You are on page 1of 8

15

THE THEORY OF DYEING

The Theory of Dyeing


F. JONES
Department of Colorir Chemistry and Dyeing, University of Leeds, Leeds LS2 9JT
Introduction
Previous reviews of research work directed toward a greater
understanding of the way in which dye molecules are transferred
from the dyeing medium to the polymer or substrate phase have
stressed that a unified fundamental theory, applicable to all
dyeing processes, is still far from a reality and may never be
attained. The main reason for this is that in any dyeing process
there are many variable parameters, some or all of which are
mutually dependent. To achieve any progress in such studies, it is
necessary to control these parameters so that the effect of each on
the dyeing system can be determined. This may not always be
possible. Thus, altering the dye concentration within an aqueous
dyebath in order to study the concentration changes of dye
within a substrate may, even where all other conditions can be
maintained constant, produce a change in the structure of the
solvent and a possible alteration in the nature of the dye species
partaking in dyeing.
Some of these variable parameters in the molecular dyeing
theory for ionic and non-ionic systems have been discussed
recently by McGregor and Peters (1). Structural features of both
dye and polymer which may influence the thermodynamics and
the kinetics of dyeing include:
(I) the nature, concentration, distribution and degree of ionisation
of ionisable groups in the dye in the solvent and substrate phases
(2) the molecular and ionic interactions of all the species present in
both phases
(3) the volume fraction, configuration and distribution of both
the crystalline and the amorphous regions, and the degree of ionisation of ionisable groups, in the substrate
(4) the existence of reversible and non-reversible stresses within
the polymer before and during coloration
( 5 ) structural changes in the solvent distributed betwaen the two
phases.
Since these changes are not independent, the researcher has to
adopt several simplifications and use model experiments where
variables can be controlled.
The results obtained in this manner can be compared with
calculated results obtained from theoretical models utilising the
same number of variables. Model systems can therefore be used
only to illustrate specific points in dyeing theory. One such model
(2), used to illustrate the equilibrium values and kinetics of
dyeing of non-ionic dyes, is based on simple mixing theory in
which dye molecules may be treated as occupying mean positions
in a quasi-crystalline or liquid (substrate) lattice. The fundamental concept is identical with that (3) describing the thermodynamic behaviour of non-ionic solutes in which the solute
structure in the solvent is comparable with that of a supercooled
liquid. It can thus be shown that the partial molar enthalpy of
solution, AE, of disperse dyes in the fibre is directly related to the
melting point of the dye according to the expression
where 4 and 4, represent the site fraction of the dye at temperatures T and Tm, the dyeing temperature and melting point,
respectively. The value 0,,, may be obtained by extrapolation.
Eqn 1 is equally valid for the solution of disperse dyes in water.
As certain non-ionic dyes can exist as metastable liquids at
temperatures as low as 20C (4), the possibility of interpretation
inherent within this model may fruitfully bear further examination. The model is, however, subject to serious limitations. It
applies only to ideal systems, i.e. those of low equilibrium dye
concentration, and can be applied only when it is known that

water plays no part in determining the equilibrium saturation


value of the dye in the substrate. More recent work (5) on the
adsorption of azobenzene and p-nitroaniline vapours by substrate films in the presence and in the absence of unsaturated
water vapour has shown that the equilibrium saturation values
of these compounds in secondary cellulose acetate are inversely
related to the amount of water vapour absorbed. The simple
binary mixing theory of dyeing may be more successfully
applied to hydrophobic polymers such as polypropylene, where
water plays an insignificant role in the absorption of dye.
By using the same basic assumptions (2). with regard to the
state and distribution of dye within the polymer, and applying
phenomenological equations, it can also be shown from this
model that the kinetics of dyeing of hydrophobic fibres can be
expressed by
C,= S[I -exp(-8r)l
. . . . (2)
where Cr is the total amount of dye absorbed at time t, S is the
solubility of the dye within the fibre and 8 is a rate constant.
Comparison of experimental rates of dyeing with Eqn 2 for
assumed values of B shows good agreement, although this
expression does not take into account a n y localised variation of
activities existing at surfaces and phase boundaries.
Much published research on dyeing and coloration is directed
towards the further understanding of practical commercial
processes, and comparatively little is published on what is
sometimes regarded as academic dyeing theory. Many experimental results from the former group would be of more value to
the theoretician if the chemical structures and purity of additives
in the dyeing process, dyes and polymers could be disclosed and
the conditions during dyeing systematically defined. Although
much is published, this review has therefore had to be limited to
salient papers in which these restrictions have been met.
Dyeing theory is concerned with the thermodynamics and
kinetics of processes occurring within the dyebath, the interaction
between dyeing species and the internal and external surfaces of
fibres, transfer from these surfaces by diffusion processes to
equilibrium positions within the substrate and the behaviour of
dyes once this type of equilibrium has been achieved. It is
proposed therefore to discuss recent advances in dyeing theory
under these headings and to include recent work carried out in
less conventional systems, such as heat-fixation methods. Since
coloration with reactive dyes entails simultaneous physical
absorption and chemical reactions within the dyebath and within
the substrate, this topic will be considered separately, although
this division is somewhat unreal. By this means it is hoped that
the reader will be given a more comprehensive view of dyeing
theory without any artificial division based on substrate types.

