmax /ran 345 456 351 420 347 -<br />

0.12 0.10 0.014 0.006* 0.060* -<br />

Hexane FluorXmax/nm 319 316 334 - 348 -<br />

4)f 0.09 0.14 0.14 - 0.30 -<br />

MCH / iF Fluor Xmax mm 322 320 340 - 335 -<br />

0.03 0.12 0.16 - 0.10 -<br />

Of<br />

PhosXmax/nm 440 421 451 - 445 -<br />

Water FluorXmax 353 366 500 346 -<br />

/TI<br />

0.07 0.02 0.015* 0.009* 0.022* -<br />

Of<br />

Neutral FluorXmax 349 354 370 - 365 -<br />

0.05 0.02 0.009 - 0.025* -<br />

Anion Fluor Xmax /nin 341 - 356 490<br />

Of<br />

0.102 - 0.018 0.011<br />

-.<br />

-<br />

-<br />

-<br />

Note on table b.4 - denotes a quantity wducn varies witn concentration<br />

Table 5.4 Excited state emission characteristics <strong>of</strong> <strong>the</strong> 4-<strong>aminobenzoic</strong>, 4-<br />

N,N-di<strong>methyl</strong><strong>aminobenzoic</strong> <strong>acids</strong> <strong>and</strong> <strong>the</strong>ir <strong>methyl</strong> <strong>esters</strong> as a<br />

function <strong>of</strong> solvent <strong>and</strong> pH<br />

51<br />

Unlike <strong>the</strong> ground state absorption spectra, <strong>the</strong> excited emission<br />

spectra <strong>of</strong> <strong>the</strong> four pa Ta- compounds vary in an expected manner. In all <strong>the</strong><br />

solvents used, <strong>the</strong> wavelength <strong>of</strong> maximum emission for <strong>the</strong> b* b<strong>and</strong> is red<br />

106


shifted with increasing solvent polarity. It increases even more with <strong>the</strong><br />

ability <strong>of</strong> <strong>the</strong> solvent to form hydrogen bonds. This trend is also seen for <strong>the</strong><br />

a* fluorescence observed for 4DMABA <strong>and</strong> its <strong>methyl</strong> ester. The<br />

fluorescence b<strong>and</strong> <strong>of</strong> 4DMABA <strong>and</strong> M4DMAB in hexane is similar to that<br />

found in all o<strong>the</strong>r <strong>solution</strong>s where <strong>the</strong> anomalous fluorescence occurs <strong>and</strong><br />

can <strong>the</strong>refore be attributed to <strong>the</strong> normal b* fluorescence. The fluorescence<br />

spectral <strong>and</strong> quantum yield data as a function <strong>of</strong> solvent are presented in<br />

Table 5.4 <strong>and</strong> <strong>the</strong> data previously reported in <strong>the</strong> literature are shown in<br />

Table 5.4a. Literature data is surprisingly lacking for <strong>the</strong>se compounds, but,<br />

where available it is in reasonable agreement with our results.<br />

Solvent Measured 4ABA 4DMABA M4DMAB<br />

quantity b* a*<br />

Acetonitrile Fluor Xmax /nm 346 46 480 46 480 46<br />

0.022 46 0.001346<br />

0.0029<br />

Of<br />

46<br />

Ethanol Fluor Xmax mm 338<br />

MCH / FluorXmax ,/rim 331 45 32346 r3324<br />

0.21 46 0.24<br />

Neutral Fluor 350 45<br />

Anion Fluor Xmax 11rim 338<br />

Table 5.4a Excited state emission characteristics <strong>of</strong> <strong>the</strong> four para<br />

compounds which have been reported in <strong>the</strong> literature<br />

The anomalous fluorescence <strong>of</strong> 4DMABA <strong>and</strong> M41)MAB is observed in<br />

acetonitrile, butyronitrile / isobutyronitrile (9:1) <strong>and</strong> chlorobutane / isopentane<br />

(9:1) with <strong>the</strong> acid also showing this anomalous fluorescence in<br />

water <strong>and</strong> ethanol. However as shown in Figure 5.3 <strong>the</strong> anomalous<br />

fluorescence observed in water is very pH dependent usually only showing<br />

very weak intensities in <strong>the</strong> a* b<strong>and</strong>. In our h<strong>and</strong>s, <strong>the</strong> fluorescence spectra<br />

<strong>and</strong> quantum yields for 4DMABA <strong>and</strong> M41)MAB in <strong>the</strong> solvents where <strong>the</strong><br />

anomalous fluorescence is observed are concentration dependent, <strong>and</strong><br />

consequently quantum yields for <strong>the</strong>se solute / solvent combinations are<br />

107


quoted in Table 5.4 at a st<strong>and</strong>ard concentration <strong>of</strong> 1x10 4 mol dm7 3 to provide<br />

an indication <strong>of</strong> <strong>the</strong> relative strength <strong>of</strong> <strong>the</strong> fluorescence in <strong>the</strong>se systems.<br />

We have also attempted to calculate individual quantum yields for <strong>the</strong> a *<br />

<strong>and</strong> b* fluorescence b<strong>and</strong>s. This is relatively straightforward in solvents<br />

such as nitriles, where <strong>the</strong>re is little overlap <strong>of</strong> <strong>the</strong> spectra. In chiorobutane<br />

<strong>and</strong> ethanol, <strong>the</strong> two individual quantum yields have to be viewed with<br />

caution as <strong>the</strong> two b<strong>and</strong>s are ra<strong>the</strong>r closer toge<strong>the</strong>r <strong>and</strong> separation <strong>of</strong> <strong>the</strong><br />

individual spectra to produce <strong>the</strong> calculated quantum yields is more<br />

difficult. Cowley <strong>and</strong> co-workers 46'47 also quote some quantum yields for<br />

4DMABA <strong>and</strong> M4DMAB (Table 5.4a), which, given <strong>the</strong> concentration<br />

dependence <strong>of</strong> <strong>the</strong>se values, are in reasonable agreement with our values in<br />

acetonitrile.<br />

C<br />

oj6<br />

.4S<br />

C<br />

1<br />

C<br />

@14<br />

U<br />

@1<br />

0<br />

-<br />

IL.<br />

I<br />

320 340 360 380 400 420 440<br />

Wavelength (nm)<br />

Figure 5.4 fluorescence emission spectra <strong>of</strong> 4ABA <strong>and</strong> M4AB in<br />

acetonitrile<br />

The anomalous fluorescence observed for 4DMABA <strong>and</strong> M41DMAB is<br />

shown in Figures 5.6, 5.7, 5.11, 5.14 <strong>and</strong> 5.15, in much more detail. By way <strong>of</strong><br />

contrast, <strong>the</strong> emission properties <strong>of</strong> 4ABA <strong>and</strong> M4AB are completely<br />

normal. A single emission b<strong>and</strong> is observed peaking between 330 <strong>and</strong> 355<br />

nm (depending on solvent) <strong>and</strong> <strong>the</strong>re is absolutely no evidence <strong>of</strong> any


anomalous emission. Typical fluorescence spectra for 4ABA <strong>and</strong> M4AB are<br />

shown in Figures 5.4 <strong>and</strong> 5.5 <strong>and</strong> <strong>the</strong> emission data for <strong>the</strong>se compounds is<br />

summarised in Table 5.4.<br />

6<br />

I<br />

C<br />

1<br />

0<br />

290 310 330 350 370 390 410 430 450<br />

Wavelength (nm)<br />

Figure 5.5<br />

fluorescence emission spectra <strong>of</strong> 4ABA <strong>and</strong> M4AB in hexane<br />

5.4 Effect <strong>of</strong> excitation wavelength on <strong>the</strong> fluorescence properties <strong>of</strong><br />

4DMABA <strong>and</strong> M4DMAB<br />

The fluorescence observed from 4DMABA <strong>and</strong> M4DMAB in hexane<br />

<strong>and</strong> M4DMAB in ethanol <strong>and</strong> water is independent <strong>of</strong> excitation<br />

wavelength, apart from a change in <strong>the</strong> overall intensity. These, <strong>of</strong> course,<br />

are <strong>the</strong> solvents in which no anomalous fluorescence is observed for <strong>the</strong>se<br />

compounds. In all <strong>the</strong> solvents for which anomalous fluorescence is<br />

observed for 4DMABA <strong>and</strong> M4DMAB, <strong>the</strong> intensity distribution between<br />

<strong>the</strong> a* <strong>and</strong> b* b<strong>and</strong>s is dependent upon excitation wavelength. The results <strong>of</strong><br />

measuring fluorescence spectra at different excitation wavelengths for <strong>the</strong>se<br />

samples are given in Table 5.5.<br />

109


Xex<br />

(nm)<br />

Intensity<br />

Acetonitrile _ Ethanol __ __ Water ______<br />

Intensity<br />

a *<br />

Ratio<br />

Intensity<br />

b<br />

Intensity<br />

a<br />

Ratio<br />

Intensity<br />

if<br />

Intensity<br />

a*<br />

Ratio<br />

b/a<br />

4-N,N-Di<strong>methyl</strong><strong>aminobenzoic</strong>_acid<br />

230 56 - - 412 48 8.58 6.2 4.1 1.51<br />

240 56 - 244 48 5.08 5.8 4.8 1.12<br />

250 104 - 214 72 4.45 8.2 5.0 1.64<br />

260 180 - 228 100 2.28 14.0 3.8 3.68<br />

270 252 12 21.0 272 140 1.94 17.5 5.0 3.50<br />

280 252 16 15.8 308 156 1.93 17.4 5.6 3.10<br />

290 200 24 8.33 276 140 1.97 15.8 5.8 2.72<br />

300 128 28 4.60 169 100 1.69 11.3 5.5 2.10<br />

310 68 28 2.40 - - - 7.8 1 4.5 1.73<br />

320 22 16 138 - - - - - -<br />

__ Methyl_4-N,N-di<strong>methyl</strong>aminobenzoate<br />

240 - - - 3 - - 40 - -<br />

250 - - - 100 - - 44 - -<br />

260 22.5 - - 300 - - 46 - -<br />

270 62.5 40 1.56 312 - - 105 - -<br />

280 100 60 1.66 256 - - 75 - -<br />

290 130 120 1.08 310 - - 25.7 - -<br />

300 135 156 0.86 260 - - 18.0 - -<br />

310 105 170 0.62 190 - - 12.3 - -<br />

320 50 100 0.50 - - - 14.3 - -<br />

Table 5.5<br />

fluorescence intensity <strong>of</strong> both emission peaks observed for<br />

4DMABA <strong>and</strong> M4DMAB as a function <strong>of</strong> excitation<br />

wavelength<br />

2<br />

7.00<br />

6.00<br />

5.00<br />

4.00<br />

'<br />

La 27Onm<br />

-----La2SOnm<br />

Lex29Onm<br />

Lex300nm.<br />

La3l0nm<br />

3.00<br />

2.00 .<br />

/ : .----- --- .-<br />

1.00 :_. -- -----<br />

0.00! I I I I I<br />

310 360 410 460 510 560<br />

Wavelength (nm)<br />

Figure 5.6 fluorescence emission spectrum <strong>of</strong> 4DMABA in acetonitrile as<br />

a function <strong>of</strong> excitation wavelength<br />

110


An example <strong>of</strong> <strong>the</strong> observed variations is given in Figure 5.6 where<br />

<strong>the</strong> b* intensity has been normalised to <strong>the</strong> same value for all <strong>the</strong> spectra.<br />

For 4DMABA in acetonitrile, <strong>the</strong> excitation spectrum for <strong>the</strong> b*<br />

fluorescence peaks at 275 nm <strong>and</strong> that for <strong>the</strong> a* fluorescence at<br />

approximately 305 nm. The latter corresponds well to <strong>the</strong> ground state<br />

absorption spectrum. The values observed for <strong>the</strong> benzoate are 295 nm (b*)<br />

<strong>and</strong> 305 nm (a*). For <strong>the</strong> acid in acetonitrile, if <strong>the</strong> excitation wavelength is<br />

moved to a shorter value (270 nm), <strong>the</strong> intensity <strong>of</strong> <strong>the</strong> b* fluorescence<br />

increases approximately fourfold whilst <strong>the</strong> a* fluorescence almost<br />

disappears (Figure 5.6). In all cases, <strong>the</strong> a* fluorescence appears to be excited<br />

preferentially by longer wavelengths than <strong>the</strong> b* fluorescence (Table 5.5).<br />

For 4DMABA in ethanol <strong>and</strong> water a similar effect is observed,<br />

although in ethanol <strong>the</strong> two peaks appear to have very similar maximum<br />

excitation wavelengths. In water <strong>the</strong> maximum excitation wavelengths are<br />

280 rim (b*) <strong>and</strong> 295 nm (a). From <strong>the</strong>se data, it seems likely that <strong>the</strong> species<br />

which is forming <strong>the</strong> anomalous fluorescence is already formed in <strong>the</strong><br />

ground state, a feature which has not been reported before. A <strong>study</strong> <strong>of</strong> <strong>the</strong><br />

effect <strong>of</strong> concentration on <strong>the</strong> fluorescence spectra was <strong>the</strong>refore thought to<br />

be appropriate.<br />

5.5 Variation <strong>of</strong> fluorescence intensity <strong>of</strong> 4DMABA <strong>and</strong> M4DMAB as a<br />

function <strong>of</strong> concentration<br />

The fluorescence emission spectra <strong>of</strong> both 4DMABA <strong>and</strong> M4DMAB<br />

in Ocetonitrile vary with concentration as depicted in Table 5.7.<br />

111


Table 5.6<br />

Concentration Intensity Intensity Ratio<br />

mold& b<br />

a*<br />

4-N,N-Di<strong>methyl</strong><strong>aminobenzoic</strong>acid<br />

1.09x10 5 0.261 0.671 0.38<br />

5.42x10 6 0.205 0.515 0.39<br />

1.09x10 6 0.063 0.133 0.46<br />

5.42x10 7 0.042 0.082 0.50<br />

1.09x10 7 0.032 0.052 0.60<br />

1.09x10 0.024 0.011 1.00<br />

Methyl 4-N,N-dimeth 'laminobentoate<br />

5.28x10 6 0.602 1.181 0.51<br />

1.05x10 6 0.171 0.304 0.55<br />

5.28x10 7 0.082 0.149 0.55<br />

1.05x10 7 0.057 0.037 1.53<br />

5.28x10 8 0.066 0.011 6.01<br />

1.09x10 8 0.047 0.008 6.00<br />

fluorescence ethission ratios <strong>of</strong> 4DMABA <strong>and</strong> M4DMAI3 as a<br />

function <strong>of</strong> concentration in acetonitrile<br />

Figure 5.7, where <strong>the</strong> intensity <strong>of</strong> <strong>the</strong> b* b<strong>and</strong> has once again been<br />

normalised, illustrates how <strong>the</strong> fluorescence <strong>of</strong> 4DMABA in acetonitrile<br />

varies with concentration. As <strong>the</strong> concentration is reduced <strong>the</strong> fluorescence<br />

intensity <strong>of</strong> <strong>the</strong> a* b<strong>and</strong> becomes less <strong>and</strong> less until it is debatable whe<strong>the</strong>r it<br />

is still <strong>the</strong>re or not (although this minimum intensity is not depicted in<br />

Figure 5.7).<br />

0.9<br />

0.8<br />

0.7<br />

'-5<br />

6<br />

>26<br />

- 0.6<br />

C<br />

3<br />

.E o.s<br />

4J<br />

U<br />

C<br />

0.4<br />

D.8<br />

D-8<br />

0.3<br />

0.2<br />

0.1<br />

0<br />

310 360 410 460 510 560 610<br />

Wavelength (nm)<br />

Figure 5.7 fluorescence emission spectra <strong>of</strong> 4DMABA as a function <strong>of</strong><br />

concentration (mol dm 4) in acetonitrile<br />

112


There are a substantial number <strong>of</strong> mechanisms which could be<br />

invoked to account for <strong>the</strong> occurrence <strong>of</strong> this anomalous fluorescence:<br />

1) A TICT mechanism in which <strong>the</strong>re is no ground state molecule<br />

which is already twisted <strong>and</strong> in which <strong>the</strong> excited TICT molecule<br />

decays into a non-twisted ground state.<br />

2) A TICT mechanism with an equilibrium being established between<br />

TICT <strong>and</strong> planar molecule in <strong>the</strong> excited state, but still no ground<br />

state twist.<br />

3) A TICT mechanism with equilibrium in <strong>the</strong> ground state between<br />

planar <strong>and</strong> twisted species.<br />

4) Ground state dimer formation, with decay <strong>of</strong> excited dimer to ground<br />

state dimer or with decay <strong>of</strong> excited dimer into an excited monomer<br />

<strong>and</strong> a ground state molecule.<br />

5) Reversible or non-reversible excimer formation<br />

6) Reversible or non-reversible exciplex formation<br />

7) Two or more <strong>of</strong> <strong>the</strong> above happening simultaneously such as a nonreversible<br />

excimer formation, <strong>and</strong> exciplex formation.<br />

Of <strong>the</strong>se mechanisms, only 4, 5 <strong>and</strong> 7 should show a concentration<br />

dependence such that a variation <strong>of</strong> <strong>the</strong> fluorescence properties with<br />

concentration would be expected. We have, however, attempted to gain data<br />

to investigate all <strong>the</strong> mechanisms <strong>and</strong> this is reported later in <strong>the</strong> <strong>the</strong>sis.<br />

113


Exploration <strong>of</strong> <strong>the</strong> possibility <strong>of</strong> excimer formation is based on <strong>the</strong><br />

following mechanisms<br />

Non reversible excimer formation<br />

M<br />

h,<br />

+ M_ K, )'Excimer<br />

K1<br />

)M+ht1<br />

K<br />

)3M(isC)+M(1C)<br />

'Excimer<br />

1Excimer<br />

K 'ht + 2M<br />

4 3Excimer + isc or 2M(ic)<br />

Reversible excimer formation<br />

M<br />

h<br />

+ M_ K,, )'Excimer<br />

x 1<br />

K<br />

)3M(isc)+M(iC)<br />

1Excimer'<br />

'Excimer'<br />

K,,., M+'M'<br />

K ht4 + 2M<br />

1Excimer'_ ' ) 3Excimer + isc or 2M(ic)<br />

Applying <strong>the</strong> steady state approximation to <strong>the</strong>se two mechanisms,<br />

equations 5.1 <strong>and</strong> 5.2<br />

1 & + Kex[M]<br />

Of sr — 1(1 K 1 (5.1)<br />

1 - (K +K)(K 1 +Knr) K +K, (5.2)<br />

Lx -<br />

lI KK CX [M] K lf<br />

are obtained for <strong>the</strong> variation <strong>of</strong> <strong>the</strong> fluorescence quantum yields for <strong>the</strong><br />

mechanism involving non-reversible excimer formation.<br />

114


If excimer formation is reversible, <strong>the</strong> relationships are more<br />

complicated but <strong>the</strong>ir basic form is unchanged <strong>and</strong> equations 5.3 <strong>and</strong> 5.4 are<br />

obtained<br />

1 1 K nr (K+K ' nil<br />

\K rev<br />

[M]<br />

K f (5.3)<br />

1 - (K + + K )(K 1 + K s,) + K + (5.4)<br />

•Ex<br />

91 ICK eX EM]<br />

In both cases an excimer mechanism (reversible or non-reversible) predicts<br />

that plots <strong>of</strong> 1 / qM vs [M] <strong>and</strong> 1 / 4" vs 1/ [M] will be linear. The first <strong>of</strong><br />

<strong>the</strong>se plots is shown in Figure 5.8 for 4DMABA <strong>and</strong> M4DMABA in<br />

acetonitrile, whilst <strong>the</strong> latter is shown in Figure 5.9.<br />

45<br />

40<br />

35<br />

C<br />

0<br />

25<br />

t 20<br />

I.<br />

S<br />

01 I I I<br />

Q.QQ+O0 2.000-06 4000.06 6.000-06 8.00E-06 1000.05 1.200-05<br />

Concentration (mol dm-3)<br />

Figure 5.8 Plot <strong>of</strong> 1/4 vs [M] for both 4DMABA <strong>and</strong> M4DMAB in<br />

acetonitrile<br />

115


1.20E.02<br />

11K3E+02<br />

a<br />

8.00E+0I<br />

2 4.00E+0I<br />

C<br />

0 2.00E+O1<br />

O,00E.W 2E.W 4.E.07 6,E.W 8E,W<br />

I / concnetration (1/mat dm-3)<br />

Figure 5.9 Plot <strong>of</strong> 1 / 4" vs 1 / [M] for both compounds in acetonitrile<br />

Unfortunately, <strong>the</strong> possibility that <strong>the</strong> anomalous fluorescence arises<br />

from dimer formation in <strong>the</strong> ground state is difficult to test. The absence <strong>of</strong><br />

any obvious change in <strong>the</strong> absorption spectra <strong>of</strong> 4DMABA <strong>and</strong> M4DMAB in<br />

acetonitrile with concentration makes it impossible to quantify any ground<br />

state equilibrium that is in existence. The effect <strong>of</strong> excitation wavelength on<br />

<strong>the</strong> fluorescence clearly suggests <strong>the</strong> presence <strong>of</strong> two ground state species<br />

<strong>and</strong> it is interesting to note that a plot <strong>of</strong> <strong>the</strong> quantum yield <strong>of</strong> <strong>the</strong> "normal"<br />

fluorescence against concentration for both 4DMABA <strong>and</strong> M4DMAB gives a<br />

near linear graph as shown in Figure 5.10. This may be simply fortuitous.<br />

0.7<br />

S.<br />

0.6<br />

03<br />

0.4<br />

0.3<br />

t 0.2<br />

&<br />

0.I<br />

0 I<br />

0.00E+00 2.00E06 4.00E-06 6000.06 6006.06 2.006-05 2.206-05<br />

Concentration (pm' dm.3)<br />

Figure 5.10 Plot <strong>of</strong> 4<br />

vs EM] for both compounds in acetonitrile<br />

5.6 fluorescence lifetime data as a function <strong>of</strong> solvent <strong>and</strong> pH<br />

1111


Fluorescence decay pr<strong>of</strong>iles may provide fur<strong>the</strong>r evidence to assist in<br />

assigning <strong>the</strong> origin <strong>of</strong> <strong>the</strong> anomalous fluorescence in 4DMABA <strong>and</strong><br />

