<<

THE FADING OF BASIC IN

POLYMER SUBSTRATES

A THESIS SUBMITTED FOR THE DEGREE

OF

DOCTOR OF PHILOSOPHY

AT THE UNIVERSITY OF NEW SOUTH WALES

BY SUSAN ELIZABETH MATTHEWS

·ft.

SCHOOL OF TEXTILE TECHNOLOGY

· ···JONE .. 1980 i

I hereby certify that the work described in this thesis was performed by me in the School of Textile Technology, The University of New South Wales, and has not been submitted previously for any other degree or award. ii

ACKNOWLEDGEMENTS

I wish to express my sincere thanks to my supervisors, Associate Professor C.H.

Nicholls and Dr. M.T. Pailthorpe, for their guidance and support throughout the course of this work.

I thank my colleagues and friends for their continued encouragement and interest.

I also thank my friend Anne for the typing and presentation of this thesis.

Finally, I particularly thank my fiance,

Paul, for his invaluable support during this time. iii

ABSTRACT

The reasons for the higher lightfastness of basic dyes on acrylic substrates compared to other substrates were investigated. Acrylic, nylon and poly(vinyl alcohol) substrates in film and fibre form were dyed and irradiated in order to study this phenomenon directly in the solid state. It was found that there is a difference in the initial quantum yields of fading of at least two orders of magnitude between the acrylic and the non-acrylic fibres.

Temperature studies indicated that the high rigidity of acrylic polymers, by virtue of their inherently compact structures and their high glass transition temperatures, contributes significantly to the high lightfastness of basic dyes present in them. This is postulated to be due to the relative inhibition of movement in the polymer, which affects both diffusional and conformational processes that are involved in the fading reactions.

Oxygen and moisture were found to have little effect on basic dye fading.

Very little difference could be found between the rate of fading of dyes bound to sulphonate and carboxylate dye sites of the same pKa. However, the presence of acid caused a significant reduction in the rate of dye fading which appears to indicate that the pK of the dye site affects a basic dye fading. This pH effect on dye fading rates suggests that electron transfer from substrate to dye could contribute iv

to the dye degradation. Thus the reduced fading in acrylics might be attributed in part to their being poor electron

donors compared with other substrates studied.

The presence of -1% cyano groups in an inert

substrate was found to significantly reduce basic dye fading, with the effect increasing with an increased cyano concen­

tration of -8%. Since cyano groups comprise approximately

50% by weight of an acrylic polymer this effect would appear

to be the single most important factor contributing to the high lightfastness of basic dyes present in them. It was postulated that the formation of an excited complex between the excited dye and one or more cyano groups would either inhibit the usual fading reactions or facilitate deactivation of the excited dye molecule. V

CONTENTS

DECLARATION i

ACKNOWLEDGEMENTS ii

ABSTRACT iii

CONTENTS V

CHAPTER 1 INTRODUCTION

1.1 Preamble 1

1.2 Photochemical Principles 2

1.2.1 Excited States 4

1.2.2 Photophysical Decay Pathways 5

1.2.3 Photochemical Decay Pathways 6

1.3 The of Dyes 9

1.3.1 Typical Photochemical Reactions of Dyes 10

1.3.1.1 Azo Dyes 11

1.3.1.2 Anthraquinone Dyes 15

1.3.1.3 Triphenylmethane Dyes 18

1. 4 Factors Influencing Dye Fading 22

1.4.1 Intensity and Spectral Distribution

of Radiation 22

1. 4 .2 The Physical State of the Dye 24

1.4.3 The Nature of the Substrate 26

1.4.4 The Dye-Fibre Bond 29

1.4.5 Oxygen 30

1.4.6 Moisture 31

1.4.7 Temperature 33

1.5 Scope of Present Work 35 vi

CHAPTER 2 QUANTUM YIELDS OF BASIC DYE FADING

2.1 Introduction 41

2.2 Fabrics 42

2.2.1 Acrilan 16 42

2.2.2 Nylon 6.6 43

2.2.3 Basic-Dyeable Nylon 43

2.2.4 Chemical Compositions 43

2.3 Dyes 44

2.4 Methods 45

2.4.1 Acrilan 16 46

2.4.2 Nylon 6.6 47

2.4.3 Basic-Dyeable Nylon 47

2.5 Determination of Dye Concentration 48

2.6 Irradiation Technique 51

2.6.1 Calibration of Incident Intensity 52

2.7 Determination of Initial Quantum Yields of

Basic Dye Fading 57

2.7.1 Measurement of Fading Rates 57

2.7.2 Determination of Quanta Absorbed by Dye 59

2.7.3 Calculation of Initial Quantum Yields of

Basic Dye Fading 60

CHAPTER 3 THE EFFECT OF THE PHYSICAL STATE OF THE

SUBSTRATE ON BASIC DYE FADING

3.1 Introduction 64

3.2 Polymer Film Preparation 65

3.2.1 Nylon 6 66

3.2.2 Poly(vinyl alcohol) 66

3.2.3 Acrilan 16 67 vii

3.3 Measurement of Glass Transition Temperatures 69

3.3.l Experimental Technique 70

3.3.2 Results 72

3.4 Irradiation Technique 74

3.4.1 Method of Mounting and Heating the

Film Sample 74

3.4.2 Method of Irradiation 76

3.5 Determination of the Initial Quantum Yields

of Basic Dye Fading 77

3.5.1 Measurement of Fading Rates 77

3.5.2 Measurement of Quanta Absorbed 81

3.5.3 Calculation of the Initial Quantum

Yields of Dye Fading 83

3.6 Experimental Results 83

3.6.1 The Effect of Temperature 84

3.6.2 The Effect of Oxygen and Moisture 90

CHAPTER 4 THE INFLUENCE OF THE CHEMICAL NATURE OF

THE SUBSTRATE ON THE FADING OF BASIC DYES

4.1 Introduction 96

4.2 Irradiation Technique 97

4.3 The Identification of Reaction Intermediates

in Basic Dye Fading 98

4.3.1 The Use of Scavenging and Quenching

Agents 98

4.3.1.1 Quenching 99

4.3.1.2 Electron and Scavenging 99

4.3.1.3 Experimental Technique 100

4.3.1.4 Results 102

4.3.2 Luminescence Studies 104 viii

4.4 The Influence of Dye Sites on Basic Dye Fading 104

4.4.1 Experimental Technique 105

4.4.2 Results 106

4.5 The Influence of the Cyano Group on Basic Dye Fading 109 4.5.1 Experimental Technique 109

4.5.2 Results 111

CHAPTER 5 CONCLUSION 114

REFERENCES 120 1

CHAPTER l

INTRODUCTION

1.1 Preamble

The fading of dyed materials by light, or photo­ fading, refers to the loss of colour or change of hue of the dye caused by its exposure to light. Being of obvious commercial interest in relation to the durability of dyed fabrics, such phenomena have long been the subject of investigation. References 1 to 8 give reviews of this research.

There are, however, still vast areas of the subject about which relatively little is known. This is due to the extremely complex nature of the photochemistry of such dye­ fibre systems and the number of different variables which can influence it. Further, as the concentration of dyes on fabrics is usually very low (no more than -3%) and because fading reactions have low quantum yields and occur only at the surface of the fabric only small yields of decomposition products are obtained. Consequently isolation and identification of these products is so tedious that other less direct methods have to be used to establish the mode of reaction. In addition, the reaction sequence of fading may be so complex that identification of the final products may not shed much light on the preceding reactions.

Model systems using organic solutions instead of solid substrates and simple instead of the more complex dyes have been used to simplify the systems and much information has 2

been gained in this way. However, great care must be taken

in this respect as it is not always possible to apply such

results directly to solid substrates. The fading behaviour may not necessarily be the same in the two environments due

to differing physical and/or chemical effects.

The aim of the present work was to elucidate some

aspects of the fading behaviour of one class of dyes in which large differences in lightfastness properties are

found between different fibre substrates. Solid substrates in the form of transparent polymer films and fabrics were used in order to study the problem directly and avoid the interpretation difficulties of solution models.

1.2 Photochemical Principles

The first step in a photochemical process is the

absorption of light by a molecule. This causes an electron to be promoted from its lowest energy level or ground state

to a higher energy level or "excited" state. This can be

represented by D + hv ~D.* A molecule in this excited

state is much more reactive than in its ground state having more energy, a different electron distribution and different molecular geometry. Consequently photochemical reactions of

various types, often unusual in ordinary "thermal" reactions,

can occur readily.

The lifetime of this reactive excited state is

critical in determining whether or not a photochemical

reaction, leading to dye fading, will occur. The longer

the molecule remains in its excited state the greater the 3 probability of such a reaction occurring. There are numerous alternative deactivation pathways by which the molecule can return to its ground state without any chemical change taking place and therefore the relative rates of these "photophysical" processes compared with the "photo­ chemical" processes will determine the rate at which the dye will fade.

The initial processes involving the electronically excited molecule are known as the primary photochemical and photophysical processes. These can then be followed by secondary or "dark" reactions of the various chemical species produced, the nature of such chemical reactions depending on the system in question.

The efficiency of the absorbed light in promoting any particular process is expressed by the quantum yield

(~) of that process and is defined as:

~=number of undergoing a particular process number of quanta absorbed by the system

The sum of the quantum yields of all the different primary photochemical and photophysical processes must be equal to unity, but the quantum yield of any particular process may vary from ~ << 1 to ~ >> 1, for chain reactions. The fading reactions of dyes generally have ~ << 1 indicating that most of the absorbed energy is dissipated in other ways. 4

1.2.1 Excited States

A molecule in an electronically excited state con­

tains two unpaired electrons, one of which has been excited by the absorption of light and promoted to a higher energy

level orbital leaving the second electron of the pair in the ground state orbital. When the two electrons of an excited state have the same (or parallel) spin the excited state is known as a triplet excited state, and when they have opposed

(or antiparallel) spin the excited state is called a singlet excited state (abbreviated as T and S respectively).

Excited states of organic molecules produced directly by the absorption of light are nearly always singlets. This is because virtually all their ground states are singlets

(denoted S 0 ) and conservation of spin during absorption is strongly favoured. Excited triplet states are therefore only formed via an intermediate excited singlet state. A

triplet state has a lower energy than the corresponding singlet state since according to the Pauli Exclusion

Principle its two electrons, having parallel spin, cannot occupy the same space and are therefore more separated and exert a smaller repulsive force on each other.

Depending on the energy of the quantum of light

absorbed by a molecule, an electron may be excited to the

lowest excited singlet state S 1 or a higher singlet state

S2, S 3 etc. However, if the electron is excited to a higher

state it very rapidly loses energy by radiationless decay

processes until it reaches the lowest vibrational level of

the lowest excited singlet state S 1 • From here it can be 5

converted to a triplet state. This process is known as

intersystem crossing and, although it is formally forbidden,

does occur due to a phenomenon known as spin-orbit coupling.

Again, if the state formed is a higher triplet state (T2,

T 3 etc.), rapid decay to T 1 will occur.

The extremely short lifetimes of the higher excited

states resulting from this mean that very few chemical

reactions can compete with the rapid deactivation processes.

Thus very few photochemical reactions originate from these higher excited states. In contrast, the lowest excited singlet and triplet states have relatively long lifetimes

(10- 9 to 10- 7s for S 1 , and 10- 4 to 10 2s for Ti). As a consequence it is these states that are the starting points for almost all photochemical reactions.

1.2 .2 Photophysical Decay Pathways

Photophysical deactivation processes return an electronically excited molecule to its ground state without

any chemical change occurring, and photochemical reactions have to compete with these processes. The photophysical

decay pathways which are available to an excited molecule

can be summarised as follows:

(1) Vibrational Relaxation. This involves transitions within the one excited state from higher to lower vibrational energy levels with an accompanying loss of heat.

(2) Emission of Radiation. This is known as fluorescence when the transition is from an excited singlet state to the

ground state, and phosphorescence when from an excited tri­

plet state to the ground state. As triplet to singlet C a

d

So

FIGURE 1.1 Jablonski diagram illustrating energy decay

pathways for a molecule raised to its second

excited singlet state.

a absorption

b vibrational decay

C internal conversion

d fluorescence

e intersystem crossing

f phosphorescence 6

transitions are spin-forbidden requiring an inversion of spin, phosphorescence has a longer lifetime than fluorescence.

(3) Radiationless conversion. This is accompanied by a loss of heat and includes two types of processes:

(i) Internal Conversion which involves transitions

between like states i.e., singlet-singlet or

triplet-triplet.

(ii) Intersystem Crossing which involves transitions

between unlike states. The transition from singlet

to triplet occurs isoenergetically to an upper

vibrational level of the triplet state.

(4) Energy transfer to another molecule. This process is known as quenching. The second molecule may gain the energy as vibrational, rotational and translational energy or it may be raised to an excited state just as if it had absorbed a .

A Jablonski diagram (Figure 1.1) illustrates these various photophysical deactivation pathways. Straight lines represent emission or absorption of radiation and wavy lines are used to indicate radiationless processes.

1.2.3 Photochemical Decay Pathways

If, instead of decaying to its ground state by photophysical processes, the excited dye molecule takes part in a chemical reaction, the following types of reactions are typical of those possible: 7

A 0 + B 0 radical formation

C + D intramolecular decomposition

BA rearrangement

photoisomerisation

R 0 H-atom abstraction AB + hv -+ photodimerisation

+ e- photoionisation

B intramolecular electron transfer

intermolecular electron } transfer

The excess electronic energy of the excited state alone is often the only reason for the particular course of a photochemical reaction. At other times the change in the electron distribution and molecular geometry of the excited state also contribute.

In any particular photochemical system only a selection of these reactions is involved and minor changes in the nature of the system may mean substantial changes in

the relative rates of the various reactions. The direct involvement of outside species in some of these reactions

indicates in particular that the nature of the medium in which a molecule exists will have an influence on the type

of reactions that will occur. 8

The types of excited states of most importance in

dye fading reactions are the (n, TT)* [i.e., produced by excitation of an electron from an nor non-bonding orbital to a TT * or antibonding TT orbital, denoted an (n-TT)* transi- . * tion] , the (TT, TT ) and charge-transfer states. (The latter involves the transfer of an electron from a donor to an acceptor group.) a and a * orbitals are rarely involved because much larger energies are required to excite electrons from or to these orbitals respectively. Each of these excited states can exist in both the singlet and the triplet form.

Which of these transitions are observed in any particular molecule will depend on the basic chemical structure of the molecule. The energies of the resulting excited states will also largely depend on structure but may in addition be influenced by the nature of the medium in which it exists. For example in acid or polar media the interaction between then electron and the solvent lowers the energy of then electron and thus an (n-TT)* transition will require more energy than in non-polar and non-acid media. The polarity of solvents also affects

(TT-TT)* transitions and transitions involving charge-transfer states. In the first case the election charge distribution is more extended and more polarisable in the (TT,TT)* state

than the ground state and so polar solvents stabilise the excited state more than the ground state:

excited state ground state =c:-- non- -- I polar polar solvent solvent 9

Charge-transfer states are affected even more because they have very large dipole moments.

The basic chemical structure of a molecule will usually play a major role in determining the energy levels of the excited states, but the nature of the substituents present, especially those with electron-withdrawing or releasing characteristics, may also influence the energies, sometimes even taking over as the major determinant. Sub­ stituents can alter the relative positions of the excited states, possibly changing the nature of the lowest excited state. The nature of this lowest excited state will largely determine the photochemical reactivity of the molecule.

1.3 The Photochemistry of Dyes

The photochemical degradation of dyes usually occurs either by photooxidation or photoreduction reactions. Intra­ molecular photoreaction i.e., photodecomposition simply as a result of the instability of the excited dye molecule is

rare particularly when the dyestuff is adsorbed on rigid polymer substrates. 6

Photooxidation has usually been found to require the presence of moisture as well as oxygen. Various entities have been postulated as the oxidising agents involved,

including HO", H0 2 ", 9 H2 0 2 5 and oxygen itself which has been

shown to react with the triplet state of the dye. 10

Photoreduction requires the presence of a suitable

hydrogen or electron donating substance. The two mechanisms 10

possible can be represented as follows:

(i) hydrogen abstraction: D* + RH + DH + R

(ii) electron abstraction: o* +RH+ D 0 + RH+

Some dyes also undergo reversible fading reactions, with the dye returning to its original form when removed from the light. This phenomenon, known as phototropy, can be the result of a number of different reactions including tautomerism, cis-trans isomerisation, ring opening, free radical formation and the formation of aggregates. While not as serious a problem as permanent fading, phototropy is clearly undesirable, and dyes which are phototropic have mostly been eliminated from commercial ranges.

The fact that a dye exhibits phototropy may lead to either an increase or a decrease in its overall lightfast­ ness. It may decrease lightfastness if the reversible process is followed by an irreversible one:

* + D +- X + Y

or it may increase lightfastness if the transition from the excited dye state to the phototropic species is so rapid that there is little time for any other photochemical

reaction to occur from the excited state.

1.3.1 Typical Photochemical Reactions of Dyes

The types of photochemical reactions that a dye may

undergo will depend mainly on its chemical structure, as

this will determine the nature of the excited state from 11 which a reaction may proceed, the position of this state relative to other states, its reactivity and lifetime, and the probability of its formation and deactivation. With respect to their basic chemical structure the majority of dyes belong to three classes - azo, triphenylmethane and anthraquinone. A smaller number of dyes belong to a variety of other classes including methine, oxazine and xanthene.

Each of these chemical dye classes have their own characteristic photochemical reactions. Then, in addition to its basic structure, the nature of the substituents present on a particular dye will also influence its photo­ chemistry, sometimes even outweighing the influences of the basic . Evans and Stapleton have produced a comprehensive review 8 of these substituent effects on each chemical and dyeing class of dye where it can be seen that it is difficult to reach any general conclusions about substituent effects on dye fading.

1.3.1.1 Azo Dyes

Azo dyes {general formula ~ 1 -N = N-~2) form the largest group of all the synthetic dyestuffs, making up approximately half of the structures disclosed in the Colour

Index. A number of the different photochemical reactions which azo dyes undergo have been reviewed by Griffiths. 11

By measuring the photocurrents produced when a platinum electrode coated with an azo dye was irradiated,

Hillson and Rideal 12 found that these dyes could be either oxidised or reduced. They postulated that similar reactions 12

occurred in dyed fibres which later researchers have

verified.

Photoreduction has been found to occur either by hydrogen transfer from the substrate to an excited dye molecule or from a photochemically excited substrate to

the dye in its ground state.

The former pathway was proposed by Blaisdell 13 for

the fading of azobenzene in various solvents on the basis of the spectroscopic identification of hydrazobenzene and aniline as fading products. Hashimoto and coworkers 14 later confirmed these results by isolating and chemically identifying the same two reaction products. In addition,

under acid conditions they found that the hydrazobenzene was converted to benzidine. A more complex azo compound,

a red dye, has also been found to produce an unstable hydrazo compound on irradiation. 15 Further reaction of the hydrazobenzenes by either thermal or photochemical dispro­ portionation can occur and results in the irreversible

cleavage of the azo linkage:

2 q,NH-NH

The second pathway was demonstrated by Van Beek

et al., 16 who found that azo dyes in anaerobic solutions

can be reduced to amines by hydrogen donors produced by

photoexcitation of suitable simple molecules. The mechanism

proposed involves the formation of unstable hydrazyl radicals

[1] which then disproportionate to form hydrazo compounds 13

[2] and these in turn decompose to the corresponding amines

[3]. Regeneration of the dye also occurs in the last two steps: . ~1-N=N-~2. + H" + ~1-NH-N-~2 (1] 2~1-NH-N-~2 + ~1-NH-NH-~2 + ~1-N=N-~2 [2]

Oxygen was found to inhibit fading by reoxidation of the reduced dye radicals.

Photoreduction of azo dyes has also been postulated in polymeric substrates. Giles et al., 17 suggested that reductive fading of azo dyes may occur in hydrocarbon media in the absence of air, and in wool, silk and gelatin in the presence of air although this appears far from conclusive.

Kamel et al. , 18 found substituent effects of azo dyes on nylon that indicated that the fading mechanism here too is reductive.

Similar substituent investigations on other substrates have been indicative of oxidative fading e.g., the work of

Matsuoki and coworkers 19 on polyester films. Various photo­ oxidative mechanisms have been proposed on the basis of studies of the photolysis of various dye-solution systems.

Naumova et al., 20 studied the fading of hydroxybisazo dyes in carbon tetrachloride and proposed that cleavage of the hydroxyl hydrogen atom occurred, followed by destruction of the azo group with the formation of quinones and aromatic anhydrides. Griffiths et al., 10 proposed a mechanism in 14

which attacks the hydrazone form of an azo

dye in methanol. They found that fading only occurred in

the presence of oxygen and that the introduction of a singlet oxygen quencher reduced the rate. Photoejection of a hydro­ gen atom from an amino group to form a radical which then disproportionates has been postulated as the mechanism of

the photooxidation of the dye Congo Red. 21

Other types of photochemical reactions have also been observed for azo dyes. Photocyclisation of azobenzene occurs in the presence of acid or a ferric chloride catalyst. 22 ' 23

It requires cis-azobenzene as the starting product and the

following mechanism has been proposed:

+ H + /H N=N/ hv

Ob 0-0H H

N=N /

0 b +H+

N=N 0--0

The final product, benzo[c]cinnoline, was identified as the

major fading product by Lewis. 22 A similar heterocyclic

decomposition product has been found following the rapid

fading of an o-aminoazo dye on cellulose 24 : 15

N=N ~NHO'--::: hv ~ 8

It is obvious from this latter fading mechanism that the ability of a dyestuff to undergo cis-trans isomerism can increase the probability of it fading rapidly. Largely to eliminate the phototropy caused by such photoisomerisation, most modern azo dyes have structures which either inhibit photoisomerisation or reduce the lifetimes of the cis-isomers.

Such structures either exist predominantly in the hydrazone form (~ 1 =N-N-~2) or have strongly electron-withdrawing groups para to the azo linkage.

1.3.1.2 Anthraquinone Dyes

Photodegradation of anthraquinone dyes (base 0 formula may also occur by either

0 reductive or oxidative mechanisms, again usually depending on the nature of the solution or substrate in which it exists.