State of Dyes in Solutions and Dispersions


Concepts concerning the structural nature of water (6), the
presence of cavities and ice-like clusters of water molecules that
are not inter-cluster hydrogen-bonded, and the influence of
solutes in changing these structures have meant that ideas on the
structure of aqueous dye solutions have been modified in recent
years. This has been particularly relevant to studies of the
phenomenon of association of dye species in the solution.
Coates (7) has surveyed the forces of interaction between like
ionic species which can lead to dimerisation and higher degrees of
association and has discussed this process from its thermodynamic and kinetic aspects. Although the standa.rd free-energy

16
change involving both enthalpy change and entropy change in the
dimerisation for any dye can be determined from the equilibrium
constant governing the reversible formation of dimer from two
monomeric species and its temperature dependence, the values
obtained in practice depend on the accuracy of the means by
which the equilibrium constant is obtained.
Thus, to a first approximation the dimerisation of a number of
ionised monosulphonated dyes and of some positively charged
basic dye cations is accompanied by standard entropy changes of
from - 10 to -20 cal deg-1 mole-1. This decrease is made up in
part by a loss in translationalentropy of the monomeric ion and in
part by a gain in entropy of the water. This gain of entropy by the
solvent is due to the probable decrease in the structural order of
the water promoted more effectively in the vicinity of a monomeric ion than in the region of a dimeric species. Conversely,
entropy values obtained for non-ionic dye vapours, which exist
mainly in the form of dimers (a), show that, under anhydrous
conditions, the decrease in entropy related to dimer formation is
only of the order of -1.0 to -2.0 cal deg-1 mole-1. This
decrease is therefore much less than that generally observed for
dimerisation in solution. It must, however, be pointed out that
association in the vapour is between molecules that, although
polarisable, do not possess a fully developed charge. It appears
then that the structural nature of water promotes association and
explains the large equilibrium shift towards monomer formation
at high temperatures, since the structure of the solvent is strongly
dependent on temperature. If association of dye molecules were
dependent only on the forces of interaction contributing to
hydrophobic bonding, then the phenomenon would also be
observed in other solvents. Since association is much less marked
in organic solvents of low dielectricconstant, it can be concluded,
at least for basic dyes possessing a non-localised positive charge
(9), that increased water-water interactions overcome the
repulsion forces acting between dye ions of similar charge.
The possibility that dye ions associate in solution is very real
and, if this is not taken into account, errors in determining such
parameters as ionisation constants can be considerable. The
problem can be overcome to some extent by using very dilute dye
solutions, where the probability of collision may reasonably be
expected to be low. Even so, self-associationoccurs at concentrations as low as one milligram of dye anion per litre of solution
(10). In determining ionisation constants for a number of
monosulphonated 00'-dihydroxyazo mordant dyes, it was
necessary to use mixtures of dioxan and water to overcome
association effects and extrapolate the ionisation constants to
values for pure water. It was then possible to compare the effect
of association on the ionisation constant related to the ionisation
of the first hydroxyl group in these dyes. This is increased when
association occurs. Under more alkaline conditions both
hydroxyl groups are ionised and association is minimised or
eliminated since the repulsion forces between the trivalent fully
ionised entities are much larger.
The influence of additives such as urea ( I I ) , formamide (12),
N-alkylacylamides (13) and alcohols (9) on the structure and
properties of dye solutions has recently been studied. It is
generallyaccepted that additives of this type induce disorientation
of the water structure in the vicinity of the dye ion, thereby
reducing aggregationattributed to loss of hydrophobic interaction.
It has been pointed out (14) that this mechanism has not been
conclusively proved, although no doubt a decrease in dye-dye
interaction by the action of urea on dye solutions does occur, and
leads to increased rates of dyeing. The swelling action on protein
substrates and reduction of hydrophobic interaction in the
substrates by urea also contribute to an increase in the rates of
dyeing, thus illustrating the interdependence of parameters
mentioned earlier.
Non-ionic disperse dyes possess very low solubilitiesin aqueous
dispersions at the dyeing temperature, and association, which