M4DMAB. The fluorescence kinetics <strong>of</strong> both <strong>the</strong> b* <strong>and</strong> <strong>the</strong> a* b<strong>and</strong>s have<br />

been measured in a variety <strong>of</strong> solvents at room temperature <strong>and</strong> analysed in<br />

terms <strong>of</strong> a sum <strong>of</strong> up to three exponential components. Parallel<br />

measurements have also been made for 4ABA <strong>and</strong> M4AB <strong>and</strong> <strong>the</strong>se results<br />

are presented in Tables 5.7, 5,8 <strong>and</strong> 5.9.<br />

Solvent<br />

Species<br />

Xem<br />

(nm)<br />

X2 Ti (ns) a1 t2(ns) a2<br />

4-Aminobenzoic acid<br />

Acetonitrile 331 1.08 1.14<br />

BC! / iP 315 1.22 0.49 0.7995 4.08 0.2005<br />

Ethanol 345 1.09 0.28 0.9673 1.45 0.0327<br />

Hexane 319 1.02 1.01 0.9174 2.49 0.0826<br />

Water 349 0.63 0.79<br />

Anion 349 1.24 0.72<br />

Neutral 341 1.33 0.82<br />

Methyl &aminobenzoate<br />

Acetonitrile 333 1 1.03 11.13 1<br />

BC1/iP 330 1 1.19 11.27 1<br />

Ethanol 346 1.11 0.14 0.9959 1.48 0.0041<br />

Hexane 318 1.09 2.06<br />

Water 353 1.07 11.34<br />

Table 5.7 fluorescence decay data for 4ABA <strong>and</strong> M4AB as a function <strong>of</strong><br />

solvent, pH <strong>and</strong> temperature<br />

In every case <strong>the</strong> observed pr<strong>of</strong>iles rise promptly <strong>and</strong> <strong>the</strong>re is no<br />

evidence for an excited precursor to <strong>the</strong> fluorescent state within <strong>the</strong><br />

experimental limits (say 100 ps) <strong>of</strong> <strong>the</strong> apparatus used. Although this in<br />

itself does not rule out <strong>the</strong> possibility <strong>of</strong> exciplex <strong>and</strong> excimer mechanisms<br />

which involve a very rapid (>1010 s 1 ) formation step, one would expect to<br />

see this reflected in <strong>the</strong> decay <strong>of</strong> <strong>the</strong> b* state. In virtually every case, <strong>the</strong> b*<br />

decay times are at least an order <strong>of</strong> magnitude longer lived (i.e. > 1.0 ns)<br />

than this. Additionally, for an excimer mechanism operating at<br />

concentrations <strong>of</strong> <strong>the</strong> order <strong>of</strong> 10 mol dm 3 <strong>and</strong> less, Kex would have to be<br />

>1015 dm3 mol1 1; five orders <strong>of</strong> magnitude greater than <strong>the</strong> diffusion<br />

controlled rate. We <strong>the</strong>refore believe that <strong>the</strong> fluorescence decay data back<br />

117


up <strong>the</strong> earlier argument that <strong>the</strong> anomalous fluorescence in 4DMABA <strong>and</strong><br />

M4DMAB does not originate from an excimer.<br />

Solvent<br />

Xern V ti(ns) a1 t2(ns) a2 t3(ns) a3<br />

Species<br />

4-N,N-di<strong>methyl</strong><strong>aminobenzoic</strong> acid<br />

Acetonitrile 350 1.29 1.72<br />

480 1.01 2.23<br />

BuCN / iBuCN 350 1.11 1.51 0.9590 5.86 0.0410<br />

485 11.54 2.39<br />

BC1 / IP 355 1.24 1.15 0.4470 3.78 0.5510<br />

420 1.30 3.25 0.2218 4.82 0.1814 0.04 0.5968<br />

Ethanol 351 1.01 2.86 0.8603 17.61 0.1397<br />

MCH / iP 334 1.21 1.25 0.7275 1 6.96 0.0333 0.45 0.2392<br />

370 1.43 1.29 0.9919 1 6.03 0.0081<br />

Hexane 334 1.01 1.28<br />

Water 345 1.05 0.77 0.9411 5.07 0.0589<br />

480 1.27 0.17 0.9580 2.39 0.0420<br />

Neutral 370 1.89 0.85<br />

480 1.27 0.17 0.9437 2.34 0.0563<br />

Anion 356 1.44 1 0.76<br />

480 1 1.19 1 0.25 0.9659 0.82 1 0.0341<br />

Table 5.8<br />

Lifetime data for 4DMABA as a function <strong>of</strong> pH <strong>and</strong> solvent<br />

Solvent<br />

Aern V 'tj(ns) a1 T2ns) a2<br />

Methyl_4-N,N-di<strong>methyl</strong>aminobenzoate<br />

Acetonitrile 350 1.08 1.65 0.8415 2.45 0.1585<br />

480 1.04 2.38<br />

BuCN / 1BuCN 340 1.23 1.07 0.9930 3.54 0.0070<br />

480 0.93 0.40 0.1721 3.00 0.8279<br />

BC1 / iP 345 1.11 1.12 0.3416 3.65 0.6584<br />

420 1.06 3.52 0.8429 5.27 0.1571<br />

Ethanol 347 1.94 0.26 0.9972 6.83 0.0028<br />

MCH / iP 334 1.19 0.95 0.9986 7.61 0.0014<br />

370 1.40 0.98 0.9525 6.98 0.0475<br />

Hexane 348 1.30 1.32 0.5000 2.12 0.5000<br />

Water 345 1.40 0.11 0.9915 5.30 0.0085<br />

Table 5.9<br />

Lifetime data for M4DMAB as a function <strong>of</strong> pH <strong>and</strong> solvent<br />

The fluorescence decay pr<strong>of</strong>iles for 4ABA <strong>and</strong> M4AB are for <strong>the</strong> most<br />

part adequately fitted by a single exponential. Where this is not <strong>the</strong> case, a I<br />

mean value (= E1a1t/E1a1 ) is also presented (Tables 5.13) to allow<br />

comparisons to be drawn amongst <strong>the</strong> data. In <strong>the</strong> absence <strong>of</strong> any<br />

118


comparable literature data on <strong>the</strong>se two compounds <strong>and</strong> having no kinetic<br />

reasons to expect multi-exponential fluorescence kinetics from <strong>the</strong>m, it is<br />

difficult to decide whe<strong>the</strong>r <strong>the</strong> double exponential fits are real or not.<br />

Fur<strong>the</strong>r measurements are needed to clarify this question. Assuming that<br />

<strong>the</strong> fluorescence kinetics are uniformly first order <strong>and</strong> so single exponential<br />

decays should be observed, fluorescence lifetimes for 4ABA <strong>and</strong> M4AB <strong>of</strong> 1 -<br />

2 ns are observed in virtually all solvents. Combination <strong>of</strong> <strong>the</strong>se lifetimes<br />

with <strong>the</strong> quantum yields given in Table 5.4 allows <strong>the</strong> calculation <strong>of</strong><br />

radiative <strong>and</strong> non-radiative rate constants. These are presented in Table<br />

5.10.<br />

Solvent O f 'Cf Kf<br />

x10 8<br />

Knr<br />

xlO-8<br />

4-Aminobenzoic acid<br />

Acetonitrile 0.27 1.20 1.88 6.08<br />

BC1 / ip 0.28 1.21* 2.31 5.95<br />

Ethanol 0.12 0.32* 3.75 27.5<br />

Hexane 0.09 1.13* 0.80 8.05<br />

Water 0.08 082 0.61 11.61<br />

Methyl_4-aminobenzoate<br />

Acetonitrile 0.34 1.05 3.24 6.29<br />

BCl / ip 0.40 1.27 3.15 4.72<br />

Ethanol 0.10 0.15* 6.67 60.00<br />

Hexane 0.14 2.3 0.61 3.74<br />

Water 0.02 1.34 0.15 7.31<br />

* t mean calculated<br />

Table 5.10 Calculated radiative <strong>and</strong> non-radiative rate constants for 4ABA<br />

<strong>and</strong> M4AB<br />

It is clear from Tables 5.7 <strong>and</strong> 5.10 that, in terms <strong>of</strong> <strong>the</strong> interaction <strong>of</strong><br />

4ABA <strong>and</strong> M4AB with <strong>the</strong> various solvents used in this <strong>study</strong>, <strong>the</strong>re<br />

appears to be an anomaly with ethanol. This is <strong>the</strong> one solvent which leads<br />

to fluorescence lifetimes substantially below <strong>the</strong> 1 - 2 ns range <strong>and</strong> this<br />

appears to be due to an enhancement <strong>of</strong> <strong>the</strong> non-radiative decay process in<br />

119


<strong>the</strong> two molecules. It is possible to speculate about <strong>the</strong> nature <strong>of</strong> this<br />

interaction, but it is not possible to be sure as to its exact nature without<br />

fur<strong>the</strong>r data. Herbich et al.85 have recently reported fluorescence quenching<br />

<strong>of</strong> 2-(2'-pyridyl)indoles by alcohols, speculating that this arises through<br />

enhanced internal conversion. The importance <strong>of</strong> cyclic inter-molecular,<br />

hydrogen-bonded structures in this enhancement is stressed (acyclic<br />

structures are found not to quench <strong>the</strong> fluorescence). It is difficult to see<br />

how such cyclic structures could be involved here, without involvement <strong>of</strong><br />

several solvent molecules, but <strong>the</strong> idea <strong>of</strong> enhanced internal conversion via<br />

a specific solute I solvent interaction would seem a logical way to explain<br />

our data. The alternative to internal conversion is intersystem crossing, but<br />

it is hard to envisage how a molecule such as ethanol would increase spinorbit<br />

coupling between S1 <strong>and</strong> T n <strong>and</strong> thus increase ISC. In <strong>the</strong> case <strong>of</strong> <strong>the</strong><br />

di<strong>methyl</strong>amino compounds 4DMABA <strong>and</strong> M4DMAB, a small amount <strong>of</strong><br />

literature data is available <strong>and</strong> <strong>the</strong> fluorescence lifetime obtained for<br />

M4DMAB in hexane agrees excellently with values <strong>of</strong> 1.1 - 1.3 as for a range<br />

<strong>of</strong> non-polar solvents quoted by Howell et al.49 .<br />

The fluorescence decay pr<strong>of</strong>ile <strong>of</strong> <strong>the</strong> anomalous fluorescence<br />

observed in ethanol for 4DMABA was not measured. The a* <strong>and</strong> b* peaks<br />

could not be adequately resolved <strong>and</strong> so <strong>the</strong> proximity <strong>of</strong> <strong>the</strong> two<br />

fluorescence b<strong>and</strong>s leads to a complex decay pr<strong>of</strong>ile requiring <strong>the</strong> sum <strong>of</strong><br />

three exponentials to adequately fit <strong>the</strong> data. We are also able to confirm<br />

results reported by Howell et aL. 49 who observed dual exponential decays for<br />

M4DMAB in mixed alkane solvents. However in none <strong>of</strong> <strong>the</strong> cases have we<br />

observed a negative pre-exponential factor which would be expected if <strong>the</strong><br />

a* fluorescence were to originate from <strong>the</strong> b* state as would be <strong>the</strong> case for<br />

both TICT <strong>and</strong> excimer formation. This points to <strong>the</strong> anomalous<br />

fluorescence being due to a ground state dimer.<br />

120


In those solvents where anomalous fluorescence is not observed, <strong>the</strong><br />

quantum yield <strong>and</strong> lifetime data for 4DMABA <strong>and</strong> M4DMAB has been<br />

combined to produce rate constants (Table 5.1).<br />

The values obtained are very similar to those for 4ABA <strong>and</strong> M4AB.<br />

Once again <strong>the</strong>re appears to be enhancement <strong>of</strong> <strong>the</strong> non-radiative decay<br />

route in ethanol as well as in water. There was some suggestion <strong>of</strong> <strong>the</strong> latter<br />

for 4ABA in water (Table 5.10), but not for M4AB even though it has a low<br />

fluorescence quantum yield. This leads to <strong>the</strong> suggestion that <strong>the</strong> lifetime<br />

for M4AB in water may be shorter than <strong>the</strong> quoted value <strong>and</strong> clearly points<br />

to <strong>the</strong> need for fur<strong>the</strong>r data.<br />

Solvent O f tf Kf<br />

Knr<br />

x10 8<br />

x10 8<br />

4-N,N-Di<strong>methyl</strong><strong>aminobenzoic</strong> acid<br />

Hexane 0.14 1.28 1.10 6.72<br />

MCH I ip . 0.16 1.25* 1.28 6.72<br />

Methyl 4-N,N-di<strong>methyl</strong>aminobenzoate<br />

MCH / ip 0.10 0.96* 1.04 9.38<br />

Ethanol 0.06 0.28* 2.14 33.57<br />

Hexane 0.30 1.72* 1.74 4.07<br />

Water 0.03 0.15* 1.67 65.00<br />

* tmean calculated<br />

Table 5.11 Calculated radiative <strong>and</strong> non-radiative rate constants for<br />

M4DMAB <strong>and</strong> 4DMABA in various solvents<br />

In view <strong>of</strong> <strong>the</strong> changes in <strong>the</strong> fluorescence spectra <strong>of</strong> 4DMABA <strong>and</strong><br />

M4DMAB with excitation wavelength, concentration <strong>and</strong> pH discussed<br />

earlier in this chapter, fluorescence decay measurements were made for <strong>the</strong><br />

two compounds in acetonitrile as a function <strong>of</strong> excitation wavelength <strong>and</strong><br />

concentration (Tables 5.12 <strong>and</strong> 5.13).<br />

121


XExI XEni I x I115I a12i a2<br />

4-N,N-Di<strong>methyl</strong><strong>aminobenzoic</strong> acid<br />

280 355 1.035 2.13<br />

1.<br />

480 1.243 2.19<br />

310 355 1.170 2.12<br />

480 1.356 2.15<br />

340 355 1.566 2.27<br />

480 1.309 2.18<br />

Methyl 4-N,N-di<strong>methyl</strong>aminobenzoate<br />

270 355 1.156 1.27 0.8947 2.31 0.1053<br />

270 480 1.439 2.68<br />

305 355 1.159 1.26 0.8140 2.49 0.1860<br />

305 480 1.467 2.64<br />

340 355 1.044 1.38 0.3072 2.77 0.6928<br />

340 480 1.340 2.67<br />

Table 5.12 fluorescence decay lifetimes <strong>of</strong> 4DMABA <strong>and</strong> M4DMAB in<br />

acetonitrile as a function <strong>of</strong> excitation wavelength<br />

Concentration XEm V 'C1 ins a1 t2 ins a2<br />

moldnr3<br />

mean<br />

4-N,N-Di<strong>methyl</strong><strong>aminobenzoic</strong> acid<br />

5x10 5 350 1.61 1.66<br />

1.<br />

480 1.48 2.30<br />

5x10 6 350 1.26 1.29 0.8420 2.74 0.1580 1.52<br />

480 1.01 2.23<br />

5x10 7 350 1.23 1.20 0.6869 1.85 0.3131 1.98<br />

480 1.14 2.27<br />

Methyl_4-.N,N-di<strong>methyl</strong>aininobenzoate<br />

5x10 6 355 1.11 1.70 0.9926 2.21 0.0074 1.70<br />

_____ 480 1.04 2.36<br />

5x10 7 355 1.08 1.65 0.9826 2.45 0.0174 1.66<br />

_____ 480 0.97 2.49<br />

5x10 355 1.00 1.69 0.9598 2.49 0.0402 1.72<br />

480 0.96 2.71<br />

Table 5.13 Fluorescence decay lifetimes <strong>of</strong> 4DMABA <strong>and</strong> M4DMAB as a<br />

function <strong>of</strong> concentration in acetonitrile (excitation<br />

wavelength 305 nm)<br />

For 4DMABA, <strong>the</strong> lifetime <strong>of</strong> <strong>the</strong> a* b<strong>and</strong> is independent <strong>of</strong><br />

concentration <strong>and</strong> excitation wavelength <strong>and</strong> <strong>the</strong> same is true for <strong>the</strong><br />

benzoate. A lifetime in <strong>the</strong> range 2.1 - 2.3 ns for <strong>the</strong> acid <strong>and</strong> 2.3 - 2.7 ns for<br />

<strong>the</strong> benzoate is obtained with nei<strong>the</strong>r showing evidence <strong>of</strong> a rise time which<br />

would be indicative <strong>of</strong> <strong>the</strong> excited state precursor which would be expected<br />

for TICT or excimer formation. These observations confirm our earlier<br />

122


conclusion about <strong>the</strong> presence <strong>of</strong> a ground state species producing <strong>the</strong> a *<br />

fluorescence. They also imply that this is <strong>the</strong> only source <strong>of</strong> <strong>the</strong> a* species.<br />

The b* excited state cannot be <strong>the</strong> precursor <strong>of</strong> a* although <strong>the</strong> reverse may<br />

possibly be true given <strong>the</strong> more complex decay pr<strong>of</strong>iles observed for <strong>the</strong> b*<br />

fluorescence b<strong>and</strong>. The decay <strong>of</strong> <strong>the</strong> b* fluorescence is not so simple with <strong>the</strong><br />

ben.zoate showing dual lifetimes <strong>of</strong> =1.7 ns <strong>and</strong> =2.4 ns, <strong>the</strong> latter values<br />

being in a similar range to those found for <strong>the</strong> a* emission although it only<br />

contributes a small percentage <strong>of</strong> <strong>the</strong> total emission. The acid on <strong>the</strong> o<strong>the</strong>r<br />

h<strong>and</strong> shows single exponential decays at a concentration <strong>of</strong> 1x10- 5 mol dm'3<br />

<strong>and</strong> this is independent <strong>of</strong> excitation wavelength. However at<br />

concentrations <strong>of</strong> lx10" 6 mol dm-3 or less, this single exponential changes to<br />

a dual exponential with lifetimes <strong>of</strong> 1.3 <strong>and</strong> 2.8 ns, <strong>the</strong> latter figure being<br />

similar to <strong>the</strong> value <strong>of</strong> 2.3 ns observed for <strong>the</strong> a* fluorescence.<br />

Multi-exponential decays are also observed in ethanol <strong>and</strong> water.<br />

Given <strong>the</strong> hydrogen bonding ability <strong>of</strong> both <strong>the</strong>se solvents, it is not<br />

surprising that, more than one excited state solute / solvent arrangement<br />

may be formed <strong>and</strong> so <strong>the</strong> complex kinetics may be real. Indeed, <strong>the</strong> multiexponential<br />

fits may be due to a distribution <strong>of</strong> lifetimes from an array <strong>of</strong><br />

slightly different solute I solvent interactions, as has been proposed by<br />

Howell et al.49 for M4DMAB in mixtures <strong>of</strong> non-polar solvents. The<br />

possibility <strong>of</strong> a non-fluorescent TICT state in <strong>the</strong>se solvents is not ruled out.<br />

Unless <strong>the</strong>re exists a very short lived precursor whose lifetime is so short<br />

that we are unable to separate it from <strong>the</strong> instrumental response function<br />

when we model <strong>the</strong> fluorescence decays <strong>of</strong> <strong>the</strong> a* <strong>and</strong> b* states, both a* <strong>and</strong><br />

b* are excited directly from <strong>the</strong> ground state. Once <strong>the</strong> two excited states<br />

have been formed, it is possible that a* converts to b*, but <strong>the</strong>re is no<br />

evidence for <strong>the</strong> reverse process.<br />

Varma <strong>and</strong> co-workers 50 have proposed that <strong>the</strong> anomalous<br />

fluorescence <strong>of</strong> ethyl 4-N,N-di<strong>methyl</strong>aminobenzoate <strong>and</strong> 4-NM-<br />

123


di<strong>methyl</strong>amino-benzonitrile in acetonitrile is due to exciplex formation 45 .<br />

Our conclusion from <strong>the</strong>se data is that <strong>the</strong> fluorescence spectrum <strong>and</strong> decay<br />

pr<strong>of</strong>ile <strong>of</strong> <strong>the</strong> b* state is comprised <strong>of</strong> contributions from a distribution <strong>of</strong><br />

excited solute / solvent arrangements, <strong>the</strong> decay kinetics <strong>of</strong> which are<br />

adequately described by a sum <strong>of</strong> two exponentials. Unlike Howell et al.,49<br />

we have no evidence to suggest that a* derives from b*. The additional<br />

observation <strong>of</strong> <strong>the</strong> a* fluorescence b<strong>and</strong> leads us to conclude that <strong>the</strong> a* is an<br />

excited dimer or possibly a higher aggregate which is populated directly<br />

from <strong>the</strong> ground state by absorption <strong>of</strong> a photon <strong>of</strong> light. Given <strong>the</strong> apparent<br />

lack <strong>of</strong> concentration dependence <strong>of</strong> <strong>the</strong> ground state absorption spectrum,<br />

<strong>the</strong> ground state dimer may not involve a strong interaction between<br />

molecules. However a loose association is sufficient to organise <strong>the</strong><br />

molecules such that <strong>the</strong>y are in approximately <strong>the</strong> correct position to form<br />

an excited dimer when a photon is absorbed by one <strong>of</strong> <strong>the</strong> molecules. We are<br />

conscious that <strong>the</strong> time re<strong>solution</strong> <strong>of</strong> our time-correlated, single photon<br />

counting system would not allow us to measure a rapid rise time (perhaps<br />


fluorescence properties <strong>of</strong> <strong>the</strong> two molecules is summarised in Table 5.14<br />

<strong>and</strong> typical spectra (for 4DMABA) are shown in figure 5.11.<br />

In both cases, although <strong>the</strong>re is no obvious change in <strong>the</strong> absorption<br />

spectrum, <strong>the</strong> b* fluorescence in hexane is quenched by low concentrations<br />

<strong>of</strong> acetonitrile, but at slightly higher concentrations it increases in intensity<br />

<strong>and</strong> red shifts some 10 - 20 nm. At <strong>the</strong> same time <strong>the</strong>re is a steady increase in<br />

a* fluorescence as <strong>the</strong> acetonitrile concentration is increased (Figure 5.11 <strong>and</strong><br />

5.12). It would appear that <strong>the</strong>re are some specific solvation effects taking<br />

place in addition to <strong>the</strong> mechanism producing <strong>the</strong> a* fluorescence. The red<br />

shift in <strong>the</strong> b* fluorescence probably results from <strong>the</strong> formation <strong>of</strong> a specific<br />

4DMABA / acetonitrile solute / solvent interaction, but could also arise<br />

because <strong>of</strong> <strong>the</strong> increase in <strong>the</strong> polarity <strong>of</strong> <strong>the</strong> overall solvent mixture as a<br />

result <strong>of</strong> <strong>the</strong> addition <strong>of</strong> <strong>the</strong> polar acetonitrile molecule.<br />