Photoreduction may occur by an electron transfer process to produce radical ions: 16

Alternatively hydrogen abstraction can occur, in which case

semiquinone radicals (DH") are produced. 25 ' 26 These may subsequently react by a dismutation process to form the hydroquinone 27 ' 28 :

DH 0 + DH 0 + DH2 + D

In alkaline solution, depending on the substituents present on the dye, dianions (D 02 -) or the semiquinone radical anion o·- have been identified29 :

DH" + OH

D* + X + o· + x·

It is well established that the excited state of

anthraquinone dyes that is most reactive towards hydrogen or electron abstraction is the (n, TT)* triplet state. 2 ' 6

The movement of the electron away from the carbonyl oxygen on excitation makes the oxygen atom more electrophilic than

in its ground state making hydrogen or electron abstraction easier. Both the (TT,TT)* state and possible charge-transfer

states are less reactive in this respect. Consequently molecules having either of these states as their lowest

excited state will have higher lightfastness than if they

have the (n, TT)* state as the lowest energy state. As the

(n, TT)* and CT states have similar energies small differences

in substituent or solvent properties can alter their relative 17

positions and thereby greatly affect the reactivity of the excited molecule. For example, p -arninobenzophenone in iso­ propanol has the unreactive CT state as the lowest state while in cyclohexane the {n, TT*) state is the lowest. This

leads to hydrogen abstraction occurring in the cyclohexane but not in isopropanol. 2 Similarly the introduction of 0 X electron-donating substituents {X): X

0 change the relative positions of the energy levels so that the {TT, TT*) becomes the lowest excited state instead of the

{n, TT*) with a consequent loss of photoreactivity towards hydrogen abstraction. 6

High light stability in anthraquinones is often

achieved by 1-hydroxy substitution where intrarnolecular hydrogen bonding with the carbonyl group leads to rapid

deactivation of the excited states. 30 ' 31 ' 32 However in polar substrates, such as nylon, this substitution gives a

low lightfastness and 1,2-dihydroxy substitution is needed to stabilise the dye. 33 Allen et al., 34 found that sub­ stitution in the 2 and 3 positions by an amino and a hydroxy

group respectively leads to extremely rapid deactivation of

the excited states. This results in much higher light

stability than the corresponding individually 2-substituted

compounds in which deactivation is slower.

Photoreduction has also been shown to occur in poly­

meric media as well as in solution. The photolysis of arnino­

anthraquinones in nylon films has been shown to be 18

reductive. 35 ' 36 ' 37

Photooxidative fading of anthraquinone dyes has also

been found to occur, for example in polyester film. 38139

The photolysis of 1,4-diaminoanthraquinone inn-butyl acetate

is also oxidative involving a radical mechanism and photo­

oxidation by molecular oxygen. 40 The fading of this dye and

its N-alkyl derivatives has also been investigated in poly-

(ethylene tereph thalate) films. 3 8 ' 3 9 The photoreactions

postulated included:

N-dealkylation

nuclear hydroxylation

substitution of amino groups by hydroxyl groups

introduction of oxygen at the carbon atom adjacent

to the nitrogen of N-alkylamino groups.

This led Wegerle 38 to suggest that oxidative photodegra­

dation involving attack by some electrophilic agent on the

unshared electron pair of the nitrogen atom occurs with

subsequent dealkylation and the formation of a carbonyl

group at the a-carbon atom.

1.3.1.3 Triphenylmethane Dyes The two triphenylmethane dyestuffs most widely

studied are Malachite Green (C.I. Basic Green 4) and

Crystal Violet (C.I. Basic Violet 3): 19

Cl

where R = H for Malachite Green

and R = N(CH 3 ) 2 for Crystal Violet.

From the investigation of the photoproducts of these dyes

formed in solution and in various solid substrates, infor­ mation about the mechanisms of and the structural factors affecting the lightfastness of triphenyl­ methane dyes has been obtained.

Photooxidation to the corresponding diphenylketones and leuco compounds has been observed in solution 8 ' 41 ' 42 ' 43 and in solid substrates. 42 ' 44 Porter and Spears 42 identified benzophenone, 4-methylamino- and 4-dimethylamino-benzophenone as photoproducts of Malachite Green in aqueous solutions and on modified cellulose substrates, together with p -dimethyl­

aminophenol in solution only. On the basis of these photo­ products they proposed a photodegradation mechanism involving

absorption of light by the leuco carbinol form of the dye which then either undergoes radical fragmentation followed by reaction with oxygen and water, or reacts directly with 20 oxygen and water: QH,)2 Q-c+x--9

excited * [ J + state

(CH3) 2N -0-oH +

+ C-o-~ N/H O II - '-cH3 0

Evidence in support of the involvement of the carbinol base

(~ 3-COH) rather than the carbonium ion (~ 3-C+) included:

The rate of fading in solution was found to

decrease as the acidity of the solvent is increased. 25

When light of wavelengths absorbed by the

carbinol base (and not the coloured carbonium ion) 21

is filtered out, either by filtering the incident

light45 or by introducing a uv absorber into the

system25 the triphenylmethane dye is more stable

to light.

Bangert et al., 44 proposed similar reaction mechanisms to account for the photoproducts they identified by thin layer chromotography from the photolysis of Malachite Green and

Crystal Violet. The reactions included dealkylation via hydroxy methyl groups and oxidation to benzophenone (with no intermediate postulated).

Leaver 46 and Owen and Allen47 using spectroscopic and electron spin resonance techniques have detected triaryl­ methyl radicals on the photolysis of certain triphenylmethane dyes in de-aerated solution and in poly(vinyl alcohol) 46 and poly(methyl methacrylate) 47 films. Pak et al., 48 studied the leuco compounds of Crystal Violet Ar 3C-X where

Ar= o-N(CH3)2 and X = H, OH, CN,and maintained

that free radicals, cation radicals or cations can be produced from the excited triphenylmethane molecule, with the radical being formed in one of the following ways:

Ar 3C-X + e + Ar 3C 0 + X + Ar3C + e + Ar 3 C 0

and that the formation of the cation depends on the surrounding medium and conditions. 22

N-dealkylation of Malachite Green has been observed by a number of workers. 8 ' 42 ' 49 This reaction has also been observed in the photodegradation of dyes in other dye classes. 8

1.4 Factors Influencing Dye Fading

The chemical structure of a dye will basically determine the various types of photochemical reactions it is capable of undergoing. Which of these alternative reaction pathways are in fact followed in any particular system, and the rate of these reactions will depend on numerous other factors characteristic of the dye-substrate system as a whole, the surrounding atmosphere and the incident radiation.

1.4.1 Intensity and Spectral Distribution of Radiation

As the first prerequisite for a photochemical reaction to proceed is the absorption of light, obviously the dye (or the substrate, if energy transfer is possible) must absorb light of wavelengths emitted by the radiation source. That is,there must be an overlap of the absorption spectrum of the dye (and/or substrate) and the emission spectrum of the light source.

In addition, the different wavelengths of light have been found to vary in their efficiency in initiating photochemical reactions. The results of numerous investigations 43 ' 50 - 54 have shown that the quantum efficiency of fading decreases with increasing wavelength of the light - the higher energy and blue light 23 having the greatest effect and the lower energy red light the least. McLaren 52 using filtered on 117 Qyes

found that in general absorbed uv and violet light will

cause fading while red light will only be active if the dye is fugitive and not if it is moderately fast. Further, he

found that while fugitive dyes are faded by absorbed light of any wavelength, lightfast dyes are faded by absorbed

light only if it is below a certain critical wavelength that is different for each dye. This effect has also been observed by other researchers. Irick and Boyd54 studied disperse dyes on various hydrophobic substrates and in solution irradiated with both monochromatic and polychromatic light and found that no dye degradation occurred with light above 400nm. The quantum yield of dye degradation for monochromatic light below 400nm generally increased with decreasing wavelength while with polychromatic radiation the dye degradation rates were generally consistent with the

fraction of short wavelength emission of the light source.

Yamada et al., 43 and Maerov and Kobsa 45 observed similar behaviour in other dye-substrate systems.

Since each photon of light of the active wavelengths

can excite one molecule the rate of fading will increase with increasing intensity of this active radiation. Thus,

although Evans 55 found that uv light was the most effective

in promoting the fading of Rhodamine B, because of the very

low proportion of uv in sunlight he found that visible light

of wavelength 400-600nm although having a much lower quantum

efficiency was responsible for 84% of fading. 24

Also,it has been demonstrated that there is no threshold level of intensity below which fading will not occur56 as was originally believed.

1.4.2 The Physical State of the Dye

It is generally accepted that the more aggregated the dye the greater its lightfastness. This view was originally put forward by Giles and his coworkers 57 ' 68 on the basis that fading only occurs on the surface of a dye particle where the dye is accessible to radiation and to those agents such as oxygen and moisture which were thought to be involved in fading. More recently another mechanism for this effect has been proposed by Terenin, 59 namely that with increased aggregation there is a greater possibility of self-extinction i.e., the transfer of excitation energy between adjacent dye molecules, which results in the subsequent reduction in the quantum efficiency of dye fading.

This aggregation effect has been used to explain a variety of phenomena, including the improvement in light­ fastness with increasing dye concentration; 56 the improvement in lightfastness of azoic dyes on soaping; 58 ' 60 the decrease in the lightfastness of disperse dyes in the presence of certain carriers which are proposed to act as disaggregating agents; 61 and the improvement in lightfast­ ness of dyes in substrates with increasing pore size such as in viscose rayon compared to cotton. 56

This trend of increasing lightfastness with 25 increasing aggregation does seem to be followed in the case of dyes in hydrophilic fibres. 58 ' 61 The lightfastness of dyes was found to increase with the increasing relative moisture regain of the fibres. However, in the case of hydrophobic fibres, the opposite trend has been observed, 61 with the lightfastness of dyes decreasing with the moisture regain of the fibres. It has been postulated61 ' 62 that the significantly smaller pore sizes in hydrophobic fibres cause a second effect to predominate, the smaller pores restricting the diffusion of agents such as oxygen and moisture through the fibre to the excited dye molecule. Therefore, the smaller the pore size the greater this restriction and hence the slower the dye will fade. The higher lightfastness of reactive dyes on drawn compared to undrawn nylon has also been attributed to this "diffusion restriction effect". 63

The actual degree of aggregation of disperse dyes on hydrophobic substrates has been the subject of some controversy with Giles et al., 6 ~' 65 postulating that the dyes are aggregated, while Wegerle 38 and others 66 ' 67 maintain that they are present in monomolecular form. Both arguments are centred around the Beer-Lambert law and the ratio of the short and long wavelength bands of the dyes. Giles found that this ratio increased both with increasing concentration and on exposure to light which he attributed to an increase in the degree of aggregation of the dye. Wegerle on the other hand observed no change in the ratio with concentration, nor on fading when the absorbance of the photoproducts is taken into account. Direct physical measurements of the aggregation of these dyes gives conflicting results. 7 26

Giles et al., 56 ' 57 also maintain that the shape of a

dye's fading curve is entirely determined by the state of

aggregation of the dye which in most cases exists in a wide range of particle sizes. Kissa68 disagreed with Giles maintaining that this oversimplified the situation and that the slope of the fading curve was also dependent on the distribution and location of the dye in the fibre. Thus he

also contests the validity of the deductions Giles has made about the degree of aggregation of dyes based on the shape of their fading curves. Blair and Boyd69 contend that the increase in fastness with dye concentration is due to a

filter effect due to an increase in the amount of light

absorbed, as well as to dye aggregation.

Thus there remain some doubts as to the influence the aggregation of a dye has on its lightfastness.

1.4.3 The Nature of the Substrate

It was discovered as early as 1813 that the nature of

the substrate had a marked effect on the rate of fading of a

dye 70 and since then this has been observed by numerous

researchers. 8 ' 71 ' 72 As a consequence, it is always necessary

to specify the substrate involved when quoting lightfastness

data.

Both the chemical and physical nature of the substrate

can influence fading. Chemically, the substrate can itself

participate in the fading reactions, commonly by donating

either a hydrogen atom or an electron to the excited dye

molecule. Also, certain groups in the substrate may be able 27

to quench the excited dye molecule thereby reducing its fading.

The chemical nature of the groups which act as dye sites may also influence fading as may the nature of the dye-substrate bond. The physical characteristics of the substrate which may influence fading include its rigidity and porosity, and its permeability to various entities such as water and oxygen.

The various mechanisms involved in a number of these cases are dealt with in other sections of this review.

Chipalkatti et al., 73 divided fibre substrates into two classes on the basis of dye fading rates, namely protein and non-protein. It was suggested that the more rapid fading found on non-proteins was probably an oxidation process while on proteins fading was probably reductive involving reaction with the substrate.

As discussed in Section 1.4.2 the porosity of fibres has been postulated to influence dye fading either by limiting the degree of dye aggregation possible, or by restricting the diffusion of gases such as oxygen and moisture through the fibre.

The rigidity of the substrate has also been found to affect the photochemical reactions occurring74 ' 75 as rigidity will largely determine the possible motions, both diffusional

and conformational, of the various species present. This phenomenon is discussed further in Section 1.4.7.

It has been observed that fading behaviour can differ quite markedly in solution and in solid substrates. 27 ' 76 28

Egerton et al., 76 observed that the near uv-visible irradiation of the vat dye 1-benzamidoanthraquinone in ethanol readily produces the anthrahydroquinone whereas the same dye in cellulose film shows little or no tendency to reduction of this kind. Similarly Egerton et al., 27 found that H-abstraction occurred with 1,5-diaminoanthra­ quinone in organic solvents containing an abstractable H and in N-methoxy-methyl nylon films. However it did not occur in carbon tetrachloride solution, nylon and cellulose acetate films, or in the solid dye itself. They postulate that hydrogen abstraction in a polymer is dependent on the accessibility of the dye reactive group to an abstractable hydrogen with the aggregation of the dye also possibly influencing the degree of dye-fibre interaction.

Porter and Spears 42 found that altering the dye bonding site in cellulose to a sulphonate group improved the lightfastness to give a 5 to 10 times slower fading rate compared to carboxylate groups.

The polarity of the substrate has also been shown to affect fading. Using flash photolysis and luminescence spectroscopy Allen et al. , 77 ' 78 identified the excited species formed on irradiation of certain amino- and amino-

0 carboxy-anthraquinone dyes as either o· or DH • Strong transient absorption was observed in polar solvents but no transients were detected in non-polar solvents. This effect was postulated to account for the low lightfastness of these dyes on nylon 6.6 fabric, which is highly polar, compared with the high fastness on the non-polar polyester fabric. 29

The presence of other chemicals in a dyed substrate can also affect the lightfastness of the dye e.g., titanium dioxide delustrant is known to accelerate fading 7 as does the presence of certain carriers used in dyeing which are thought to act as disaggregating agents. 61

1.4.4 The Dye-Fibre Bond

The effect of covalent bonding between reactive dyes and substrate on lightfastness remains unresolved. Several workers have found improved lightfastness on fixation 79 -ai while others have obtained inconclusive results, 63 ' 82 - 8 ~ with some dye-fibre systems showing improvement in light­ fastness on fixing, some showing reductions and others no significant change.

The improvement of lightfastness on fixation has been attributed to the greater possibility of energy transfer away from the dye to the substrate through the covalent bond. 79 ,ao However this ignores the fact that energy transfer is unlikely because of the energy levels involved.

Giles et al., 82 attributed reductions in lightfastness with covalent bonding to the accompanying restriction to the aggregation of the dye. Krichevskii et al., 83 took a similar view but maintained that two opposing effects occurred.They postulated that closer contact between dye molecules lead to greater self-extinction in addition to the formation of aggregates reducing the surface area of the dye accessible to fading. 30

1.4.5 Oxygen

It was originally believed that oxygen always promoted the fading of dyes, with fading being retarded or even entirely prevented in its absence. 85 ' 86 Giles 56 for example found that lightfastness increased linearly with the concentration of oxygen in the atmosphere. However, other investigators have found that sometimes dyes fade more rapidly in the absence of oxygen. 3 ' 43 ' 87 Egerton et al., 3 for example found that the fading of a particular disperse dye on cellulose acetate film was negligible in dry nitrogen but significant in dry oxygen while the same dye in nylon film faded appreciably in dry nitrogen. Again it becomes obvious that observations for a particular dye-fibre system are specific to that system and are not necessarily generally applicable.

The accelerating effect of oxygen may be caused by the direct reaction of the oxygen with the excited dye molecule. Alternatively the oxygen may first quench the excited dye molecule thereby becoming excited itself and then react with the dye in either its ground or excited state. It has been shown that the excited state of oxygen involved is likely to be singlet excited oxygen. 5

Various explanations have been advanced for the retarding effect that oxygen can have. Van Beek et al., 87 postulated that oxygen inhibits the fading of azo dyes in solution with hydrogen-donor substrates by competing with the dye for hydrogen from the excited substrate molecules and by reoxidation of the reduced dye radicals back to the 31

original dye by the peroxide radicals formed. Yamada et al.,~ 3 studied triphenylmethane dyes in solution and found that the quantum yields in anaerobic solutions were ten times greater

than those in aerobic solutions. They proposed that these dyes undergo both oxidation and reduction and that the dissolved oxygen restricts the latter reactions which have higher quantum yields than the former.

1.4.6 Moisture

Lightfastness normally decreases with increasing atmospheric humidity 71 ' 88 ~ 0 with the extent of the effect being dependent on the dye-substrate system in question.

An increase in fading rate of up to sixteen times when damp compared to dry has been reported by McLaren9 0 for some dyed samples while Giles et al.,9 1 on the other hand found that dye fading on polyester film was not sensitive to relative humidity at all. There are basically two explanations that have been advanced to explain this phenomenon. The first is that the water itself or some compound formed from it takes part in the fading reactions. The second involves the physical effect that moisture has on either the substrate by swelling it and thereby increasing the rates of diffusion within it or on the dye, possibly by displacing adsorbed dye or by disaggregating it.

Giles et al.,9 2 confirmed previous observations that

the fading of dyes on cellulosic substrates was particularly

sensitive to changes in the relative humidity of the

surrounding atmosphere while dyes on proteins are relatively

insensitive to changes in humidity. Giles proposed that the 32 reason for this was that water swelled the substrates thereby increasing the diffusion of atmospheric oxygen to the excited dye molecule. In cellulosic substrates where the major degradation mechanism is oxidation the moisture content of the fibre therefore has a significant effect. However in protein substrates where reduction is the major mechanism of

fading, with hydrogen being extracted from the substrate itself, this increase in oxygen diffusion has little effect.

Disperse dyes on hydrophobic synthetic polymers have also been examined for relative humidity sensitivity in

fading.9 1 On cellulose acetate and triacetate and on nylon the dyes showed an increase in fading rate with increasing relative humidity. On polyester film on the other hand the dye fading was not sensitive to humidity, which is under­ standable since water does not significantly swell polyester.

The action of water in increasing fading was attributed to a

combination of the following factors:

- displacement of adsorbed dye by the water, thereby

interfering with the dye-substrate interactions,

- increased accessibility of adsorbed dye molecules

to the oxygen of the atmosphere caused by the

swelling action of water on the substrate.

Datyner et al., 84 investigated the effect of liquid

water on the quantum yields of fading of reactive dyes on

cellulose. For all the dyes examined the quantum yields

were higher wet than dry. With one dye the poor lightfast­

ness in water was found to be due to the participation of 33 hydroxyl radicals produced directly from water in the fading reactions. For another dye the major fading reaction involved the photoejection of electrons which was greatly enhanced in an aqueous environment.

Thus there appears to be a number of ways in which moisture can affect the fading of dyes and it is possible that in any situation a number of these mechanisms could be operating simultaneously.

1.4.7 Temperature

At constant humidity a rise in temperature increases the rate of fading of dyes. 8 8 ' 9 1 ' 9 3 The effect is small but significant with measured activation energies ranging from approximately 4 to 76 kJmol- 1 and is more marked at high rather than low moisture regains. As photochemical reactions are activated by light rather than heat they have virtually no activation energy. Therefore, the temperature dependence of fading must be due to the temperature dependence of "dark" reactions in the fading sequence subsequent to the initial excitation of the dye molecule.

The extent of this temperature effect depends again on the dye-substrate system in question. Giles et al. ,9 1 found that the activation energies of dye fading were dependent more on the nature of the substrate and on the relative humidity than on the nature of the dye molecule. They determined activation energies ranging from approximately

8 kJmol- 1 for anionic dyes on wool to 76 kJmol- 1 for disperse dyes on nylon. From Hedges' results 88 for anionic dyes on 34 wool Giles also calculated activation energies of

approximately 5 to 13 kJmol- 1 • Shah and Srinivasan 93 on the

other hand found that activation energies were determined both by the substrate and the class of dye and gave the

following values:

Apparent Activation

Energies of Fading

(kJmol- 1 )

anionic dyes on gelatin film 13 - 19

reactive dyes on cellophane 31 - 43

disperse reactive dyes on nylon 49 - 55 disperse dyes on nylon 52 - 76

When a material is exposed to light, the temperature of the surface is increased significantly above that of its surroundings. Temperature rises of the order of 35 to 55°c were observed by Nordharnrnar 94 on exposure to sunlight.

However the moisture content of the exposed surface decreases

concurrently and so the rate of fading in practice will be

influenced by the combination of these two effects.

Temperature has also been found to influence photo­

chemical reactions in another way, through its effect on

the rigidity of solid substrates. 75 ' 9 5 This effect has been extensively reviewed by Williams and Daly95 where the importance of free volume and the glass transition of polymers is emphasized. The temperature, and whether it is

above or below the polymer's glass transition temperature, will determine the amount of free volume available for the 35 movement of a reactive site. The amount of movement necessary to form the geometry of the excited or transition state and to allow some separation of the products will determine the amount of free volume required for a given photochemical reaction to take place. Typical motions which may be involved include conformational changes of a cyclohexane ring, cis­ trans isomerisation, rotations of a methyl or phenyl group and diffusion of a small molecule to a large molecule where reaction can take place. Depending on the amount of movement required some reactions will be unaffected by substrate rigidity while others will be considerably affected.

Thus it becomes obvious from the preceding discussion that each dye-fibre system must be considered individually as so many different factors can affect dye fading that it is impossible to generalise to any great extent.

1.5 Scope of the Present Work

The aspect of dye fading studied in this work involves a substrate effect, the aim of the research being to elucidate the anomalously high lightfastness of basic dyes observed on acrylic fibres compared to the fastness on cotton, nylon, wool and silk. The reason for particular interest in this area lies in the very desirable properties of this class of dyestuff.

By definition, basic dyes are those dyes in which the coloured chromophore is positively charged. For this reason 36 they are also known as cationic dyes and in fact this

classification has become used more frequently, especially with the more recently developed dyestuff ranges. The particular advantages of dyes of this type are their very bright colours and their very high extinction coefficients.

A large range of bright colours is available within the basic dye class and less dye is required for a given depth of shade compared with other dye classes.

Their use originally was on wool and tannin-mordanted

cotton but by the 1950's this had been virtually discontinued because of their very poor lightfastness. About this time it was discovered that newly-developed acrylic fibres could be dyed with basic dyes and that, unexpectedly, the dyes displayed very good fastness to light. This led to their reintroduction and the subsequent development of many new basic dyes with improved properties.

Acrylic fibres contain at least 85% polyacrylonitrile

(-CH 2 -CH-) . Usually the remainder consists mainly of a I n CN

copolymer with a bulky side chain such as vinyl acetate or methyl acrylate which opens up the structure to allow penetration of dyes. In addition a further copolymer

containing either acidic or basic groups to provide dye

sites is often incorporated in very low concentrations.

Of most interest for basic dyes are the cationic-dyeable

acrylics which contain acidic dye site groups, typically

as sodium styrene sulphonate. 96 37

Other synthetic fibres including polyester and nylon have also been chemically modified to produce cationic­ dyeable fibres. However the use of basic dyes on these substrates has been less successful because they display lower fastness to light, and, as alternative dyes are available for these substrates (unlike the acrylics} basic dyes are not often used on them.