REVIEW

OF PROGRESS IN COLORATION; JONES

may occur in.the absence of formally charged structures, may be


very difficult to detect by conventional means. Anomalies in
rates of dyeing found by McDowell and Weingarten (15) in
applying four pure disperse dyes to polyester material have, on
the other hand, been interpreted in terms of an increase in
particle size of the dye, leading to lower aqueous solubility. This
reasoning assumes that the rate of dyeing is directly related to the
concentration of dissolved dye molecules. When the pure dye is
pretreated in boiling water, the rate of dyeing in some cases
decreases and in others increases. These authors had earlier (16)
drawn attention to the classical equation relating solubility to
particle size and molecular weight, viz.

.. . (3)
where S, is the mean solubility of a particle of radius r, y is the
free surface energy, p is the density of the solid and S is the
minimum solubilitywhen the particle size is increased. There must
be a maximum value of r for the condition Sr =S and this can be
shown (17) to be approximately 10-2 pm, which is much less than
the mean radius of disperse dye particles. Further, the anomalies
in rates of dyeing were inconsistent, and it is now suggested that
the inconsistency may be explained by the formation of different
structural modifications of the solid dye during dyeing. These
modifications could be verifiable from X-ray diffraction data.
That this possibility has not been put forward by the authors is
surprising, since in another context (18) it is stated that X-ray
diffraction data are obtained as a matter of routine.
Solid-state transitions and polymorphic changes occurring
through solution and recrystallisation mechanisms are well
established in non-ionic dyes and pigments. In azo pigments
transitions occur on heating the pigment in an aqueous environment (19), and Apperley (20) has recently studied the influence of
surface-active agents on the morphology of C.I. Disperse Yellow
3 at the coupling stage. If such changes are occurring in dyebaths,
then further research on specific systems from this aspect may
throw considerable light on anomalous results obtained in
dyeing research.
Jhetics of Dyeing

Diffusion processes in dyeing are essentially those describing


the mass transfer of dye from the external aqueous phase to the
interior of the substrate, and the distribution of dye within this
substrate up to its saturation equilibrium value. A complete
description of transfer mechanisms should therefore include
possible formation of a boundary layer in proximity to the fibre
within the aqueous phase, the boundary conditions at the interface between this layer and the fibre, a possible diffusion layer
between the interface and the fibre interior, and the mechanism by which the dye is transferred to the centre of the fibre.
When possible changes with respect to concentration and the
effect of additives other than dyes in the above phases are considered, in relation to structural changes occurring within the
substrate, it can be seen that the overall dyeing mechanism is very
complex. Fundamental studies of rates of dyeing are therefore,
where possible, limited to conditions that will allow assessment of
transfer processes with a minimum of dependent parameters.
Weisz and Zollinger (21) have considered the transfer of dye in
solution within substrate capillaries in which partial sorption or
immobilisation of the diffusant occurs. In this model, where
sorption of this type is reversible,the apparent diffusion coefficient
DA is defined by

.. ..(4)

where P is the fraction of the substrate accessible to diffusion


processes, b is a tortuosity constant, which is less than 4 3 , Cf/C,
is the ratio of the total concentration of dye, Cj; per unit volume
of fibre to an external dyebath concentration C,,, at equilibrium.