4DMABA<br />

M4DMAB<br />

%CH3CN mt b* Tnt a * a* / b* mt b* Tnt a* a* / b*<br />

0.0 3030 0.00 0.000 571 64.70 0.113<br />

0.2 29.00 5.00 0.172 570 127 0.222<br />

0.4 20.00 8.50 0.425 497 176 0.354<br />

0.6 8.00 7.00 0.875 433 212 0.490<br />

0.8 3.50 4.25 1.214 362 235 0.649<br />

1.0 9.20 10.10 1.097 341 252 0.739<br />

1.2 5.95 9.80 1.647 282 235 0.833<br />

1.4 14.00 22.40 1.600 280 247 0.882<br />

1.8 17.50 35.40 2.000 206 218 1.058<br />

2.0 14.00 29.30 2.092 194 229 1.180<br />

Table 5.14 Fluorescence intensities (Int) <strong>of</strong> both <strong>the</strong> normal <strong>and</strong><br />

anomalous fluorescence peaks <strong>of</strong> both 4DMABA <strong>and</strong><br />

M4DMAB in hexane (=1x104 mol dm 3) as a function <strong>of</strong> added<br />

acetonitrile<br />

125


45.00<br />

40.00..<br />

i I<br />

300<br />

O.M. CH3CN<br />

-----0A%CH3CN<br />

0.6% C}-I3CN<br />

350 400 450 500 550<br />

Wavetength (nm)<br />

Figure 5.11 Fluorescence emission spectra <strong>of</strong> 4DMABA in hexane as a<br />

function <strong>of</strong> added acetonitrile<br />

The new b<strong>and</strong> at 440 nm may be an exciplex or may be <strong>the</strong> same<br />

species as is seen in pure acetonitrile but with a blue shifted fluorescence<br />

due to <strong>the</strong> non-polar solvent environment in which it is situated.<br />

Whatever <strong>the</strong> nature <strong>of</strong> <strong>the</strong> species producing <strong>the</strong> 440 nm fluorescence, it is<br />

dear that <strong>the</strong> acetonitrile plays a key role in its formation.<br />

This effect <strong>of</strong> adding a polar solvent to a solvent <strong>of</strong> lower polarity was<br />

first noted by Ch<strong>and</strong>ros 66 who observed similar results to those above for<br />

DMABN in <strong>methyl</strong>cyclohexane I acetonitrile mixtures. He felt that <strong>the</strong><br />

addition <strong>of</strong> <strong>the</strong> polar solvent facilitated <strong>the</strong> formation <strong>of</strong> an exciplex. Varma<br />

et al.68 invoking a similar mechanism, considered that <strong>the</strong> anomalously<br />

emitting species is ei<strong>the</strong>r formed from <strong>the</strong> reaction <strong>of</strong> a single polar<br />

molecule (P) with a molecule <strong>of</strong> DMABN in its first excited singlet state or<br />

arises from an already existing ground state complex between solute <strong>and</strong><br />

solvent. If <strong>the</strong> latter was <strong>the</strong> case however, it would be expected that <strong>the</strong><br />

ratio <strong>of</strong> <strong>the</strong> two peaks would depend on <strong>the</strong> initial concentration <strong>of</strong><br />

DMABN, which has not been observed. In <strong>the</strong> case <strong>of</strong> 4DMABA <strong>and</strong><br />

M4DMAB we have observed a concentration dependency in pure solvents<br />

iPLi


ut have not yet repeated this experiment in solvent mixtures at different<br />

solute concentrations. Rettig et al. 57 have also postulated that <strong>the</strong> affect <strong>of</strong><br />

<strong>the</strong> addition <strong>of</strong> a polar solvent to an inert solvent helps to relax <strong>the</strong><br />

activation energy <strong>of</strong> <strong>the</strong> TICT state <strong>and</strong> so facilitates TICT state formation.<br />

-2.50<br />

2.00<br />

150<br />

Cd<br />

1.00<br />

0.50<br />

i I I I I<br />

0.00 0.20 0.40 0.60 0.80 1.00 120 1.40 1.60 1.80 2.00<br />

Concentration <strong>of</strong> added acetonitrile %<br />

Figure 5.12 Plot <strong>of</strong> <strong>the</strong> ratio <strong>of</strong> a* / b* for both 4DMABA <strong>and</strong> M4DMAB in<br />

hexane as a function <strong>of</strong> added acetonitrile<br />

0% --- -- 0.25T---------0.40% - - -- -- 0.50% - ---- 75%<br />

1.00% ----- 2.0Ot- -------- 3.00% ------4.00%<br />

0.60<br />

0.50<br />

a,<br />

v 0.40<br />

e 0.30<br />

0<br />

0.20<br />

0.10<br />

0.00<br />

220 240 260 280 300 320 340<br />

Wavelength (nm)<br />

Figure 5.13 Cround state absorption spectra <strong>of</strong> 4DMABA in acetonitrile<br />

(5.03x1O mol dma) as a function <strong>of</strong> added water<br />

127


Jouvet 63 has presented evidence that <strong>the</strong> anomalous fluorescence in<br />

this series <strong>of</strong> compounds derives from <strong>the</strong> formation <strong>of</strong> solute / water<br />

complexes. We have <strong>the</strong>refore undertaken some preliminary experiments<br />

on <strong>the</strong> effect <strong>of</strong> water concentration on <strong>the</strong> absorption <strong>and</strong> fluorescence<br />

properties <strong>of</strong> 4DMABA <strong>and</strong> M4DMAB in acetonitrile. Although <strong>the</strong><br />

addition <strong>of</strong> acetonitrile to a <strong>solution</strong> <strong>of</strong> ei<strong>the</strong>r 4DMABA or M4DMAB in<br />

hexane causes no change in <strong>the</strong> ground state absorption spectra, addition <strong>of</strong><br />

water to a <strong>solution</strong> <strong>of</strong> 4DMABA in acetonitrile (Figure 5.13 <strong>and</strong> Table 5.14)<br />

does result in a dramatic change in <strong>the</strong> ground state absorption spectra. This<br />

is not <strong>the</strong> case for <strong>the</strong> corresponding benzoate.<br />

1.07x10 7 mol dm 3 5.03x10 6 mol din-3<br />

%H20 hit hit hit Mt<br />

308nm 282nm 308nm 282nin<br />

0 0.5510 0.2973 02662 0.0967<br />

0.25 0.4555 0.3319 0.2408 0.2621<br />

0.40 0.4164 0.3451 0.1517 0.1720<br />

0.50 0.4077 0.3610 1 0.1494 0.1743<br />

0.75 0.3950 0.3781 0.1374 0.1814<br />

1.0 0.3723 0.4204 0.1025 0.2051<br />

2.0 0.2307 0.4749 0.0776 0.2264<br />

3.0 0.2196 0.4959 0.0757 0.2304<br />

5.0 0.1939 0.4964 - -<br />

Table 5.15 Ground state absorption data for 4DMARA in acetonitrile as a<br />

function <strong>of</strong> added water at two concentrations<br />

On addition <strong>of</strong> > 5% water to <strong>the</strong> above concentrations <strong>of</strong> <strong>solution</strong>,<br />

<strong>the</strong> absorption spectrum becomes unreadable because <strong>of</strong> <strong>the</strong> turbidity <strong>of</strong> <strong>the</strong><br />

<strong>solution</strong> <strong>and</strong> so only values up to 5% were taken. The spectra show that at<br />

concentrations <strong>of</strong> 2% water, <strong>the</strong> ground state species absorbing at 308 nm<br />

becomes <strong>the</strong> minor component <strong>and</strong> <strong>the</strong> 282 nm species becomes <strong>the</strong> major.<br />

This corresponds well to <strong>the</strong> fluorescence excitation spectrum <strong>of</strong> 4DMABA<br />

in acetonitrile which shows <strong>the</strong> anomalous fluorescence having a<br />

wavelength <strong>of</strong> maximum absorption around 305 nm <strong>and</strong> <strong>the</strong> normal<br />

fluorescence a value <strong>of</strong> 275 nm. This implies that <strong>the</strong> addition <strong>of</strong> water<br />

128


destroys <strong>the</strong> ground state species which is potentially forming <strong>the</strong><br />

anomalous fluorescence. There was no observable change in <strong>the</strong> ground<br />

state absorption spectra <strong>of</strong> M4DMAB as a function <strong>of</strong> water added to a<br />

<strong>solution</strong> <strong>of</strong> acetonitrile.<br />

Fluorescence spectra were also measured for <strong>the</strong>se <strong>solution</strong>s. The<br />

fluorescence spectra for <strong>the</strong> 4DMABA <strong>solution</strong>s were excited at <strong>the</strong> isosbestic<br />

point but as no change in <strong>the</strong> ground state absorption spectrum could be<br />

observed for M41DMAB an excitation wavelength <strong>of</strong> 306 nm was used. As<br />

water was added to a <strong>solution</strong> <strong>of</strong> M41DMAB in acetonitrile, <strong>the</strong> anomalous<br />

fluorescence decreased whilst <strong>the</strong> normal fluorescence increased (Figure 5.14<br />

<strong>and</strong> Table 5.16).<br />

6<br />

5<br />

C<br />

0)4<br />

C<br />

0)<br />

u3<br />

C<br />

0)<br />

V<br />

0)<br />

Li<br />

0<br />

1z 1<br />

f<br />

I<br />

( \\<br />

0% Water<br />

------1.0% Water<br />

2.0% Water<br />

-----3.0% Water<br />

- 5% Water<br />

10% Water<br />

15% Water<br />

0 4-<br />

-S<br />

310 360 410 460 510 560 610<br />

Wavelength (rim)<br />

Figure 5.14 Fluorescence spectra <strong>of</strong> M4DMAB in acetonitrile as a function<br />

<strong>of</strong> percentage added water<br />

129


%H20 mt a * hit b* Ratio<br />

0 61.0 52 1.17<br />

1.0 33.0 49 0.67<br />

2.0 23.0 42 0.54<br />

3.0 152 39 0.39<br />

4.0 14.5 36 0.40<br />

5.0 9.0 23 0.39<br />

10.0 5.3 15 0.35<br />

20.0 3.0 12 0.25<br />

Table 5.16 Variation <strong>of</strong> fluorescence intensity <strong>of</strong> a* <strong>and</strong> b* peaks <strong>of</strong><br />

M4DMAB in acetonitrile as a function <strong>of</strong> added water<br />

Similar data are presented in Figure 5.15 <strong>and</strong> Table 5.17 for 4DMABA.<br />

7<br />

0% Water<br />

6<br />

U, 5<br />

0)<br />

I .<br />

I<br />

•<br />

-----0.50% Water<br />

1.00% Water<br />

------1.50% Water<br />

.4.<br />

01<br />

U<br />

C<br />

013,<br />

_<br />

rn<br />

I'<br />

I<br />

-- - 3.0% Water<br />

6.0% Water<br />

I- 2<br />

iz<br />

1<br />

.<br />

"S<br />

-----10.0% Water<br />

o i --- I I<br />

310 360 410 460 510<br />

Wavelength (rim)<br />

560 610<br />

Figure 5.15 fluorescence spectra <strong>of</strong> 4DMABA in acetonitrile as a function<br />

percentage added water<br />

ilkis]


%H 20 Ab mt b* Xmaxa mt a *<br />

0 350 160 488 117<br />

0.25 354 218 487 84<br />

0.50 354 270 490 64<br />

0.80 353 295 490 53<br />

1.00 355 343 490 47<br />

1.50 354 429 490 44<br />

2.00 356 538 490 20<br />

3.00 358 591 430 170<br />

4.00 358 473 438 212<br />

6.00 360 263 450 211<br />

10.0 360 114 456 163<br />

15.0 358 81 460 121<br />

20.0 355 62 462 95<br />

30.0 352 49 470 61<br />

50.0 352 46 475 34<br />

75.0 353 33 500 10<br />

90.0 354 31 500 8<br />

Table 5.17 fluorescence spectra <strong>of</strong> 4DMABA in acetonitrile as a function<br />

<strong>of</strong> added %water<br />

The ratio <strong>of</strong> decrease <strong>of</strong> <strong>the</strong> a* over increase <strong>of</strong> <strong>the</strong> b* intensity is 1 2,<br />

implying that <strong>the</strong> anomalous fluorescence is due to an interaction between<br />

at least two molecules <strong>of</strong> 4DMABA ra<strong>the</strong>r than interaction with a polar<br />

acetonitrile molecule. The anomalous fluorescence disappears at =2.0% v/v<br />

water in acetonitrile (Figure 5.16). A new fluorescence peak is <strong>the</strong>n observed<br />

at approximately 430 nm . With fur<strong>the</strong>r addition <strong>of</strong> water, <strong>the</strong> a *<br />

fluorescence starts to decrease as <strong>the</strong> intensity <strong>of</strong> <strong>the</strong> new peak grows,<br />

starting at 430 nm <strong>and</strong> becoming substantially red shifted to 500 run when<br />

<strong>the</strong> solvent composition is biased in favour <strong>of</strong> water. This could be due to<br />

<strong>the</strong> change in <strong>the</strong> dielectric constant <strong>of</strong> <strong>the</strong> water / acetonitrile mixture. For<br />

a concentration <strong>of</strong> 1.06x10 -5 mol dm 3, a plot <strong>of</strong> <strong>the</strong> %H 20 up to 2% against<br />

both <strong>the</strong> a* <strong>and</strong> b* intensity yields a straight line plot (Figure 5.16). It is a<br />

point worthy <strong>of</strong> note that although <strong>the</strong> absorption spectra were unreadable<br />

at water concentration> 5%, <strong>the</strong> fluorescence spectra were satisfactory.<br />

131


.b500<br />

'0<br />

C<br />

400<br />

1 300<br />

200<br />

100<br />

0<br />

0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2<br />

Percentage added water<br />

Figure 5.16 Plot <strong>of</strong> b* <strong>and</strong> a* emission from 4DMABA in acetonitrile as a<br />

function <strong>of</strong> added water<br />

We infer that if <strong>the</strong> anomalous fluorescence <strong>of</strong> 4DMABA in<br />

acetonitrile deriyes from a dimer, <strong>the</strong> addition <strong>of</strong> water preferentially forms<br />

<strong>the</strong> anion <strong>and</strong> consequently <strong>the</strong> emergence <strong>of</strong> <strong>the</strong> new peak is probably a<br />

solvent I solute complex which changes position due to <strong>the</strong> change in <strong>the</strong><br />

dielectric constant <strong>of</strong> <strong>the</strong> mixture. Although this evidence is not conclusive<br />

it does appear that <strong>the</strong> addition <strong>of</strong> water to acetonitrile destroys <strong>the</strong> dimer. if<br />

<strong>the</strong> anomalous fluorescence observed in water does not originate from a<br />

ground state dimer, <strong>the</strong> proposal <strong>of</strong> Jouvet 63 may be supported by <strong>the</strong>se data.<br />

Finally, a few fluorescence decay pr<strong>of</strong>iles have been measured for<br />

4DMABA in acetonitrile as a function <strong>of</strong> added water content (Table 5.18)<br />

132


%H20<br />

XEx XEm<br />

Intensity<br />

x 2<br />

Ir l<br />

al 'C2 a2 'C<br />

mean<br />

o 308 350 24.13 1.04 1.63<br />

490 24.00 1.13 2.09<br />

0.5 307 350 20.04 1.07 1.49<br />

490 15.03 1.19 1.36<br />

1.0 283 350 130.0 1.19 1.70 0.5513 3.03 0.4487 2.30<br />

J18<br />

3.0 279 350 115.0 1.09 0.35 0.4259 3.02 0.5741<br />

430 40.00 1.18 3.08<br />

480 25.00 2.29 3.13<br />

10.0 281.6 350 22.00 1.71 2.02<br />

_<br />

450 36.00 1.09 0.33 0.1055 2.09 0.8945<br />

Table 5.18 Variation <strong>of</strong> fluorescence intensity <strong>of</strong> a* <strong>and</strong> b* peaks <strong>of</strong><br />

M4DMAB in acetonitrile as a function <strong>of</strong> added water toge<strong>the</strong>r<br />

with fluorescence decay data<br />

The lifetime <strong>of</strong> <strong>the</strong> b* fluorescence remains roughly constant in <strong>the</strong><br />

1.49 - 2.02 ns range. The lifetime <strong>of</strong> <strong>the</strong> a* b<strong>and</strong> initially decreases from 2.09<br />

to 1.36 ns on <strong>the</strong> addition <strong>of</strong> 0.5% v/v water, but increases to approximately<br />

3.1 ns in <strong>the</strong> presence <strong>of</strong> 3% v/v water. The latter correlates with <strong>the</strong><br />

observed shift in <strong>the</strong> a* peak from 490 to 430 nm. The data are admittedly<br />

sparse <strong>and</strong> incomplete, but tend to suggest an initial quenching <strong>of</strong> <strong>the</strong> a*<br />

state on <strong>the</strong> addition <strong>of</strong> water, followed by <strong>the</strong> formation <strong>of</strong> a new a* species<br />

(at water concentrations > 2% v/v) with a 3.1 ns lifetime. From <strong>the</strong>se data, it<br />

is suggested that fluorescence decay pr<strong>of</strong>iles for <strong>the</strong>se mixed solvent systems<br />

could be a useful source <strong>of</strong> data to help explain <strong>the</strong> observed fluorescence<br />

changes.<br />

5.8 Effect <strong>of</strong> temperature on <strong>the</strong> emission properties <strong>of</strong> 4-<strong>aminobenzoic</strong><br />

acid, 4-N,N-di<strong>methyl</strong><strong>aminobenzoic</strong> acid <strong>and</strong> <strong>the</strong>ir <strong>methyl</strong> <strong>esters</strong><br />

Rettig <strong>and</strong> co-workers 86 have highlighted <strong>the</strong> useful information<br />

which may be obtained from variable temperature fluorescence spectral <strong>and</strong><br />

kinetic measurements, particularly when <strong>the</strong> photochemistry involves<br />

TICT states. Measurements <strong>of</strong> <strong>the</strong> fluorescence properties <strong>of</strong> 4DMABA <strong>and</strong><br />

133


M4DMAB in <strong>the</strong> temperature range 77-298K have <strong>the</strong>refore been<br />

undertaken. Some data have also been recorded for 4ABA <strong>and</strong> M4AB which<br />

do not exhibit <strong>the</strong> anomalous fluorescence since <strong>the</strong>y may provide a<br />

baseline from which to judge <strong>the</strong> 4DMABA <strong>and</strong> M4DMAB data.<br />

Fluorescence spectra for <strong>the</strong> four compounds in <strong>the</strong> solvent mixtures<br />

<strong>methyl</strong>cyclohexane / isopentane (4:1), chlorobutane / isopentane (9:1) <strong>and</strong><br />

butyronitrile I isobutyronitrile (9:1) were measured in <strong>the</strong> temperature<br />

range 77 - 298K. Based on <strong>the</strong> quantum yield at ambient temperature,<br />

fluorescence quantum yields were calculated at temperatures down to 77K.<br />

However, no allowance has been made for absorption changes as <strong>the</strong><br />

temperature is reduced 87 <strong>and</strong> <strong>the</strong> quantum yield changes which are<br />

recorded in Table 5.17 must be viewed with this in mind.<br />

CMD Solvent Quantum yield as a function <strong>of</strong> temperature<br />

293K 230K 193K 153K 133K 113K<br />

4ABA BuCN / iBuCN 0.36 0.62 0.94 1.00 1.00 1.00<br />

BC! / iP 0.28 0.28 0.35 0.80 1.00 1.00<br />

MCH / 11' 0.03 0.05 0.08 0.09 0.09 0.08<br />

M4AB BuCN / IBuCN 0.38 0.63 0.93 0.99 1.00 1.00<br />

BC! / iF 0.40 0.54 0.75 0.90 1.00 1.00<br />

MCH / iF 0.12 0.18 0.31 0.43 0.35 0.12<br />

4DMABA BuCN / iBuCN 0.003 0.003 0.004 0.004 0.006 0.02<br />

BC! / iP 0.04 0.05 0.08 0.12 0.18 0.25<br />

MCH / iF 0.16 0.15 0.026 0.019 0.013 0.03<br />

480nm BuCN / 1BuCN 0.031 0.022 0.015 0.08 0.08 0.04<br />

BC! I iF 0.12 0.07 0.04 0.04 0.03 0.03<br />

M4DMAB BuCN / IBuCN 0.005 0.007 0.008 0.010 0.012 0.008<br />

BC1 / iP 0.09 0.09 0.10 0.10 0.14 0.17<br />

MCH / iF 0.10 10.14 10.14 0.12 0.08 10.16<br />

480nm BuCN / 1BuCN 0.022 10.014 10.008 0.004 10.004 1 P<br />

BC! / iF 0.17 10.11 10.09 0.06 10.04 1 0.04<br />

Table 5.19<br />

fluorescence quantum yields <strong>of</strong> 4-<strong>aminobenzoic</strong> acid, 4-N,Ndi<strong>methyl</strong><strong>aminobenzoic</strong><br />

acid <strong>and</strong> <strong>the</strong>ir <strong>methyl</strong> <strong>esters</strong> as a<br />

function <strong>of</strong> solvent <strong>and</strong> temperature<br />

For 4ABA <strong>and</strong> M4AB, <strong>the</strong> shape <strong>and</strong> position <strong>of</strong> <strong>the</strong> fluorescence<br />

spectra are largely invariant with temperature in <strong>the</strong> solvents used. There is<br />

some evidence for a slight red shift as <strong>the</strong> temperature is decreased, as<br />

134


typified by M4AB in <strong>methyl</strong>cyclohexane / isopentane (4:1) (Figure 5.17), but<br />

<strong>the</strong> shift is only <strong>of</strong> <strong>the</strong> order <strong>of</strong> 10-20 nm. The apparent quantum yield<br />

changes reported in Table 5.19 are much more significant <strong>and</strong> <strong>the</strong>re is a clear<br />

general trend for <strong>the</strong> quantum yields to increase as <strong>the</strong> temperature is<br />

decreased. One would <strong>the</strong>refore anticipate that <strong>the</strong> fluorescence lifetime<br />

values would perhaps exhibit a similar trend, <strong>the</strong> quantum yield increase<br />

possibly being a consequence <strong>of</strong> a decrease in <strong>the</strong> non-radiative decay rate.<br />

40000<br />

35000<br />

30000<br />

C<br />

25000<br />

20000<br />

15000<br />

,a10000<br />

5000<br />

¼'<br />

293K<br />

-- - --230K<br />

193K<br />

------153K<br />

133K<br />

\<br />

113K<br />

-----77K<br />

0<br />

280 300 320 340 360 380 400 420 440<br />

Wavelength (nm)<br />

Figure 5.17 Fluorescence spectra <strong>of</strong> M4AB in <strong>methyl</strong>cyclohexáne /<br />

isopentane (4:1) as a function <strong>of</strong> temperature.<br />

The data that are available at this point is conflicting. The<br />

fluorescence decays at reduced temperature have been recorded at <strong>the</strong><br />