Various explanations for the extremely high light­ fastness of basic dyes on acrylics compared with other substrates have been proposed but very little research has been carried out on the phenomenon. One postulate is that acrylics are less permeable to oxygen and moisture. 61 , 9 7

Schwen and Schmidt 71 found that for the range of basic dyes they studied, the presence of moisture slightly increased the rate of fading on the various substrates tested. In most cases the dyes also faded more rapidly in oxygen than in nitrogen but there were exceptions. However, even with both oxygen and moisture excluded (i.e., in a dry nitrogen environment} the lightfastness of the dyes on the non­ acrylic fibres, which included cotton, silk and wool, was generally not as high as on the acrylic fibre. In fact, differences in lightfastness of up to more than 6 points on the standard ISO lightfastness scale (which ranges from 1 to 8) were observed. It would appear that permeability to oxygen and moisture probably does play some role, which could be different for each dye.

It has also been postulated that the nature of the bond between the basic dye and the substrate is a reason 38 for their relatively high lightfastness on acrylics.

Wegmann98 found that the lightfastness of basic dyes increased with decreasing basicity of the dye cation which he attributed to the changing character of the dye­ fibre bond. He postulates that as the basicity of the dye decreases, the dye-fibre interaction becomes less electrostatic in character which in turn means increased interaction between dye and fibre. This means that the absorbed energy is more easily conducted away from the dye and therefore slows down its photodegradation. Evans and Stapleton 8 on the other hand found that while this correlation between basicity and lightfastness held for

Orlon it does not hold for wooLwhich suggests that different fading mechanisms may be operating on the two fibres.

On the basis of Wegmann's explanation Bitzeret al., 25 tried to modify cotton by the introduction of acidic groups so as to produce similar binding conditions between dye and fibre as those in acrylic fibres. Their treatments included carboxylation, cyanoethylation, carboxymethylation and treatments with butane sulfone and acrylic acid. The results obtained were variable, with each treatment giving slight improvements in the lightfastness of some dyes but not others. In all cases the improvements did not bring the lightfastness up to that on acrylic. They attributed the lower lightfastness on cotton to be due to its higher water content and to the dye being present in a different physical state. 39

Porter and Spears 42 also modified cotton, producing sulfoethyl and carboxymethyl ethers to determine the effect of sulphonate and carboxylate dye sites. They found that the triphenylmethane dye, Malachite Green, faded five to ten times slower on the cotton containing -SO 3 sites compared with that containing -COO sites or untreated cotton but that the photodecomposition products on all three were identical. They deduced that the reactions involved were the same in each case and proceeded via the carbinol form of the dye. The difference in rate was postulated to be related to the pK of the dye site a which would determine the concentration of the carbonium ion (Ar3C+) and therefore the concentration of the carbonol base present. Porter and Spears also suggest that differences in dye sites could be a major reason for the lower lightfastness of basic dyes on nylon compared with acrylics as cationic-dyeable acrylics generally have sulphonic acid groups incorporated into the polymer while the nylon utilises carboxyl groups for fixation.

Zollinger ,9 9 .. suggests an alternative reason for the difference between carboxyl or sulphonate sites. The excited dye molecule is postulated to abstract an electron from the carboxylate group to form a carboxyl radical which then decomposes to carbon dioxide and a residual radical within the fibre. Thus a recombination reaction with the dye radical is prevented and it will decompose irreversibly. Sulphonate ions on the other hand have little tendency to form radicals and so it is possible for the excited state of the dye to decay to its ground 40 state without decomposing.

Zollinger99 and his coworkers 96 disagreed with

Wegmann's basicity arguments pointing out that only dyes of similar chemical structure should be compared in such an investigation and that when discussing so-called basicity the measurable quantity pK should be used. When Johnson used Wegmann's lightfastness values and pK values determined himself he found the opposite mechanism to that proposed by

Wegmann with lightfastness increasing as the ionic character of the dye-fibre interaction increases. Subsequently

Johnson found that this relationship was also valid for a series of triphenylmethane dyes whose structure was systematically varied. 96

From the preceding discussion it can be seen that the explanation for the relatively high lightfastness of basic dyes on acrylic fibres is still far from being resolved.

It was the aim of the present work to examine the phenomenon more closely and from different angles to those already employed, to attempt to elucidate the mechanisms and factors involved. Solid substrates only, in the form of fabric and transparent polymer films were used and the factors studied included the effects of oxygen and moisture, the effect of dye sites and other groups present in the polymers and the effect of the rigidity of the substrate on fading. The latter was studied by varying the temperature of fading.

Also it was hoped to identify the excited species involved by the use of appropriate scavenging and quenching agents. 41

CHAPTER 2

QUANTUM YIELDS OF BASIC DYE FADING

2.1 Introduction

The initial task was to quantify the phenomenon to be studied. This was done by determining the quantum yields of fading of basic dyes on an acrylic substrate compared to other polymer substrates. The photochemical quantum yield of dye fading (~) gives a measure of the efficiency of the absorbed light in causing fading and is defined as:

number of dye molecules degraded ~ = number of quanta of light absorbed

Dyed fabrics were used in these quantum yield determinations as this is the dye-substrate form of most interest in practice. The difficulties involved in measuring dye fading on fabrics has meant that most recent research on solid substrates has been carried out on transparent polymer films rather than on fabrics. The technique for following the fading of dye in fabric involves the measurement of the reflectance of the fabric, a property which is not directly related to dye concentration. Recent improvements in instrumentation have greatly increased the speed and accuracy of reflectance measurements.

Preliminary fading trials were carried out on a variety of fabrics and dyes to determine those most suitable for use. Dyes chosen were ones that showed a marked difference in lightfastness on acrylic fibres compared with 42

the other fibres tested. As far as possible dyes were selected that faded "on-tone" i.e., with no accompanying change of hue and that faded within a reasonable time so

that experimental fading times would not be inconveniently long. This latter condition effectively precluded the use of the more recently developed basic dyes.

Cotton, wool, nylon and acrylic fabrics were used in these trial fadings, but as it was found that the dyed acrylic fabric had by far the superior lightfastness it was decided to restrict investigations to two fibre types i.e.,

acrylic and one other. Nylon was chosen as the second

fabric as the dyes on nylon displayed the lowest lightfast­ ness of the three dyed non-acrylic fabrics examined.

2.2 Fabrics

2.2.1 Acrilan 16

The acrylic polymer used was Acrilan 16, a basic­ dyeable acrylic manufactured by Monsanto. The exact

chemical structure of Acrilan 16 would not be disclosed, but being an acrylic polymer it would contain at least 85% polyacrylonitrile with the remainder being made up of a number of copolymers whose functions would be to open up

the structure and provide negative dye sites for the basic

dyes. It was obtained as a plain weave fabric of weight

300 ± .6 gm- 2 from Bradmill. The fabric was sufficiently

dense to be self-supporting and virtually opaque. 43

2.2.2 Nylon 6.6

Nylon 6.6 [NH(CH2) 6 NH-CO(CH ) ~CO-] was used in the 2 n form of a knitted fabric. The fabric was of locknit . F2-3/l-0 construction (Bl-O/l-2 ) produced from 44 decitex nylon 6.6 filament yarn (Fibremakers) on a "Hobley", 2 bar, 28 gauge warp knitting machine (J.Hobley & Co. Ltd., England) in these

laboratories. As the fabric construction was a loose one

(fabric weight 107 ± 0.5 gm-2) it was folded in four and

attached to a cardboard backing for irradiation in order to

reduce light penetration.

2.2.3 Basic-Dyeable Nylon

A nylon polymer modified by the introduction of

anionic dye sites to increase its substantivity to basic

dyes was obtained in yarn form from ICI Fibremakers. The yarn was a heavy one (610 ± 40 tex = g.km- 1 ) and samples

for fading were produced by winding the dyed yarn around

a cardboard backing.

2.2.4 Chemical Compositions

In an attempt to elucidate the chemical structure

of the polymers, an elemental analysis was carried out on

them by the Australian Microanalytical Service (C.S.I.R.O.

Division of Applied Organic Chemistry, Melbourne). Table

2.1 gives the results obtained together with the calculated

quantities for the polyacrylonitrile and nylon 6 homopolymers

(all expressed as percentages by weight).

In addition, the presence of carboxyl groups in

Acrilan 16 was tested by a method described by Heinkel; 00 TABLE 2.2 Structures and Lightfastness Data

of the Basic Dyes Studied.

C.I. Basic Blue 4 (oxazine)

lightfastness cotton 2-3

C.I. Basic Red 18 (azo) Cl

02N ;;--{ N=N-o~-N/C2Hs 1 \d. _ \ +/CH3 CH2CH2N'-. CH3 Cl CH3

light fastness acrylic 6-7

C.I. Basic Violet 14 (triphenylmethane)

lightfastness cotton 1, wool 1-2, acrylic 4-5 44

According to Heinkel, acrylic fibres containing carboxyl groups are very pH sensitive and by treatment for 2 minutes

0 at loo c with the dye Maxilon Black Tat pH 2 using H2 S0 4 , acrylic fibres containing carboxyl groups are stained pale blue whereas fibres which do not contain carboxyl groups vary between khaki and dark olive green. When Acrilan 16 was tested in this manner it was stained a dark olive colour indicating that carboxyl groups are not present.

The results from the staining test and the elemental analyses indicate that -so 3 and not -coo sites are present in each polymer.

2.3 Dyes

Dyes were chosen from three different chemical classes that met the required criteria i.e., azo, oxazine and triphenylmethane. Table 2.2 gives their chemical structures together with the limited lightfastness data available in the dye manufacturers' pattern cards and the

Colour Index. The preliminary fading trials with these dyes demonstrated the extremely different fading rates on the nylon and acrylic fabrics, with the difference being clearly visible by eye.

It was subsequently found that the fading rates for the triphenylmethane dye, C.I. Basic Violet 14, on nylon and acrylic fabrics could not be measured since this dye did not fade "on-tone" and the change in hue which accompanied fading was found to interfere with the determination of dye concentration. (The other dyes of the 45

triphenylmethane class that were tested in the preliminary trials also exhibited similar behaviour.) However, in subsequent work involving transparent polymer films this dye could be studied because the determination of concentration involved a different technique using a single wavelength of light rather than a wavelength band.

In order to determine the extinction coefficients of the dyes for concentration determinations dye samples were purified by recrystallisation from 50/50 (v/v) ethanol/water until the absorbance of a constant weight/ volume solution was maximised.

2.4 Dyeing Methods

All dyeings were carried out in beakers with heat supplied by hotplates. Dyestuffs were used in their commercial form (i.e., without purification) as the dyeing operation results in only the pure dye being absorbed and bound by the fibres while any impurities (which are typically salts) either remain in the dyebath or are removed from the fabric on rinsing. A range of four concentrations of each dye on each fabric was prepared. The actual concentrations used depended on the extinction coefficient of the particular dye and the amount of dye that could be adsorbed by the fabric. Very high concentrations of dye had to be avoided as these concentrations often resulted in changes in the hue of the dyeing, probably caused by aggregation of the dye. 46

2.4.1 Acrilan 16

A dyeing recipe typical of those given in dyestuff manufacturers' pattern cards for basic dyes on acrylics was followed:

Dyebath:

0 . 0 5 to 1 • 0 % dye

2.6% Astragal PAN (Bayer)

1% sodium acetate

0.9% acetic acid

10% sodium sulphate

Liquor Ratio 40:1 [=volume of dyebath (mls):

weight fabric (g)]

All percentages are on weight of goods. Astragal PAN is a cationic retarder which is used to control the rate of dyeing and thereby obtain level dyeings. Sodium acetate/ acetic acid are used to buffer the dyebath to pH 4.5-5.

Method: The fabric was entered at 50-60°c and run for a few minutes. The temperature was then raised gradually to

80-85°c over approximately 20 minutes, then to the boil over

40 minutes. Dyeing was then continued for approximately 2 hours at the boil. Finally the fabric was rinsed thoroughly in cold water.

The concentration of dye adsorbed by the fabric was determined by a mass balance technique: The absorbance and volume of the dyebath were measured before dyeing, and that of the dyebath plus rinse water after dyeing. The dye 47

concentration on the fabric thus obtained was expressed as

µmol (dye)/gram (fabric).

2.4.2 Nylon 6.6

Since unmodified nylon has a low substantivity for basic dyes, a very high concentration dyebath had to be used and the concentration range obtainable was fairly limited.

No other additives were present in the dyebath apart from the dye and again a liquor ratio of 40:1 was used. The

fabric was entered at 40-60°c, raised to the boil and then dyed at the boil for approximately 1 hour. This was followed by a thorough rinsing in cold water.

The concentration of dye on the fabric could not be determined in the same way as for the acrylic fabric as only

a very small fraction of the very large amount of dye in the dyebath was adsorbed by the fabric. Instead, the dye was

subsequently extracted from a dyed fabric sample of known weight by refluxing with 50/50 (v/v) pyridine/water. The

absorbance of the extracted dye was then measured and the

concentration of dye on the fabric calculated. Extinction coefficients of purified dye in 50/50 (v/v) pyridine/water were determined for this purpose.

2.4.3 Basic-dyeable Nylon

The acrylic dyeing recipe was found to give

satisfactory results with the modified nylon fabric. The

cationic retarder was necessary to control the rate of

dyeing and so obtain level dyeings. 48

2.5 Determination of Dye Concentration

The reflectance of the dyed fabrics was measured on a Hunterlab D25D2M Colour and Colour Difference Meter. This instrument measures reflectance in terms of the three C.I.E. Tristimulus Values X, Y and z. These X, Y and z readings give the reflectances of the sample at three different wavelength bands.

The diameter of the specimen port of the Hunterlab was reduced from 56mm to 32mm by the use of a specially constructed black annular plate, to enable the small fabric samples (37mm x 40mm exposed area) used in the fading experiments to be measured. The X, Y and Z readings obtained were corrected to account for this reduced area as the instrument is calibrated using the larger port:

a (R - b) R = reading

where R = X, y or z C. I.E. Tristimulus Value

R = x, y or z reading obtained using the reading 32mm port

a = Area correction

= 1.2 for X and Y

= 1.5 for z

b = Zero correction

= 1.4 for x, y and z

These correction factors were obtained empirically. The

X, Y and z values so obtained were then converted to percentage values as follows: Z% R = ITO 6

5

4

3

2

1 X% R = I'ITTf

1 2 3 4 5

Dye Concentration (µmol.g- 1 )

(l-R) 2 FIGURE 2.1 The Kubelka-Munk function 2R X% Y% Z% ( for R , and versus dye = 100 100' 100> concentration for C.I. Basic Red 18 on

Acrilan 16 fabric. 49

Y% = y

X X% = . 9 80 41

Z% z = 1.18103 where 100% reflectance indicates a perfect white reflector, for the light source used.

As stated previously, there is no direct relationship between the reflectance of a dyed fabric and its dye concentration. However the empirical Kubelka-Munk relation 101

[2.1] has been found to give good agreement between the two properties:

(1 - R) 2 (1 - Ro) 2 + kC [2.1] 2R = 2Ro where R = fractional reflectance of dyed fabric X% Y% Z% (in this case, R or = 100' 100 100 Ro = fractional reflectance of undyed fabric (as above)

C = concentration of dye k = the absorption constant for the given dye-substrate system.

X, Y and Z values of duplicate samples of four dye concentrations of each dye-fabric system were measured, the dye concentration determined as described in Section (1 - R) 2 2. 4 and versus dye concentration plotted. 2R A typical set of Kubelka-Munk plots, for C.I. Basic Red 18 on Acrilan 16 and on nylon 6.6 are shown in Figures 2.1 and 4.0 Z% R = 100

3.5

3.0

2.5 (-.j -p:: p:: IN r-1 2.0

1.5

1.0

X% R 0.5 = 100

1 2 3 4 5

Dye Concentration (µmol.g- 1 )

(1-R) 2 FIGURE 2.2 The Kubelka-Munk function 2R X% · Y% Z% (for R = t5o, 100 .and 100) versus dye concentration for C.I. Basic Red 18 on

Nylon 6.6 fabric. 50

2.2 respectively.

The Kubelka-Munk relation was found to be valid for

2 all the dye-fabric systems tested, (1 - R) 2R versus C giving straight line plots with correlation coefficients

{r) greater than 0.9. However, the intercept on the (1 - R) 2 {l-Ro) 2 axis {i.e., when C = 0) was not equal to 2R 2Ro - where Ro corresponds to the reflectance of the "blank dyed" fabric i.e., fabric which has been subjected to the dyeing treatment but without the dye. It was considered that the (1 - R) 2 actual intercept of the plot of 2R versus C was the more appropriate value to use in calculating dye concentrations from reflectance measurements. This was a reasonable approach providing the concentrations being determined were within the range for which the Kubelka-Munk relation had been demonstrated to be valid. Therefore equation [2.1] can be expressed in the following way:

f f. + kc = 1 f - f. and C = 1 k

(1 R)2 X% Y% Z% where f - for R = 2R = 100' 100' 100 f. value of f at C 0 axis, which is a constant 1 = = for a given dye-substrate system

C = concentration of dye k = the absorption constant for a given dye-substrate system

Hence the progress of fading was determined as follows:

drain drain

Coils Coils

to to

Cooling Cooling

II II

Tap Tap

II)..,. II)..,.

System System

Solution Solution

CuS01t CuS01t

1.0% 1.0%

Filter Filter

Sample Sample

Holder Holder

Irradiation Irradiation

Lamp Lamp

Xenon Xenon

Filter Filter

the the

Heat Heat

of of

Electrode Electrode

Negative Negative

= =

d<:)Fan d<:)Fan

Elevation Elevation

Side Side

to to

to to

I I

rP1 rP1

supply supply

power power

supply supply

power power

• •

· ·

Schematic Schematic

Cover Cover

3 3

,. ,.

2. 2.

Motor Motor

FIGURE FIGURE to to 51

Fraction of dye remaining at time t concentration of dye present at time t = concentration of dye present at zero time

f . f. time t 1 = f f. zero time 1

As with transparent media where the maximum

concentration accuracy is obtained by measurement at the wavelength of maximum absorption, so the most sensitive

reflectance region is at the wavelength of minimum reflectance. Consequently the Tristimulus Value used for each dye was the one whose wavelength band was nearest the absorption band of that dye. This gave the most accurate determination of concentration.

2.6 Irradiation Technique

A 1600W Xenon lamp (Wotan XBO 1600W -free

lamp) was used as the radiation source since it has a

spectral output similar to sunlight. Its high output

ensured short fading times, but had the disadvantage of

considerable heat generation, and hence it was found necessary to filter out the heat to prevent thermal

degradation of the acrylic fabrics.

A schematic side elevation of the lamp system is

given in Figure 2.3. The Xenon lamp is located at the

centre of the system. The heat filter, consisting of two

sealed pyrex glass cylinders with the space between the

cylinders filled with 1% CuSO4 solution, surrounds the 52

lamp. This filter absorbs the heat from the lamp without

absorbing the lower wavelength radiation. The CuSO 4 solution is continuously pumped through the glass filter from a water­ cooled bath.

Although CuSO 4 inhibits organic growth its presence alone was not sufficient to completely eliminate all growth in the heat filter system and so a few drops of m-cresol, a strong fungicide, were also added. The filter was also regularly cleaned by pumping through 6M HNO 3 , followed by

10% NaCl, and finally rinsing it with fresh water before refilling it with freshly prepared CuSO 4 solution.

The samples for fading were clamped in metal frames and placed in vertical slots mounted on a carousel sample carriage positioned outside the heat filter. The carriage rotates slowly (1.2 r.p.m.) around the Xenon lamp to average the radiation received by each sample. The samples were positioned level with the centre of the lamp where the intensity of the incident light was a maximum. The distance of the sample from the lamp, and the area of sample illuminated were chosen so that the variation in the incident light intensity across the sample was less than 5%.

2.6.1 Calibration of Incident Light Intensity

Because of the variation in the radiant output of the Xenon lamp with time, due to changes in the city voltage supply and to aging of the lamp, it was necessary to

continuously monitor the light intensity output during

fading and a light flux integrator was constructed for this

v v

Counter Counter

I I

-

-

I0.12µF I0.12µF

I I

-

-

< <

-1-

16µF 16µF

2.2Mn 2.2Mn

0.12µF 0.12µF

RL-A RL-A

1akn 1akn

l.8Mn l.8Mn

. .

-

Integrator Integrator

50µF 50µF

L L

_ _

+ +

-

-

-

RL-B/2 RL-B/2

-

1

Flux Flux

lOkn lOkn

;J ;J

Light Light

-

21okn 21okn

100n 100n

. .

-

the the

RL-B RL-B

of of

-

-

I I

10,000µF 10,000µF

Diagram Diagram

+ +

+ +

lOOµF lOOµF

Circuit Circuit

2.4 2.4

6x4 6x4

4x0A202 4x0A202

G~ G~ FIGURE FIGURE 53 purpose. The light was detected by a photocell (RCA 935) placed in one of the sample slots on the rotating sample

carriage, and arranged so that the light dose incident on

the sample was measured. The terminals of the photocell were connected to two concentric copper tracks running

around the carriage base. Two copper contacts fixed to the

lamp cover relayed the signal from the photocell to a voltage to frequency converter and from there a cumulative

frequency count was displayed on a frequency meter (a Fluke

1900A Multi-Counter). The circuit diagram for the integrator

is given in Figure 2.4.

The cumulative light "count" obtained from the light

flux integrator was calibrated by chemical actinometry using

an acid solution of potassium ferrioxalate [K 3Fe(C2O~)3]. 102,103 The potassium ferrioxalate is reduced during

irradiation to produce Fe2+ with a known quantum efficiency.

The concentration of Fe 2+ present after exposure was

determined by adding 1,10-phenanthroline to form an intensely

red coloured complex whose concentration was determined

spectrophotometrically.

The number of Fe 2+ ions produced during irradiation

(NFe2+) was calculated as follows:

V1 . V3 2 NFe2+ = NA X X [Fe +] V2

where NA = Avogadro's Number V1 = the volume of actinometer solution irradiated (20mls) 54

V2 = the aliquot volume taken for analysis (10 mls)

V3 = the volume of the flask used in the analysis (25 mls) and [Fe 2 +] = As 1 o E: • .R, where As 1 o = the measured absorbance of the F e z+ - phenanthroline solution at the wavelength

of its maximum absorbance, 510nm.

£ = molar extinction coefficient of the

Fe 2 +-phenanthroline at 510nm •

.R, = optical path length of spectrophotometer cell used for the measurement of the

absorbance.

The value obtained for As 10 was 8.82 x 10- 3 count- 1 cm- 2 for a 1 cm cell. The value of E: = 1.06 x 10 4 .R, mol- 1 cm- 1 obtained was in good agreement with literature values

[(1.10 - 1.15) x 10 4 .R, mol- 1 cm- 1 ]. Using this data the value for NFe 2 + was calculated to be:

= 2.76 X

Now, number of quanta =

where ~a = quantum yield of the actinometer.

As ~a varies with wavelength it was necessary to determine 55

the spectral distribution of the radiation before the number of quanta could be calculated.