THE THEORY

OF DYEING

n is the equilibrium partition coefficient between the mobile


portion of the internal dye concentration and the external dye
concentration, and D is the true diffusion coefficient of the
mobile dye molecules within that region of the substrate accessible
to diffusive motion. The constant a depends on the particular
form of the absorption isotherm and falls within the range
1*&l a6 depending on the affinity of the dye in the system.
Eqn 4 is of general applicability and has been applied to an
earlier model for diffusion in cellulose (22) when the dye is
considered to be migrating along water-filled pores with simultaneous adsorption. By assuming that the accessibility parameter,
P,can be equated with a fractional uptake of water by the fibre
and that the dye concentration is the same in the internal and
external aqueous phases, i.e. n = 1, for dyes of high affinity where
a approaches 1 6, the equation has been successfully applied to
previously published results on diffusion. Although there must be
some reservation on the assumed values of n and a, it is interesting to observe that, for a direct dye, the value of the true diffusion
coefficient, D, of the mobile dye molecules within the fibre is very
similar in magnitude to the value for the diffusion coefficient of
dye molecules within the bulk aqueous phase. It is concluded that
for this dye the pore model with partial dye sorption is adequate
to account for the kinetics of dye transport. Agreement is not as
good for acid dyes on nylon, but, with further assumptions about
the aqueous solubility of disperse dyes, values for D obtained
from the general Eqn 4 for disperse dyes applied to polyesters and
cellulose acetates closely follow the D values for diffusion in
water alone.
The general applicability of this equation cannot, however, be
substantiated until further quantitative data on dye solubility and
free aqueous diffusion coefficients have been obtained. Very
little recent information is available on the latter. Murfet (23) has
determined the relative diffusivity of C.I. Acid Red 1 using a
vertical diffusion cell with a sintered-glass membrane. Although
the true aqueous diffusivity of the dye cannot be determined,
since the tortuosity and effective area of the sinter are not
amenable to absolute measurement, it is interesting to note that
relative diffusivity decreases in the presence of urea, in contrast
with the fact that urea causes an increase in the rate of dyeing of
this dye on wool (14). One possible interpretation is that urea
accelerates dyeing by influencing other parameters such as the
structure of the boundary layer and the substrate, while having an
opposing effect on the migration of dye in the internal aqueous
phase.
Although Ficks equations are often used as a basis for
diffusion studies, it is now accepted that theoretically determined
rate curves are comparatively insensitive to initial and boundary
conditions and to differential equations used in their derivation.
To gain more detailed information on the diffusion mechanisms,
it is necessary to obtain concentration-distance profiles, usually
by cross-sectioning fibres or films and determining dye distribution by microdensitometry. This has been done (24) for the
dyeing of various polymer films with disperse dyes in the presence
and in the absence of carriers. That in conventional dyeing
processes there are at least three components-dye, substrate and
solvent-leads to a complex situation. The addition of carrier is a
further complication. Even with the simplest systems rates of
dyeing cannot be adequately dealt with in Fickian terms, since
the latter are expressly concerned with a binary system.
According to McGregor et al. (24), each component will
occupy a definite fraction of the fibre phase, so that, when
restrictions are imposed on this total volume, we might expect
that an inward flow or volume flux of dye would be accompanied
by an outward flow of dyebath medium and that these opposing
flows might interact. This relative motion of molecules in a
multi-component system can cause a change in volume or a
hydrodynamic bulk flow of the system. It becomes necessary
then, particularly at high diffusant concentrations, to correct the