BESS? synchrotron source in Berlin <strong>and</strong> restricted access to <strong>the</strong> source<br />

toge<strong>the</strong>r with instrumental problems have limited <strong>the</strong> quantity <strong>and</strong> quality<br />

<strong>of</strong> data which have been ga<strong>the</strong>red. The most complete dataset is for 4ABA in<br />

chiorobutane / isopentane (9:1) (Table 5.20) which suggests that <strong>the</strong><br />

fluorescence lifetime is effectively constant in <strong>the</strong> temperature range 77 - 298<br />

K.<br />

135


Solvent<br />

Species<br />

Temp<br />

K<br />

I<br />

Xem<br />

(nm)<br />

%2<br />

I<br />

t1(ns)<br />

I<br />

a1<br />

I<br />

t2(ns)<br />

mean<br />

4-Aminobenzoic acid<br />

Acetonitrile 293 331 1.08 1.14<br />

BC! / iP 293 315 1.22 0.49 0.7995 4.08 0.2005 1.21<br />

193 315 Poor 0.52 0.8404 3.94 0.1596 1.07<br />

153 315 1.26 1.36 0.9805. 3.04 0.0195 1.39<br />

113 315 1.66 0.93 0.7009 1.78 0.2991 1.18<br />

77 315 1.11 0.76 0.6495 1.58 0.3505 1.05<br />

Ethanol 293 345 1.09 0.28 0.9673 1.45 0.0327 0.32<br />

Hexane 293 31.02 1.01 0.9174 2.49 0.0826 1.13<br />

Water 293 0.63 0.79<br />

Anion 293 1.24 0.72<br />

Neutral 293 0.82 1<br />

Methyl 4-aminobenzoate<br />

Acetonitrile 293 333 1.03 1.13<br />

BCl/iP 293 330 1.19 1.27<br />

193 330 Poor 2.43 2.43<br />

153 330 Poor 2.38 2.32<br />

113 330 1.25 2.91 2.68<br />

77 330 1.12 0.75 0.7998 2.39 0.2002 0.97<br />

Ethanol 293 346 1.11 0.14 0.9959 1.48 0.0041 1.02<br />

Hexane 293 318 1.09 2.06<br />

Water 293 1353 1 1.07 1 1.34 1<br />

Table 5.20 fluorescence decay data for 4ABA <strong>and</strong> M4AB as a function <strong>of</strong><br />

solvent, pH <strong>and</strong> temperature<br />

I<br />

a2<br />

Although most <strong>of</strong> <strong>the</strong> decay data presented appears to be double<br />

exponential, <strong>the</strong> decays are usually dominated by a component with a<br />

lifetime <strong>of</strong> approximately 1.0 ns. This may be confirmed by comparing <br />

with t mean when it is noticeable that <strong>the</strong>re is only a slight difference<br />

between <strong>the</strong> two values. The decays are largely mono-exponential with a<br />

small, longer lived tail. A similar conclusion regarding <strong>the</strong> invariance <strong>of</strong><br />

<strong>the</strong> fluorescence lifetime with temperature can be drawn for 4ABA in<br />

<strong>methyl</strong>cyclohexane I isopentane (4:1) although <strong>the</strong> data are sparse. The<br />

<strong>methyl</strong> ester, on <strong>the</strong> o<strong>the</strong>r h<strong>and</strong>, does appear to have an increased<br />

fluorescence lifetime at reduced temperature in both chiorobutane I<br />

isopentane (9:1) <strong>and</strong> <strong>methyl</strong>cyclohexane I isopentane (4:1) (Table 5.20). It is<br />

136


possible that 4ABA <strong>and</strong> M4AB behave differently in terms <strong>of</strong> <strong>the</strong>ir<br />

fluorescence properties as a function <strong>of</strong> temperature, but in <strong>the</strong> absence <strong>of</strong><br />

<strong>the</strong> absorption data highlighted earlier, it is impossible to say, or to confirm<br />

that <strong>the</strong> apparent quantum yield increases are real. This is particularly so for<br />

4ABA.<br />

The corresponding N,N-di<strong>methyl</strong>amino compounds, 4DMABA <strong>and</strong><br />

M4DMAB, exhibit anomalous fluorescence in polar solvents. The<br />

fluorescence properties <strong>of</strong> both compounds were studied in <strong>the</strong> three<br />

solvent mixtures <strong>methyl</strong>cyclohexane I isopentane (4:1), chlorobutane /<br />

isopentane (9:1) <strong>and</strong> butyronitrile / isobutyronitrile (9:1); three mixtures<br />

which represent low, intermediate <strong>and</strong> high polarity respectively <strong>and</strong> which<br />

form good quality glasses at low temperature. Fluorescence spectra <strong>and</strong><br />

quantum yields were measured for both compounds in all three solvents in<br />

<strong>the</strong> temperature range 113-298K <strong>and</strong> <strong>the</strong> quantum yield values are given in<br />

Table 5.19. Both compounds produce anomalous fluorescence in<br />

butyronitrile / isobutyronitrile (9:1) <strong>and</strong> chlorobutane / isopentane (9:1), but<br />

in <strong>the</strong> low polarity alkane mixture only normal fluorescence is observed at<br />

all temperatures. This contrasts with <strong>the</strong> reports <strong>of</strong> anomalous fluorescence<br />

from ethyl 4-N,N-diethylaminobenzoate (E4DEAB) in low polarity<br />

solvents 51 . It is perhaps surprising that <strong>the</strong> replacement <strong>of</strong> <strong>the</strong> three <strong>methyl</strong><br />

groups in M4DMAB with ethyl groups should have such a substantial effect.<br />

Like <strong>the</strong>ir amino counterparts, 4DMABA <strong>and</strong> M4DMAB appear to<br />

exhibit an increasing quantum yield <strong>of</strong> normal fluorescence as <strong>the</strong><br />

temperature decreases. Once again, some or all <strong>of</strong> this increase may be<br />

attributed to increased sample absorbance at lower temperatures. However,<br />

it is clear that <strong>the</strong> steady decrease in <strong>the</strong> quantum yield <strong>of</strong> <strong>the</strong> anomalous<br />

fluorescence with decreasing temperature cannot be explained in a similar<br />

manner. Indeed, if <strong>the</strong> quantum yield values are being enhanced by<br />

increased sample absorbance, <strong>the</strong>n correcting for this will cause <strong>the</strong><br />

137


plots are found to be reasonably linear (Figure 5.19 <strong>and</strong> 5.20) <strong>and</strong> <strong>the</strong> slopes<br />

<strong>of</strong> <strong>the</strong> plots yield activation energies in <strong>the</strong> range 2.5 - 6.0 kJ mol 1 (table<br />

5.22). Rettig 88 has also derived similar linear plots for 4DMABN in<br />

propionitrile <strong>and</strong> capronitrile solvents giving values <strong>of</strong> 7.5 <strong>and</strong> 11.0 kj mo1 1<br />

respectivly.<br />

4DMABA<br />

M4DMAB<br />

Temp lIT BuCN / 1BuCN BC! lip BuCN / iBuCN BCI lip<br />

Lfl(4 a /4b) Ln(4a/4b) Ln®a/Gb) Ln(a/4b)<br />

293 0.003413 2.335 1.100 1.482 0.636<br />

230 0.004348 1.992 0.336 0.693 0.201<br />

193 0.005183 1322 -0.693 0.000 -0.105<br />

153 0.006536 2.996<br />

1<br />

-1.100<br />

1<br />

-0.916 -0.511<br />

133 10.007519 2.590 1 -1.790 1 -1.099 -1.253<br />

113 10.008849 0.069 1 -2.120 1 - -1.447<br />

Table 5.21 Variance <strong>of</strong> <strong>the</strong> natural log <strong>of</strong> <strong>the</strong> ratio <strong>of</strong> <strong>the</strong> two fluorescence<br />

peaks <strong>of</strong> 4DMABA <strong>and</strong> M4DMAB in chlorobutane /<br />

isopentane (9:1) <strong>and</strong> butyronitrile / isobutyronitrile (9:1)<br />

Solvent<br />

Compound BuCN / IP<br />

kJ mol-1<br />

BC1 / ip<br />

kJ mo1 1<br />

4DABA 4.15 5.91<br />

M4DMAB 6.47 3.96<br />

Table 5.22 Activation energies calculated from Figure 5.18 for 4DMABA<br />

<strong>and</strong> figure 5.19 M4DMAB in chlorobutane / isopentane (9:1)<br />

<strong>and</strong> butyronitrile / isobutyronitrile (9:1)<br />

139


3 —<br />

.<br />

2<br />

CY<br />

•:<br />

.5 •<br />

0.( 3 0.004 0.00..._<br />

-<br />

0.006 0.007 0.008 0.009<br />

- BCN/IBCN --<br />

-2 BestELBCN/1BCN<br />

• SCl/Ip<br />

-.-- .<br />

--ScstfitBCl/ip ---<br />

1/ T (IC-i)<br />

Figure 5.18 Arrhenius type plots <strong>of</strong> Ln(4/$b) versus liT for 4DMABA in<br />

chiorobutane / iso pentane (9:1) <strong>and</strong> butyronitrile /<br />

isobutyronitrile (9:1)<br />

-o<br />

a'<br />

a'<br />

C<br />

1.5<br />

1<br />

0.5<br />

0<br />

.t -0.5<br />

-1<br />

-1.5<br />

-2<br />

-2.5<br />

N<br />

e<br />

0.004 ~<br />

• BCN/ IBCN<br />

Best fit BCN / IBCN<br />

• BCL/il'<br />

-<br />

Best fit BCI / ip<br />

0.007 0.008 0.009<br />

_Szczzz<br />

-<br />

ii T (K-i)<br />

Figure 5.19 Arrhenius type plots <strong>of</strong> Ln(4a/$b) versus 1ff for M4DMAB in<br />

chlorobutane I iso pentane (9:1) <strong>and</strong> butyronitrile I<br />

isobutyronitrile (9:1)<br />

140


The activation energies calculated by Rettig 71 are quite close to <strong>the</strong><br />

activation energies for <strong>the</strong> viscous flow <strong>of</strong> <strong>the</strong>se two solvents, allowing <strong>the</strong><br />

conclusion that <strong>the</strong> rate determining step, <strong>and</strong> hence <strong>the</strong> main determinant<br />

<strong>of</strong> <strong>the</strong> temperature dependence, is <strong>the</strong> viscosity-dependent intramolecular<br />

twisting step (rate constant K1) which produces <strong>the</strong> a* state. The activation<br />

energies derived from our plots are generally considerably lower than <strong>the</strong><br />

activation energies for viscous flow <strong>of</strong> chlorobutane (7.1 kJ mol 1 ) <strong>and</strong><br />

butyronitrile (9.3 kJ mol 1). In view <strong>of</strong> our proposal that <strong>the</strong> anomalous<br />

fluorescence derives from ground state dimer formation, <strong>the</strong> activation<br />

energies derived from our plots <strong>of</strong> ln®a / tb) versus lIT perhaps relate to<br />

<strong>the</strong> monomer 4—> dimer equilibrium in <strong>the</strong> ground state<br />

An alternative interpretation <strong>of</strong> our observations <strong>and</strong> many "TICT -<br />

like" phenomena is that <strong>of</strong> Cazeau-Dubroca et al. 87, who proposed that <strong>the</strong><br />

anomalous fluorescence from molecules such as 4DMABN derives from a<br />

twisted ground state, <strong>the</strong> formation <strong>of</strong> which is promoted by <strong>the</strong> presence <strong>of</strong><br />

small amounts <strong>of</strong> water in <strong>the</strong> solvent. Absorption <strong>and</strong> fluorescence spectral<br />

changes at ambient <strong>and</strong> reduced temperature are cited to back up <strong>the</strong><br />

proposal. We have been conscious throughout <strong>the</strong> work reported in this<br />

<strong>the</strong>sis that water contamination <strong>of</strong> <strong>the</strong> solvents may lead to spurious<br />

results, but we have not observed any obvious differences between samples<br />

where <strong>the</strong> solvent has been scrupulously dried <strong>and</strong> those where it has not.<br />

In addition, a mechanism involving a twisted ground state would not<br />

account for <strong>the</strong> concentration dependence that is observed for <strong>the</strong><br />

anomalous fluorescence <strong>of</strong> 4DMABA <strong>and</strong> M4DMAB. We <strong>the</strong>refore believe<br />

that this mechanism is not operating in our systems.<br />

Fluorescence decay pr<strong>of</strong>iles have been measured for 4DMABA <strong>and</strong><br />

M4DMAB in <strong>the</strong> three solvent systems in <strong>the</strong> temperature range 77 - 298K.<br />

The results <strong>of</strong>'<strong>the</strong>se measurements agree given in Tables 5.23 - 5.25.<br />

141


Solvent Tern Aem V t1(ns) a1 12(ns) a2 t(ns) a3<br />

4-N,N-Di<strong>methyl</strong><strong>aminobenzoic</strong> acid<br />

BuCN / iBuCN 293 350 1.11 1.51 0.9590 5.86 0.0410 __<br />

230 345 1.99 0.021 0.9456 1.65 0.0334 6.67 0.0162<br />

193 345 1.17 0.095 0.9605 1.71 0.0382 9.76 0.0013<br />

153 345 1.32 0.39 0.5289 1.38 0.4641 4.71 0.0007<br />

113 350 1.29 1.71 0.9806 6.98 0.0194<br />

77 360 1.90 1.82 0.7479 11.11 0.2521<br />

293 485 1.54 2.39<br />

193 490 1.49 1.05 0.4477 5.78 0.5523<br />

153 495 1.45 0.25 0.7807 1.88 0.2193<br />

133 490 0.97 2.30 0.5869 4.58 0.4131<br />

77 440 1phos<br />

BC! / iP 293 355 1.24 1.15 0.4470 3.78 0.5510<br />

193 365 n/c<br />

153 365 1.48 0.06 0.9731 1.58 0.0268 9.63 0.0001<br />

133 350 1.12 0.36 0.7274 1.40 0.2717 4.35 0.0009<br />

77 355 n/c<br />

293 420 1.30 3.25 0.2218 4.82 0.1814 0.04 0.5968<br />

193 420 1.17 2.55 0.1648 15.24 0.0967 0.19 0.7385<br />

153 420 1.05 1.77 0.5096 23.82 0.4904<br />

133 420 1.19 1.05 0.5522 5.18 0.3444 70.6 0.1042<br />

77 420 0.99 0.26 0.7652 5.06 0.0841 53.8 0.1507<br />

MCH / iP 293 334 1.74 1.18 0.9923 4.93 0.0077<br />

230 334 2.46 0.38 0.6751 2.56 0.3249<br />

193 334 1.11 1.04 0.8276 2.52 0.1724<br />

153 334 Ct> 0.93<br />

133 334 1.17 0.82 0.6275 3.08 0.3725<br />

113 334 1.48 1.06 0.6543 6.78 0.4566<br />

77 334 1.37 1.38 0.4210 4.97 0.0053 1.37 0.5737<br />

293 370 1.43 1.29 0.9919 6.03 0.0081<br />

193 370 1.22 1.68 0.9581 3.67 0.0419<br />

153 370 1.34 2.10 0.1909 4.48 0.0680 0.17 0.7411<br />

133 370 n/c<br />

113 370 1.48 1.06 0.2134 6.78 0.0161 0.31 0.7706<br />

77 370 1.15 1.53 0.6904 7.46 0.0075 0.34 1 0.3021<br />

Phos = Phosphorescence occurring <strong>and</strong> fluorescence not measurable, <strong>and</strong><br />

n/c means that <strong>the</strong> decay pr<strong>of</strong>ile could not be fitted<br />

Table 5.23 Lifetime data <strong>of</strong> 4DMABA as a function <strong>of</strong> pH, solvent <strong>and</strong><br />

temperature<br />

142


Solvent<br />

I Tern Acm<br />

pK<br />

(Nfl)<br />

XI I 11(ns) a1 t2(ns) a2 t3(ns) a3<br />

ethyl 4-N,N-dimethvlaminobenzoate<br />

BuCN / IBuCN 293 340 1.23 1.07 0.9930 3.54 0.0070<br />

193 340 1.37 1.75 0.4089 3.51 0.5911<br />

153 340 1.05 1.68 0.9956 9.94 0.0044<br />

133 340 1.33 1.71 0.9802 8.62 0.0198<br />

113 340 1.29 1.85 0.9644 6.40 0.0356<br />

77 340 1.28 1.81 0.9901 8.02 0.0099<br />

293 480 0.93 0.40 0.1721 3.00 0.8279<br />

193 480 2.26 3.96 0.5239 7.32 0.4761<br />

153 480 1.03 2.41 0.9464 6.01 0.0536<br />

133 480 1.04 3.88 0.9295 6.90 0.0705<br />

113 480 1.07 4.25 0.7604 51.75 0.2396<br />

77 480 n/c<br />

BC! / iP 293 345 1.11 1.12 0.3416 3.65 0.6584<br />

193 345 1.18 1.68 0.9895 5.81 0.0105<br />

153 345 1.41 0.10 0.9137 1.76 0.0863<br />

133 345 1.59 1.44 0.9840 5.55 0.0160<br />

77 345 1.29 1.57 0.8808 5.03 0.1182<br />

293 420 1.06 3.52 0.8429 5.27 0.1571<br />

193 420 1.42 4.22 0.3333 8.18 0.0944 0.35 0.5723<br />

153 420 1.07 2.92 0.6511 8.75 0.3489<br />

133 420 1.03 2.37 0.7568 7.83 0.2432<br />

77 420 1.00 2.32 0.3000 256.9 0.2619 0.04 0.4381<br />

MCH / iP 293 334 1.19 0.95 0.9986 7.61 0.0014<br />

193 334 1.06 2.06 0.9955 9.09 0.0045<br />

153 334 Poor 2.47 0.8726 4.43 0.1274<br />

133 334 1.35 1.12 0.0496 3.25 0.0270 0.23 0.9234<br />

113 334 1.14 0.98 0.3361 5.42 0.0045 0.35 0.6594<br />

77 334 1.60 1.37 0.9852 4.46 0.0148<br />

293 370 1.40 0.98 0.9525 6.98 0.0475<br />

193 370 1.28 3.99 0.8643 9.75 0.1337<br />

153 370 1.20 2.30 0.3625 9.53 0.6375<br />

133 370 1.18 1.52 0.3312 3.54 0.3364 0.32 0.3324<br />

113 370 1.42 0.78 0.5023 2.11 0.4941 12.7 0.0036<br />

___ ____<br />

77 370 1.09 1.40 0.2210 1 2.77 0.7748 9.07 0.0042<br />

Phos = Phosphorescence occurring <strong>and</strong> fluorescence not measurable, <strong>and</strong><br />

n/c means that <strong>the</strong> decay pr<strong>of</strong>ile could not be fitted<br />

Table 5.24 Lifetime data <strong>of</strong> M4DMAB as a function <strong>of</strong> pH, solvent <strong>and</strong><br />

temperature<br />

143


Solvent Temp Xem 4DMABA M4DMAB<br />

K tmean 'U mean<br />

BuCN / iBuCN 293 340 1.69 1.09<br />

193 340 1.97 1.45<br />

153 340 1.85 1.72<br />

133 340 1.81 1.85<br />

77 340 4.16 1.87<br />

293 480 2.39 1.27<br />

193 480 3.66 2.28<br />

153 480 0.61 1.30<br />

133 480 3.24 2.05<br />

BC1 / iF 293 345 2.60 2.78<br />

193 345 - 1.78<br />

153 345 1.61 1.51<br />

133 345 1.41 1.98<br />

77 345 - 1.52<br />

293 420 1.62 3.79<br />

193 420 2.03 2.38<br />

153 420 12.58 4.95<br />

133 420 9.72 3.70<br />

77 420 8.73 6.80<br />

MCH / iP 293 334 1.25 0.96<br />

193 334 0.08 2.09<br />

153 334 2.45 2.72<br />

133 334 3.30 0.36<br />

113 334 3.15 0.58<br />

77 370 1.39 1.42<br />

Table 5.25 Mean lifetimes <strong>of</strong> 4DMABA <strong>and</strong> M4DMAB as a function <strong>of</strong><br />

solvent <strong>and</strong> temperature<br />

All <strong>the</strong> pr<strong>of</strong>iles rise promptly <strong>and</strong> <strong>the</strong>re is no evidence for precursors<br />

to <strong>the</strong> emitting state in any <strong>of</strong> <strong>the</strong> recorded pr<strong>of</strong>iles. The decay pr<strong>of</strong>iles are<br />

largely mono-exponential at room temperature, but become more complex<br />

as <strong>the</strong> temperature is reduced. As <strong>the</strong> fluorescence spectra <strong>of</strong> <strong>the</strong> compounds<br />

only effectively change in terms <strong>of</strong> <strong>the</strong>ir overall intensity as <strong>the</strong> temperature<br />

is reduced, <strong>the</strong> reason for <strong>the</strong> increased complexity <strong>of</strong> <strong>the</strong> decay pr<strong>of</strong>iles is<br />

not obvious. In many cases <strong>the</strong> pr<strong>of</strong>iles are still largely mono-exponential,<br />

but in all cases where more than one exponential component is needed to<br />

adequately fit <strong>the</strong> data, <strong>the</strong> mean lifetime has also been calculated to aid<br />

comparison <strong>of</strong> data at different temperatures.<br />

144


The fluorescence lifetime ('C or 'C mean) <strong>of</strong> 4DMABA in<br />

<strong>methyl</strong>cyclohexane / isopentane (4:1) appears to be effectively independent<br />

<strong>of</strong> temperature. There is one anomalous dataset at 113K, but virtually all <strong>the</strong><br />

o<strong>the</strong>r values lie in <strong>the</strong> region 0.5 - 2.5 ns. It may be possible to draw <strong>the</strong> same<br />

conclusion for M4DMAB in this solvent, but <strong>the</strong> data are much more<br />

scattered. Figure 5.21 gives a picture <strong>of</strong> <strong>the</strong> variation <strong>of</strong> <strong>the</strong> lifetime values ('t<br />

or I mean) for M4DMAB with temperature in <strong>methyl</strong>cyclohexane /<br />

isopentane(4:1) <strong>and</strong> lends some credence to <strong>the</strong> possibility that <strong>the</strong> lifetime<br />

is independent <strong>of</strong> temperature. This conclusion would certainly be in accord<br />

with <strong>the</strong> M4DMAB quantum yield which also appears to be temperatureindependent.<br />