Light from the lamp passed through the heat filter and onto the entry slit of a Farrand Grating Monochromator

Cat. No. 117180 (Farrand Optical Co. Inc. New York) which was coupled with an RCA 1P28 Photomultiplier. The plot of intensity versus wavelength obtained was corrected for the spectral response of the photomultiplier and the percentage transmission of the monochromator grating. This information was obtained from the manufacturer's catalogues where the data was given at 20nm intervals, and interpolated to obtain values every 5nm.

The relative intensity distribution thus obtained was then calibrated in units of quanta cm- 2 count- 1 using the actinometry result obtained previously

= 2.76 X in the following manner:

Now =

where I>.. = Intensity of light at wavelength >.. (quanta cm - 2 count -i)

IR Relative intensity of light at wavelength >.. >.. = (in arbitrary units)

k conversion factor (constant) and = I ...j I 0 12 r-1 X

Q) r-1 ~ 10 rtl Cl) ...- ~ I 0 .µ .µ § 8 0 ~ u ~ ·r-1 N u I ~ I:: H u 6 ~ rtl 0 .µ •r-1 .µ ~ rtl ::, ·r-1 01 ,0 rtl 4 p::;

4-1 0

.µ:>"t ·r-1 2 !/l ~ Q) .µ ~ H

350 400 450 500 550 600 650

Wavelength (nm)

FIGURE 2.5 Intensity Distribution of Radiation from the

Xenon Lamp Incident on Sample Surface. 56

where ~aA = quantum yield of the actinometer at

wavelength A.

The upper and lower limits involved in this summation are

= 580nm, above which the quantum efficiency of the

actinometer is zero, and Al = 320nm, below which the lamp output intensity is zero.

Therefore

A=580 A=580

= ~ = ~ IR k NFe2+ l aA IA I aA . A . A=320 A=320

NFe2+ and hence k = A=580 IR I ~aA A A=320

The value of k could thus be calculated and then used to convert the relative intensity distribution to an absolute one in units of quanta cm-2 count- 1 . The summation was performed over 5nm intervals which gave adequate accuracy

(± 2%). The values for ~aA required were obtained by graphical interpolation of the values determined by

Hatchard and Parker.1021103 The resulting intensity distribution output from the lamp incident on the samples is given in Figure 2.5. It can be seen that the intensity begins to fall above approximately 590nm. This is due to the absorption of light by the copper sulphate filter. 2.0

- i::: ·r-1°' i::: •r-1 rtl s Q) p:::

Q) ::>i 1.9 Cl r .µ i::: Q) .99 u ~ .99 Q) -Ill tJ'I .99 0 r-1

• 9 7

1.8

.99 • 99

1 2 3 4

FIGURE 2.9 Semi-log Plot of the Fading Curves presented in Figure 2.8, for C.I. Basic Red 18 in:

(i) Nylon 6.6 Fabric for c 0 =: (ii) Basic-dyeable Nylon .& 1 • 18 µmo 1 • g - 1 Yarn for Co=:

1:::,. 2. 59 µmol. g - 1 • 0 • 9 2 µmo 1 • g - 1 • 4 • 0 0 µmo 1 • g - 1 D 1.91 µmol.g- 1 o 5 • 8 8 µmo 1 • g - 1 100

90

s::O'l ·r-1s:: ·r-1 "'s ~ 80 0) ~ 0 .µ s:: 0) t) 1-1 0) PI

70

60

0 1 2 3 4

Total Incident Quanta x 10-17 (cm2 )

FIGURE 2.8 Fading Curves for C.I. Basic Red 18 in: (i} Nylon 6.6 Fabric for Co=: (ii} Basic-dyeable Nylon • 1.18 µmol_.g-~ Yarn for Co=: b. 2 .59 µmol .g- 1 • 0 • 9 2 µmo 1 • g - 1 • 4 . 0 0 µmo 1 • g - 1 D 1 • 91 µmol • g - 1 o 5. 8 8 µmol. g - 1 96 s::bl ·r-1s:: ·r-1 Id El Q) ~ Q) 94 >i 0 .µ s:: Q) 0 1-1 Q) Pl 92

0 1 2 3

Total Incident Quanta x 10-17 (cm-2 )

FIGURE 2.7 Fading Curves for C.I. Basic Blue 4 in Acrilan 16 Fabric for Co=: h. 0 • 4 9 µrno 1 . g - 1 o O. 9 8 µrnol. g - 1 . -1 .A 2 .24- µrnol.g • 3.86 µmol.g-1 80 tJl i:: ·r-1 i:: ·r-1 rd I::: Q) ~ 70 Q) >i Cl .µ i:: Q) u 1-1 Q) Al 60

0.5 1.0 1.5 2.0 2.5

FIGURE 2.6 Fading Curves for C.I. Basic Blue 4 in:

{i) Nylon 6.6 Fabric, for Co=: {ii) Basic-dyeable Nylon ... 0.51 µmol.g- 1 Yarn, for Co=: D. 0. 80 µmol.g -1. • 0 • 5 9 µmol. g - 1 • 1.11 µmol.g- 1 D 0 • 9 3 µmol. g - 1 0 1.29 µmol.g- 1 57

2.7 Determination of Initial Quantum Yields of Basic

Dye Fading

2.7.1 Measurement of Fading Rates

Fabric samples were irradiated on the rotating sample carriage for an appropriate light exposure dose, then removed from their holders and their reflectance measured immediately. The samples were then returned to the lamp for a further period of irradiation, and this procedure repeated until at least five readings were obtained for each sample. Duplicates of each dyed fabric were faded and gave extremely good agreement, with a coefficient of variation of the percent dye remaining of less than 1.5% for the fading of acrylic fabrics and better than 3% for nylon fabrics.

The concentration of dye remaining relative to the initial concentration was determined after each period of irradiation using the Kubelka-Munk relation as described in Section 2.5. The results were plotted as percent dye remaining versus the exposure dose (total incident quanta) and these plots are presented in Figures 2.6 to 2.8.

It was found that plots of the log of percent dye remaining against total incident quanta gave straight lines, with correlation coefficients from linear regression, r ~ 0.96, indicating that the fadings follow apparent first order kinetics. Figure 2.9 illustrates this first order test for C.I. Basic Red 18 on nylon 6.6 and basic-dyeable nylon. 58

The slope of this line gives the apparent first

order rate constant, [ k1, cm 2 (total incident quanta)- 1]:

k 1 = 2.303 x slope

k1 was then used to calculate the initial rate of dye

degradation:

dC dio = where Co= initial dye concentration

To obtain the initial dye concentration in terms of the number of dye molecules per cm 2 of fabric surface irradiated

required an estimation of the depth of fabric involved since

the incident light does not penetrate right through the

fabric. It was decided that an appropriate fabric thickness would be to that depth of fabric by which 95% of the incident

light is absorbed. This led to the following expression:

Rate of dye degradation,

[molecules. (total incident quanta)- 1 ]

dN = dio

where N = number of dye molecules degraded

Io = total incident quanta of incident A95% = Absorbance at which 95% light is absorbed TABLE 2.3 Apparent First Order Rate Constants of Fading and

Initial Rates of Dye Degradation Determined.

(Values calculated in terms of total incident

quanta.) dN k 1 dI 0

C 0 * [cm2 (total (dye molecules Dye/Fibre (µmol.g- 1 ) incident degraded per quanta)- 1 ] total incident quanta)

C.I.Basic Blue 4 in: Acrilan 16 0.49 3.37xl0-23 2.72xl0- 7 0.98 2.40xl0-23 l.93xl0- 7 2.24 3.35xl0-23 1. 89xl0- 7 3.86 2.33xl0-23 l.88xl0- 7

Nylon 6.6 0.51 2.40xl0-21 l.93xl0- 5 0. 80 2.0lxl0-21 l.62xl0- 5 1.11 l.98xl0-21 l.59xl0- 5 1.29 l.69xl0-21 1. 36xl0- 5

Basic-dyeable 0.59 3 .62 xl0- 2 1 2.9lxl0- 5 Nylon 0. 9 3 3 .2 8 xl0- 2 1 2.65xl0- 5

C.I.Basic Red 18 in: Nylon 6.6 1.18 7.78xl0-23 2.29xl0- 6 2.59 6. 39xl0- 2 3 l.89xl0- 6 4.00 5.50xl0-23 l.62xl0- 6 4.88 5.90xl0-23 l.74xl0- 6

Basic-dyeable 0.92 l.49xl0-22 4.29xl0- 6 Nylon 1. 91 l.33xl0-22 3.92xl0- 6

* Dye concentrations were found to vary by± 1% on Acrilan 16 and basic-dyeable nylon and by± 5% on

nylon 6.6. 59

-A i.e., Fraction of light absorbed = 1 - 10 95% = 0.95 and hence = 1. 301

NA = Avogadro's Number

k1 = apparent first order rate constant

of fading [cm2 .(total incident

quanta) -i]

E = molar extinction coefficient of

dN The values obtained for k and for each of the dye/ 1 dio fibre systems are presented in Table 2.3.

2.7.2 Determination of Quanta Absorbed by Dye

The fraction of the incident quanta that is absorbed by the dye initially (F) was calculated as follows:

>-2 1 F = = Io l

where IA = total quanta absorbed by dye initially Io = total incident quanta

>. = wavelength. The summation was carried out

at 5nm intervals over the wavelength range

of the radiation incident on the sample

i.e., >. 1 = 320nm and >-2 = 650nm

= Intensity of the incident light at wavelength >. I >. RA = Initial fractional reflectance of the dyed fabric at wavelength >. TABLE 2. 4 Fraction of the Incident Quanta that is Absorbed

by the Dyes Initially

Dye/Fibre Co F (µmol. g- 1 ) (quanta absorbed per total incident quanta)

C.I.Basic Blue 4 in:

Acrilan 16 0.49 0.538

0.98 0.616

2.24 0.713

3.86 0.784

Nylon 6.6 0.51 0.510

0. 80 0.564

1.11 0.631

1.29 0.665

Basic-dyeable 0.59 0.467

Nylon 0. 9 3 0.536

C.I.Basic Red 18 in:

Nylon 6.6 1.18 0.650

2.59 0.744

4.00 0.799

5.88 0.846

Basic-dyeable 0.92 0.731

Nylon 1. 91 0.805 60

and 1 - RA = Fraction of the incident light of

wavelength A that is absorbed

Since subsequent studies of polymer films showed that the polymers being studied have a very small absorbance above

320nm, it was assumed that all the light being absorbed was absorbed by the dye.

The required reflectances were determined by measuring the reflectance spectrum of each dyed fabric on a Beckman DK2 Recording Spectrophotometer prior to fading.

The samples were measured on a white background to ensure that all the incident light was either reflected or absorbed (none transmitted). The white tile used as the reference was calibrated against freshly smoked magnesium oxide which was used as the primary reference standard.

The values obtained for the quanta absorbed by the dye initially, as a fraction of the total incident quanta, in each of the dye/fibre systems are given in Table 2.4.

2.7.3 Calculation of Initial Quantum Yields of Basic

Dye Fading

As the calculation of the initial rate of dye degradation was based on 95% of the absorbed light, so too the value of the fraction of the incident quanta that is

absorbed by the dye is 95% of the total calculated. The

use of initial rates minimises the interference from any

absorbing fading products and means that the concentration

of dye present can be accurately determined. TABLE 2.5 Initial Quantum Yields of Basic Dye

Fading

Dye/Fibre Co (µmol.g -i)

C.I.Basic Blue 4 in:

Acrilan 16 0.49 5.3 X 10- 7

0. 9 8 3.3 X 10- 7

2.24 2.8 X 10- 7

3.86 2.5 X 10- 7

Nylon 6.6 0.51 4.0 X 10- 5

0. 80 3.0 X 10- 5

1.11 2.7 X 10-5

1.29 2.2 X 10- 5

Basic-dyeable 0.59 6.6 X 10-5

Nylon 0.93 5.2 X 10- 5

C. I .Basic Red 18 in:

Nylon 6.6 1.18 3.7 X 10-6

2.59 2.7 X 10-6

4.00 2.1 X 10-6

5.88 2.2 X 10-6

Basic-dyeable 0.92 6.3 X 10-6

Nylon 1.91 5.1 X 10- 6 61

The quantum yields obtained for each dye/fibre combination are presented in Table 2.5. The quantum yields are estimated to be accurate to± 10%.

The results clearly illustrate the trend being examined. There is a hundred-fold lower quantum yield on

Acrilan fabric compared to that on the nylons for C.I. Basic

Blue 4, while for C.I. Basic Red 18 the fading rate on

Acrilan was too slow to be detected on the time scale of these experiments. No fading was detected in the latter case after an exposure dose of 3.6 x 10 20 total incident quanta.cm- 2 which corresponds to approximately seven hours exposure, while the same dye on the polyamide substrates displayed significant fading. Also, it can be seen that both dyes display significantly higher quantum yields on basic-dyeable nylon than on the conventional nylon 6.6, but within the same order of magnitude.

The values of the quantum yields given in Table 2.5 are typical of quantum yields of dye fading quoted in the literature. For example,Irick and Boyd54 found values ranging from less than 5 x 10-6 to 1.1 x 10- 3 for various disperse dyes in 90/10 (v/v) methanol/isopropyl alcohol.

Shah andSrinivasan5 3 found a quantum yield range of 10- 4 to

10- 6 for disperse dyes on various polymer films. They attributed the low values to a combination of:

dye aggregation

low rates of diffusion of gases to and from the

dye molecules 6

5

3

• 2

0 1 2 3 4 5 6

Initial Dye Concentration (µrnol.g- 1 )

FIGURE 2.11 The Initial Quantum Yields of Fading of C.I. Basic Red 18 versus Initial Dye Concentration in: • Basic-dyeable Nylon Yarn • Nylon 6. 6 Fabric ( a} 6

5

1/) 0 4 r-1

3

2

0.4 0.6 0.8 1.0 1.2

Initial Dye Concentration (µmol.g- 1 )

(b} 5

4

r-- 0 r-1 X ,e, 3

2 0 1 2 3 4

Initial Dye Concentration (µmol.g- 1 )

FIGURE 2.10 The Initial Quantum Yields of Fading of C.I. Basic Blue 4 versus Initial Dye Concentration in: ( a) • Basic-dyeable Nylon Yarn • Nylon 6 .6 Fabric (b) .a. Acrilan 16 Fabric 62

- energy transfer from the dye through the dye­

substrate bonds

internal energy conversion

- the "cage" effect of the surrounding polymer

substrate

Datyner et al., 84 found quantum yields for reactive azo dyes

on cellulose of (1 to 3) x 10- 6 dry and (1.3 to 5.4) x 10-6 wet.

Figures 2.10 and 2.11 show plots of the quantum

yields against dye concentration. It can be seen that there

is a definite concentration dependence of quantum yield, with the quantum yield decreasing with increasing dye

concentration. This phenomenon has frequently been observed,

and the usual explanation is that increasing dye aggregation

accompanies increasing concentration and that this causes

the slower fading (see Section 1.4.2). However, it is

unlikely that dye aggregation explains the trends observed here as dye concentrations are quite low (in the range 0.02

to 0.25% dyeing) which means that there would probably be

sufficient dye sites available for each dye molecule to be

individually adsorbed onto a dye site thus limiting

aggregation. On the other hand, since dyeing would only take

place in the amorphous regions of the polymers it is possible

that the dye molecules may be close enough to allow self­

quenching to occur by a resonance energy transfer mechanism

(see Section 1.4.2) and the probability of this occurring

would increase with increasing concentration of dye. 63

In conclusion, the initial quantum yields of basic dye fading determined demonstrate a substantial substrate effect on the fading of basic dyes. A difference of at least two orders of magnitude was observed between the quantum yields on an acrylic polymer and on the non-acrylic polymer studied, with a concentration effect also being observed. It would appear therefore that there exists some major physical or chemical difference between the two types of substrate that gives rise to such large differences in the lightfastness of basic dyes present in them. The remainder of this thesis therefore deals with the identifi­ cation of the physical and/or chemical factors that

influence basic dye lightfastness. 64

CHAPTER 3

THE EFFECT OF THE PHYSICAL STATE OF

THE SUBSTRATE ON BASIC DYE FADING

3.1 Introduction

The influence of the physical state of the polymer

substrate on the fading of the three basic dyes being

studied was investigated by studying the rate of fading of

dyed polymer films over a range of temperatures. Film

rather than fabric substrates were examined as the size of

a fabric sample necessary for a satisfactory measurement of

its reflectance could not be accommodated in the experimental

system required for this work. The polyamide used here was nylon 6, general formula [NH-(CH 2 ) 5 -CO-] , as nylon 6 .6 was n not available in film form. The dyes on nylon 6 film display

a relatively low lightfastness similar to that on nylon 6.6

fabric.

Varying the temperature of the polymer alters its

rigidity by changing the internal motion of the polymer

chains. A corresponding change in the "free volume" of the

polymer also occurs. In particular, the rate at which these properties change with temperature will increase at the

glass transition temperature of the polymer system i.e.,

the temperature at which the polymer changes from a "glassy"

to a viscoelastic state. Consequently it was of particular

interest to examine the fading at temperatures both above

and below the glass transition temperature of each polymer. 65

The physical state of the substrate also affects its permeability to vapours. Accordingly the fading of basic dyes in various substrates in different atmospheres containing oxygen and moisture was examined.

3.2 Polymer Film Preparation

Transparent polymer films made from Acrilan 16, nylon 6 and poly(vinyl alcohol) were used in this section of the work. Unlike dye fading on fabric, the dye fading rates of very small film samples can be determined since film size is only limited by the slit size of the spectro­ photometer used for the measurement of its absorbance (in this case approximately 1 x 10mm). The small sample size used also ensured that the variation in its temperature during irradiation could be kept to a minimum.

The initial concentration of the dyes present in the films (C 0 ,µmol.g- 1 ) was calculated from the initial absorbance of the dye at the wavelength of its maximum absorbance (Ao), the extinction coefficient of the dye at this wavelength (E, cm 2mol- 1 ), the thickness of the film

(£, cm) and the density of the polymer (p, gml- 1 ):

Co = X

The film absorbance was measured in a number of places using a Perkin-Elmer Coleman 55 single beam spectrophotometer which was found to be accurate to ±.002A, and the values obtained averaged. To determine the 66

absorbance due to the dye, the absorbance due to the film at the wavelength in question was subtracted from the total absorbance measured.

The extinction coefficients used were those determined previously in aqueous solution (Section 2.3).

This approach is justified as the absorbance spectra of the dyes in the polymer substrates are not significantly different from their spectra in aqueous solution. The concentrations of dye in the films were adjusted to give an initial maximum absorbance of approximately 1.

Film thicknesses were measured on a Minicom High

Speed Electric Micrometer.

3.2.1 Nylon 6

Extruded undrawn nylon 6 film, nominally 30µ thick and free of delustrant and plasticiser, was obtained from the Nihon Rayon Co. Ltd., of Japan. The film was dyed at the boil in a dyebath containing only dye, with no additives.

3.2.2 Poly(vinyl alcohol)

Poly(vinyl alcohol), (PVA), repeat unit (CH2-CH-) I n OH was selected as being a suitable solid model substrate. It is photochemically stable, transparent above 250nm and is easily prepared into films by casting from aqueous solution.

Consequently water soluble additives, such as dyes, can be incorporated into the substrate on casting, thus ensuring an homogeneous distribution of additive throughout the film. 67

A 4% (w/w) solution of PVA (Matheson, Coleman and

Bell, 99% hydrolysed) containing the appropriate amount of

dye was cast by pipette onto glass microscope slides placed

on specially levelled preparation tables and allowed to dry.

4 mls of a 4% solution cast onto a 37.5 x 50mm slide gave a

film of thickness 55 ± 10µ. The edges of the film, which were thicker than the central portion, were discarded.

3.2.3 Acrilan 16

It was not possible to obtain Acrilan 16 (or any other basic-dyeable acrylic) in film form. Consequently

films had to be prepared by dissolving the Acrilan 16 fibre

in a suitable solvent and casting the solution to form a

film. In fact, a great many difficulties were encountered

in producing transparent acrylic films. The following preparation technique was found to give the most satisfactory

results but unfortunately, not consistently:

1) Acrilan 16 fibre was dissolved in freshly distilled

dimethylformamide to produce a 4% (w/w) solution.

2) This solution was centrifuged using a Sorvall Superspeed

RC2-B Automatic Refrigerated Centrifuge to remove the

titanium dioxide delustrant, and the required amount

of dye powder added. (Attempts to dye clear film

proved unsatisfactory as the films turned opaque during

the dyeing process.)

3) 4 mls of the solution were cast onto a glass microscope

slide (50 x 75mm) placed in a levelled vacuum oven

(Virtis Unitrap Freeze-Dryer). 0 4) The oven was heated to 60 C under a vacuum of 68

approximately 1mm Hg for one hour, in order to evaporate

the solvent.

Films of approximately 65µ thickness were obtained, from which the thicker edge portion was discarded.

Difficulties persisted with producing clear uniform acrylic films until finally all the films being produced were unusable due to uneven ripples over their entire surface. The uneven thickness of the film resulted in a variable absorbance across the film as well as causing reflections of incident light at the surface. Two factors were considered to contribute to the production of these unsuitable films, viz.,

1) the adhesion of the film to the glass slide, and

2) the rate of removal of the solvent.

A skin was observed to form over the surface of the film as the dimethylformamide evaporated and this could have contributed to the problem.

Numerous variations in method were tried in an attempt to overcome these factors and included variations in the following conditions:

1) the temperature of drying, 2) the temperature program - both constant and variable

temperatures were tried,

3) the level of the vacuum applied, 69

4) the method of cleansing the slides,

5) the age of the acrylic solution used,

6) the amount of oxygen present in the acrylic solution,

7) the type of surface on which the films were cast, e.g.,

glass, stainless steel, silicone oil.

However, none of these measures were found to solve the problem and eventually the attempt to produce satisfactory films had to be abandoned. Fortunately sufficient clear films or portions of films had been produced to enable the temperature work to proceed, although insufficient films dyed with C.I. Basic Violet 14 meant that the effect of different atmospheres and concentrations could not be examined for this dye on Acrilan 16.

3. 3 Measurement of Glass Transition Temperatures

At the glass transition temperature, changes occur in the temperature dependence of various properties of the polymer that are influenced by its rigidity. These properties include its specific volume, specific heat, refractive index, elastic modulus and permeability.

Therefore, by measuring one of these properties while continuously varying the temperature an indication of the glass transition temperature can be obtained. Unfortunately there are numerous difficulties associated with such measurements:

The changes occurring at T may be too small to detect. 1) g This can be caused by a high degree of crystallinity since the physical changes are essentially associated with the 70 non-crystalline regions. Alternatively, the particular physical property being monitored may merely show only a

small change at T . g 2) The experimental conditions under which the determination

is carried out can influence the T observed including both g the temperature at which it occurs and the extent of the

associated property changes. Critical factors include the

rate at which the temperature is varied, the size of the

sample and its previous thermal history. This means, for example, that the reproducibility of a T determination g cannot necessarily be determined by repeating the

determination on the same sample, especially if it has been heated beyond its melting point during the first measurement.

3) Foreign substances present in the polymer may act as plasticisers and therefore depress T • Such substances g include various agents which may have been added to the polymer to give it specific properties, and water vapour

from the atmosphere. The latter means that the atmospheric humidity conditions during measurement can affect the Tg

found.

3.3.1 Experimental Technique

Differential thermal analysis (DTA), which is based

on the change in the specific heat of a substance with

temperature, was used here to determine the glass transition

temperatures of the polymers. This technique consists of measuring the temperature difference that occurs between

the substance under investigation and a thermally inert

substance when both are subjected to identical heating at a

The detection of T relies carefully controlled slow rate. g Exo Crystallisation { rOxidation

~ Glass b.T Transition

Decomposition Melting

Endo

Temperature (T)

FIGURE 3.1 Schematic DTA Curve of a typical polymer. 104 71

on the fact that a glass transition involves an endothermic

phase change and will therefore be seen as a change in the

baseline of a recording of 6T versus T. A schematic DTA

curve of a typical polymer, including a glass transition is

given in Figure 3.1. It should be noted that first order

changes such as melting and crystallisation give rise to

quite sharp, distinct peaks, whereas the second order glass

transition is not nearly as easy to distinguish as it

involves only a shift in the baseline which, in addition, may not be a large one.

Two different instruments were used which employed

slightly different techniques. The first was a du Pont 900

Thermal Analyzer using a conventional DTA cell, and the

second, a later and a more sophisticated and sensitive

instrument, was a du Pont 990 Thermal Analyzer which

employed a differential scanning calorimetry (DSC) cell.

The essential relevant difference between the two was in

the sample mounting and temperature measurement techniques.

In the former instrument, the samples are packed into glass

tubes, a thermocouple inserted and both mounted in a heating

block. In the latter, samples are pressed into small

aluminium discs and sealed to form a pan. These pans were

then placed on top of a constantan disc which acts as one

element of the temperature measuring thermoelectric functions

as well as the primary means of heat transfer to the sample

and reference positions.

The most desirable sample form to use was the same

as that to be used in the fading experiments i.e., polymer 72

in film form. Using the pan mounting technique this could

easily be achieved by using cut snippets of polymer film.

However, with the glass tube mounting method close contact between the sample and the thermocouple was more difficult

to achieve and other types of samples were tried, including polymer in fibre form and polymer powder obtained by

grinding film or fibre at liquid nitrogen temperatures in

a Grindex ball mill (Research and Industrial Instruments

Co. England). Both types of sample can be packed more

tightly into the sample tubes to obtain better contact and

therefore improved sensitivity. However, care does need to be taken in the interpretation of results from ground

samples as sufficient heat can be generated in the grinding process to change the sample history. 105

All samples were dried before measurement in dry nitrogen. Initially undyed samples were examined with the

intention of subsequently comparing the T's of dyed and g undyed polymers in order to determine any plasticising action

that the dyes may have on the polymers.

3.3.2 Results

Acrilan 16, nylon 6 and poly(vinyl alcohol) in

undyed film and fibre form were examined on both instruments

using a number of different instrument conditions. Changes

in the baseline indicating possible second-order phase

changes were observed in some samples e.g., for Acrilan 16

there was some evidence of a transition around 80 0 C, and

for PVA at approximately 20 0 C and 50 0 to 60 0 C. However, the

transitions could not be unequivocally distinguished from TABLE 3 .2 The Variation of the Glass Transition

Temperature of Nylon 6 with Water Content. 108

RH (%) Water Content (%W/W) T (±1°C) g

dried over silica gel 0. 35 94

12 1.17 77

33 1.99 56

44 2.70 45

55 3.47 43

66 4.45 40

86 6.61 23

97 10. 3 -6 TABLE 3.1 Analysis of Key Operational Parameters in

Differential Thermal Analysis 105

Sample Size:

Large: - Useful for detecting low level transitions

- Curve peaks are broad

- Low resolution and temperature accuracy

- Requires slow heating rate

Small: - Good resolution of curve peaks

- Peaks are sharp

- Permits fast heating rate

Heating Rate:

Fast: - Increases sensitivity

- Decreases resolution

- Decreases temperature accuracy

Maximum Maximum

Resolution Sensitivity

Sample Size Small Large

Heating Rate Slow Fast

Surface/Volume of Sample Large Small 1

~T

0 50 100

Temperature (°C)

FIGURE 3.2 Experimental DTA Curves obtained for: 1 and 2: Poly(vinyl alcohol) on the du Pont 900 Thermal Analyzer. 1 Ground film sample 2 Commercial Powder sample

3 and 4: Acrilan 16 on the du Pont 990 Thermal Analyzer - both cut film samples. 73 the variations due to instrument noise in the recordings and further, they were not sufficiently reproducible.

Figure 3.2 shows two of the DTA curves obtained for two samples of Acrilan 16 film on the du Pont 990 Thermal

Analyzer and for PVA on the du Pont 900 Thermal Analyzer, which are typical of the curves obtained. The difficulties with instrument noise and lack of reproducibility are readily apparent.

This lack of success could be due to any of the difficulties mentioned earlier. It is possible that some of the polymers were too crystalline for sufficiently large

T effects to be exhibited. Alternatively the various g experimental conditions employed may not have been optimum for the detection of T e.g., the correct combination of g sample size and heating rate in particular may not have been found. An idea of the compromises faced experimentally can be obtained from the analysis of key operational parameters given in Table 3.1.

Since this research did not yield the required glass transition temperatures, literature values were used instead to determine the approximate temperature ranges required so that fading measurements covered temperatures both above and below the T of the polymers. The following values were g representative of those reported:

Polyacrylonitrile 87° to 95°c, and 140°c 1 0 6

Poly(vinyl alcohol) asoc i o 6 , 71 oc i o 7

Nylon 6 70°c (lowered by moisture)1 0 6 -6° to 94°c depending on water content (see Table 3.2) Lamps

Unit

BE

Tube

FL

Sample

at

E •

Ice-water

G.

Reference

Film

Control

Vacuum

Heater

Films

~

Dyed

Fading

Atmospheres.

Thermocouple

r--~------4Voltmeter

for

Chamber

Irradiation

used

r------1

a------+-1--+--

Different

System

under

Heater

and

l

I

I

I

I

I

l

I

I

for

ice

room)

used

Experimental

nitrogen

or

Exchanger

the

below

..__._~..__,

temperatures

Temperatures

t~I I

(only &.------

r-

Heat

of

liquid

ethanol/dry

I

I

I

I

I I

I

I

I

_J

used

Different

Diagram

moist

lf//l'A

____

l

I

(only

,----

for

Humidifier

atmospheres)

Schematic

Flow

Meter

Water

Scinter

3

3.

Gas

FIGURE 74

The discrepancies between the literature values reported

underline again the difficulties and uncertainties associated with the measurement of T and the number of factors which g can affect the measured T of a polymer. g

Based on these values the range of temperatures

selected for use were:

Acrilan 16 o0 c to 140°c Poly(vinyl alcohol) -4o 0 c to 95°c Nylon 6 -40°c to 90°c

3.4 Irradiation Technique

3.4.1 Method of Mounting and Heating the Film Sample

The equipment used was part of the Varian Intracavity

Variable Temperature Apparatus as used in a V4531 Reflection

Cavity of an electron spin resonance spectrometer. A

schematic diagram of the experimental set-up is given in

Figure 3. 3.

The film sample to be faded was placed around the

inside of a quartz vacuum tube which was positioned in the

centre of the irradiation chamber. This tube was connected

via a heater and a heat exchanger to a gas bottle, and the

various temperatures were obtained by heating or cooling

the gas which then passed over the sample. The heater

temperature was controlled by means of a Varian EPR Heater

Control Unit. This was used in conjunction with liquid

nitrogen or an ethanol-dry ice mixture in the heat exchanger 75

to obtain temperatures below room temperature, and used on

its own to achieve temperatures above room temperature. The

temperatures obtained were measured by means of a copper­

constantan thermocouple which was placed in the tube near

the top of the film sample and connected to a millivoltmeter

and an ice-water reference. The Hewlett-Packard 34703A

DVC/DCA/n meter used was highly accurate and was found to

give readings which corresponded to a temperature accuracy

of ±½0 c. The experimental set-up used was found to give a maximum variation in temperature over the sample of ±5°c.

Both dry nitrogen and dry oxygen gases were used

and their flow rate, controlled by a Fischer and Porter

Variable Area Flow Meter (Model lOA 3135M.35R2110), was set

at 2. 5 .Q, /min. To obtain temperatures below room temperature

for oxygen, an ethanol-dry ice slurry rather than liquid

nitrogen had to be used in the heat exchanger as liquid

nitrogen liquefied the oxygen. Water vapour was introduced

by bubbling the gas through liquid water. By measuring the

amount of water evaporated by a known volume of gas, as

calculated from the flow rate for a given time, it was

found that the gas became completely saturated with moisture.

The moist gases were only used at room temperature as it was

considered that sufficient information on the effect of

moisture could be obtained from this one temperature,

especially since it is the temperature of most relevance in

practice. 10

...- 8 I C) Q) Ul

<'I I 6 6 Ill .µ s:: Ill ~ 01

~... I 4 0 r-1 X :>-t .µ ·r-1 Ul s:: 2 Q) .µ s:: H

0

300 350 400 450 500

Wavelength (nm)

FIGURE 3. 4 Emission Spectrum of Radiation from sixteen

G.E. FL8E lamps incident on sample. 76

3.4.2 Method of Irradiation

An Oliphant Irradiation Chamber, Model PCR-128W which can house up to sixteen 8 watt fluorescent tubes was used as the irradiation chamber. The chamber is cylindrical with 12 inch long lamps attached vertically at equal spacings

around a reflecting wall. The vacuum tube containing the

film sample to be irradiated was mounted vertically in the

centre of the chamber, with the sample situated approximately halfway down the length of the lamps to ensure uniform

irradiation of the sample.

The lamps used were General Electric FL8E fluorescent

lamps. Their spectral distribution was determined in the

same manner as for the Xenon lamp system, using the same

actinometry method. The emission spectrum of the lamps thus

obtained is presented in Figure 3.4. It can be seen that these lamps have a much higher uv component than the Xenon

lamp system, with virtually all their emission below 450nm.

This contrasts significantly with the Xenon lamp system which has the majority of its output above 400nm, with quite

a large portion above 500nm. Obviously results obtained

from the two different lamp systems could not be directly

compared with each other unless an "action spectrum" of the particular dye's fading was measured to determine the

relative efficiency of each wavelength of light in producing

fading. However such comparisons were not necessary in this work since its purpose was to compare the effect of different

substrates on dye fading for a particular illuminant.

Despite the low power output of the fluorescent 77

lamps compared with the Xenon lamp's output (128 W compared

to 1600 W) very similar fading times were found to be

necessary. This is probably due to a much greater portion

of the emitted energy reaching the sample in the Oliphant

system and due to its much higher uv component which has been

shown in the literature to be the most effective spectral

region in producing dye fading (Section 1.4.1).

It had been found that the Oliphant lamps emit

virtually constant radiant flux during their lifetime.

Consequently it was not necessary to continuously monitor

the flux as was the case with the Xenon lamp system. The

fading intervals were therefore simply determined by the

length of time for which the lamps were switched on. There was a small error associated with this technique as all the

lamps did not always switch straight on. However the error

in timing involved was only approximately 5 seconds for

only a few of the twelve lamps at the one time, and this in

terms of irradiation times of between 2 and 60 minutes for

twelve lamps was considered to be negligible.

3.5 Determination of the Initial Quantum Yields of Basic

Dye Fading

3.5.1 Measurement of Fading Rates

Each film sample was irradiated for four consecutive

time intervals. The progress of fading was followed by

measuring the absorbance of the dyed film at the wavelength

of the dye's maximum absorbance after each period of

irradiation. In this manner a dye's fading curve was 78 obtained in the form of a plot showing the proportion of the original dye remaining versus the total irradiation time.

Prior to the first irradiation each sample was subjected to a flow of gas at the temperature and humidity at which irradiation was to take place to allow the dyed film to come to equilibrium with the desired conditions before its initial concentration was determined. This allowed any shrinkage of the film due to either the heating or drying effects of the passing gases. This phenomenon was found to occur with both of the cast films, PVA and

Acrilan, and resulted in an increase in the thickness of the film and a corresponding increase in the measured absorbance. The effect also increased with increasing temperature.

Following this pretreatment, the absorbance (at the wavelength of maximum absorbance) and the thickness of the dyed film were measured and these values used to calculate the initial dye concentration in the film as described in Section 3.2. The absorbance of the dye obtained here was also taken as the initial absorbance for the determination of the fading curve. In all cases, the absorbance of the dye was obtained by subtracting from the absorbance of the dyed film the corresponding absorbance of the undyed film, which was for all films relatively smal 1 ( A ~ 0 . 0 4) •

In order to allow the film samples to equilibrate to the conditions before each period of irradiation, the 2.0

1.95 30°c

61 °C 1.90

1.85 -t,'l s::: 97°c •r-i s::: •r-i rd s Q) 1.80 P::

Q) ~ 0 .µ s::: 1.75 Q) 0 1-1 Q) r:i.. - 1.70 t,'l 0 ,-f

1.65

1.60

138°C 1.55

0.5 1.0 1.5 2.0 2.5

FIGURE 3 .6 Semi-log plot of the Fading Curves, given in

Figure 3.5, for C.I. Basic Blue 4 in Acrilan 16

Film at various temperatures. 100

-10°C 90 30°c

61°C

80

t,'l s:: ·r-1 s:: ·r-1 m E: 70 Q) 97°c p:;

Q) >t 0 .µ s:: Q) 60 u 1-1 Q) Ill

50 ll8°C

40 138°C

0.5 1.0 1.5 2.0 2.5

FIGURE 3 .5 Fading Curves for C.I. Basic Blue 4 in

Acrilan 16 Film at various temperatures. 79 films were again subjected to the appropriately conditioned gas flow for between 10 and 30 minutes (depending on the temperature) prior to each irradiation.

In order to completely encircle the vacuum tube the film samples had to be approximately 16mm wide and for ease of measurement on the spectrophotometer at least 18mm long.

The technique developed to measure the absorbance of a film involved mounting the film at a fixed position between two glass microscope slides in order to flatten it. A number of absorbance readings were taken across the sample and averaged. The results were then expressed as follows:

Percent Dye Remaining at time t

absorbance of dye at time t = X 100% absorbance of dye at zero time

where absorbance of dye = absorbance of dyed film

absorbance of undyed film

All absorbances were measured at the wavelength of maximum absorbance of the dye.

A typical set of fading curves, for C.I. Basic Blue

4 in Acrilan 16 film, is shown in Figure 3.5. The same results are presented in Figure 3.6 plotted on a semi-log scale to demonstrate the test for the existence of first order fading kinetics. It was found that most of the fading curves followed apparent first order kinetics although a few exhibited apparent zero order kinetics. The latter were 80

assumed to be the initial apparent linear portion of a first

order reaction. This is consistent with the results obtained

for fabric fading presented in Chapter 2.

The apparent first order rate constants, k 1 , were

determined by linear regression on the log (percent dye

remaining) versus total incident quanta curve as calculated

on a Hewlett-Packard HP25 Calculator. The slope of the line

of best fit (a 1 ) determined in this way was used to calculate

the apparent first order rate constant (k 1 ): k 1 = 2.303 ai­

The square of the correlation coefficients (r2 ) obtained were all better than 0.9 with most around 0.99, indicating

the extremely good fit of the results to first order plots

and better than 99% confidence in the slope of the line.

The initial rate of dye degradation,~~ (molecules

of dye degraded cm- 2 sec- 1 ), was then calculated from the

apparent first order rate constant as follows:

dN dt = = where N = dye molecules degraded per cm 2

t = time

k1 = apparent first order rate constant

Co = initial concentration of dye

Ao = initial maximum absorbance of dye

NA = Avogadro's Number

E: = extinction coefficient of dye 81

3.5.2 Measurement of Quanta Absorbed

Since the film curves around the inner wall of the

tube during irradiation the sample receives light from both

sides. This must be taken into consideration when

calculating the amount of energy absorbed by the film:

where I 0 = incident intensity

IT = transmitted

intensity

Amount of light absorbed by the

sample from outside the tube where A = absorbance of the film

Amount of light transmitted

through the sample = = I 0 I 0 (1 - 10-A) = Io X 10-A = light incident on

the back of film

Total amount of light absorbed by the film

= Io(l - 10-A) + Io X 10-A (1 - 10-A) = I o ( 1 - 10-A) (1 + 10-A) = I o (1 - 10-2A)

Thus the amount of energy absorbed in this situation is

equal to a film of double the thickness (and therefore double

the absorbance) absorbing energy from only one side. 82

The initial rate of quanta absorption by the dye dia (quanta absorbed cm- 2 sec- 1 ) was calculated as follows: dt

dI >..=520nm _..£! dt = I >..=275nm

where: Ia = quanta absorbed by dye per cm 2

t = time (sec.)

>.. = wavelength. The wavelength range used was

275nm to 520nm, which covered the total

emission of the fluorescent lamps used.

The summation was carried out at 5nm

intervals over this range.

= intensity of the incident light at

wavelength>.. (quanta cm- 2 sec- 1 ). = absorbance due to the dye only at wavelength >... = total absorbance of the dyed film at wavelength >... = + where: = absorbance due to polymer film at wavelength>... T 1 - l0-2 A>.. = fraction of incident light that is absorbed

by the dyed film.

The absorbance spectra of each dyed and undyed film was measured on a Varian Techtron 634 double beam spectro­ photometer which has an absorbance accuracy of better than ±.002A (at l.OA) and a wavelength accuracy of better than O. 5nm. 83

To facilitate the calculation of quanta absorbed, normalised spectra of each type of dyed and undyed film were stored on a cassette tape and a calculator program devised to use this stored data. This meant that only the maximum initial absorbance of a dyed film and its thickness

(after pretreatment with the appropriately conditioned gas but prior to irradiation) had to be fed into the calculator in order to calculate the initial rate of quanta absorption by the dye. A Hewlett-Packard 9821A Calculator was used for this purpose.

3.5.3 Calculation of the Initial Quantum Yields of

Dye Fading

The initial quantum yields of dye fading(~) were calculated as follows:

dN dt ~ =

3.6 Experimental Results Each of the three dyes being studied, c.r. Basic Blue 4, c.r. Basic Red 18 and C.I. Basic Violet 14 was examined in each of the three films, poly(vinyl alcohol), nylon 6 and Acrilan 16. Where sufficient samples of nylon

6 and Acrilan 16 films were available, dry and wet nitrogen, and dry and wet oxygen atmospheres were used in order to concurrently determine the effects of oxygen and moisture on fading. With PVA films only a dry nitrogen atmosphere was used since PVA films are relatively impermeable to oxygen and are soluble in water. Also, in some cases two 1.5

1.4

1.3

II)- 0 r-1 :< 1.2 -,e, O'I 0 r-1

1.1

1.0

0.9

FIGURE 3 .14 Arrhenius Plot of the Initial Quantum Yield of Fading of C.I. Basic Violet 14 in Acrilan 16 Film for an initial dye concentration of 2.60 µmol.g- 1 in dry nitrogen. 1.2

1.0

1/)- 0 .-I >< 0.8 -,e, O"I 0 .-I

0.6

0.4

0.2

3.0 3.5 4.0 1 T

FIGURE 3.13 Arrhenius plots of the Initial Quantum Yields of Fading of C.I. Basic Violet 14 in Nylon 6 Film for an initial dye concentration of 2.63 µmol.g- 1 • in dry nitrogen o in dry oxygen 1.6

1.4

1.2

II) 0 r-1 >< -,e, 1.0 tJ'I 0 r-1

0.8 •

0.6

0.4

3.0 4.0 1 X T

FIGURE 3.12 Arrhenius plots of the Initial Quantum Yield of Fading of C.I. Basic Violet 18 in Poly(vinyl alcohol) ·Film in dry nitrogen for initial dye concentrations of: • 2 • 4 2 µmo 1 . g - 1 • 1 • 2 8 µmo 1 • g - 1 1.6

1.4

1.2 "'- 0 .-I >< -,e, 1.0 O'I 0 .-I

0.8

0.6

0.4

2.5 3.5 1 X T

FIGURE 3 .·11 Arrhenius plots of the Initial Quantum Yield of Fading of C.I. Basic Red 18 in Acrilan 16 Film for initial dye concentrations of: 3.29 µmol.g -1 , in dry nitrogen • ' -1 0 3.29 µmol.g , in dry oxygen • 1.62 µmol.g -1 , in dry nitrogen 1.8

1.6

1.4

1.2 - "'0 r-1

~ ,e, 1.0

t,'I 0 r-1 0.8

· 0.6

0.4

0.2

3.0 3.5 4.0 1 T X

FIGURE 3.10 Arrhenius Plots of the Initial Quantum Yield of Fading of C.I. Basic Red 18 in Nylon 6 Film for initial dye concentrations of:

• 5 .18 µmol. g - 1 , in dry nitrogen o 5 .18 µmol. g- 1 , in dry oxygen • 2 • 9 4 µmol. g - 1 , in dry nitrogen and in Poly(vinyl alcohol) Film for initial dye concentrations of: .._ 4. 38 µmol. g -1 , in dry nitrogen .6. 2 .66 µmol.g- 1 , in dry nitrogen 1.6

1.4

1.2

3- 0 r-1 1.0 >< -,e, t,'I 0 r-1 0.8

0.6

0.4

FIGURE 3 .9 Arrhenius Plots of the Initial Quantum Yield of Fading of C.I. Basic Blue 4 in Acrilan 16 Film for initial dye concentrations of: • 1.21 µmol.g -1 , in dry nitrogen -1 0 1.21 µmol.g , in dry oxygen • 2.11 µmol.g -1 , in dry nitrogen 1.8

1.6

.;2-- 0 r-1 1.4 ~ -,e, tJ'I 0 r-1

1.2

1.0

0

3.0 4.0 1 X T

FIGURE 3. 8 Arrhenius Plots of the Initial Quantum Yield of Fading of C.I. Basic Blue 4 in Nylon 6 Film for Initial Dye Concentrations of: . . -1 • 1.39 µmol.g , in dry nitrogen o 1.39 µmol.g -1 , in dry oxygen . -1 • 2.60 µmol.g , in dry nitrogen 1.6

1.4

3 0 r-i >< 1.2 -,e, tJ'I 0 r-i

1.0

0.8

• 3.0 4.0 1 X T

FIGURE 3. 7 Arrhenius Plots of the Initial Quantum Yield of Fading of C.I. Basic Blue 4 in Poly(vinyl alcohol) Film in dry nitrogen for Initial Dye Concentrations of: D 1. 31 µmol. g - 1 • 2 • 9 3 µmo 1 . g - 1 84

different concentrations of dye were studied in order to

clarify the effect of dye concentration on fading.

3.6.1 The Effect of Temperature

The results obtained are graphed in the form of an

Arrhenius plot i.e., the logarithm of the initial quantum yield of dye degradation (log 10 ~) versus the reciprocal of 1 the absolute temperature (T, K- 1). In this form the apparent

activation energies of fading (E) can be calculated from the a slope of the graph, and any change in the apparent activation energy with temperatures is readily observed:

E a ~ Aexp ( Arrhenius Equation = - RT

E a log10~ = + constant 2. 30 3RT

and hence E 2. 30 3 X R x slope a = -

The graphs obtained for each dye/film combination are shown

in Figures 3.7 to 3.14. The quantum yields determined are 1 accurate to ±10% and T to ±1.5%.