Given <strong>the</strong> constant lifetime for 4DMABA in this low polarity<br />

solvent mixture, <strong>the</strong> large decrease in quantum yield (Table 5.19) appears<br />

out <strong>of</strong> place, especially when compared with <strong>the</strong> behaviour <strong>of</strong> <strong>the</strong> ester. The<br />

decrease in <strong>the</strong> 4DMABA quantum yield could reflect <strong>the</strong> lack <strong>of</strong> solubility<br />

<strong>of</strong> this compound in non-polar solvents - a point which has been made<br />

earlier about <strong>the</strong> room temperature absorption measurements. Lowering<br />

<strong>the</strong> temperature would be expected to exacerbate <strong>the</strong>se solubility problems.<br />

The behaviour <strong>of</strong> <strong>the</strong> two compounds in <strong>the</strong> butyronitrile mixture appears<br />

to be quite similar. As Figures 5.21 - 5.24 illustrate, <strong>the</strong> lifetime <strong>of</strong> <strong>the</strong><br />

normal fluorescence appears to be largely independent <strong>of</strong> temperature<br />

whilst that <strong>of</strong> <strong>the</strong> anomalous fluorescence increases somewhat as <strong>the</strong><br />

temperature is reduced.<br />

145


4.50<br />

MKII<br />

2.50<br />

ii :<br />

"S<br />

77 113 133 153 193 233 293<br />

Temperature (K)<br />

Figure 5.20 Variation <strong>of</strong> lifetime values observed for M4DMAB in<br />

<strong>methyl</strong>cyclohexane / isopentane (4:1)<br />

Rotkiewicz et al. 71 have found that 4DMABN <strong>and</strong> some related<br />

compounds form dimers <strong>and</strong> higher aggregates in non-polar solvents at<br />

reduced temperatures. These cause changes in <strong>the</strong> fluorescence spectra <strong>and</strong><br />

lead to multi-exponential decay pr<strong>of</strong>iles. In view <strong>of</strong> <strong>the</strong> fact that we also<br />

observe <strong>the</strong> latter effect, <strong>the</strong> formation <strong>of</strong> <strong>the</strong>se polymeric species in our<br />

systems cannot be ruled out. However, <strong>the</strong>se dimers or higher aggregates<br />

ei<strong>the</strong>r do not fluoresce (unlike <strong>the</strong> species reported by Rotkiewicz et al.) or<br />

<strong>the</strong>y fluoresce weakly in <strong>the</strong> same region <strong>of</strong> <strong>the</strong> spectrum as <strong>the</strong> monomers.<br />

The large Stokes shift for <strong>the</strong> anomalous fluorescence attributed to dimer<br />

formation in more polar solvents is not repeated here. A substantial<br />

amount <strong>of</strong> fur<strong>the</strong>r work is clearly needed to clarify <strong>the</strong> behaviour <strong>of</strong><br />

4DMABA in non-polar solvents.<br />

146


The comments made above about <strong>the</strong> complexity <strong>of</strong> <strong>the</strong> fluorescence<br />

decays for 4DMABA <strong>and</strong> M4DMAB in non-polar solvents also apply to <strong>the</strong><br />

o<strong>the</strong>r two solvent mixtures used here (Tables 5.23 - 5.25).<br />

too<br />

• a' -----Best fin' • V ------Bestfltb'<br />

£3.00<br />

2.50<br />

U<br />

2.00<br />

1.50<br />

g 1.00<br />

0.50<br />

0.00<br />

70 120 170 220 270<br />

Temperature (K)<br />

Figure 5.21 The variation <strong>of</strong> <strong>the</strong> fluorescence lifetime with temperature for<br />

normal <strong>and</strong> anomalous fluorescence <strong>of</strong> 4DMABA in<br />

butyronitrile / isobutyronitrile (9:1)<br />

'I<br />

C'<br />

C<br />

I)<br />

U<br />

3.50 E-02<br />

3.00E-02<br />

' 2,50E-02<br />

2.00E.02<br />

1.50&02<br />

2 1.00E-02<br />

0<br />

• a'<br />

-----Best fit a'<br />

-<br />

• V<br />

- - - -<br />

- - - - -- Best fit V - ---<br />

1<br />

- - -<br />

--- •<br />

--<br />

•<br />

W.<br />

5.00E-03<br />

• -----<br />

O.00E+®<br />

70 120 170 220 270<br />

Temperature (K)<br />

Figure 5.22 The variation <strong>of</strong> <strong>the</strong> fluorescence quantum yield with<br />

temperature for normal <strong>and</strong> anomalous fluorescence <strong>of</strong><br />

4DMABA in butyronitrile I isobutyronitrile (9:1)<br />

147


4.50<br />

4.00<br />

3.50<br />

S<br />

• a•<br />

Bnt (It .<br />

• b'<br />

— Bt(Itb'<br />

.5 3.00<br />

E<br />

2.50<br />

—1<br />

4,<br />

2.00<br />

e<br />

- - —<br />

S<br />

—a<br />

a 1.50<br />

EC<br />

1.00<br />

0.50<br />

0.00<br />

70<br />

120 170 220 270<br />

Temperature (K)<br />

Figure 5.23 The variation <strong>of</strong> <strong>the</strong> fluorescence lifetime with temperature for<br />

normal <strong>and</strong> anomalous fluorescence <strong>of</strong> M4DMAB in<br />

butyronitrile / isobutyronitrile (9:1)<br />

0.025<br />

•o 0.02<br />

• a<br />

----Best fit a<br />

6<br />

-<br />

4, • b'<br />

-<br />

E<br />

------ Bestfitb<br />

0.015<br />

.--<br />

0.01<br />

4<br />

0.005<br />

.<br />

--<br />

o I I<br />

70 120 170 220 270<br />

Temperature (K)<br />

Figure 5.24 The variation <strong>of</strong> <strong>the</strong> fluorescence quantum yield with<br />

temperature for normal <strong>and</strong> anomalous fluorescence <strong>of</strong><br />

M4DMAB in butyronitrile / isobutyronitrile (9:1)<br />

The latter effect is not large <strong>and</strong> <strong>the</strong> scatter in <strong>the</strong> data for <strong>the</strong> ester<br />

certainly invites scepticism as to <strong>the</strong> validity <strong>of</strong> this observation. However,<br />

148


an increasing lifetime for <strong>the</strong> a* fluorescence would certainly fit in with <strong>the</strong><br />

observed decrease in <strong>the</strong> fluorescence yield <strong>and</strong> would point to a<br />

temperature dependence for <strong>the</strong> radiative rate constant for <strong>the</strong> a* excited<br />

state. In view <strong>of</strong> <strong>the</strong> low fluorescence quantum yields which are observed in<br />

<strong>the</strong> mixed butyronitrile solvent, <strong>the</strong> vast majority <strong>of</strong> <strong>the</strong> excitation energy is<br />

dissipated non-radiatively. Therefore it is possible to get large changes in 4,<br />

yet much smaller changes in <strong>the</strong> lifetime, since K, (or K) >> K 1 (or K).<br />

There will presumably also be a change in <strong>the</strong> monomer


concentration effects. Taken all toge<strong>the</strong>r, <strong>the</strong> measurements made in this<br />

project <strong>and</strong> reported in this <strong>the</strong>sis point to a ground state origin for <strong>the</strong><br />

anomalous fluorescence from M4DMAB <strong>and</strong> 4DMABA.<br />

5.9 Summary <strong>of</strong> conclusions<br />

The absorption <strong>and</strong> fluorescence spectra <strong>of</strong> 4ABA <strong>and</strong> M4AB do not exhibit<br />

any obviously anomalous properties. They are always some 20 nm blue<br />

shifted from <strong>the</strong>ir N, N- dime thylamino counterparts. The latter, however,<br />

exhibit anomalous fluorescence properties in some solvents. Despite <strong>the</strong><br />

lack <strong>of</strong> concentration effects on <strong>the</strong> absorption spectra <strong>of</strong> 4DMABA <strong>and</strong><br />

M4DMAB in acetonitrile in <strong>the</strong> concentration range 10 - iO mol dm 3, <strong>the</strong><br />

anomalous fluorescence is attributed to ground state dimer formation.<br />

Evidence for this was derived from fluorescence excitation spectra,<br />

fluorescence decay pr<strong>of</strong>iles <strong>and</strong> <strong>the</strong> effect <strong>of</strong> temperature on <strong>the</strong> fluorescence<br />

properties <strong>of</strong> <strong>the</strong>se molecules. O<strong>the</strong>r mechanisms such as excimer, exciplex<br />

or TICT state formation were clearly at odds with <strong>the</strong> experimental data.<br />

However, some preliminary experiments on <strong>the</strong> effects <strong>of</strong> adding ei<strong>the</strong>r<br />

small quantities <strong>of</strong> acetonitrile to hexane <strong>solution</strong>s <strong>of</strong> 4DMABA <strong>and</strong><br />

M4DMAB or water to acetonitrile <strong>solution</strong>s <strong>of</strong> <strong>the</strong>se two compounds suggest<br />

that more than one route to <strong>the</strong> anomalous fluorescence may be operating.<br />

150


Chavter 6


References<br />

1 R.B.Cundall <strong>and</strong> A.Gilbert, "Studies in modern photochemistry",<br />

p.2, Nelson, London, (1970)<br />

2 A.Streitweiser, "Molecular Orbital Theory for Organic Chemists",<br />

John Wiley <strong>and</strong> Sons, London, (1961)<br />

3 L.C.Jones <strong>and</strong> L.W.Taylor, Anal., Chem., 2Z, 228, (1955)<br />

4 J.A.Barltrop <strong>and</strong> J.D.Coyle, "Principles <strong>of</strong> Photochemistry", p.30,<br />

John Wiley <strong>and</strong> Sons, London, (1978)<br />

5 H.L.McMurry, J.Chem.Phys., 9, 231, (1941)<br />

6 D.F.Evans, J.Chem.Phys.. fl 1424, (1955)<br />

7 G. Briegleb, "Eletronen-Donator-Komplexe", Springer, Berlin, (1961)<br />

8 J.D.Coyle, "Introduction to Organic Photochemistry", John Wiley<br />

<strong>and</strong> Sons, Bristol, (1986)<br />

9 J.G.Calvert <strong>and</strong> J.N.Pitts, "Photochemistry", John Wiley <strong>and</strong> Sons,<br />

London, (1976)<br />

10 K.W.Hodgson, PhD <strong>the</strong>sis C.N.A.A., (1986)<br />

151


11 J.Franck, Trans.Faraday Soc., a 536, (1926)<br />

12 P.M.Morse, Phys.Rev., L4. 57, (1929)<br />

13 K.K.Rohatgi-Mukherjee, "Fundamentals <strong>of</strong> Photochemistrv", John<br />

Wiley <strong>and</strong> Sons, India, (1977)<br />

14 R.S.Mulliken, T.Chem.Phys., 1 14, (1939)<br />

15 J.B.Birks, "Photophysics <strong>of</strong> aromatic molecules", John Wiley <strong>and</strong><br />

Sons, London, (1969)<br />

16 I.B.Berlman, "H<strong>and</strong>book <strong>of</strong> Fluorescence Spectra <strong>of</strong> Aromatic<br />

Molecules", Academic Press, New York, (1971)<br />

17 B.M.Wepster, Rec.Trav.chim.Pays-Bas. Zñ 335, (1957)<br />

18 B.M.Wepster, Rec.Trav.chim.Pays-Bas. a 357, (1957)<br />

19 B.M.Wepster, Rec.Trav.chim.Pays-Bas. ZZ, 491, (1958)<br />

20 J.Burgers, M.A.Hoefnagel, P.E.Verkade, H.Visser <strong>and</strong> B.M.Wepster,<br />

Rec.Trav.chim.Pays-Bai, fl 269, (1958)<br />

21 E.Drent, Chem.Phys.Lett., 2. 526, (1968)<br />

22 C.A.Parker <strong>and</strong> C.G.Hatchard, Proc.Roy.Soc., A269. 574, (1962)<br />

152


23 C.A.Parker, Adv.Photochem., 9, 1, (1974)<br />

24 Th.Forster <strong>and</strong> K.Kasper, Z.Phys.Chem. N.F.), 1, 19, (1954), <strong>and</strong><br />

Bunsenges.. Phys.Chem., a 977, (1955).<br />

25 E.Lippert, W.Luder, H.Boos,: "Advances in Molecular<br />

Spectroscopy", Ed. A.Mangini, Pergamon Press, Oxford, (1962)<br />

26 Z.R.Grabowski, K.Rotkiewicz <strong>and</strong> K.H.Grellman, Chem.Phys.Lett,<br />

19 315, (1973)<br />

27 W.Rettig, J.Phys.Chem.. , 1970, (1982)<br />

28 W.Rettig, "EPA Newsletter" No 41, p.3, (1991)<br />

29 G.R.Alms, D.R.Bauer, J.LBrauman <strong>and</strong> R.Pecora, T.Chem.Phys., a<br />

5310, (1973)<br />

30 W.Rettig, Angew.Chem.lnt.Ed.Engl. fl 971 (1986)<br />

31 R.P.Wayne, "Principles <strong>and</strong> Applications <strong>of</strong> Photochemistry",<br />

Oxford Science Publications, (1988)<br />

32 D.L.Dexter, T.Chem.Phys., a, 838, (1953)<br />

33 F.Wilkinson, Adv.Photochem.. 3. 241, (1965)<br />

153


34 J.G.Calvert <strong>and</strong> J.N.Pitts, 'Photochemistry", p.88, JohnWiley <strong>and</strong><br />

Sons, London, (1966)<br />

35 E.V<strong>and</strong>er Donckt, Progr.Reaction Kinetics 5. 273, (1970)<br />

36 L.Doub <strong>and</strong> J.M.V<strong>and</strong>enbelt, T.Am.Chem.Soc., 59, 2714, (1947)<br />

37 L.Doub <strong>and</strong> J.M.V<strong>and</strong>enbelt, T.Am.Chem.Soc., 71, 2414, (1949)<br />

38 L.Doub <strong>and</strong> J.M.V<strong>and</strong>enbelt, Organic <strong>and</strong> Biological Chemistry, ZZ,<br />

4535, (1955)<br />

39 W.H.Melhuish, T.Phys.Chem. 55, 229, (1961)<br />

40 N.Mataga <strong>and</strong> M.Ottolenghi, "Photochemical Aspects <strong>of</strong><br />

Exciplexes",<br />

41 P.Leggate <strong>and</strong> G.E.Dunn, Can.T.Chem., 41 1158,(1965)<br />

42 A.Tramer, J.Mol.Structure., £ 313, (1969)<br />

43 R.Kuhn <strong>and</strong> K.Geider, Chem.Ber., ]&j, 3597, (1968)<br />

44 A.Tramer, T.Phys.Chem., 74 887, (1970)<br />

45 D.V.S.Jian, F.S.N<strong>and</strong>el, P.Singla <strong>and</strong> D.J.Keiv, Indian I.Chem. 25A.<br />

15,(1986)<br />

154


46 D.J.Cowley <strong>and</strong> P.J.Healy, Proc.IR.Acad.. 77B. 397, (1977)<br />

47 D.J.Cowley, P.J.Healy <strong>and</strong> A.H.Peoples, T.Photochem., 9, 240, (1978)<br />

48 J.A.T.Revill <strong>and</strong> R.G.Brown, Chem.Phys.Lett.. it 433, (1992)<br />

49 R.Howell, A.C.Jones, A.G.Taylor <strong>and</strong> D.Phillips, Chem.Phys.Lett.,<br />

163, 282,(1989)<br />

50 R.J.Visser, P.C.M.Weisenborn <strong>and</strong> C.A.G.O.Varma, T.Chem.Phys.<br />

113,330, (1984)<br />

51 A.Nag, TdCundu <strong>and</strong> K.Bhattacharya, Chem.Phys.Lett.. lift. 257,<br />

(1989)<br />

52 J.A.T.Revill <strong>and</strong> R.G.Brown, 1.Fluorescence. L 107, (1992)<br />

53 P.C.M.Weisenborn, A.H.Huizer <strong>and</strong> C.A.G.O.Varma, Chemical<br />

Physics. 1Z51 425, (188)<br />

54 Z.R.Grabowski, K.Rotkiewicz, A.Siemiarczuk, D.J.Cowley <strong>and</strong><br />

W.Baumann. Nouveau T.Chimie. 3., 443, (1979)<br />

55 E.M.Kosower <strong>and</strong> H.Dodiuk, J.Am.Chem.Soc., 98. 924, (1976)<br />

56 G.Wermuth, W.Rettig <strong>and</strong> E.Lippert, Ber.Bunsenges.Phys.Chem..<br />

~85 64, (1981)<br />

155


57 W.Rettig <strong>and</strong> R.Gleiter, J.Am.Chem.Soc. 9, 4676, (1985)<br />

58 F.Heisel, J.A.Miehe <strong>and</strong> J.M.G.Martinho, Chemical Physics., 243,<br />

(1985)<br />

59 F.Heisel <strong>and</strong> J.A.Miehe, Chemical Physics, 9, 233, (1985)<br />

60 D.Huppert, S.D.R<strong>and</strong>, P.M.Rentzepis, P.F.Barbara, W.S.Struve <strong>and</strong><br />

Z.R.Grabowski, T.Photochem., fl 89, (1982)<br />

61 K.Rotkiewicz, Z.R.Grabowski <strong>and</strong> J.Jasny, T.Chem.Phys., 5.0 55,<br />

(1975)<br />

62 T.Kobayashi, M.Futakami <strong>and</strong> O.Kajimoto, Chem.Phys.Lett..<br />

63,(1986)<br />

63 J.Augüst, T.F.Palmer, J.P.Simons, C.Jouvet <strong>and</strong> W.Rettig,<br />

Chem.Phys.LetL 145,273,(1988)<br />

64 N.Nakashima <strong>and</strong> N.Mataga, Bull.Chem.Soc.Tapan. 4.5, 3016, (1973)<br />

65 O.S.Khahil, R.H.H<strong>of</strong>eldt <strong>and</strong> S.P.McGlynn, Chem.Phys.Lett., U, 479,<br />

(1972)<br />

66 E.A.Ch<strong>and</strong>ross, "The Exciplex. Conf. Proc.". New York Academic<br />

Press, (1975)<br />

67 P.Suppan, Chem.Phys.Lett. 1 28,160,(1986)<br />

156


68 P.C.M.Weisenborn, A.H.Huizer <strong>and</strong> C.A.G.O.Varma, Chemical<br />

Physics. 126, 425,(1988)<br />

69 P.C.M.Weisenborn, A.H.Huizer <strong>and</strong> C.A.G.O.Varma, Chemical<br />

Physics. ifl 437, (1989)<br />

70 R.J.Visser, P.C.M.Weisenborn <strong>and</strong> C.A.G.O.Varma,<br />

Chem.Phys.Lett. ID, 330, (1985)<br />

71 K.Rotkiewicz, H.Leismann <strong>and</strong> W.Rettig, T.Photochem. <strong>and</strong><br />

Photobiol., A. 49, 347, (1989)<br />

72 W.Rettig, Ber.Bunsenges.Phys.Chem.. 25, 259, (1991)<br />

73 O.Noboru, Bull.Chem.Soc.Tapan. a 2401, (1978)<br />

74 G.Hallas, i.Am.Chem.Soc., 5, 271, (1936)<br />

75 A.Albert <strong>and</strong> E.P.Serjeant, "The Determination <strong>of</strong> Jonisation<br />

Constants", p94, Chapman <strong>and</strong> Hall Ltd, Edinburgh, (1971)<br />

76 R.A.Velapoldi <strong>and</strong> K.D.Mielenz, NBS Special Publication. 260-64,<br />

(1980)<br />

77 J.N.Demas, T.Phys.Chem., fl, 991, (1971)<br />

157


78 D.Phillips <strong>and</strong> D.V.O'Connor, "Single Photon Counting",<br />

Academic Press, New York, (1984)<br />

79 R.Sparrow, R.G.Brown, E.H.Evans <strong>and</strong> D.Shaw, J.Chem.Soc.,<br />

Faraday Trans II., fl 2249, (1986)<br />

80 H.H.Jaffe <strong>and</strong> M.Orchin, "Theory <strong>and</strong> Applications <strong>of</strong> Ultraviolet<br />

Spectroscopy", Wiley, New York, (1963)<br />

81 A.Weller, Progr. Reaction Kinetics.!, 189, (1961)<br />

82 P.Sykes, "Mechanisms in Organic Chemistry", p56, Longman,<br />

London, (1981)<br />

83 W.T.Stacey <strong>and</strong> C.E.Swenberg. T.Chem.Phys.,, 1962, (1970)<br />

84 T.P.Carsey, G.L.Findley <strong>and</strong> S.P.McGlynn, J.Am.Chem.Soc., J&IJ<br />

4502, (1979)<br />

85 J.Herbich, W.Rettig, R.P.Thummel <strong>and</strong> J.Waluk, Chem.Phys.LflL<br />

195,556,(1992)<br />

86 W.Rettig, W.Majenz, R.Lapouyade <strong>and</strong> M.Vogel, T.Photochem. <strong>and</strong><br />

Photobiol., A,4, 49, (1992)<br />

87 C.Cazeau-Dubroca, S.A.Lyazidi, P.Caubou, A.Perigina, Ph.Cazeau<br />

<strong>and</strong> M.Pesquer, LPhys.Chem., fl 2347, (1989)<br />

LM;]