It can be seen that some of the Arrhenius plots are

straight lines, implying a constant activation energy while

others are curves indicating that a change in activation energy occurs with temperature. For a curved plot it is

assumed that two activated processes are operating, and

their activation energies are estimated graphically from the

limiting slopes of the curve as indicated in the following

schematic diagram: TABLE 3.3 Inflection Temperatures and Activation Energies

of Dye Fading in Dry Atmospheres, determined

from Experimental Arrhenius Plots.

Dye/Film Co Atm. t. E (kJmol - 1 } 1. a ( µmol. g - 1 } (OC} >t.

C.I.Basic Blue 4 in:

Nylon 6 1.39 N2 9. 8 2.60 N2 10.5 1.39 02 6.9

Poly(vinyl alcohol} 1.31 N2 36 21 4.7 2.93 N2 31 25 7.1

Acrilan 16 1.21 lh 65 33 6.9 2.11 N2 69 25 4.5 1.21 02 7.3

C.I.Basic Red 18 in: Nylon 6 2.94 N2 7.4 5.18 N2 7.6 5.18 02 2.8

Poly(vinyl alcohol} 2.66 N2 13 15 4.8 4.38 N2 ]

Acrilan 16 1.62 N2 85 54 8.5 3.29 N2 97 66 5.4 3.29 02 76 34 4.4

C.I.Basic Violet 14 in: Nylon 6 2.63 N2 9 19 5.1 2. 6 3 02 11.2

Poly(vinyl alcohol} 1.28 N2 12 23 0.7 2.42 N2 24 34 4.1

Acrilan 16 2.60 N2 8.4 85

log~

t. 1 i T

The point of inflection of the curve (t., 0 c) is determined i as the intersection of the two lines as shown. The greater the difference between the two activation energies the sharper will be the inflection seen in the curve.

Obviously the errors involved in the determination of the activation energies and the inflection temperatures will depend largely on the shape of the curve. Two well defined slope regions will give the more accurate values.

The accuracy of the inflection temperature is also affected by the nature of the plot itself since a small error in

~(K- 1 ) can represent quite a large error when converted to T . 0 t . in c. The maximum errors found were ±7 kJ mol- 1 in the i activation energies, and ±16°c in the inflection temperature, but most errors were approximately ±5 kJ mol- 1 and ±5°c respectively. Table 3.3 lists the activation energies and inflection temperatures determined for each dye/film/ atmosphere combination.

The fact that dye fadings exhibit an activation energy at all implies that a non-photochemical process is the rate-determining step, as purely photochemical reactions have essentially no activation energy because they are 86

activated by light rather than heat. This rate-controlling

step may involve a diffusion or a conformation process as both would be affected by the rigidity of the substrate and

therefore by temperature. The species which could be involved

include the excited dye molecule, partially degraded dye

fragments or reactive species within the substrate such as oxygen and water. The diffusion of partially degraded dye

fragments away from each other is often necessary to prevent

recombination reactions occurring, and species such as oxygen

and moisture may play either a reactive or a quenching role

in fading. Various conformational changes of the dye may

facilitate certain reactions and the ability of the dye to

alter its conformation will be dependent on the amount of

free volume available and therefore also be dependent on substrate rigidity.

The fact that the apparent activation energy of

fading in some cases alters with temperature would seem to

imply that a change in the substrate rigidity is influencing

fading. It is probable that the points of inflection in the

Arrhenius plots correspond to structural changes within the polymer similar to, or the same as, the glass transitions.

At temperatures above this transition temperature polymer

chain mobility and therefore free volume increase more

rapidly with increasing temperature than they do below T g where the polymer is essentially in a glassy state. This may mean that the diffusion of some species held immobile below the transition temperature becomes possible above this

temperature, or that rates of diffusion of species already mobile below the transition temperature increase more rapidly 87 with temperature above the transition temperature.

Consequently the increases in activation energies above the transition temperatures that are observed may be due to either the introduction of a new mechanism with a higher activation energy, or due to an increase in the temperature sensitivity of the existing mechanism, above the transition temperature.

The actual temperature at which an inflection is observed will depend on the amount of free volume necessary for the diffusion or mobility of the species involved in the rate-controlling step of the reaction. This in turn will depend on the dye fading mechanism(s) involved as this will determine either the size of the diffusing species or the amount of movement necessary for the required confor­ mational change. The temperature at which this free volume becomes available in the polymer will be closely related to the glass transition temperature of the polymer. In addition, the presence of dye in the polymer is likely to have a plasticising effect on the polymer which also means that each dye will probably lower the glass transition temperature to a different extent. Consequently the inflection temperature observed is characteristic of each particular dye-film system. This is seen to be the case from Table 3.3.

The absence of an inflection in a number of the

Arrhenius plots including C.I. Basic Blue 4 and C.I. Basic

Red 18 in nylon, C.I. Basic Violet 14 in Acrilan and various

fadings in an oxygen atmosphere may be due to one of the 88 following reasons:

(1) only one fading mechanism that is not inhibited by the rigidity of the polymer either above or below T operating. g Such a mechanism may be an intramolecular one that is not affected by conformational or steric effects, or a mechanism involving the diffusion of relatively small species.

(2) where oxygen atmospheres give rise to straight line

Arrhenius plots the presence of oxygen may be quenching the second mechanism that would operate in its absence. This phenomenon is seen clearly for C.I. Basic Blue 4 in Acrilan and C.I. Basic Violet 14 in nylon where the plots in dry nitrogen do show an inflection.

(3) the inflection temperature being outside the temperature range examined.

The fading of two concentrations of dye, one being approximately double the other were examined for a number of the dye-film systems. It can be seen from Figures 3.7 to 3.14 that in general there is a decrease in quantum yield at the higher concentration which is the same trend as noted for the fading of the dyed fabrics. In nearly all cases the differences between the quantum yields of the two concentrations are not large, with the exception of C.I.

Basic Violet 14 in PVA where the difference is quite marked.

In all cases however the shapes of the two Arrhenius plots are similar and hence the activation energies and inflection temperatures are also similar for the two dye concentrations.

As it is the effect of substrate on basic dye fading that is of most interest here, the results have been regrouped 1.6

1.4

1.2

II)- 0 ..-I 1.0 X -,e, O'I 0 ..-I 0.8

0.6

0.4

0.2 2.5 3.0 4.0 1 T X

FIGURE 3.17 Arrhenius plots of the Initial Quantum Yield of Fading of C.I. Basic Violet 14 in dry nitrogen in:

• Acrilan 16 Film (C 0 =2.60 µmol.g- 1 )

D Poly.(vinyl alcohol) Film (C 0 =2.42 µmol.g- 1)

• Nylon 6 Film (C 0 =2.63 µmol.g- 1 ) 1.8

1.6

1.4

1.2 &t)- 0 r-i >< ,e, 1.0 tJ'I 0 r-i

0.8

0.6

0.4

2.5 3.0 4.0 1 X T

FIGURE 3.16. Arrhenius Plots of the Initial Quantum Yield of Fading of C.I. Basic Red 18 in:

• Acrilan 16 Film (C 0 =3.29 µmol.g- 1 )

o Poly(vinyl alcohol) Film (Co=2.66 µmol.g- 1 ) . . - l • Nylon 6 Film (Co= 2 .94 µmol.g ) 1.8

1.6

1.4

~- 0 r-l 1.2 >< ,e,

01 0 ..-i 1.0

0.8

0.6

0.4 2.5 3.0 3.5 4.0

FIGURE 3 .15 Arrhenius Plots of the Initial Quantum Yield of Fading of C.I. Basic Blue 4 ~n:

• Acri~an 16 Film (Co=l.21 µmol.g- 1 )

o Poly(vinyl alcohol) Film (C 0 =1.31 µmol.g- 1 ) • Nylon 6 Film (Cq=l. 39 µmol. g -i) 89

1 so that log~ versus T for similar concentrations of each dye on all three substrates are given on the one graph.

(Only dry nitrogen results are presented.) These graphs are given in Figures 3.15 to 3.17.

One factor that becomes very obvious is that each dye/polymer system has its own individual fading behaviour.

Each dye displays quite different Arrhenius curves on each polymer with differing quantum yields and activation energies. It can be seen that the dyes also have differing effects on the apparent glass transition of the polymer.

Further, the significant effects that temperature is seen to have on dye fading quantum yields demonstrates the important role that the rigidity of the substrate has in dye fading.

With the exception of C.I. Basic Violet 14, the log~ versus} curve for Acrilan 16 is below the curves for

PVA and nylon which is to be expected from the results in

Chapter 2 and confirms that these basic dyes do have a higher lightfastness on Acrilan 16 than on nylon or PVA.

(The C.I. Basic Violet 14 result is at the present time inexplicable.)

In addition, at room temperature (} ~ 3.3 x 10- 3K- 1 ) the dyed Acrilan 16 films can be seen to exist below their apparent glass transition temperature whereas the other dyed films at room temperature exist above their apparent glass transition temperatures. This means that the fading of basic dyes in Acrilan 16 under normal circumstances occurs

3 3

5 5

2 2

0

82xl0-i+ 82xl0-i+

wet wet

7.7xl0-i+ 7.7xl0-i+

l.2xl0-

5.2xl0-i+ 5.2xl0-i+

0. 0.

5.6xl0-

Films Films

3 3

5 5

Dyed Dyed

02 02

Oxygen Oxygen

8xl0-i+ 8xl0-i+

for for

8xl0-

dry dry

4. 4.

l.3xl0-

l.2xl0-i+ l.2xl0-i+

3°c). 3°c).

6.Sxl0-1+ 6.Sxl0-1+

5. 5.

and and

(q>) (q>)

q> q>

± ±

3 3

5 5

5 5

2 2

(30 (30

N

Fading Fading

.2xl0-i+ .2xl0-i+

.2xl0-

Nitrogen Nitrogen

wet wet

l l

l.Sxl0-

l.4xl0-

l.lxl0-1+ l.lxl0-1+

7.lxl0-i+ 7.lxl0-i+

6 6 5.8xl0-i+ 5.8xl0-i+

Dye Dye

Wet Wet

3 3

5 5

5 5

of of

N2 N2

and and

Temperature Temperature

Sxl0-1+ Sxl0-1+

.Sxl0-1+ .Sxl0-1+

dry dry

7.7xl0-i+ 7.7xl0-i+

l.Sxl0- l l

l.6xl0-

Dry Dry

5.7xl0-i+ 5.7xl0-i+

6.2xl0-

1. 1.

Yields Yields

Room Room

i) i)

at at

-

under under

g g

Co Co

Quantum Quantum

3.29 3.29

2.63 2.63

2.94 2.94 1.21 1.21

1.39 1.39

2.60 2.60

5.18 5.18

µmol. µmol.

( (

Irradiated Irradiated

Initial Initial

14 14

Atmospheres Atmospheres

4 4

4 4

18 18

18 18

.4 .4

3 3

FILMS: FILMS:

Red Red

Red Red

Blue Blue

Violet Violet

Blue Blue

16 16

FILMS: FILMS:

Film/Dye Film/Dye

TABLE TABLE

.Basic .Basic

I I

C. C.

C.I.Basic C.I.Basic

C.I.Basic C.I.Basic

C.I.Basic C.I.Basic

ACRILAN ACRILAN

C.I.Basic C.I.Basic NYLON NYLON 90 in the region of lower activation energies and fading is therefore less affected by temperature. It also means that the greater rigidity of the acrylic polymer, because it exists below its T, is a significant factor in explaining g the greater lightfastness of basic dyes on this substrate.

It is also probable that the greater inherent rigidity of acrylic polymers also contributes to the overall effect.

3.6.2 The Effect of Oxygen and Moisture

The physical state of a polymer will also determine its permeability to species such as oxygen and water which may influence the fading of dyes present in the polymer.

The fading of the three basic dyes being studied was therefore examined in nylon 6 and Acrilan 16 films in dry and wet oxygen and nitrogen atmospheres. The results obtained from fading the various dyed films in oxygen atmospheres at various temperatures are included in the

Arrhenius plots in the preceding section (Figures 3.7 to

3.14). Table 3.4 gives the initial quantum yields of dye fading for the same dye/film systems in wet oxygen and wet nitrogen at room temperature together with the corresponding results in dry nitrogen and dry oxygen for comparison.

It can be seen from the Arrhenius plots that the fading rates of basic dyes in all the cases studied is reduced by the presence of oxygen. This applies to C.I.

Basic Blue 4 and C.I. Basic Red 18 in nylon 6 and Acrilan

16 films and C.I. Basic Violet 14 in nylon 6 film which represent all the dye/film systems studied under dry oxygen.

This could be due to oxygen acting as a quencher of the 91

excited dye. Alternatively the presence of oxygen may

restrict photoreduction reactions of the dye, with the photooxidation reactions that take over in their stead having a lower quantum efficiency. (This mechanism was proposed by Yamada43 to explain the higher quantum yields

of fading of triphenylmethane dyes in anaerobic compared

to aerobic solutions.)

All the dry oxygen Arrhenius plots except for C.I.

Basic Red 18 in Acrilan 16 give straight lines. Where the plot in dry nitrogen is also a straight line as for C.I.

Basic Blue 4 and C.I. Basic Red 18 in nylon 6 the explanation

for the lack of inflection would be the same as that given

in the preceding section i.e., that probably only one mechanism is involved and it is not inhibited by the rigidity of the substrate. On the other hand, where the dry nitrogen plot is curved, the straight line oxygen plot would seem to indicate that the presence of oxygen prevents a second mechanism from operating.

From the results given in Table 3.4 it can be seen

that the presence of moisture in a nitrogen atmosphere

decreases the fading rate of all three dyes in nylon films

and of two of the three dyes i.e., C.I. Basic Blue 4 and

C.I. Basic Red 18, in an oxygen atmosphere. The effect is most pronounced for C.I. Basic Blue 4 where a decrease of

approximately 20 to 26% in quantum yield is observed in

nitrogen and approximately 32% in oxygen, while the other

reductions are much smaller. This decrease in quantum yield

in the presence of moisture is the opposite effect to that 92

generally reported in the literature for other dye/fibre systems (see Section 1.4.6).

The presence of water in a polymer will swell and plasticise it, as is evidenced by the decrease in the glass transition temperature of nylon 6 with increasing water content shown in Table 3.2. In wet nitrogen the decrease in quantum yield must in some way be due to an increase in the rate of deactivation processes that occur, possibly involving an increase in self-quenching due to an increase in the dye's mobility. In wet oxygen on the other hand the decreases in rate observed may be due to the increased rate of diffusion of oxygen which is seen to quench fading in the dry polymer.

For C.I. Basic Violet 14, an increase in quantum yield is observed in wet compared to dry oxygen. In this case water itself may take part in the fading reaction either alone or in conjunction with the oxygen present.

Alternatively water's effect may be a physical one either on the dye itself by displacing the dye from its site or causing disaggregation (although aggregation is unlikely at the low concentrations used here), or on the substrate with its plasticising effect increasing the rates of diffusion of various species within the substrate. In the latter case the species involved could not be oxygen as its presence alone was found to decrease fading rather ti1an increase it.

The presence of moisture has little effect on the

fading of C.I. Basic Blue 4 and C.I. Basic Red 18 in Acrilan 93

16. This is probably due to the lack of penetration of the water into this polymer since its moisture regain is very low - 1.4% under standard conditions (20°c, 65% RH) compared to 4.2% for nylon.

Having observed some moisture effects on dyed films it was decided to extend investigations to dyed fabric.

Only nylon 6.6 fabric was used, as dyed Acrilan 16 fades too slowly for effects to be easily detected and in any case its moisture absorption is very low.

One concentration of each dye on nylon fabric was used and quadruplicate samples of each dyed fabric were faded under each of three atmospheres:

(1) dry nitrogen

(2) dry air (contains oxygen)

(3) moist air (contains oxygen and moisture)

All samples were conditioned in the appropriate atmosphere and then sealed in polyethylene bags for irradiation in the

Xenon lamp system. The polyethylene is transparent over the whole wavelength range emitted by the lamp system and did not deteriorate with irradiation.

To obtain the two dry gas atmospheres the fabric samples, attached to cardboard backings, were placed in the bags open at one end and flushed with the appropriate gas for at least two hours before being sealed. The samples to be faded in moist air were simply allowed to condition TABLE 3 .5 Percent of Initial Concentration of Dye Degraded

after a given Incident Light Exposure Dose in

Different Atmospheres.

Percent Dye Degraded Atmosphere C.I.Basic Blue 4 C.I.Basic Red 18

Dry nitrogen 18.4 (±.4) 16.9 (±1.0)

Dry air 14.4 (±.7) 16 .5 (± .9)

+ oxygen)

Moist air

(+oxygen+ water) 16 .o (±1.8) 15. 3 ( ±. 5)

Initial Dye

Concentration (µmol.g- 1 ) 0.54 1.72

Incident Light Exposure

Dose (quanta crn- 2 ) 7.7 X 10 15 1.7 X 10 17

The errors quoted are the 95% confidence limits. 94

under room conditions for at least two hours before sealing.

The samples were irradiated in the Xenon lamp system for an appropriate light exposure dose, then removed from the plastic bags and their reflectance measured on a Hunterlab

D25D2M Colour Difference Meter for determination of the dye degraded by the method given in Section 2.5.

The experimental results obtained are given in Table

3.5. They are expressed as the percentage of the original dye present that is degraded after a given incident light exposure dose.

It can be seen that oxygen decreases the fading rate of C.I. Basic Blue 4 by approximately 20% at this exposure dose but has no significant effect on C.I. Basic Red 18.

The decrease in fading is consistent with the dyed film results presented earlier. The effect of oxygen and moisture together are seen to decrease the fading of C.I. Basic Red 18 by approximately 10% at this exposure dose but have no significant effect on C.I. Basic Blue 4. Again this decrease is consistent with the dyed film results. Thus these nylon fabric results confirm the results obtained on nylon film substrates.

It can be seen that the effects of oxygen and moisture on basic dye fading are not large especially when compared to the differences in fading rates that occur between the different substrates. Moreover, both oxygen and moisture were usually found to decrease basic dye fading rates rather 95 than increase them which means that the low permeability of acrylic polymers to these two species can hardly account for the high lightfastness of basic dyes present in them. 96

CHAPTER 4

THE INFLUENCE OF THE CHEMICAL NATURE OF

THE SUBSTRATE ON THE FADING OF BASIC DYES

4.1 Introduction

The results obtained in Chapter 3 showed that the physical nature of the substrates does influence the fading of basic dyes quite significantly, and that the relatively high rigidity of acrylic compared to other substrates plays a role in reducing the rate of fading of the basic dyes present. However, it is quite possible that factors of a chemical nature also contribute to the substantially higher lightfastness of basic dyes in acrylic substrates. In this section attention was focused on various chemical factors which may influence basic dye fading.

Firstly, it was hoped that the elucidation of the reaction mechanisms operating in the fading of basic dyes would give an indication of the possible types of substrate interactions. To this end the identification of various types of reaction intermediates was attempted by the use of scavenging and quenching agents in a dye-polymer system, and by a study of the luminescence of the dyes.

Also, the effect on basic dye fading of various

chemical groups present in the different substrates was investigated. In particular, the influence of the different

dye site groups that may be present, and the influence of the cyano group which is a major constituent of acrylic 97 polymers, were studied with the aid of model systems. Any

information gained from this line of research should indicate

chemical factors which contribute to the particularly high

lightfastness displayed by basic dyes in acrylic polymers.

4.2 Irradiation Technique

The Xenon lamp system described in Section 2.6 was used for the irradiation of the fabrics, yarns and films.

Films were left on the microscope slides on which they were prepared and irradiated in the same metal holders that were

used for fabric mounting. The progress of fading was

followed in the same manner as described in Sections 2.5 and

3.5, i.e., by measuring the reflectance of fabrics and yarn,

or the absorbance of films, before exposure and after each

successive irradiation. With the films, the absorbance of

the dyed films at the wavelength of maximum absorbance of

the dye was measured in eight separate and reproducible positions over their area. After each period of irradiation

the proportion of dye remaining was calculated for each of

the eight positions, and then averaged for the whole film.

Variations in film thickness over its area caused a

corresponding variation in the absorbance of the dye and

this averaging technique gave reliable results. In all

cases duplicate samples were irradiated and were found to

be reproducible to within ±1.5% in percent dye remaining

for films and to within ±3% for nylon fabric and yarn. 98

4.3 The Identification of Reaction Intermediates

in Basic Dye Fading

4.3.1 The Use of Scavenging and Quenching Agents

Some indication of the types of reactions involved

in dye fading can be obtained by incorporating into the dye-

substrate systems certain agents which will selectively

either scavenge or quench any intermediates present.

Excited states of molecules can be quenched with an

accompanying transfer of energy from the excited molecule

to the quencher molecule, while species such as electrons

and free radicals can be scavenged (or "trapped") by

appropriate agents. By interfering with the fading reactions

in this manner the presence of scavengers and quenchers can produce a change in the rate of the reaction.

Since the presence of agents which act as excited state quenchers in the system may also introduce additional photochemical pathways, only a decrease in rate on the

addition of such an agent can be taken as evidence of the presence of a particular intermediate. An increase in rate may be due to an added reaction involving the agent itself

or a by-product, but a decrease in rate cannot be attributed

to this effect since if the rate of the additional reaction

introduced is slower than that of the normal fading reaction,

no change in fading rate will be evident.

Scavenging agents may also produce an increase in

rate. If such an agent is not photoreactive itself then

this effect can be assumed to be due to interference with 99 the normal reactions, for example by scavenging species which prevent dark or back reactions occurring. As such, an increase in the dye's fading rate in the presence of these agents can also be taken as an indication of the presence of the particular type of intermediate being scavenged.

4.3.1.1 Excited State Quenching

Various types of chemical agents have been found to be efficient excited state quenchers. They include

2-hydroxybenzophenones, hydroxybenzotriazoles, transition metal chelates especially chelates of nickel (II), metal salts and alkali metal halides, all of which are usually quoted as being excited triplet state quenchers. 109 Singlet excited oxygen quenchers include 3 -carotene, 1 1 0 sodium azide1 11 and 1, 4-diazabicyclo ( 2, 2 ,2) -octane •11 0 Nickel ( II) chelates have also been reported to be effective quenchers of singlet excited oxygen, particularly those chelates with sulphur donor ligands. 109

4. 3 .1.2 Electron and Radical Scavenging

A variety of compounds have been found to function as electron and radical scavengers. Transition metal complexes have been found to be efficient electron scavengers, with Cd2 + compounds being the most efficient. 112 A number of metal chelates have been found to be efficient scavengers of "OH and "OR radicals. 113 ' 11 ~ Other radical scavengers include phenols and dialkyldithiocarbamates.

The major difficulty encountered with this line of 100

research is that most of the chemicals involved can act in

more than one way. For example, nickel chelates are known

to be highly efficient triplet state quenchers, 10 9 but have

also been observed to quench singlet excited oxygen115 and

may also act as radical scavengers.10 9 Manganese salts have been postulated to be capable of destroying hydroperoxides

as well as quenching excited triplet states. 116 Whereas

potassium thiocyanate has been shown to scavenge hydroxyl

radicals, 117 it is unlikely to be specific to this one type

of radical and probably reacts with other radicals as well.

Consequently a cautious approach to the interpretation of

scavenging and quenching experiments is required. The

possible multifunctional mode of operation of most of these

agents means that no one result can be conclusive.

4.3.1.3 Experimental Technique

The choice of scavenging and quenching agents was

restricted by a number of factors. The use of dyed fabrics

in this work meant that only substances of very low vapour

pressure could be used in order to avoid evaporation losses

during preparation and irradiation. Coloured compounds

whose absorbance would interfere with the measurement of

fabric reflectance used to calculate the dye's concentration,

and substances which would compete with the dye for the

incident light were not suitable.

The scavengers and quenchers examined were:

1) Manganous salts (e.g., MnS04) are generally recognised

as excited triplet state quenchers. 109

2) Nickel (II) diethyldithiocarbamate{[(C2Hs)2NCS2]2Ni} is 101

typical of the nickel chelates which have been found to be

particularly efficient triplet excited state quenchers. 115

Unfortunately, the slight green colour of this compound was found to interfere with the measurement of the

concentration of both of the dyes used here and so its

use was discontinued.

3) Sodium diethyldithiocarbamate [(C 2 H5 ) 2 NCS 2 Na] has been observed to act as a radical scavenger.

4) Potassium thiocyanate (KCNS) has been demonstrated to

scavenge hydroxyl radicals. 117

5) Cd2 + ions (e.g., CdS0 4 ) are highly efficient electron

scavengers. 11 z

6) Sodium azide (NaN 3 ) is known to quench singlet excited

oxygen. 111 However, as it was found to hydrolyse to hydrazoic acid during its application its use was

discontinued.

Since the very slow fading rates of basic dyes on

acrylic fabrics would make it difficult to detect decreases

in their fading rates it was decided to apply the various

scavenging and quenching agents to nylon 6.6 fabric where basic dye fading is relatively rapid. Changes in rate, both increases and decreases, would therefore be more

readily detected. The nylon 6.6 fabric was dyed with c.r.

Basic Blue 4 and C.I. Basic Red 18 as described in Section

2.4.2. For all the additives, the dyed fabric was soaked

in a 1% (w/v) solution of the agent for approximately 30

minutes at room temperature. (The solution could not be

heated to facilitate penetration because of the very low

wet fastness of these dyes on nylon.) The wet fabric was 100

95

tJ'I s:: 90 ·r-i s:: •r-i Ill E:l Q) ~ Q) 85 KCNS >t Cl .µ s:: Q) 0 Dye 1-1 Q) 80 only A.. CdSOi. and MnSOi.

75

70

0 1.0 2.0 3.0

FIGURE 4.2 Fading Curves for C.I. Basic Red 18 in Nylon 6.6 Fabric containing: • Dye only (1.72 µmol.g- 1 ); plus: o Potassium Thiocyanate (79.5 µmol.g- 1 ) • Cadmium Sulphate (33.7 µrnol.g- 1 ) o Manganous Sulphate (46 .4 µrnol.g- 1 ) • Sodium diethyldithiocarbarnate (58.5 µrnol.g- 1 ) ( a)

b'I 90 i:: ·r-1 i:: ·r-1 rtl s Q) 80 p:;

Q) :>t 0 .j.J 70 i:: Q) u 1--1 Q) 111 60

0 0.5 1.0 1.5 2.0 2.5

(b) 100

b'I i:: 90 ·r-1 i:: ·r-1 rtl E Q) 80 p:;

Q) ::,... 0 .j.J 70 i:: Q)u J..I Q) 111 60

0 0.5 1.0 1.5 2.0 2.5

Total Incident Quanta x 10-16 (cm- 2 )

FIGURE 4.1 Fading Curves for C.I.Basic Blue 4 on Nylon 6.6 Fabric containing: (a) Dye only (b) Dye only • (0 .54 µmol.g- 1 ); plus: • (0. 80 µmol.g- 1 ); plus: D Potassium thiocyanate D Cadmium sulphate ( 79. 5 µmol. g - 1 ) ( 33. 7 µmol. g -i) ... Sodium diethyl dithio- .... Manganous sulphate carbamate (5 8 .5 µmol .g - 1 ) ( 4 6 • 4 µmo 1 • g - 1 ) 102

then passed through a padding mangle, which resulted in a

70% liquor retention, and then allowed to dry. Hence a minimum concentration of 0.7% (w/w) of each agent on the fabric was obtained. Duplicate samples of each dye/additive combination were then faded for five progressive irradiation doses.

4.3.1.4 Results

The fading curves obtained for the various scavenging and quenching agents applied to nylon 6.6 fabric dyed with

C.I. Basic Blue 4 and C.I. Basic Red 18 are given in Figures

4.1 and 4.2.

The radical scavenger sodium diethyldithiocarbarnate increased the fading rate of both dyes quite significantly.

This would seem to imply that radicals are involved in the photochemical reactions occurring. By scavenging these radicals sodium diethyldithiocarbamate may be preventing a back reaction occurring or perhaps interfering with some dark reaction to cause the observed increase in fading rate.

However, in the light of subsequent results, it is possible that this could be purely a pH effect.

Potassium thiocyanate, another radical scavenger also markedly affected the fading rates of both dyes. In the case of C.I. Basic Blue 4 the rate was increased while for C.I. Basic Red 18 it caused a substantial decrease in rate. In the latter case it is probable that the radicals being scavenged are reactive intermediates in the fading process so that their subsequent trapping causes a decrease 100

MnS011 90

CdS01t

80 tri i::: ·r-4 i::: ·r-4 rtj ~ Q) ~ Q) 70 !>-1 Cl .µ i::: Q)u 1-1 Q) P-4 60

50 Dye only

40 0 0.25 0.5 0. 75 1.0 0.25

FIGURE 4 .4 Fading Curves for C.I. Basic Violet 14 in Poly(vinyl alcohol) Films containing: • Dye only (2.97 µmol.g- 1 ); plus: o Cadmium Sulphate ( 38 µmol. g -i)

• Manganous Sulphate (150 µmol.g- 1 ) 100

95

MnSOii

90 tJ"I s:: ·r-i s:: ·r-i rtj s CdSOii (1) p::

(1) 85 >i Cl .µ s:: (1) CJ H (1) Ill 80

Dye only

75

70

0 0.5 1.0 1.5 2.0 2.5

FIGURE 4. 3 Fading Curves for C.I. Basic Blue 4 in Poly(vinyl alcohol) Films containing: • Dye only (1.85 µrnol.g- 1 ); plus: o Cadrni~ Sulphate (38 µrnol.g- 1 ) • Manganous Sulphate (150 µrnol.g- 1 ) 103 in the rate of fading.

It can be seen that both manganese sulphate and cadmium sulphate had little effect on either dye's fading.

These results would seem to imply that neither triplet excited states nor electrons are involved in either fading reaction. Alternatively the agents may not be located sufficiently close to the dye molecules to have an effect on their fading reactions.

In order to see the effect of an homogeneous distribution of these two agents throughout a dyed substrate, films of the photochemically inert polymer, poly(vinyl alcohol), were prepared containing both dye and quenching agent, and the rate of dye fading determined. The fading of C.I. Basic Blue 4 and C.I. Basic Violet 14 was investigated but unfortunately C.I. Basic Red 18 reacted with both of the agents in solution and so could not be used.

Figures 4.3 and 4.4 show the fading curves obtained for C.I. Basic Blue 4 and C.I. Basic Violet 14 in PVA films containing 38 µmol.g- 1 CdSO 4 and 150 µmol.g- 1 MnSO 4 •

(38µmol.g- 1 CdSO 4 is the limit of the solubility of CdSO 4 in

PVA.) Dye concentrations were chosen to give a maximum absorbance of 1.

It can be seen that the homogeneous distribution of both agents has a marked effect in reducing the fading rates of both dyes indicating the involvement of both electrons and excited triplet states in this particular system. 104

However, this does not necessarily mean that the same species

are involved in the fading of these dyes in other substrates

as different substrates may favour different mechanisms. It

does demonstrate that electrons and excited triplet states

are involved in one of the possible fading reactions of these

dyes and that the lack of any observable effect of these

agents in dyed nylon 6.6 fabric may be due to their uneven

distribution rather than due to the absence of electrons and

excited triplet states.

4.3.2 Luminescence Studies

The possibilities of using luminescence studies of

the dyed films to give further insight into the excited

species involved in the dye fading were investigated on a

spectrofluorophosphorimeter constructed in these

laboratories. 118 However, in common with other dyes, any

fluorescence observable was very weak and at quite long wavelengths (peaks in the range 590 - 690nm) while no phosphorescence could be detected within the wavelength

range of the instrument used.

4.4 The Influence of Dye Sites on Basic Dye Fading

Model compounds containing the anionic groups that

are used as dye bonding sites for basic dyes were introduced

into an inert solid model substrate to determine how the

nature of the bonding site affects the fading of the basic

dyes attached to them. Essentially only two types of dye

sites appear to be used i.e., sulphonate groups (SO 3 -) and

carboxylate groups (C00-). 96 It has been suggested42 (see

Section 1.5) that the reason for e1e higher lightfastness of 105 basic dyes on acrylics is that these fibres contain

sulphonate dye sites instead of carboxylate sites and that

it is the sulphonate sites that cause the improved

lightfastness of the basic dyes attached to them.

4.4.1 Experimental Technique

Poly(vinyl alcohol) films were chosen as the model

substrate for this work as they contain no potential sites

for basic dyes other than those specifically introduced.

Also, because the films are cast, additives such as the dye

and the dye site models can be introduced homogeneously

through the films.

The model compounds chosen to simulate the dye sites, benzene sulphonic acid and trichloroacetic acid,have the same

dissociation constant in water (K = 2 x 10- 1 ). a

In addition, sodium benzene sulphonate and sulphuric

acid were each used to determine the effect of the -SO 3 and

H+ components of benzene sulphonic acid. Sodium hydroxide was used to determine the effect of base and sodium sulphate

to determine the effect of the sulphate group.

Each compound was added to the film-forming

poly(vinyl alcohol) solution at a concentration that gave

a ratio of additive molecules to dye molecules of 33:1 in

the film, thereby ensuring that most dye molecules would

be in close proximity to an additive molecule.

Only C.I. oasic Blue 4 was used in this work since 100

95

°'s:: ·r-i 90 s:: H2S04 ·r-i ro I::: & Q) !>t Q .µ s:: Q) 85 u 1-1 Q) Al

. 80

Dye only NaOH

75

0 0.5 1.0 1.5 2.0 2.5

FIGURE 4 .5 Fading Curves for C.I. Basic Blue 4 in Poly(vinyl alcohol) Films (1.85 µmol.g- 1 ) containing various additives in the molar ratio additive : dye = 33 : 1 106

C.I. Basic Red 18 was found to precipitate from PVA on the

addition of acid and C.I. Basic Violet 14 changed colour in

the presence of acid. An initial dye concentration of

1.85 µmol.g -1 C.I. Basic Blue 4 was used, to give an initial maximum absorbance of 1.

4.4.2 Results

The resulting fading curves for 1.85 µmol.g -1 C.I.

Basic Blue 4 in poly(vinyl alcohol) films containing the various additives are given in Figure 4.5. It can be seen

that there is virtually no difference in the dye's fading

rate between the films containing the two different dye sites i.e., there appears to be no significant difference in fading of C.I. Basic Blue 4 caused by its bonding to sulphonate or

carboxylate sites. Both additives do cause a decrease in

the fading rate of C.I. Basic Blue 4. However this is likely to be simply due to their acidic properties as it can be seen that a stronger acid, sulphuric acid (K = 1.20 x 10-2 ) a has an even greater effect in decreasing the rate, whereas the sodium salt of benzene sulphonic acid has no significant effect. Sodium hydroxide and sodium sulphate are seen to have very little effect on the dye's fading. The sodium

sulphate result would seem to imply that the effect of

sulphuric acid in decreasing the fading rate is due to the

+ 2- effect of the H rather than the SO 4 ions.

Hence these results for C.I. Basic Blue 4 in the

model substrate PVA provide no evidence that the type of

dye site affects the lightfastness of the dye bonded to it

providing that their pKa's are the same. 100

95

s::tJ'I ·r-1s:: ·r-1 ro El Q) p:; Q) 90 >t 0 .µ s:: Q) u 1-1 Q) Ill

85

80

0 0.5 1.0 1.5 2.0 2.5 Total Incident Quanta x 10-16 (cm- 2 ) FIGURE 4.6 Fading Curves for C.I. Basic Blue 4 in Poly(vinyl alcohol) Films containing various amounts of Sulphuric Acid. -1 • Dye only (0.93 µmol.g ) ; • Dye only (1.85 µmol.g- 1 ); plus: plus: -1 D 61.8 µmol.g- 1 H2SO~ • 5.6 µmol.g H2S0 4 6 15 µmol.g- 1 H2S0~ o 30.6 µmol.g- 1 H2S04 T 61.8 µmol.g- 1 H2SO~ V 91.8 µmol.g- 1 H2S04 107

In order to determine the effect of the pH of the

system on the fading of C.I. Basic Blue 4 in PVA, various

concentrations of sulphuric acid were added to dyed films

and the films irradiated. The sulphuric acid concentration

in the film was varied from 5.6 to91.8µmol/g PVA while the

concentration of dye was kept constant at 1.85 µmol/g PVA.

This represents a range in the molar ratio of acid to dye of approximately 3:1 up to 50:1. The concentration of dye was also halved for one acid concentration (0.93 µmol.C.I.

Basic Blue 4/g PVA +61.8µmol.H 2 SO~/g PVA) which represents a molar ratio of acid to dye of approximately 70:1. The

resulting fading curves are given in Figure 4.6.

It can be seen that the rate of fading of C.I. Basic

Blue 4 in PVA does decrease with increasing sulphuric acid

concentration and that the effect is quite a significant

one. This decrease in rate is also seen for the lower

concentration of dye examined.

Having noted the effect of acid on the fading of

C.I. Basic Blue 4 in the model substrate PVA, it was decided

to examine the effects of acid and base on the fading of

dyed fibre substrates. Nylon 6.6 fabric and basic-dyeable nylon yarn dyed with one concentration of C.I. Basic Blue 4

and C.I. Basic Red 18 were examined. Unfortunately only the basic-dyeable nylon yarn could be used as attempts to apply

acid to the dyed nylon 6.6 fabrics resulted in excessive

leaching of the dye out of the fabric. There was a slight

leaching of dye out of the basic-dyeable nylon but it was

possible by restricting treatment time to keep this leaching { a}

t,"Is:: ·r-1s:: •r-1 Ill 80 s & Q) ~ +l s:: 60 Q) u l--1 Q) Al

40.,__ ___--+-----+------f-----4------1- 0 0.5 1.0 1.5 2.0 2.5

{b} 10

tJ'I s:: ·r-1s:: ·r-1 Ill s & Q) >t Q +l s:: Q)u l--1 Q) Al

0 0.5 1.0 1.5 2.0 2.5 Total Incident Quanta x 10- 17 (cm- 2 )

FIGURE 4. 7 Fading Curves for {a) C.I. Basic Blue 4 {0.84 µmol.g- 1 ) {b) C.I. Basic Red 18 (1.9 µmol.g- 1 ) in Basic-dyeable nylon yarn containing: • Dye only; plus: 1 D H 2 SO 4 { 410 µmo • g 1 - )

1:::,,. NaOH ( 2 7. 3 µmol. g -i) 108

to an acceptable level while still obtaining sufficient acid

take-up.

The dyed yarn was soaked for 15 minutes in lM NaOH

or 1 minute in 0.5M H2 SO 4 at room temperature. The sample was then briefly immersed in ice-cold water to remove the non-absorbed acid or base from the fibre's surface. The yarn was allowed to dry and then wound onto a cardboard backing

for irradiation. The amount of acid or base absorbed by the

fibres was determined by the back titration of similarly treated samples using a Sargent-Welch Model PBL pH meter to detect the end point.

The fading curves obtained for 0.84 µmol.C.I. Basic

Blue 4 per gram and 1.9 µmol.C.I. Basic Red 18 per gram of basic-dyeable nylon containing 410 µmol(H+)/g as sulphuric

acid and27.3µmol(OH )/gas sodium hydroxide are given in

Figure 4.7.

It can be seen that the presence of acid decreases

the fading rate of C.I. Basic Blue 4 in basic-dyeable nylon which corresponds to the effect seen for this dye in poly(vinyl alcohol) films. The presence of base is seen

to markedly increase its rate of fading here whereas in poly(vinyl alcohol) films no change in rate was observed.

For C.I. Basic Red 18, acid has no significant effect

on its fading but like C.I. Basic Blue 4 it displays a marked

increase in fading rate in the presence of base. 109

These results demonstrate that pH can have a marked effect on the fading of basic dyes both in model polymer

films and in fibre substrates. It appears that in general

the presence of acid tends to reduce the dye's fading rate

and that of base to increase it. The extent of the effect appears to depend on the dye/substrate system in question.

This pH effect observed means that the pK of the dye site a is likely to influence the fading of basic dyes attached to

them. As it is quite probable that the pK 's of the same a site in different polymers or of different sites in the one polymer will differ, it is therefore possible that the nature

of the dye site may by virtue of its pK influence the fading a of basic dyes.

4.5 The Influence of the Cyano Group on Basic Dye Fading

The influence of the main side group present in

acrylic fibres, the cyano group (-C=N), on dye fading was

investigated by incorporating a model compound containing

this group into dyed poly(vinyl alcohol) films and examining

the effect on the dyes' fading.

4.5.1 Experimental Technique

The model compound chosen was succinonitrile

(NCCH2CH2CN) which is similar to the repeat unit of the polyacrylonitrile polymer (CH 2-CH-) and contains only the I n CH

CN group as the possible active component. Since acrylic polymers consist of at least 85% polyacrylonitrile, at least

49% by weight of an acrylic polymer consists of -CN. A maximum of 1% CN by weight could be added to poly(vinyl 110

alcohol) films, as this represented the limit of solubility

of succinonitrile in poly(vinyl alcohol). Above this

concentration the succinonitrile precipitated out producing

an opaque film. Malononitrile (NCCH 2 CN) was found to be more

soluble in the film, but due to its volatility evaporated

from the film during the drying.

Consequently, the use of other polymer substrates was

investigated and it was found that films of methyl cellulose

could be prepared containing up to 8% CN by weight as

succinonitrile without precipitation occurring. The method of preparation of these films involves first swelling the polymer by the addition of a little boiling water, followed by dissolution with the added dye and succinonitrile in cold water. A 2% (w/w) solution of methyl cellulose (Fluka

Methocell MC 400cP) was prepared and 6 mls of this solution

cast on a 50 x 37.5mm slide. This produced a film of

thickness 30 ± 5 µ.

It was found subsequently that the succinonitrile

gradually evaporated from the polymer films. The dyed films were therefore irradiated as soon as possible after they had

dried but unfortunately this evaporation meant that the

concentration of succinonitrile present in the dyed film at

the time of fading could not be determined accurately.

Consequently the values given are only approximate and, being the concentration of succinonitrile introduced into

the film on casting, they represent the maximum possible

concentration that could be present. This evaporation effect was much greater in the relatively porous methyl cellulose (a) 100

t,"I 95 s:: ·r-1s:: ·r-1 =IO 90 ~ Q) >t Q 85 .µ s:: Q) CJ 1-1 Q) 80 ~

0 0 .• 5 1.0 1.5 2.0 2.5 Total Incident Quanta x 10- 16 (cm-2 )

(b)

0 0.5 1.0 1.5 2.0 2.5

FIGURE 4 .9 Fading Curves for C.I. Basic Blue 4 (1.85 µmol.g- 1 ) in: (a) Poly(vinyl alcohol) Film (b) Methyl Cellulose Film, containing: • Dye only; plus: a Up to 1% CN (as succinonitrile) • Up to 8% CN (as succinonitrile) ( a) 100

s::t,"I 98 ·r-is:: ·r-i rt! s 96 & (I) ~ 0 94 .j.J s:: (I)u 1-1 (I) Ill 92

90 0 0.5 1.0 1.5 . 2.0 2.5 3.0

(b) 100

t,"I 90 s:: ·r-is:: ·r-i rt! s 80 & (I)

£: 70 .j.J s:: (I)u 1-1 (I) Ill 60

50 0 0.25 0.5 0.75 1.0 1.25

FIGURE 4.8 Fading Curves for (a} C.I. Basic Red 18 (1.69 µmol.g- 1 ) (b) C.I. Basic Violet 14 (1.85 µmol.g- 1 ) in Poly(vinyl alcohol) Films containing: • Dye only; plus: t::. up to 1% CN (as succinonitrile} 111

films than in the poly(vinyl alcohol). Consequently, the effects of succinonitrile on the lightfastness of all three dyes were determined in poly(vinyl alcohol), but in methyl

cellulose only the dye C.I. Basic Blue 4 could be studied as

it required a much shorter irradiation time than the other

two dyes. (Irradiation does result in a slight temperature

rise of the sample which over the longer exposure times

caused marked evaporation of the succinonitrile.) The use of less volatile models for the cyano group such as phthalo-

N ni trile ( O( ) was investigated but due to their low N solubility in water they could not be incorporated into the

films by the common solvent technique of forming films. In any case, the use of succinonitrile does demonstrate the influence of the cyano group on fading even though the effect cannot be strictly quantified.

4.5.2 Results

The resulting fading curves obtained for similar

concentrations of C.I. Basic Blue 4, C.I. Basic Red 18 and

C.I. Basic Violet 14 in PVA film and for C.I. Basic Blue 4

in methyl cellulose film for the dye alone and in the presence of succinonitrile are presented in Figures 4.8 and

4. 9.

It can be seen that in all cases the presence of

cyano groups decreased the rate of fading of the dye

concerned. For C.I. Basic Blue 4 and C.I. Basic Violet 14

in PVA the effect is small but for Basic Red 18 it is quite

100% 100%

X X

4 4

.6 .6

11.9 11.9

21.6 21.6

10. 10.

58 58

54.6 54.6

l l

kD+S kD+S

k? k?

l l -

D D

k

alone, alone,

alcohol), alcohol),

18 18

18 18

Dye Dye

) )

8 8

2

9 9

1

1 1

18 18

lQ-

10-

the the

fading fading

cm

1

X X

X X

10-

10-

10-

presence presence

for for

X X

dye dye X X

X X

poly(vinyl poly(vinyl

k?+S k?+S

8 8

4.48 4.48

2.07 2.07

0 0

in in

the the

of of

(quanta-

4.92 4.92

8. 8.

1.44 1.44

groups. groups.

by by

(i) (i)

(ii) (ii)

Fading Fading

) )

Dye Dye

maximum maximum

7 7

2

8 8

constant constant

Cyano Cyano

l l

19 19

18 18

1 1

caused caused

cm

of of

CN CN

1

of of

succinonitrile succinonitrile

10-

10- 10-

10-

rate rate

1% 1%

rate rate

X X

X X X X

X X

k? k?

of of

3 3

to to

0 0

00 00 in in

(quanta-

3.17 3.17

5.58 5.58

1. 1.