88 W.Rettig, J.Lumim, a 21, (1981)<br />

159


Appendix A


2 msc = 3100 REM value = to 1 volt<br />

4 pfin=28<br />

6 gyfin=350<br />

20 DIM 1(700)<br />

30 DIM WAVENUMBER(400)<br />

40 DIM FLUOR(400)<br />

50 MARK=0<br />

60 PRT=0<br />

90 GOSUB ten:<br />

120 CLSO:<br />

125 PRINT<br />

130 PRINT '* MENU 6/9/92<br />

140 PRINT "* (1) LOAD SAMPLE DATA "<br />

150 PRINT "* (2) CALCULATE AREA<br />

160 PRINT "* (3) RUN SAMPLE<br />

170 PRINT '* (4) QUANTUM YIELD<br />

180 PRINT "* (5) PRINT PARAMETERS<br />

215 PRINT "* e(6) DIRECTORY *"<br />

217 PRINT "* (7) QUIT *<br />

220 PRINT<br />

240 INPUT "YOUR CHOICE PLEASE"; C$<br />

250 if C$ - "1" THEN GOSUB 2180<br />

260 if C$ - "2" THEN GOSUB 2720<br />

270 if C$ - "3" THEN GOSUB 1470<br />

280 if C$ - "4" THEN GOSUB 3700<br />

290 if C$ - "5" THEN PRT - 0 GOSUB 920<br />

325 if C$ - "6" TI-lEN GOSUB 3850<br />

327 if C$ "7" THEN END<br />

330 GOTO 120<br />

410 REM SUB GRAPH<br />

420 if MARK = 0 THEN PRINT "SORRY NO DATA LOADED ": CLS:<br />

GOTO 120<br />

430 CLS GOSUB ten<br />

440 PSET (50, 0)<br />

450 LINE (50, 5)-(610, 295), 1, B<br />

451 LOCATE 6, 1 PRINT "I"<br />

452 LOCATE 7, F PRINT 'n"<br />

453 LOCATE 8, 1: PRINT "t"<br />

454 LOCATE 9, 1: PRINT "e"<br />

455 LOCATE 10, 1: PRINT "n'<br />

456 LOCATE 11, F PRINT "s'<br />

457 LOCATE 12, 1: PRINT "i'<br />

458 LOCATE 13, 1: PRINT "F'<br />

459 LOCATE 14, 1: PRINT "y"<br />

465 LOCATE 21, 30: PRINT "Wavelength (nm)' 470 FOR x = 50 TO 610<br />

STEP 40<br />

480 LINE (x, 295)-(x, 290), 1<br />

1


485 LINE (x, 5)-(x, 10), 1<br />

490 NEXT x<br />

500 FORy=3OTO28OSTEP2S<br />

510 LINE (50, y)-(55, y), 1: LINE (605, y)-(610, y), 1 520 NEXT y<br />

530 x=0<br />

550 FORI=0T07<br />

600 LQCATE20,6+(I*10) : PR1NTINT(START+x+.5)610x=x+<br />

(FINISH - START) / 7<br />

620 NEXT I<br />

625 x=0<br />

630 FORI=9TOOSTEP4<br />

640 LOCATE(I2)+1,2:PR1NTINT(x*10+.5)/ 10<br />

650 x=x+sen/9<br />

660 NEXT<br />

720 LOCATE pfin - 6, 1: PRINT "Date "; DT,, , "File "; title$<br />

730 LOCATE pfin - 4, 1: PRINT "Name "; NAM$<br />

750 LOCATE pfin - 5, 1: PRINT "Solvent "; SOLVENTs,,,<br />

"Concentration "; CONC 770 RETURN<br />

920 REM SUB PARA<br />

930 IF MARK = 1 THEN 990<br />

940 SLIT = 2.5: START = 380<br />

950 SCANSPEED = 240<br />

960 EXT = 2.5: WEXT = 340<br />

970 FINISH = 650<br />

980 sen=1<br />

985 errr = .935<br />

990 CLS<br />

1000 LOCATE 3, 5: PRINT" Parameter settings "1010 LOCATE 4, 5:<br />

PRINT<br />

1020 LOCATE 6, 1: PRINT" All following must be in order"<br />

1030 LOCATE 8, 1: PRINT "Emission slit ----------------- "; SLIT;" nm'<br />

1040 LOCATE 9, 1: PRINT "Emission wavelength start ----- "; START;"<br />

run" 1050 LOCATE 10, 1: PRINT "Emission wavelength finishing -<br />

FINISH;" nm'<br />

1060 nreadings = FINISH - START<br />

1070 LOCATE 11, 1: PRINT "Excitation slit --------------- "; EXT; " nm"<br />

1080 LOCATE 12, 1: PRINT "Excitation wavelength --------- "; WEXT;"<br />

nm"<br />

1090 LOCATE 13, 1: PRINT "Scan speed -------- ------"; SCANSPEED;"<br />

sec"<br />

1100 LOCATE 14, 1: PRINT "Sensitivity (Scale) ----------- "; sen<br />

1120 if PRT = 1 THEN GOTO 1310<br />

1130 MARK = 10<br />

1140 LOCATE 16, 1: PRINT "Do you wish to alter any parameters?<br />

(YIN)"<br />

1141 A$ = INKEYS<br />

1142 IF A$ = "Y" OR A$ = "y" THEN 1210


1143 IF A$ = "N" OR A$ = 'n" THEN 1280<br />

1144 GOTO 1141<br />

1210 LOCATE 8,32: INPUT SL: if SL c> 0 THEN SLiT = SL<br />

1220 LOCATE 9,32: INPUT ST: IF ST c> 0 THEN START = ST<br />

1230 LOCATE 10, 32: INPUT Fl: IF Fl 0 THEN FINISH = FT<br />

1240 LOCATE 11,32: INPUT EX: IF EX 0 THEN EXT = EX<br />

1250 LOCATE 12, 32: INPUT WE: IF WE 0 THEN WEXT = WE<br />

1260 LOCATE 13, 32: INPUT sc: if sc 0 0 THEN SCANSPEED = sc<br />

1270 LOCATE 14, 32: INPUT SE: IF SE o 0 THEN sen = SE<br />

1280 GOSUB 2060:<br />

1285 LOCATE 18, 1: PRINT "DO YOU REQUIRE A PRINT OUT<br />

1286 A$ = INTCEY$<br />

1287 IF A$ - "1" OR A$ - "y" THEN GOSUB 5000<br />

1288 IF A$ = "N" OR A$ = "n" THEN 1310<br />

1290 GOTO 1286<br />

1300 C1.S<br />

1310 MARK = 1<br />

1320 RETURN<br />

1470 REM SUB DATA<br />

1480 GOSUB 3210<br />

1490 GOSUB 920<br />

1500 GOSUB 410<br />

1510 nreadings = FINISH - START<br />

1520 YSCALE = 290 / (msc - 2048)<br />

1530 XSCALE = (610 - 50) I nreadings<br />

1540 DLAY = SCANSPEED I 60<br />

1545 GOSUB 2720<br />

1550 LOCATE pfin - 1, 10: PRINT "Align parameters on Fluorimeter<br />

with computer"<br />

1560 LOCATE pfin, 13: PRINT "When ready start Fluorimeter" : A = 0<br />

1564 than = arrayl(1)<br />

1570 OUT (&H303), 1<br />

1572 OUT (&H302), 0<br />

1574 FORI=1TO100<br />

1576 NEXT I<br />

1578 10 = INP(&H300) AND &HFF<br />

1580 hi = INP(&H301) AND &HF<br />

1582 SPIKE = ((hi * 256) + 10) AND &HFFF<br />

1584 F SPIKE> 3000 THEN 1564<br />

1586 IF SPIKE > 2000 THEN A = A + 1<br />

1588 IFSPIKEc2000ANDA


1622 OUT (&H303), 0<br />

1623 OUT (&H302), 0<br />

1624 FORI=1TO100<br />

1625 NEXT I<br />

1626 10 = INP(&H300) AND &HFF<br />

1627 hi = INP(&H301) AND &HF<br />

1628 s<strong>of</strong>t = ((hi * 256) + lo) AND &HFFF<br />

1630 I(count) = (s<strong>of</strong>t - 2048) / (msc * sen)<br />

1640 LINE -(50 + count * XSCALE, 295 - (s<strong>of</strong>t - 2048) * YSCALE),<br />

1645 NEXT count<br />

1650 GOSUB 1830<br />

1660 GOSUB 3600<br />

1670 RETURN<br />

1830 REM SUB FILEIN<br />

1840 LOCATE pfin -1,2: PRINT "DATA COLLECTION COMPLETE<br />

LOCATE pfin, 2: PRINT "SAVING DATA" 1850 OPEN "0', #1,<br />

title$ +<br />

1860 PRINT #1, NAM$<br />

1861 PRINT #1, SOLVENT$<br />

1862 PRINT #1, CONC<br />

1863 PRINT #1, SLIT<br />

1866 PRINT #1, EXT<br />

1864 PRINT #1, START<br />

1865 PRINT #1, FINISH<br />

1867 PRINT #1, WEXT<br />

1868 PRINT #1, SCANSPEED<br />

1869 PRINT #1, sen<br />

1870 FOR count = 0 TO nreadings<br />

1880 PRINT #1, count + START; CHR$(9); I(count) 1890 NEXT count<br />

1900 CLOSE #1<br />

1905 LOCATE pfin - 1, 2: PRINT "Do you wish to save a shorter data set<br />

(YIN)"; 1907 INPUT ans$<br />

1908 IF ans$ = "1" OR ans$ = "y" THEN GOTO parse:<br />

1909 IF ans$ = "N" OR ans$ = 'n" THEN 1911<br />

1910 GOTO 1907<br />

1911 RETURN<br />

2060 REM SUB BLANK<br />

2080 LOCATE pfin - 3, 1: PRINT"<br />

"2090 LOCATE pfin - 2, 1: PRINT"<br />

"2091 LOCATE pfin - 1, 1: PRINT"<br />

"2095 LOCATE pfin, 1: PRINT"<br />

"2100 RETURN<br />

2180 REM SUB FILERET<br />

2200 CLS : LOCATE 1,4: PRINT "ENTER NAME OF FILE TO BE<br />

RETRIEVED": INPUT title$<br />

2210 OPEN "I", #1, titleS + ".XLB"<br />

2211 INPUT #1, NAM$<br />

El


2212 INPUT #1, SOLVENT$<br />

2213 INPUT #1,CONC:<br />

2214 INPUT #1, SLIT<br />

2215 INPUT #1, EXT<br />

2216 INPUT #1, START<br />

2217 INPUT #1, FINISH<br />

2218 INPUT #1, WEXT<br />

2219 INPUT #1, SCANSPEED<br />

2220 INPUT #1, sen<br />

2225 nreadings = FINISH - START<br />

2230 FOR count = 1 TO nreadings<br />

2240 INPUT #1, A$<br />

2242 temp$ = LEFT$(A$, 4)<br />

2243 A = LEN(A$)<br />

2244 I(count) = VAL(fflGHT$(A$, A - 5))<br />

2250 NEXT<br />

2260 CLOSE #1<br />

2270 MARK = 1<br />

2280 CLS<br />

2290 LOCATE 1, 4: PRINT "data now loaded"<br />

2300 FOR x = 1 TO 5000: NEXT<br />

2310 CLS<br />

2320 RETURN<br />

2480 REM SUB DRAW<br />

2490 GOSUB ten: PSET (150, 400), 0<br />

2500 IF MARK = 0 THEN LOCATE 24,2: PRiNT "SORRY NO DATA<br />

STORED": RETURN 2515 YSCALE = 290 I (msc - 2048)<br />

2517 XSCALE = (610-50) / nreadings<br />

2525 PSET (50, 295)<br />

2530 FOR count = 0 TO (FINISH - START)<br />

2535 s<strong>of</strong>t = I(count) * msc * sen<br />

2540 LINE -(50 + count * XSCALE, 295 - (s<strong>of</strong>t) * YSCALE), 12550 NEXT<br />

2560 RETURN<br />

2720 GOSUB ten<br />

2730 GOSUB 410<br />

2731 GOSUB 2490<br />

2759 GOSUB 2060<br />

2760 LOCATE pfin - 2, 2: INPUT "IS THIS THE CORRECT CHART<br />

(YIN)";<br />

2780 IF p$ = "1" OR p$ = "y" THEN 2800<br />

2790 LOCATE 23,2: PRINT "PLEASE LOAD THE REQUIRED DATA<br />

FIRST": FOR x = 1 TO 500: NEXT: GOTO 120<br />

2800 GOSUB 2060: LOCATE pfin - 1, 1: INPUT "INPUT INITIAL<br />

WAVELENGTH "; IN1T<br />

2805 LOCATE pfin, 1: INPUT "FINAL WAVELENGTH "; FINI<br />

2810 IFINITCSTARTORINIT>FINISHORFINI>FINISHORFINI<<br />

INIT THEN GOSUB 2060 flSE 2820 GOSUB 2060: LOCATE pfin, 1:<br />

5


PRINT "SORRY RANGE IS UNACCEPTABLE": FOR x = 1 TO 5000:<br />

NEXr: GOTO 2800<br />

2820 REM INTENSITY AT INIT=INT (VAR)<br />

2830 FORN=OTOFINTSH - START<br />

2840 FLUOR(N) = 1(N)<br />

2850 WAVENTJMBER(N) = 1E+07 / (START + N) 2860 NEXT N<br />

2870 Area = 0<br />

2880 FOR WAVENO = 40000 TO 13000 STEP -500<br />

2890 if WAVENO c 1E+07 I (FINI + 1) THEN FINTENSITY = 0: GOTO<br />

2970 2900 IF WAVENO> 1E+07 / (IN1T - 1) THEN FINTENSITY = 0:<br />

GOTO 2970 2910 FOR N = 1 TO FINISH - START<br />

2920 if WAVENTJMBER(N) c WAVENO THEN 2950<br />

2930 NEXT N<br />

2940 GOTO 3000<br />

2950 FINTENSITY = (FLUOR(N -1) + (FLUOR(N) - FLUOR(N - 1)) *<br />

(WAVENIJMBER(N - 1) - WAVENO) / 500) 2970 if WAVENO =<br />

40000 THEN Area = Area + FINTENSITY / 2: GOTO 2990<br />

2980 if WAVENO = 13000 THEN Area = Area + FINTENS1TY / 2 ELSE<br />

Area = Area + FINTENSITY<br />

2990 NEXT WAVENO<br />

3000 GOSUB 2060<br />

3010 LOCATE pfin - 1,2: PRINT" AREA BETWEEN"; INIT;<br />

"nm AND"; FINI; "nm IS"; INT(Area * 1000) I 1000 3011 LOCATE pfin,<br />

2: PRINT " Press Shift PRTSCR for Hard Copy any o<strong>the</strong>r key to return<br />

to menu"<br />

3012 A$ = INKEYS<br />

3013 if A$ - " THEN 3012<br />

3014 if A$ = 'Y" OR A$ = "y" THEN GOSUB sdump ELSE 3015<br />

3015 GOSUB 3600<br />

3017 RETURN<br />

3210 REM SUB INFO<br />

3220 PRINT : PRINT : PRINT: PRINT<br />

3230 if DT$ = " THEN 3240 ELSE 3250<br />

3240 INPUT "ENTER TODAYS DATE "; DT<br />

3250 INPUT "ENTER NAME OF FILE "; titleS<br />

3260 INPUT "ENTER NAME OF COMPOUND"; NAM$<br />

3270 INPUT "ENTER SOLVENT USED"; SOLVENTS<br />

3280 INPUT "ENTER CONCENTRATION "; CONC<br />

3290 RETURN<br />

3600 REM SUB OPTIONS<br />

3610 GOSUB 2060<br />

3620 LOCATE pfin - 3, 15: PRINT "OPTION (1) Return to Menu"<br />

3630 LOCATE pfin - 2, 15: PRINT" (2) Calculate Area"<br />

3650 LOCATE pfin, 15: INPUT "PLEASE ENTER YOUR CHOICE"; R$<br />

3660 IF R$ = "1" THEN CLS: GOTO 120<br />

3680 IF ES = "2" THEN GOSUB 2060: GOSUB 2760<br />

3690 RETURN


3700 REM SUB YIELD<br />

3710 CLS<br />

3720 PRINT: PRINT: PRINT<br />

3730 INPUT "ENTER ABSORBANCE OF QS"; ABSQ<br />

3740 INPUT "ENTER ABSORBANCE OF SAMPLE."; ABSS<br />

3750 INPUT "ENTER AREA UNDER QS"; ARQ<br />

3760 INPUT "ENTER AREA UNDER SAMPLE"; ARS<br />

3770 QY = ((ARS * ABSQ) / (ARQ * ABSS)) * .55 3780 CLS<br />

3800 PRINT<br />

3810 PRINT "* QUANTUM YIELD="; INT((QY * 100) + .5) / 100; "<br />

3820 PRINT<br />

3825 INPUT "PRESS I TO CONTINUE"; A$<br />

3826 IF A$ = " THEN 3826 ELSE 3830<br />

3830 GOSUB ten<br />

3840 RETURN<br />

3850 CLS<br />

3870 FILES A$<br />

3880 INPUT "PRESS Y TO CONTINUE"; K$<br />

3890 IF K$ = " THEN 3890<br />

3900 RETURN<br />

4000 NEXT<br />

5000 LOCATE 1, 1: LPRINT" PARAMETER SEflINGS"<br />

5001 LOCATE 2, 1: LPRINT "<br />

5002 LOCATE 3, 1: LPRTNT "EMISSION SLiT "; SLIT;<br />

n m"<br />

5003 LOCATE 4, 1: LPRINT "EMISSION WAVELENGTH START<br />

START; "nm"<br />

5004 LOCATE 5, 1: LPRINT "EMISSION WAVELENGTH FINISHING";<br />

FINISH; "nm'<br />

5005 LOCATE 6, 1: LPRINT "EXCITATION SUT "; EXT; 'nm"<br />

5006 LOCATE 7, 1: LPRINT "EXCITATION WAVELENGTH<br />

WEXT;<br />

"nm'<br />

5007 LOCATE 8, 1: LPRINT "SCANSPEED<br />

SCANSPEED; "sec"<br />

5008 LOCATE 9, 1: LPRINT "SENSITIVlTY "; sen<br />

5009 RETURN<br />

5010 ten:<br />

5020 SCREEN 12<br />

5030 RETURN<br />

5040 sdump:<br />

5050 LOCATE pfin, 1: PRINT"<br />

7000 'SCREEN DUMP ROUTINE for any 9 pinGRAPHICS<br />

printer<br />

7010 PRT:<br />

7020 LPRINT CHR$(13) 'clears print<br />

buffer<br />

7


7050 LPRINT CHR$(27); "3"; CHR$(12) 'sets line spacing to 12/216"<br />

7060 FOR y = 1 TO 350 STEP 8 'sets y scanning range (8 bit<br />

mode)<br />

7070 GOSUB scan: 'calls subroutine to scan x & print<br />

7080 NEXTy<br />

7090 RETURN<br />

7100 scan:<br />

7104 w = 3<br />

7105 . LPRINT CHR$(13) 'CR printer code<br />

7110 x = 300 'width <strong>of</strong> paper in head<br />

movements<br />

7120 nl = x MOD 256 'number <strong>of</strong> bit image data<br />

7130 n2 = INT(x / 256)<br />

7140 LPRINT CHR$(27); "L";<br />

7145 CHR$(nl); CHR$(n2); 'sets high density bit image<br />

mode 7150 EORJ = 1 TO x<br />

7160 sc = 640 / 6 'scales <strong>the</strong> screen to printer<br />

7170 sum=0<br />

7180 IFPOINT(j*sc, y) w THEN sum = sum +4<br />

7240 IFPOINT(j*sc, y +6) w THEN sum = sum +2<br />

7250 IFPOINT(j* sc, y +7)o w THEN sum = sum +1<br />

7260<br />

7270<br />

LPRINT CHR$(sum);<br />

NEXT j.<br />

7380 RETURN<br />

7400 parse:<br />

7410 LOCATE pfin - 1, 1: PRINT" "LOCATE pfin, 1: PRINT"<br />

7420 LOCATE pfin - 1, 1: INPUT "Enter number <strong>of</strong> data points required";<br />

ans<br />

j = nreadings / ans<br />

7430 jump=INT(j*10+.5)/10<br />

7450 title$ = titleS +<br />

8850 OPEN "0", #1, titleS<br />

8860 PRINT #1, "Name - "; CHR$(9); NAM$<br />

8861 PRINT #1, "Solvent - "; CHR$(9); SOLVENTS<br />

8862 PRINT #1, "[Conc] - "; CHR$(9); CONC<br />

8863 PRINT #1, "Em slit - "; CHR$(9); SLIT<br />

8866 PRINT #1, "Ex slit - "; CHR$(9); EXT<br />

8864 PRINT #1, "Start - "; CHR$(9); START<br />

8865 PRINT #1, "Finish - "; CHR$(9); FINISH<br />

8867 PRINT #1, "Ex Lmax - "; CHR$(9); WEXT<br />

8868 PRINT #1, "Scanspeed "; CHR$(9); SCANSPEED<br />

E1


8869 PRINT #1, "Sensitivity"; CHR$(9); sen<br />

8870 FOR count = 0 TO nreadings STEP jump<br />

8880 PRINT #1, count + START; CHR$(9); I(count)<br />

8890 NEXT count<br />

8900 CLOSE #1<br />

8905 LOCATE pfin - 1, 1: PRINT "You now have two data sets one";<br />

title$;<br />

".XLB containing all data"<br />

8906 LOCATE pfin, 1: PRINT "<strong>and</strong> ano<strong>the</strong>r "; title$;" containing parsed<br />

data'<br />

8907 FOR x = 0 TO 5000: NEXT x<br />

8911 GOTO 1911


Appendix B


Volume laS, number 5,6 CHEMICAL PHYSICS LZfltRS 17 Januaty 1992<br />

Excimer venus TICT state formation in polar <strong>solution</strong>s<br />

<strong>of</strong> <strong>methyl</strong> 4- (N',N-di<strong>methyl</strong>amino )benzoate<br />

John A.T. Revill • <strong>and</strong> Robert G. Drown<br />

• DePartment QIChemisay. Lancashire Po&technic, hrswn. Lancashire FRI 2712 IlK<br />

• SERCOwesbury Laruio,y, Wanington, Cbnhfre WAI lAD. (iN<br />

Ralved 29 September 1991; in final (cnn 2$ October 1991<br />

The fluorescence properties <strong>of</strong> <strong>methyl</strong> 4-(N,N4imeth4amlno)beno.te In a nricty <strong>of</strong> solvents are reported. No.n,sr nunmnce<br />

from a distribution <strong>of</strong> excited wiute/wivent arrausements Is observed In most solvent Additionally, in acetonitrile, a<br />

Iongr wavelength fluorescence b<strong>and</strong> Is observed which is auributed to excited dieter fluorescence. Itis proposed that <strong>the</strong> excited<br />

dimer is formed by dbt excitation from <strong>the</strong> pound state.<br />

1. Introdactlon<br />

The concept <strong>of</strong> <strong>the</strong> twisted intramolecular charge<br />

transfer ('ncr) state as an explanation for anomalous<br />

dual fluorescence has met with wide acceptance.<br />

In donor-acceptor molecules (typified by <strong>the</strong><br />

4-(N,N-dialkylamino)benzcnitriles), <strong>the</strong> different<br />

relative orientations which can be adopted by <strong>the</strong><br />

donor <strong>and</strong> acceptor parts <strong>of</strong> <strong>the</strong> molecule give risc to<br />

excited states which can be identified with "normal"<br />

<strong>and</strong> TIC' fluorescence (1-3 ]. It is now recognised<br />

that TICT slates may play a role in <strong>the</strong> photochemistry<br />

<strong>of</strong> a wide range <strong>of</strong> molecules <strong>and</strong> molecular systents<br />

( 4 1<br />

TICT state formation in <strong>the</strong> 4-(N,N.dialkylamino)benzonitriles<br />

has been investigated for a wide<br />

range <strong>of</strong> donor groups (act eg. ref. (] <strong>and</strong> references<br />

<strong>the</strong>rein), but variation in <strong>the</strong> acceptor system<br />

is more difficult. Dual fluorescence has been obaerved<br />

for <strong>the</strong> <strong>methyl</strong> <strong>and</strong> ethyl estcn <strong>of</strong> 4-(N,N-diethyLamino)<br />

<strong>and</strong> 4-(N,N-di<strong>methyl</strong>amin0benioic<br />

acid. These observations have been ascribed to TICT<br />

state formation 11,6-81. This assignment has been<br />

challenged by Visser et al. (9] who believe that salutc-solvent<br />

exciplexea arc <strong>the</strong> causc <strong>of</strong> <strong>the</strong> dual fluorescence<br />

<strong>and</strong> by Howell ci aL (10) who have evidence<br />

for excimer formation in <strong>solution</strong> <strong>and</strong> for<br />

dimer (or lilgber aggregate) in a supersonic jet. Excimer<br />

fluorescence has also recently been reported in<br />

polar <strong>solution</strong>s <strong>of</strong> hcxadccyl 4-(N,N-di<strong>methyl</strong>axnino)-benzoate<br />

[II). However, TICT state for'<br />

mation has also been claimed in <strong>the</strong> related system<br />

<strong>of</strong> <strong>methyl</strong> 2-meihoxy-4- (N.N-disnethylamino) henzoate<br />

[12,13], We arc investigating <strong>the</strong> photophysica<br />

<strong>of</strong> <strong>methyl</strong> 4-(N,N-di<strong>methyl</strong>amino)-benroate as<br />

part <strong>of</strong> a comprehensive <strong>study</strong> <strong>of</strong> <strong>the</strong> photochemistry<br />

<strong>of</strong> <strong>the</strong> amino- <strong>and</strong> (N,N-dimcthylamino )benzoic<br />

<strong>acids</strong> <strong>and</strong> <strong>the</strong>ir <strong>methyl</strong> <strong>esters</strong>. We report here <strong>the</strong> results<br />

<strong>of</strong> our measurements on this compound under<br />

various room temperature conditions.<br />

2. ExperImental<br />

Methyl 4-(N,N'di<strong>methyl</strong>amino)bcnzoate was prepared<br />

by <strong>the</strong> method <strong>of</strong> HaJIas (14]. A <strong>solution</strong> <strong>of</strong><br />

4-di<strong>methyl</strong>aminobcnzoic acid (0.06 mol) in methanol<br />

(os cnr 3 ) was added to boron trifluoridemethanol<br />

complex (0J8 mol) <strong>and</strong> <strong>the</strong> mixture boiled<br />

under reflux for 4 h. After cooling, <strong>the</strong> mixture was<br />

poured into saturated sodium hydrogen carbonate<br />

<strong>solution</strong> <strong>and</strong> <strong>the</strong> ester extracted with ethyl acetate.<br />

The ethyl acetate extract was dried <strong>and</strong> <strong>the</strong> ethyl acetate<br />

removed to yield <strong>the</strong> crude ester which was twice<br />

recrystallised from ethanol Yield 10.02 g (79%),<br />

m.p. ll0-lll°C (literaturem.p. 111-113°C (IS)).<br />

The solvents used in <strong>the</strong>se studies were ei<strong>the</strong>r<br />

spectrophotometric or spectr<strong>of</strong>luorirnetric grade <strong>and</strong><br />

00094614/9215 CL® 0 1992 Elsevler Science ?ublishcn B.V. All riJtts reserved. 433<br />

10


Volume IU. number 5,6 - CHEMICAL PHYSICS LETFEPS 17 January 1992<br />

were used as received except for acetonitrile. For <strong>the</strong><br />

fluorescence measurements (spectra <strong>and</strong> decays), <strong>the</strong><br />

acetonitrile was distilled over phosphorus pentoxide<br />

<strong>and</strong> collected under an inert atptospbert The pure<br />

solvent <strong>and</strong> all relevant samples were stored under<br />

argon to inhibit <strong>the</strong> absorption <strong>of</strong> water oroxygen.<br />

Water was doubly distilled <strong>and</strong> <strong>the</strong>n passed through<br />

an ion exchange system.<br />

Absorption spectra were measured on a Perkin-<br />

Elmer Lambda 3 spectrophotometer using quartz<br />

cells <strong>of</strong> varying pathlenghts. *11 working <strong>solution</strong>s<br />

were prepared from a stock <strong>solution</strong> on a daily basis<br />

<strong>and</strong> <strong>the</strong> integrity <strong>of</strong> <strong>the</strong> stock <strong>solution</strong> was continuously<br />

monitored. The accuracy <strong>of</strong> <strong>the</strong> measurements<br />

is estimated at ±2 am for wavelengths <strong>and</strong> ±50%<br />

for extinction coefficients.<br />

Fluorescence spectra were measured on a Perkin-<br />

Elmer LSS spectr<strong>of</strong>luorimeter in fully corrected<br />

mode. The fluorescence spectra were collected on a<br />

Rae B microcomputer which was capable <strong>of</strong> calculating<br />

<strong>the</strong> area under <strong>the</strong> spectnaan for quantum<br />

yield determinations. Quantum yields were measured<br />

on optically dilute samples (absothance<br />


Volume ItS. number 3,6 CHEMICAL PHYSIC LZFItRS 37 January 1992<br />

I<br />

1'<br />

I<br />

Wavelength (am)<br />

Fig I.Absorptioa aadfluorncencespectno(M4DMABIn seetonithle (1.63x 10'mol dnr').<br />

Tabk I<br />

Absoipsion <strong>and</strong> fluorescence properties <strong>of</strong> MIDMAB us function <strong>of</strong> solvent<br />

Solvent Absorption Fluorescence<br />

A. (cm) extinction coefficient 2,.,, (cm) 4 t, (us)<br />

(dm'molcm")<br />

hexane 212 17000 348 0.300 1.30<br />

ethanol 299 19800 341 0.067 0.23, 1.16"<br />

acntonithle 292 19100 335,480 -' - fl<br />

"titer 316 16600 346 0.022 0.11,3.30"<br />

"The quantum yield In acetonluilt niles with eoncentntlon.<br />

"Double exponential decays observed. "Complex behaviour - see text.<br />

different solute/solvent interactions 1211 as has been<br />

proposed by Howell et al. for M4DMAB in mixtures<br />

<strong>of</strong>non-polarsolvents 1101. The possibitity<strong>of</strong>a nonfluorescent<br />

TICT state in <strong>the</strong>se solvents is not ruled<br />

out, however, <strong>and</strong> we are undertaking fur<strong>the</strong>r measurements<br />

in hydrogen-bonding solvents in an attempt<br />

to clarify <strong>the</strong> position.<br />

The fluorescence properties <strong>of</strong> M4DMAB in acetonitrite<br />

are anomalous in that a second fluorescence<br />

peak is observed at 480 nm. This confirms <strong>the</strong> observations<br />

<strong>of</strong> o<strong>the</strong>r worken [l,6—t0J. Although mere<br />

is no obvious change in <strong>the</strong> absorption spectrum as<br />

concentration is varied, <strong>the</strong> fluorescence spectra do<br />

change <strong>and</strong> <strong>the</strong> intensity <strong>of</strong> <strong>the</strong> long wavelength peak<br />

decrcascs with decirasing concentration (fig. 2). The<br />

anomalous ("red") fluorescence peak is <strong>the</strong>refore<br />

outwardly exhibiting <strong>the</strong> typical characteristics <strong>of</strong><br />

excimer formation 1221. However, <strong>the</strong> fluorescence<br />

quantum yields for <strong>the</strong> two fluorcscence b<strong>and</strong>s do<br />

not vary with concentration in <strong>the</strong> manner which<br />

would be expected for excimer formation <strong>and</strong> <strong>the</strong><br />

fluorescence decay pr<strong>of</strong>iles do not show <strong>the</strong> idationship<br />

that would be expected on <strong>the</strong> basis <strong>of</strong> cxcinier<br />

formation. Indeed, it is <strong>the</strong> "red" fluorescence<br />

435<br />

13


Volume ISt rnunbes 5,6 ! ;b'j.i (04 E u In 17januvy 1992<br />

I<br />

I LU<br />

I<br />

I<br />

0<br />

7 -.--<br />

loaD<br />

p<br />

F1. 2. Fluoresccncc .pectz. <strong>of</strong> M4DMAB in aeetonlirile ass function <strong>of</strong>vonautration: (—) 5.Ox 10-' rnol dnr", () 2.5x IV'mol<br />

dnr', (.--) I.OX 10 6 mol dir', (---) LOX I04moldnr1, (--) 2.5x lr'noldzr'.<br />

which exhibits a single exponential decay whilst <strong>the</strong><br />

"blue" peak is double exponential!<br />

Over a range <strong>of</strong> concentrations (5x 10'-Sx 10<br />

inol dm') <strong>and</strong> excitation wavelengths (270-340<br />

nm), <strong>the</strong> "red" fluorescence b<strong>and</strong> gives excellent<br />

single exponential fits to <strong>the</strong> expciimcntal decay pr<strong>of</strong>ile<br />

with lifetimes in <strong>the</strong> range 2.36-2.71 iii. For <strong>the</strong><br />

same <strong>solution</strong>s, <strong>the</strong> "blue" fluorescence is clearly not<br />

a single exponential but is adequately explained by<br />

a sum <strong>of</strong> two exponentials with lifetimes in <strong>the</strong> range<br />

1.27-1.70 <strong>and</strong> 2.21-2.77 as (table 2). At present we<br />

are unsure as to whe<strong>the</strong>r <strong>the</strong>re is a correlation between<br />

<strong>the</strong> "red" lifetime <strong>and</strong> <strong>the</strong> longer <strong>of</strong> <strong>the</strong> two<br />

"blue" lifetimes. It is possible that <strong>the</strong> latter represents<br />

a process whereby <strong>the</strong> 'S" excited state is<br />

convened to <strong>the</strong> "blue" excited state as proposed by<br />

Howell et al. (10]. However, <strong>the</strong> discrepancy between<br />

<strong>the</strong> two lifetimes <strong>and</strong> <strong>the</strong> fact that <strong>the</strong> a2 values<br />

in <strong>the</strong> double exponential fits are positive when<br />

a risetime should have a negative pre-exponential<br />

Table 2<br />

fluorescence decay pm'pertics <strong>of</strong> M4DMAD in acctonithlc as a function <strong>of</strong>onczntntlon<br />

Conccntntion 'Eftlltton a, r, (os) a, 1, (ns)<br />

(moldnr') wavelength(nm)<br />

3.0x10' 355 0.0403 1.70 0.0003 2.21<br />

480 0.0365 2.36<br />

LOX 10' 355 0.0395 1.65 0.0007 2.45<br />

480 0.0372 2.49<br />

Lox 10-' 355 0.0191 1.69 0.0008 2.49<br />

480 0.0375 2.71<br />

436<br />

iV


Vohuna 151, number 3.6 OIEMICAL PHYSICS LEfltRS *7 January 1992<br />

factor suests that <strong>the</strong> similarity in <strong>the</strong>se lifetimes<br />

is purely coincidentaL<br />

Adapting <strong>the</strong> accepted terminology for <strong>the</strong>se systems<br />

where <strong>the</strong> "normal" blue fluorescence is attributed<br />

to an ached state labelled be <strong>and</strong> <strong>the</strong> red fluorescencc<br />

to a state labelled C we can identify a<br />

number <strong>of</strong> difreitnt mechanisms for this fluoracence<br />

behaviour. Unless <strong>the</strong>re exists a very thort-lived<br />

precursor whose lifetime is to short that we are unable<br />

to separate it from <strong>the</strong> instrumental response<br />

function when we model <strong>the</strong> fluorescence decays <strong>of</strong><br />

as <strong>and</strong> b states, both 0 <strong>and</strong> b O are excited directly<br />

from <strong>the</strong> ground stata Once <strong>the</strong> two excited states<br />

have been formed, it h possible that a' converts to<br />

b' but <strong>the</strong>re is no evidence for <strong>the</strong> reverse process.<br />

Our conclusion from this data Is that <strong>the</strong> fluorescence<br />

spectrum <strong>and</strong> dccay pr<strong>of</strong>ile <strong>of</strong> <strong>the</strong> V state Is<br />

comprised <strong>of</strong> contributions from a distribution <strong>of</strong><br />

excited solute/solvent arrangements, <strong>the</strong> decay kinetics<br />

<strong>of</strong> which are adequately described by a sum <strong>of</strong><br />

two exponentials [2 3,24 ) Proving that <strong>the</strong> decays<br />

are multicxponcntial will require high quality data<br />

with large numbers <strong>of</strong> counts in <strong>the</strong> decay pr<strong>of</strong>iles<br />

(21) but we intend to undertake <strong>the</strong> required experiments<br />

to acquire this data in <strong>the</strong> near future.<br />

Unlike Howell a al. [10] we have no evidence to<br />

suggest that a' derives from b'. The additional obtervation<br />

<strong>of</strong> <strong>the</strong> concentration dependence <strong>of</strong> <strong>the</strong> a'<br />

fluorescence b<strong>and</strong> leads usto conclude that as is an<br />

excited dirtier (or possibly a higher aggregate) which<br />

is populated directly from <strong>the</strong> ground state by absorption<br />

<strong>of</strong> a photon <strong>of</strong> light Given <strong>the</strong> lack <strong>of</strong> concentration<br />

dependence <strong>of</strong> <strong>the</strong> ground state absorption<br />

spectrum, <strong>the</strong> ground state "dimer" may not<br />

involve a strong interaction between molecules.<br />

However, a loose association is sufficient to organize<br />

<strong>the</strong> molecules such that <strong>the</strong>y are in approximately<br />

<strong>the</strong> con'ect position to form an excited dimcr when<br />

a photon is absorbed by one <strong>of</strong> <strong>the</strong> molecules. We are<br />

conscious that <strong>the</strong> time re<strong>solution</strong> <strong>of</strong> our time-correlated,<br />

single photon counting system would not allow<br />

us to measure a rapid risetime (perhaps


Volume III. number 5.6 CHDIICAL rHYsia LEflS 17 J.nuaq £992<br />

(191DJ.S. BITSI, A. Duteb R.E. Imh<strong>of</strong>, B. Nadohki <strong>and</strong> I.<br />

Soutar.J.Pbctachenj8 (1997)239.<br />

[201 R. Snow, R.O. Bgown. Efl Ev.na <strong>and</strong> D. Sbaw.S. C,cm.<br />

Soc. Faraday Trans 1182 (1986) 2249.<br />

[21) D.R. Saint, <strong>and</strong> Wi. Ware, Gate.. Pitys. Letter, 126<br />

(1986)7.<br />

[22)S.t BirkL Photophyaa at aromaik molecule, (Wiley-<br />

Intendtnee, New Von, 1970).<br />

1231 5.9. Mad. <strong>and</strong> D. Pbiflip.. J. Otca Soc. Faraday Trw.<br />

1183 (1987) 1941.<br />

(24)W. RettiL P4. Vo,eI. E. Lippat <strong>and</strong> H. Otto, O,em. Pby.<br />

Letter, 103 (1986) 381.<br />

438<br />

16


Jaw'.aiefflanwa VOL 1 No, 2 1992<br />

Anomalous fluorescence Properties <strong>of</strong> 4-N,N-<br />

Di<strong>methyl</strong><strong>aminobenzoic</strong> Acid in Polar Solvents<br />

John A. T. RievIll' <strong>and</strong> Robert C. BrownW<br />

ReedS Much 17. IPPt n,Ud laTh 199Z• .ecqtS Is 24 1992<br />

4.?6N-Dimcthylsmlnotcnzolc acid eth1bfts anomalous fluorescence In polar <strong>and</strong> hydro;enbonding<br />

solvents. The fluorescence spectra <strong>and</strong> kinella naggest that this arises due to <strong>the</strong><br />

formation <strong>of</strong> a ground-sat; dimer or blgbcr polymer- Preliminary measurements in bent.<br />

containing small amounts <strong>of</strong> polar acstonitrila do not S. out <strong>the</strong> possibility <strong>of</strong> exnipiex fornation<br />

also ocarring.<br />

KEY WORDS; Bwawa anomalous fluotesesna; 444N-dlmttbytsStolc acM dImasSn.<br />

INTRODUCTION<br />

Derivatives <strong>of</strong> 4-a%N.4lmethybm1nobenzoic add<br />

such asS nitrile <strong>and</strong> various eaten have been known<br />

for • number <strong>of</strong> yen to show anomalous fluorescence<br />

properties (1-3]. Two fluorescence b<strong>and</strong>s are observed<br />

for <strong>the</strong>se compounds In some solvents. The shorter.<br />

wavelength emission, with a Stokes shift <strong>of</strong> a few thous<strong>and</strong><br />

wavcnwnbcra, Is Identified as "normal" fluorescent<br />

from <strong>the</strong> knily exefted Lc., Franck-Coodco (It)<br />

slngjet state to <strong>the</strong> K ground state [WC)]). In current<br />

termjnolov this is also referred to as a b' state. The<br />

anomnious. longer-wavelength emission (from a state Ishelen<br />

a) has a number <strong>of</strong> possible origins.<br />

The suggestion that <strong>the</strong> a' emission results from a<br />

twisted intramolecular change transfer (flC) state has<br />

met with quite wide acceptance (4-4). ThIs transition<br />

would be labeled S, (CT) —. $ (PC) in <strong>the</strong> notation<br />

adopted by HeidI ci al. (7). However, Vatma <strong>and</strong> w-<br />

'G.a,thoy Dspanment, Univmi?y <strong>of</strong> Cesesi Uncasbire. Pretco.<br />

Lanhire, Pit 2H1 UK.<br />

'sac Dsresbwy Ubotstmy, Waningion. aesbfte. WA4 4AD UK.<br />

To whom amtjposexz slould bc addteex&<br />

workers, at an early stage in <strong>the</strong> studies <strong>of</strong> <strong>the</strong>se mote--<br />

cults, proposed that <strong>the</strong> a' emission was due to a solutel<br />

solvent exeiplex (8) <strong>and</strong> have maintained this belief In<br />

a series <strong>of</strong> later papers [see 9 <strong>and</strong> references <strong>the</strong>rein).<br />

More recently, Phillips' group has presented evidence<br />

that <strong>the</strong> a' emission <strong>of</strong> 4-N N-dl<strong>methyl</strong>sn.invbeoaonlnil;<br />

<strong>and</strong> <strong>the</strong> <strong>methyl</strong> <strong>esters</strong> <strong>of</strong> 4491-dhnethylaznino- <strong>and</strong><br />

4-N.N-dlcthylnmlnobcnzolc acid In a supersonic Jet Is<br />

due to dimer formatIon (10,11) We have reported a<br />

concentration dependence for <strong>the</strong> anomalous fluorescence<br />

<strong>of</strong> <strong>methyl</strong> 4-N,N41i<strong>methyl</strong>srnlnobenzo3te in soltition,<br />

which also points to dirner formation [121. Although<br />

<strong>the</strong>re was no evidence from <strong>the</strong> absorption spectra <strong>of</strong><br />

ground-state dimer formation, <strong>the</strong> fluorescence properties<br />

<strong>of</strong> this ester did not conform to those expected for<br />

excirncr formation <strong>and</strong> we concluded that at least a loose<br />

ground-state a.saociadon exists.<br />

We have also studied <strong>the</strong> absorption <strong>and</strong> fluorescues<br />

properties <strong>of</strong> <strong>the</strong> parent 4-N,N4i<strong>methyl</strong><strong>aminobenzoic</strong><br />

acid. Cowley or aL (3) have reported that this<br />

molecule also exhibhs a fluorescence (in acetonitrile),<br />

but this appears to be <strong>the</strong> only recent <strong>study</strong> on <strong>the</strong> parent<br />

acid. We <strong>the</strong>refore present our findings on this compound<br />

to date in this paper.<br />

107<br />

lwafl.iqta. C nez ruaaa r.,ss.1<br />

17


log<br />

EXPERIMENTAL<br />

4-N,N-Dimetbyl<strong>aminobenzoic</strong> acid was purchased<br />

from Lancaster Syn<strong>the</strong>sis Ltd. <strong>and</strong> was reclystattlzed<br />

twice from ethanol prior to use; m.p., 244-<br />

245*C (literature value, 242-243t 113)). Infrared <strong>and</strong><br />

proton <strong>and</strong> "C NMR spectra were also acceptable. The<br />

solvents used In this work were ei<strong>the</strong>r spectrophotometric,<br />

spectr<strong>of</strong>luorimetrie. or HPLC giade<strong>and</strong> were<br />

used as received with <strong>the</strong> exception <strong>of</strong> acetonitrile.<br />

This was distilled over phosphorus pentoxide <strong>and</strong> collected<br />

<strong>and</strong> stored under an inert atmosphere until it<br />

was used. The water was doubly distilled <strong>and</strong> <strong>the</strong>n<br />

passed through an ion-exchange resin prior to use. The<br />

p11 <strong>of</strong> <strong>the</strong> unbuffered water was 615.<br />

Absorption spectra were measured on a Ferkin-Elmet<br />

lambda 3 speetrophotometer or & Hewlett-Packard<br />

HPS451A diode-tiny spectrometer using matched quart<br />

cuvettes <strong>of</strong> varying path lengths. All working <strong>solution</strong>s<br />

were prepared fresh on <strong>the</strong> day <strong>of</strong> use from a stock <strong>solution</strong><br />

which was checked at regular intervals for pudty.<br />

The accuracy <strong>of</strong> <strong>the</strong> measurements made under <strong>the</strong>se<br />

conditions is estimated as ± 2 nm for wavelengths <strong>and</strong><br />

± 10% for extinction coefficients.<br />

fluorescence spectra were measured on Perkin-Elmet<br />

LSS or LS50 speetr<strong>of</strong>luorimeters in <strong>the</strong> fully corrected<br />

mode. Excitation <strong>and</strong> emission slit widths <strong>of</strong> S<br />

nm were used unless o<strong>the</strong>rwise indicated. For <strong>the</strong> LSS<br />

<strong>the</strong> spectra were transferred to a BBC microcomputer to<br />

undertake quantum yield calculations. Quantum yields<br />

were measured on optically dilute samples (absorbance<br />


Anomalous fluorncence Properties <strong>of</strong> 4.N.N-Di<strong>methyl</strong>aminobeozolc Add 109<br />

So4nm<br />

Table I. AAwqxion <strong>and</strong> Flcsccna Ytupatin <strong>of</strong> 40MADA b Vattes Sent, it Bas Tcnwcnwtc<br />