5. 5.

order order

presence presence

Constants Constants

presence presence

the the

) )

first first

Rate Rate

1

in in

alone alone

cellulose. cellulose.

decrease decrease

-

in in

.g .g

corresponding corresponding

dye dye

Co Co

dye dye

and and

Order Order

1.85 1.85

1.85 1.85

1.84 1.84

1.69 1.69

succinonitrile succinonitrile

methyl methyl

(µmol (µmol

for for

for for

of of

percent percent

Apparent Apparent

in in

First First

= =

= =

= =

= =

S S CN CN

+ +

8% 8%

100% 100%

D D

D D

k1 k1

x x

14 14

Film: Film:

concentrations concentrations

4 4

4 4

Apparent Apparent

(ii) (ii)

18 18

Film: Film:

1 1

kD+S kD+S

l l

and and

_ _

Red Red

kD kD Blue Blue

Violet Violet

where where

Blue Blue

1 1

4.1 4.1

alcohol) alcohol)

kD kD

1% 1%

Basic Basic

Basic Basic

Cellulose Cellulose

Basic Basic Basic Basic

. .

. . .

TABLE TABLE

Dye/Film Dye/Film

(yj_nyl (yj_nyl

(i) (i)

succinonitrile succinonitrile

& &

C.I. C.I.

C.I. C.I.

C.I. C.I.

C.I. C.I.

for for

and and

Poly Poly Methyl Methyl 112

large. Further, it can be seen that by increasing the concentration of -CN by a factor of approximately 8, in methyl cellulose, the rate of fading of Basic Blue 4 is

further reduced.

Considering that the concentration of succinonitrile in the methyl cellulose films still represents only 8% CN at most, it would be expected that in an acrylic polymer where the concentration of CN is approximately 49% that a much larger effect still would occur.

The apparent first order rate constants for these curves were calculated from the slope of log (percent dye remaining) versus total incident quanta curves by linear regression. These results, together with the percentage change in the rate constant in the presence of succinonitrile are given in Table 4.1.

Expressed in this manner the significance of these results is even more apparent. Even the small effects represent an appreciable decrease in the rate constant (and therefore also the rate itself) of 12 to 22%, for C.I. Basic

Violet 14 and C.I. Basic Blue 4 respectively in the presence of a maximum of 1% CN, while for C.I. Basic Red 18 the decrease is up to 55%. Then, on increasing the concentration of CN by a factor of approximately 8, the decrease for C.I.

Basic Blue 4 changes from 10% to 59%. This effect could be explained by the CN groups acting as excited state quenchers

and deactivating the excited dye molecules. However, the very small energies of the lowest excited states of the dyes 113 compared to the relatively high energy levels of any of the groups in the polymer, including the cyano group, make energy transfer highly unlikely.

Alternatively, the cyano groups could reduce dye fading by combining with the excited dye to form an excited complex, or exiplex. Such a complex, which could involve one or more cyano groups, could have more numerous or more rapid photophysical deactivation processes than the excited dye molecule alone. This would lead to more rapid deactivation of the dye from the excited state and thereby reduce the probability of fading. Alternatively the structure of the exiplex may in some way inhibit the normal fading reactions of the dye. 114

CHAPTER 5

CONCLUSION

Published data reveals the high lightfastness of basic dyes in acrylics compared with other substrates. The three dyes studied here are typical, with quoted lightfastness of between 4 and 7 on acrylics and approximately 1 to 3 on other fibres. This information has been quantified by the determination of the initial quantum yields of the dyes' fading in Acrilan 16, nylon 6.6 and basic-dyeable nylon fibres, and differences of at least two orders of magnitude between the quantum yields of the dyes in acrylic and the non-acrylic fibres were detected.

A definite concentration dependence of quantum yield was also observed, with the quantum yields decreasing with increasing dye concentration. This is the trend that has generally been observed but the usual explanation of the phenomenon as being due to increasing dye aggregation with increasing concentration is unlikely to apply here because of the very low dye concentrations used.

The possibility that the large difference in fading rates between acrylic and non-acrylic substrates could be due to differences in the physical structure of the polymer substrates was examined by determining the effect of temperature on fading rates. A definite temperature dependence of quantum yield was observed and many dye/film systems displayed an inflection in their Arrhenius plots 115

which was postulated to correspond to a physical transition

in the polymer equivalent to its glass transition. The

presence of these inflection points indicates that the

physical structure of the polymer has a significant effect

on dye fading, the more rigid the structure, the lower the

rate of fading. Thus the greater inherent rigidity of

acrylics compared with polyamides would contribute to the

improved lightfastness.

Furthermore, the fact that in most cases the

inflection points found for the acrylic substrate were above

ambient temperature while those for the other substrates were at or below ambient temperature suggests that fading

occurs with the acrylic in its most rigid form, while the other substrates are in a more flexible state. This effect

is probably even more pronounced in practice as the glass

transition temperatures of polymers such as nylon will be

lowered by the presence of moisture much more than that of

acrylic polymers. This is demonstrated by the fact that the

T of Acrilan is lowered from 90°c dry to 57°c wet 119 while g the T of nylon 6 is 94°c dry but -6°c at 97% R.H. 108 g

Since a purely photochemical reaction has essentially zero activation energy, then activation energies associated with photochemical reactions must be due to some diffusional or conformational process involved in the degradation reaction. Hence the high rigidity of acrylics is likely to affect dye fading by restricting these motions.

The activation energies determined ranged from 116

approximately 1 to 10 kJmol- 1 below the inflection temperatures and 15 to 66 kJmol- 1 above them. These values are of the same order of magnitude as values reported for diffusion processes in polymers. For example, small dye molecules such as vat and direct dyes, and azoic coupling

components have been reported to have activation energies of diffusion, relating to the dyeing process, in the range

42 to 59 kJmol- 1 in the relatively porous polymer viscose rayon. 120 On the other hand the larger basic dyes in the relatively non-porous Orlon 42 have much higher activation energies, reported as 250 kJmol- 1 • 121 It would appear from this that any rate-determining diffusional processes involved in basic dye fading are unlikely to involve whole dye molecules.

Pace 122 related data for the diameters (d) of simple penetrant molecules to their activation energies of diffusion

(Ea), and found that, in the case of polyethylene, the very small helium atom (d = 0.22nm) had an Ed of approximately

20 kJmol- 1 • The relatively high activation energies for such a small particle would seem to indicate that activation energies below 20 kJmol- 1 are probably unlikely to be due to

diffusional processes, except in the case of even smaller particles such as electrons, and are more likely to be due

to conformational processes.

This effect of substrate rigidity and the explanation

for it have frequently been quoted in studies into photo­

degradation of polymers, 7 5 ' 95 but has not previously been linked with dye fading studies. 117

The low permeability of acrylics to oxygen and moisture has been postulated by some 61 ' 97 to contribute to the high lightfastness of basic dyes present, since some

fading reactions of dyes have been found to involve these species. However, as reported in Chapter 3, moisture and oxygen were found to have little effect on basic dye fading and furthermore, where an effect was observed, their presence generally reduced rather than increased dye fading.

Some workers 42 ' 99.have observed that the nature of the dye site in the polymer affects basic dye fading. In particular, sulphonate sites appear to give rise to slower fading for dyes adsorbed onto them than the other typical dye site, viz., carboxylate. However, in these studies no difference was detected between the two different sites providing their pK 's were the same. On the other hand, a the pH of the polymer was found to have a marked effect, with an acidic environment reducing fading and a basic environment generally increasing fading. It is quite probable that the different dye site/polymer combinations give rise to quite different pH's and that the nature of the dye site will therefore by virtue of its pK affect a basic dye fading. This pH effect on dye fading rates suggests that electron transfer from substrate to dye could contribute to the dye degradation. The reduction in fading of dyes in poly(vinyl alcohol) in the presence of an electron scavenger (cadmium ions) also supports this suggestion. Thus the reduced fading in acrylics might be attributed in part to them being poor electron donors

compared with the other substrates studied. 118

A major factor influencing basic dye fading was found to be the quenching action of the cyano group which constitutes approximately 50% by weight of an acrylic polymer.

Although experimental difficulties limited the amount of information which could be obtained, the results did show that approximately 1% concentrations of cyano groups in an inert substrate produced a significant reduction in fading rate with the effect becoming greater with a cyano concen­ tration of approximately 8%. This implies that the high cyano group concentration in acrylics would result in quite a significant effect on dye fading.

The mechanism by which the cyano group acts to reduce fading is unlikely to involve energy transfer because of the energy levels involved. Enhanced electron transfer from substrate to dye or vice versa is also unlikely since this action would result in an increase in fading rate. It has therefore been postulated that an excited complex is formed between the excited dye molecule and one or more cyano groups which may facilitate rapid deactivation of the dye or alternatively the structure of the complex may inhibit the normal fading reactions of the dye.

In summary, it can be seen from the above that several independent properties contribute to the high lightfastness of basic dyes on acrylic fibres. The physical nature of the acrylic polymers, by virtue of their high rigidity plays an important role. In addition, the low pK a of their dye sites, and their poor electron donating properties may also contribute. However, the major influence 119

would appear to be the high cyano group content of the polymer. 120

REFERENCES

1. Giles,C.H., and McKay,R.B., Text.Res.J., 31, 527 (1963).

2. Egerton,G.S., and Morgan,A.G., J.Soc.Dyers Colour., 86,

79 (1970).

3. idem, ibid., 86, 242 (1970).

4. idem, ibid., 87,223 (1971).

5. idem, ibid., 87,268 (1971).

6. Meier, H., in "The Chemistry of Synthetic Dyes Vol. IV",

Ed.K.Venkataraman, Academic Press, New York (1971),

Ch.VII.

7. Bentley,P., McKellar,J.F., and Phillips,G.O., Rev.Frog.

Color.Relat.Top., ?._, 33 (1974).

8. Evans ,N .A. , and Stapleton, I. W. , in "The Chemistry of

Synthetic Dyes Vol.VIII", Ed.K.Venkataraman, Academic

Press, New York (1978), eh.VI.

9. Blum,H.F., and Spealman,C.R., J.Phys.Chem., 37, 1123

( 19 3 3) •

10. Griffiths,J., and Hawkins,C., J.Chem.Soc., Chem.Commun.,

463 (1972).

11. Griffiths,J., Chem.Soc.Rev., 1, 481 (1972).

12. Hillson,P.J., and Rideal,E.K., Proc.R.Soc.London, Ser.A.,

216, 458 (1953) •

13 . B 1 ai s de 11 , B • E • , J . So c . Dye r s Co 1 our . , ~, 6 18 ( 19 4 9 ) .

14. Hashimoto,S., Sunamoto,J., and Fujii,H., Kogyo Kagaku

Zasshi, 7..!2_, 699 (1967).

15. Irick,G.Jr., and Pacifici,J.G., Tetrahedron Lett.,

No. 17, 1303 (1969).

16. van Beek,H.C.A., Heertjes,P.M., Houtepen,C., and

Retzloff,D., J.Soc.Dyers Colour.,~, 87 (1971). 121

17. Giles,C.H., Shah,C.D., Watts,W.E., and Sinclair,R.S.,

ibid.,~, 433 (1972). 18. Kamel,M., Issa,R.M., Abdel-Wahab,L., Shakra,S., and

Osman,A., Textilveredlung, §_, 224 (1971).

19. Matsuoka,M., Seko,H., Sugimoto,T., Yutani,S., Kitao,T.,

and Konishi ,K. , Kogyo Ka gak u z ass hi , 2_!, 1655 ( 19 71) .

20. Naumova,I.A., Zaitsev,B.E., and Belen'kii,L.I.,

Izv.Vyssh.Uchebn.Zaved.Khim.Khim.Teknol., 13, 1340 (1970).

21. Tomkiewicz,M., and Klein,M.P., J.Am.Chem.Soc., 95, 3132

(1973).

22. Lewis,G.E., Tetrahedron Lett., No. 9, 12 (1960).

23. Hugelshofer,P., Kalvoda,J., and Schaffner,K.,

Helv.Chim.Acta., i1_, 1322 (1960).

24. Krollpfeiffer,F., Miihlhausen,C., and Wolf,G.,

Annalen., 508, 39 (1934).

25. Bitzer,D., and Brielmaier,H., Melliand Textilber.,

41, 62 (1960).

26. Kato,S., Morita,M., and Koizumi,M., Bull.Chem.soc.Jpn.,

r!__, 117 (1964).

27. Egerton,G.S., Gleadle,J.M., and Roach,A.G., Nature,

202, 345 (1964).

28. Wilkinson,F., J.Phys.Chem., §_§_, 2569 (1962).

29. Davies,A.K., McKellar,J.F., and Phillips,G.O.,

Proc.R.Soc.London,Ser.A, 323, 69 (1971).

30. Allen,N.S., Bentley,P., and McKellar,J.F.,

J .Photochem., ~, 225 (1976).

31. Allen,N.S., and McKellar,J.F., ibid.,~, 317 (1976).

32. idem, ibid., ?_, 107 (1977).

33. McKellar,J.F., Radiat.Res.Rev., l, 141 (1971). 122

34. Allen,N.S., McKellar,J.F., Moghaddam,B.M., and Phillips,

G.O., J.Photochem., 11,101 (1979).

35. Naumova,I.A., Zaitsev,B.E., and Belen'kii,L.I.,

Anilino-Krasochnaya Promshlennost, 109 (1970) .

36. Egerton,G.S., and Assaad,N.E.N., J.Soc.Dyers Colour.,

~, 85 (1967).

37. Egerton,G.s., Assaad,N.E.N., and Uffindell,N.D.,

Chem.Ind. (London), 1172 (1967).

38. Wegerle ,D., J .Soc .Dyers Colour.,~, 54 (1973).

39. Giles,C.H., and Sinclair,R.S., ibid.,~, 109 (1972).

40. Jelinek,Z.K., and Kratochvil,V., Textil {Prague),

25, 335 (1970).

41. Desai,C.M., and Vaidya,B.K., J.Indian Chem.Boe.,

31, 261 (1954).

42. Porter,J.J., and Spears,S.B.Jr., Text.Chem.Color., ~,

191 ( 19 70) •

43. Yamada,K., Shosenji,H., and Gotoh,K., J.Soc.Dyers

Colour.,~, 219 (1977).

44. Bangert,R., Aichele W., Schollmeyer,E., Weimann,B., and

Herlinger,H., Melliand Textilber., 58, 399 (1977).

45. Maerov,S.B., and Kobsa,H., Text.Res.J., 31,697 (1961).

46. Leaver,I.H., Photochem.Photobiol., 16, 189 (1972).

47. Owen,E.D., and Allen,R.T., J.Appl.Chem.Biotechnol.,

~, 799 (1972) .

48. Pak,M.A., Shigorin,D.N., and Ozerova,G.A., Dokl.Akad.

Na uk • SSS R , 18 6 , 3 6 9 ( 19 6 9 ) •

49. Henriquez,P.C., Rec.Trav.Chim.Pays-Bas, 52, 991 (1933).

50. Russel ,W .J., and Abney ,W .de W., "Action of Light on

Water Colours", Report to the Science and Art Dept. of

the Committee of Council on Education C5453, London: 123

H.M.S.O., (1888).

51. Bolis,M.A., Rev.Gen.Matieres Color.,Blanchiment,Teint.,

Impress.Apprets, 12, 289 (1908).

52. McLaren,K., J.Soc.Dyers Colour., 72, 86 (1956).

53. Shah,C.D., and Srinivasan,R., Text.Res.J., 47, 625

(1977).

54. Irick,G.Jr., and Boyd,E.G., ibid., ~, 238 (1973).

55. Evans,N.A., ibid., Q, 697 (1973).

56. Giles,C.H., J.Appl.Chem., 15,541 (1965).

57. Baxter,G., Giles,C.H., and Lewington,W.J., J.Soc.Dyers

Colour, '!..]_, 386 (1957).

58. Baxter,G., Giles,C.H., McKee,M.N., and Macaulay,N.,

ibid., 71, 218 (1955).

59. Terenin,A.N., "Photonics of Dye Molecules", Moscow,

Nauka, (1977).

60. Bean,C.P., and Rowe,F.M., J.Soc.Dyers Colour.,~, 67

(1929).

61. Giles,C.H., and Rahman,S.M.K., Text.Res.J., l!_, 1012

(1961).

62. Giles,C.H., J.Soc.Dyers Colour,'!..]_, 127 (1957).

63. Shah,C.H., and Srinivasan,R., Text.Res.J., 45, 486

(1975) •

64. Giles,C.H., and Shah,C.D., J.Soc.Dyers Colour., 85,

417 (1969).

65. Giles,C.H., and Duff,D.G., ibid.,~' 405 (1970).

66. Hoffman,K., McDowell,W., and Weingarten,R., ibid.,

~' 405 (1970) •

67. McDowell,W., and Weingarten,R., ibid.,~' 406 (1970).

68. Kissa,E., Text.Res.J., 41, 715 (1971).

69. Blair,H.S., and Boyd,N.L., J.Soc.Dyers Colour., 91, 124

112 (1975).

70. Bancroft,E., "Experimental Researches concerning the

Philosophy of Permanent Colours", Cadell and Davies,

London, (1813).

71. Schwen,G., and Schmidt,G., J.Soc.Dyers Colour., Ti_,

101 (1959).

72. Evans,N.A., and Waters,P.J., Proc.5th Int.Wool Text.

Res.Conf.Vol.III, 567 (1976).

73. Chipalkatti,H.R., Desai,N.F., Giles,C.H., and

Macaulay N., J .Soc .Dyers Colour., ?..!2_, 487 (1954) •

74. Simons ,J.P., "Photochemistry and Spectroscopy",

Wiley-Interscience, London (1971).

75. Guillet,J.E., Pure Appl.Chem.,~' 249 (1977).

76. Egerton,G.S., Assaad,N.E.N., and Uffindell,N.D.,

J.Soc.Dyers Colour.,~' 220 (1968).

77. Allen,N.S., Harwood,B., and McKellar,J.F., J.Photochem.,

10, 187 (1979).

78. Allen,N.S., Harwood,B., McKellar,J.F., and Greenhalgh,

C. W. , ibid. , 10 , 19 3 ( 19 79) •

79. Ingamells ,W., J .Soc.J;Jyers Colour., 79, 651 (1963).

80. Osawa,E., Kurihara,o., and Nozawa,J., ibid.,~' 205

( 19 6 4) •

81. Zakharova,T.D., Solov'eva,V.B., and Zaitseva,V.N.,

Zh .Prikl .Khim., !i_, 2771 (1971).

82. Giles,C.H., Johari,D.P., and Shah,C.D., Text.Res.J.,

41, 83 (1971).

83. Krichevskii,G.E., Vachobov,B., Ershov,Y.A., and

Do vb ii , E . B . , i bi d . , 4 5 , 6 0 8 ( 19 7 5 ) •

84. Datyner,A., Nicholls,C.H., and Pailthorpe,M.T.,

J.Soc.Dyers Colour., 93, 213 (1977). 125

85. Laserev,P., Bull.Acad.Sci.(Petrograd), 1916, 583 (1908).

86. Desai,N.F., and Giles,C.H., Proc.Symp. on Photochem. in

Reln.to Textiles, Harrowgate (1949).

87. van Beek,H.C.A., and Heertjes,P.M., J.Soc.Dyers Colour.,

?.J_, 661 (1963).

88. Hedges,J.J., ibid., 44, 341 (1928).

89. Cunliffe,P.W., and Lambert,P.N., ibid.,~, 59 (1932).

90. McLaren,K., ibid., ]2, 527 (1956).

91. Giles,C.H., Hojiwala,B.J., Shah,C.D., and Sinclair,R.S.,

ibid., 90, 45 (1974).

92. Giles,C.H., Haslam,R., and Duff,D.G., Text.Res.J., i.§_,

51 (1976).

93. Shah,C.D., and Srinivasan,R., ibid., 46, 380 (1976).

94. Nordhammar,G., Am.Dyest.Rep., l§_, 571 (1949).

95. Williams,J.L.R., and Daly,R.C., Prog.Polym.Sci., ~, 61 (1977).

96. Johnson,R.F., Ph.D.Thesis, Swiss Federal Institute of

Technology, Zurich (1963).

97. Trotman,E.R., "Dyeing and Chemical Technology of

Textile Fibres", 4th Edn, Griffin, London. (1970), p.570.

98. Wegmann,J., Melliand Textilber., ~, 408 (1958).

99. Zollinger,H., Am.Dyest .Rep., ~, 634 (1965).

100. Heinkel,H., Textilveredlung, 12, 403 (1977).

101. Kubelka,P., and Munk,F., Z.Tech.Phys., 12, 593 (1931).

102. Parker,C.A., Proc.R.Soc.London,Ser.A, 220, 104 (1953).

103. Hatchard,C.G., and Parker,C.A., ibid., 235, 518 (1956).

104. Wendlandt,W.W., "Thermal Methods of Analysis", 2nd Edn,

Wiley-Interscience, New York, (1974), p.295.

105. Mackenzie,R.C., "Differential Thermal Analysis",

Academic Press, London (1972). 126

106. Roff,W.J., Scott,J.R., and Pacitti,J., "Fibres, Films,

Plastics and Rubbers", Butterworths, London (1971).

107. Finch,C.A., "Polyvinylalcohol", John Wiley and Sons,

London (1973).

108. Kettle,G.T., Polymer, 18, 743 (1977).

109. Allen,N.S., and McKellar,J.F., Chem.sac.Rev.,!, 533

(1975) .

110. Kearns,D.R., Chem.Rev., 71, 395 (1971).

111. Nilsson,R., Merkel,P.B., and Kearns,D.R., Photochem.

Photobiol., 16, 117 (1972).

112. Hart ,E .J., and Anbar ,M., "The Hydrated Electron",

Wiley-Interscience, New York (1970).

113. Carlsson,D.J., and Wiles,D.M., Macromolecules, 7,

259 (1974).

114. Ranaweera,R.P.R., and Scott,G., Chem.Ind., (London),

774 (1974); J.Polym.Sci., Part B, 13, 71 (1975).

115. Carlsson,D.J., Suprunchuk,T., and Wiles,D.M.,

Polym.Lett., 11, 61 (1973).

116. Marek,B., and Lerch,E., J.Soc.Dyers Colour.,~, 481

(1965) .

117. Adams,G.E., Willson,R.L., Aldrich,J.E., and Cundall,

R.B., Int .J .Radiat .Biol., 16, 333 (1969).

118. Pailthorpe ,M. T., J. Phys. E. , ~' 900 ( 19 75) .

119. Gur-Arieh,Z., and Ingarnells,W.C., J.Soc.Dyers Colour.,

~, 12 (1974).

120. Jones ,F. , in "The Theory of Col oration of Texti 1 es",

Ed.C.L.Bird and W.S. Boston, Dyers Company Publications

Trust, Bradford (1975), Ch. 5.

121. Tokaska, J.Soc.Text.Cellulose Ind. Japan, 20, 100

(1964) . 127

122. Pace,R.J., Ph.D. Thesis, University of N.S.W., Sydney,

(1979).