Absorption<br />

maxims (am)<br />

Eminedno<br />

coefficient<br />

(din' md-' or')<br />

Thjogcsctace<br />

msxüna (nm)<br />

fluaracence<br />

,xcitaIoa<br />

maxima (vim)<br />

Hmnc 302 a 340 300<br />

225 a<br />

EThanol 295(1liy - 350 vs<br />

282 20,300 420 280<br />

Alaaluile 306 22,500 350 275<br />

295(a) - 480 308<br />

Wamr<br />

(nbgffcnd) 288 a 365 270<br />

222 a 500 290<br />

Water<br />

AnIon 284 a 355 274<br />

225 a 490 306<br />

Ncun.t 314 • 370 290(th).300<br />

230 a 480.490 300<br />

Cdoc 272 a C -<br />

230<br />

• amonnd a paitlsily z1ilg In this tolvem.<br />

'sh, s)oidc.<br />

t4<strong>of</strong>luwtscencc.<br />

C<br />

a<br />

t<br />

in<br />

C<br />

S<br />

S<br />

S<br />

I'<br />

'a<br />

g<br />

Wautlenoth (vim)<br />

F!p t. fluorcscncc spews <strong>of</strong> 4DMABA (1.08 x 10 mol•dnr') In .cetooiuilc at different excitation wsvcicn8ths.<br />

19


no<br />

Revill <strong>and</strong> Brown<br />

U..<br />

251 255 311 351<br />

Wauelangth (rim)<br />

FIg. 2. Absorption spectrum <strong>of</strong> 4DMABA (53 x 10" mol-dr') in aettoniufle.<br />

that this b<strong>and</strong> corresponds to <strong>the</strong> normal V (5 1 (It) -,<br />

80 (FC) fluorescence) for 4DMABA <strong>and</strong> <strong>the</strong> longer<br />

wavelength b<strong>and</strong> to anomalous (a) fluorescence. In<br />

henna, <strong>the</strong> fluorescence spectrum <strong>of</strong> 4DMABA is independent<br />

<strong>of</strong> excitation wavelength (except for a change<br />

in overall intensity) <strong>and</strong> <strong>the</strong> excitation spectrum is independent<br />

<strong>of</strong> emission wavelength. This is not <strong>the</strong> case<br />

in all <strong>the</strong> o<strong>the</strong>r solvent systems used. The measured<br />

quantum yield <strong>of</strong> 0.18 in hexane Is in good agreement<br />

with <strong>the</strong> value <strong>of</strong> 0.21 in cyclohexane reported by Cow.<br />

leyetaL (31-<br />

In acetonitrile, excitation <strong>of</strong> a I x 10" mol'dm'<br />

<strong>solution</strong> at <strong>the</strong> absorption maximum <strong>of</strong> 306 rim produces<br />

a relatively weak fluorescence ape ctrum (rig. 1) but with<br />

<strong>the</strong> two b<strong>and</strong>s cicarly visible. If <strong>the</strong> excitation is moved<br />

to a shorter wavelength (e.g.. 270 rim), <strong>the</strong> intensity <strong>of</strong><br />

<strong>the</strong> b fluorescence increases approximately fourfold,<br />

while <strong>the</strong> a fluorescence almost disappears. The spectra<br />

at intermediate wavelengths tie between <strong>the</strong>se two cxtremes.<br />

It is not surprising <strong>the</strong>refore to find that <strong>the</strong> cx-<br />

citation spectra for <strong>the</strong> two fluorescence b<strong>and</strong>s are very<br />

different, with that for <strong>the</strong> a' fluorescence peaking at<br />

308 nm <strong>and</strong> that for <strong>the</strong> b' fluorescence at 275 rim. The<br />

latter may correspond to <strong>the</strong> asymmetry on <strong>the</strong> absorpdon<br />

spectrum <strong>of</strong> 4DMABA in acetonitrile (Fig. 2).<br />

Emission spectra <strong>of</strong> 4DMABA in water exhibit an<br />

identical trend to thou in acetonitrile but with overall<br />

smaller changes; <strong>the</strong> V fluorescence Intensity increases<br />

by only a factor <strong>of</strong> two as <strong>the</strong> excitation wavelength is<br />

changed from 300 to 210 nfl <strong>and</strong> <strong>the</strong> a intensity de.<br />

ceases by a similar amount (Fig. 3). The excitation spectra<br />

peak at approximately 290 urn (Cf. absorption maximum<br />

at 288 urn) <strong>and</strong> 270 not for <strong>the</strong> a' <strong>and</strong> V fluorescence,<br />

respectively. In water <strong>the</strong>re is no obvious absorption<br />

spectral feature at 270 nm to correspond to <strong>the</strong> latter<br />

peak. The fluorescence spectra <strong>of</strong> 4DMABA in water<br />

are also pl4 dependent (Pig. 4). At alkaline pH values<br />

both anomalous <strong>and</strong> normal fluorescence are observed,<br />

with <strong>the</strong> tatter peaking at 355 nut This must <strong>the</strong>refore<br />

correspond to normal fluorescence from <strong>the</strong> 4DMABA<br />

20


Anomalous Fluorescence Properties <strong>of</strong> 4-N,N-Dimctbyl<strong>aminobenzoic</strong> Acid 111<br />

21<br />

- ZGDM<br />

• fl*mi<br />

a Sitvn<br />

P.<br />

ls<br />

2'<br />

a<br />

S<br />

N<br />

I1<br />

I<br />

Si, l'S Si, Sit<br />

Waoilanth lam)<br />

tlg.3. Muoresanee spectra <strong>of</strong> 4DMABA In watu at diiktcnt cxcitation wavelengths.<br />

carboxylate anion, as <strong>the</strong> ground-state pK. is 4.79 <strong>and</strong><br />

<strong>the</strong>re Is insufficient acidity present to convert <strong>the</strong> excited<br />

anion to <strong>the</strong> excited neutral species despite a calculated<br />

1201 pK,' <strong>of</strong> 11.95 [19]. As <strong>the</strong> pH is reduced, <strong>the</strong> b<br />

fluorescence b<strong>and</strong> shifts to 375 am <strong>and</strong> <strong>the</strong> a fluorescence<br />

decreases in intensity until it is debatable whe<strong>the</strong>r<br />

it is still present. In strong acId <strong>the</strong> fluorescence disappean<br />

completely—<strong>the</strong> 4DMAEA cation Is nonfluorescent.<br />

In ethanol, <strong>the</strong> anomalous fluorescence is shifted<br />

less to <strong>the</strong> red than in acetonitrile <strong>and</strong> water <strong>and</strong> appears<br />

as a shoulder on <strong>the</strong> long-wavelength side <strong>of</strong> <strong>the</strong> emission<br />

b<strong>and</strong> (Fig. 5). Once again, <strong>the</strong> spectrum varies with<br />

excitation wavelength but <strong>the</strong> observed changes are <strong>the</strong><br />

smallest for <strong>the</strong> three solvents where <strong>the</strong> anomalous fluorescence<br />

is observed. These observations dearly suggest<br />

that <strong>the</strong>re is more than one ground-state species<br />

present in <strong>the</strong> polar solvents used in <strong>the</strong>se studies. Our<br />

measurements to date have not shown any change in<br />

absorption spectra with concentration but this may be<br />

due to <strong>the</strong> fact that any changes are taking place at much<br />

lower concentrations than we have studied to date - The<br />

evidence to 5upport this Is that <strong>the</strong> fluorescence spectra<br />

for 4DMABA In acetonitrile are concentration dependent.<br />

This has also been observed for M4DMAB [12).<br />

As <strong>the</strong> concentration <strong>of</strong> 4DMASA is varied between<br />

10' <strong>and</strong> 10' mol-dnr 3 (Fig. 6). <strong>the</strong> amount <strong>of</strong> a<br />

fluorescence decreases until at 10' mol-dm 3 <strong>the</strong>re is<br />

vexy little remaining, although <strong>the</strong> overall weakness <strong>of</strong><br />

<strong>the</strong> fluorescence from 4DMAEA makes measurements<br />

difficult at <strong>the</strong>se low concentrations. It is possible that<br />

absorption spectral changes may be visible at <strong>the</strong>se concentrations<br />

but <strong>the</strong> 4DMABA extinction coefficients preelude<br />

acetirate absorption measurements at concentratiotis<br />

below approximately 10' mol-dm'.<br />

The data presented so far suggest quite clearly that<br />

<strong>the</strong> anomalous fluorescence from 4DMABA arises from<br />

a ground-state species. The concentration dependence <strong>of</strong><br />

<strong>the</strong> fluorescence in acetonitrile points to this species being<br />

a dimer or higher polymer, although we have no ab-<br />

21


112 Haiti <strong>and</strong> Brown<br />

a<br />

C<br />

ii<br />

I<br />

Wavelength tarn)<br />

Pig. & fluorescence spears <strong>of</strong> 4DMAKA in water at different pH values.<br />

sorption evidence for this. The variation <strong>of</strong> <strong>the</strong> fluorescence<br />

spectra with concentration certainly does not agree<br />

with a mechanism Involving <strong>the</strong> formation <strong>of</strong> a TICT<br />

state 14], <strong>and</strong> our attempts to analyze <strong>the</strong> data using<br />

st<strong>and</strong>ard exchner kinetics 1211 have not been successful—<strong>the</strong><br />

expected dependence <strong>of</strong> fluorescence intensity<br />

<strong>and</strong> kinetics on concentration is not observed. Unfortunately,<br />

<strong>the</strong> absence <strong>of</strong> any sgnifkant changes in <strong>the</strong><br />

absorption spectra with concentration means that we are<br />

unable to verify <strong>the</strong> presence <strong>of</strong> ground state equilibrium.<br />

We have, however, studicd <strong>the</strong> fluorescence decay<br />

pr<strong>of</strong>iles for <strong>the</strong> two fluorescence b<strong>and</strong>s in some <strong>of</strong> <strong>the</strong><br />

solvents used here. In hexane <strong>the</strong> decay is single exponential<br />

with a lifetime <strong>of</strong> 1.22 ns, very similar to <strong>the</strong><br />

value <strong>of</strong> 1.30 ns found for M4DMAB in hexanc [12].<br />

In ethanol, <strong>the</strong> proximity <strong>of</strong> <strong>the</strong> C <strong>and</strong> b fluorescence<br />

b<strong>and</strong>s leads to complex decay pr<strong>of</strong>iles which require threeexponential<br />

components to fit <strong>the</strong> data adequately. We<br />

are making fur<strong>the</strong>r measurements at a variety <strong>of</strong> cxci-<br />

tation <strong>and</strong> emission wavelengths to try to resolve this<br />

complexity, but in view <strong>of</strong> <strong>the</strong> various solute/solvent<br />

interactions which are poasible in hydrogen-bonding solvents<br />

(In addition to any o<strong>the</strong>r processes taking place),<br />

decay pr<strong>of</strong>iles which are not single exponential are not<br />

surprising. In water <strong>and</strong> acetonitrile, <strong>the</strong> as fluorescence<br />

decays by a single exponential, whereas <strong>the</strong> b' fluorescence<br />

requires two exponentials for an adeuqate fit (Table<br />

U). In aectonithle, where we have accumulated a<br />

large amount <strong>of</strong> decay data, <strong>the</strong> lifetime <strong>of</strong> <strong>the</strong> C b<strong>and</strong><br />

is independent <strong>of</strong> conantration <strong>and</strong> excitation wavelength.<br />

A lifetime in a range <strong>of</strong> 2.1-13 ns is always<br />

obtained <strong>and</strong> <strong>the</strong>re is no evidence <strong>of</strong> a rise time indicative<br />

<strong>of</strong> an excited-state precursor. As Ilasselbacher eta?. [22]<br />

have shown, this does not absolutely rule out this mccl,-<br />

anism but it does indicate that at least one alternative<br />

mechanism <strong>of</strong> populating <strong>the</strong> a' state is operative. We<br />

interpret <strong>the</strong>se observations as confirming our earlier<br />

conclusion about <strong>the</strong> presence <strong>of</strong> a ground-state species<br />

producing <strong>the</strong> a' fluorescence. They also imply that this<br />

22


Anomalous fluorescence Properties <strong>of</strong> 4-N,N-Dl<strong>methyl</strong>amlnobenzolc Add 113<br />

I<br />

$<br />

S<br />

S<br />

.1<br />

5<br />

I<br />

Waslelenath Inmi<br />

Figs. Flisoruceac. aped,. <strong>of</strong> 4DMABA (1.14 x 10-' mot-dnr') In ethanol.' dirftrent excitation wavelengths.<br />

is <strong>the</strong> only source <strong>of</strong> <strong>the</strong> C species—<strong>the</strong> b' excited state<br />

does not appear to be a precinct <strong>of</strong> a', although <strong>the</strong><br />

reverse may possibly be true given <strong>the</strong> more complex<br />

decay pr<strong>of</strong>iles observed for <strong>the</strong> b* fluorescence b<strong>and</strong>.<br />

Vanna <strong>and</strong> co-workers (23,24J have proposed that<br />

<strong>the</strong> anomalous fluorescence <strong>of</strong> ethyl 4-N,N-di<strong>methyl</strong>amlnobcnzoate<br />

123] <strong>and</strong> 4-N,N4<strong>methyl</strong>aminobenzontrile<br />

[24J in acetonitrile is due to exciplex formation. We have<br />

<strong>the</strong>refore added small amounts <strong>of</strong> acetonitrile to a <strong>solution</strong><br />

<strong>of</strong> 4DMABA in hexane up to <strong>the</strong> miscibility limis<br />

(approximately 2.2% by volume) <strong>and</strong> have found that a<br />

new fluorescence b<strong>and</strong> at approximately 440 nm results<br />

(Fig. 7). The observed variations in <strong>the</strong> emission spectra<br />

do not appeer to fit a st<strong>and</strong>ard cxciplcx mechanism. The<br />

b' fluorescence In hexane Is quenched by low concentrations<br />

<strong>of</strong> scetonitrile. but at slightly higher concentralions<br />

it increases in intensity <strong>and</strong> red-shifts by some 10-<br />

20 nm. At <strong>the</strong> same time <strong>the</strong>re is a steady increase in a'<br />

fluorescence as <strong>the</strong> .cetonitrile concentration is in.<br />

creased. It would appear that <strong>the</strong>re are some specific<br />

salvation effects taking place in addition to <strong>the</strong> mccliinism<br />

producing <strong>the</strong> C fluorescence. The red shift in <strong>the</strong><br />

b fluorescence probably results from <strong>the</strong> formation <strong>of</strong><br />

a specific 4DMABA/.cetonitrile solute/solvent interac.<br />

don. The new b<strong>and</strong> at 440 urn may be an exciplex or<br />

may be <strong>the</strong> some species as Is seen in pure acetonitrile<br />

but with a blue-shifted fluorescence due to <strong>the</strong> nonpolar<br />

solvent environment in which it is situated. Whatever<br />

<strong>the</strong> nature <strong>of</strong> <strong>the</strong> species producing <strong>the</strong> 440-nm fluorescence,<br />

it is clear that acetonitrile plays a key role in its<br />

formstinn <strong>and</strong> we are undertaking fur<strong>the</strong>r work to attempt<br />

to clarify <strong>the</strong> nature <strong>of</strong> this species.<br />

ACKNOWLEDGMtNrS<br />

We thank <strong>the</strong> Science <strong>and</strong> Engineering Research<br />

Council for a studentship (SAT?.) <strong>and</strong> for access to <strong>the</strong><br />

Daresbwy Laboratory <strong>and</strong> University <strong>of</strong> Cenual Lanashire,<br />

<strong>the</strong> Ciba-Oci' Tnsst, NATO, <strong>and</strong> <strong>the</strong> British Coun-<br />

23


114 Revill <strong>and</strong> Brown<br />

Tabis 11. flootnaia Decay Peopcnie. <strong>of</strong> 4DMABA in Vuioua Solvezi,<br />

k.. Cornntjon Ti TI<br />

Solveat (na.) (net) (moltr') a, (ill)<br />

306 340 1.01 1.000 1.28 - -<br />

Ethuol 280 350 1.01 0.859 2.86 0.141 17.61<br />

Went 290 365 1.05 0.994 0.77 0.X6 5.07<br />

500 1.27 0.973 0.17 0.025 2.39<br />

Aato&vile 280 350 LOx 10a 1.04 I.000 2.13 - -<br />

233 500 LOX 10-0 1.74 LOOD 116 - -<br />

303 350 1.OxlO-' 3.61 LOOD 1.66 - -<br />

303 480 1.OxIO-' 3.48 1103 2.30 - -<br />

300 350 1.OxlD-' 1.26 0.842 1.29 0.158 2.74<br />

303 480 LOX10- 6 3.01 LODO 223 - -<br />

303 350 1.0x10' 1.23 0.687 0.20 0.313 1.85<br />

303 480 1.0x 10- 1.14 I. 127 - -<br />

310 350 1.OxlOa 1.17 1.000 2.1* - -<br />

310 503 1.OxlO" 1.36 IMD 2.15 - -<br />

340 350 I.OxlO-' 1.57 2.000 2.21 - -<br />

340 500 1.OxlO-' 1.31 1.000 2.17 - -<br />

1.3<br />

- lI.IjilI-1 ri<br />

SSSii,-7 H<br />

- 1.1I0I-1H<br />

4 I.1IeII-4M<br />

- l.lIaIU-5P4<br />

•1<br />

t<br />

5.2<br />

S<br />

a<br />

a<br />

I.'<br />

6<br />

Is<br />

525 374 42* 452 535<br />

Waustn9th (em)<br />

39,<br />

F1g.6. Flvon=m lpCtn ci 4DMADA in zcnoolnik at diffnnt COAMOUlliM.<br />

24


Anomalous Fluortscencc Properties <strong>of</strong> 4.N,N-Dimethyiamlnobenzoic Acid 115<br />

S<br />

a<br />

if<br />

t<br />

S<br />

a<br />

C<br />

S<br />

S<br />

E<br />

S<br />

a<br />

Wautlanath Inn)<br />

Elg.1. Fluonscence .pewa <strong>of</strong> 4DMABA in henna with different amounts <strong>of</strong> added acetooiufle; pure haanc (%. Wv).<br />

cii for financial suppoit. We are grateful to Pr<strong>of</strong>essor<br />

Wolfgang Pettig <strong>of</strong> <strong>the</strong> TechniCal University, <strong>of</strong> Berlin for<br />

his comments <strong>and</strong> fn4tM discussion <strong>and</strong> to <strong>the</strong> Photophysic<br />

Research Group at <strong>the</strong> University <strong>of</strong> Strathclyde<br />

for allowing us access to <strong>the</strong>ir fluorescence lifetime spectrometers.<br />

REVERENCES<br />

I. 2. R. Crabowüi. K. Roakewia. D. J. Ssmiaruk. 0.1. Cow.<br />

Icy, <strong>and</strong> W. Banmann (1979) Now. J. CAL.,. 3, 443-lU.<br />

2. 6. Wermuth. W. Renig, <strong>and</strong> K Lippert (1931) Ba'. BUAIE$9EI.<br />

Pity,. Chant 85. 64-70.<br />

3. 0.1. Co-ky. P.i. Italy. iSA. H. Peoplca(1978)J. fob.<br />

<strong>the</strong>m. P, 240-242.<br />

4. W. Rflhig (1I86) AMw. Chn. tar. Ed. Lag. 25, 971-916.<br />

S. W. ReaaI1 (1991) EPA New4eu., No. 41. 3-19.<br />

6. lii. Van dat Anwarser, Z. 2. GnbowakL <strong>and</strong> W. Reni1 (1991)<br />

J. flo'& Cheat 95. 2083-2092.<br />

7. J. HeIdi, D. Goimin. <strong>and</strong> M. (nba (1939) Chant P,s 136.<br />

321-344.<br />

& K. I. Ylsscr, <strong>and</strong> C. A. 0. 0. Yarmi (1980) 1. Chem, Soc.<br />

Fonda3, Tn,,,. 2 76, 433-476.<br />

9. P. C. M. Wcisenborn, A. H. Huizer, <strong>and</strong> C. A. G. 0. Var,,,.<br />

(1989) China. Thys. 133. 437-432.<br />

10. R. Howell, H. Petek, D. Phillips, <strong>and</strong> K. YoAiban (1991)Chem.<br />

FSn. L.M. 183, 249-253.<br />

11. R. Howell. A. C. jones, A. 0. Taylor. <strong>and</strong> D. Phillips (1989)<br />

CAn Ps In. 10.2*2-290.<br />

12. J. A. T. RaIl, <strong>and</strong> 2.0. Brown (1992) Cheat. n,& Leer. lB.<br />

433-43*.<br />

13. Lancaster, Synlhesis Lid. 1991. Orgsnie Resesrch Chemicals Catalogue.<br />

Lancaster Syn<strong>the</strong>sis.<br />

14. R. A. Vetapoldi, <strong>and</strong> K. D. Mittens (1980) PJBS Special Pubti.<br />

catIon 260-6-4.<br />

15. D. P. Eaton (1981)1. han tWL CAr,,,. 60, 1107"I1I4,<br />

16. D. V. O'Connor. <strong>and</strong> D. PWllipa (1984) rn Carnioted Sbag!e<br />

Phcia.a Cowuint Andemie Press. New York. -<br />

II. D. I. S. Birch. A. Dutch, 2. E. I.nla<strong>of</strong>. 8. NS4OI*IEI. <strong>and</strong> t. Sown<br />

(19*7)1. P*oroch.aa. II, 239-254.<br />

IS. R. Sparrow. HG. Dro.na, E. H. Evans, <strong>and</strong> D. flaw (lQS)J.<br />

Claim. Soc. Fsnhry twit 2 82, 2249-2262.<br />

19. J. A. T. RcviIl, <strong>and</strong> R. C. Brown, Unpublished work.<br />

20. Th. Farster (1950)1 Elacemehant 54, 531-539.<br />

21. J. B. SitU (1970) Fhmçtysia <strong>of</strong> Anmalic Molecules. Wiley<br />

Inrencienca, London.<br />

22. C. A. Hamlbacher, E. Waxy,..,,, L T. G.haü. P. B. Cooaino. 3.<br />

B. A. lWss, <strong>and</strong> W. R. Laws (1991)/. PhF. Chem. 93, 2993-<br />

3005.<br />

23. R. J. Visser, P. C. SI. Wthcnborn, <strong>and</strong> C. A. G. 0. Vanna<br />

(1985) Cheat Pity,. 1,14 113, 330-336.<br />

24. P. C. M. Weixobona. A. H. Huize,, <strong>and</strong> C. A. G. 0. Vanna<br />

(1988) Cheat PAys. 116.425-433.<br />

25

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!