17
measured diffusion flow to take account of any hydrodynamic
transfer of the component under investigation. This could be
achieved by taking into account a frame of reference within
which the diffusion process is measured. The reference frame may
be delineated in several ways. By conducting diffusion experiments in pre-swollen fibres or films,it can be assumed that no
further change in volume occurs on dyeing, and here the reference
frame is fixed with respect to the surface of the fibre or film.
Under these conditions Ficks laws d o appear to apply and
reference-frame effects may become important only in carrier
dyeing and in dyeing fibres that have not been pre-swollen.
Irreversible changes in polyester structures have recently been
observed (25)in carrier dyeing with disperse dyes.
A method for determining concentration-distance profiles
without cross-sectioning and thereby standardising the frame-ofreference effects still further has been developed by Blacker and
Patterson (26). The method utilises the continuous changes in
transmitted monochromatic light when a dyed filament, of
circular cross-section, is scanned across its longitudinal axis by
moving the filament across a narrow slit. A microspectrophotometer is used for this purpose, the results being suitable for
computer calculation to determine the dye distribution across the
fibre. Changes in profile shape for a number of disperse dyeings
on polyester and nylon 6.6 filaments over a range of dyeing times
were obtained. The profile shape and the observation that in all
cases a time-dependent increasing surface concentration of dye
occurs showed that the rate of transport of dye to a boundary
just within the surface of the substrate is no higher than that at
which dye is transferred to the interior. It is also interesting to
observe that for nylon 6.6 this latter rate is extremely high, even
during the initial stages of dyeing, since horizontal profiles were
obtained. This behaviour is difficult to interpret unless it is
assumed either that the substrate is behaving like a liquid or that
the driving force for diffusion is a variation in activity and not
concentration of the diffusing species.
When the dye and the substrate possess charges of opposite
sign, which usually applies to the nylon-acid dye system under
acid conditions, the charge on the fibre becomes increasingly
negative during dyeing. This happens in the dyeing of nylon
with Orange I1 (C.I. Acid Orange 7), the surface potential
becoming increasingly more negative as dye concentrations both
within (27) and on the surface (28) increase. The initial surface
potential depends on the history of the substrate. Bell (29) has
found that rates of dyeing of acid dyes on nylon 6.6 are directly
proportional to the surface area and the saturation equilibrium
value of the dye on the substrate. The latter values are considered
to be directly related to amine end-group content, but, in contrast
with Suzawa and Saitos results (28), it is assumed that there is
no possibility of adsorption of dye on the surface. Bells conclusions must therefore be taken with some reservation, particularly when it is noted that dye concentrations in the fibre phase
were estimated solely on the basis of changes in dye concentration
occurring within the dyebath.
When the ion and the substrate are oppositely charged,
interaction between species can lead to additional factors in the
interpretation of molecular diffusion processes. Some attention
has been paid to this problem by Mayer (30) in the dyeing of
acrylic fibres (negative sites) with cationic basic dyes under
commercial conditions. Diffusion is satisfactorily achieved only
below a certain temperature and strict temperature control is
necessary to achieve a reasonable rate of dyeing consistent with
levelness. Above this maximum temperature, bond formation,
represented by salt linkage, inhibits the attainment of adequate
rates of diffusion. Information to investigate this problem in a
fundamental way is, however, lacking at present.
A generalised treatment for explaining non-Fickian behaviour
has been given recently (31). This type of diffusion, usually
attributed to substrate changes, can also occur when a second

REVIEW OF PROGRESS IN COLORATION; JONES

18
independent process such as an immobilising chemical reaction
is superimposed with a comparable time scale. The presence of
blind pores acting as diffusant 'sinks' in a porous substrate can
also have this effect. Anomalous diffusion may also arise in
systems in which 'thermodynamic' diffusion coefficients, theoretically determined from practical diffusion coefficients and
diffusant solubility, depend both on activity and on penetration
distance in such a way that activity and distance variables cannot
be separately assessed. Solutions to this general problem require
a large amount of computation and are wcrthwhile only in
particular circumstances.
The molecular interpretation of diffusion of dyes in polymeric
systems is based on the strong dependence of the apparent
diffusion coefficients D A on temperature, which often follows a
simple Arrhenius equation, viz.
DA

Do exp( - E / R T )

.. . .(5)

where D,,is a pre-exponential or proportionality constant and E


is the activation energy of diffusion. This energy is required by
the diffusing molecule to enable it to 'jump' from one absorption
site to a vacant neighbour. The pre-exponential factor is related
to the jump distance and entropy of diffusion. For more hydrophobic fibres, the activation energy of diffusion is related to the
energy of 'hole' formation within the polymer which allows the
dye molecule to diffuse. I n this respect it has been observed (32)
that the activation energy of diffusion for disperse dyes in
unmodified polypropylene is higher (by approximately 14 kcal/
mole) than that observed when the same dyes are applied to
cellulose acetate. The difference may be due to the temperature
dependence of interchain bonding in polypropylene and to the
swelling of cellulose acetate caused by absorbed water. In
confirmation, it has previously been found that desorption from
vapour-dyed polypropylene film occurs simply on cooling,
whereas desorption from vapour-dyed cellulose acetate occurs
only in the presence of water vapour. I t is therefore possible that
at high temperatures diffusion into polypropylene is a process
simply of mixing.
Thermodynamics of Dyeing Processes
When diffusion is allowcd to contitwe until n o further dye is
absorbed, the dye-substrate-solvent system can be considered
to be in a state of dynamic equilibrium provided that the physicril
and chemical forces appertaining to the processes of adsorption
are completely reversible, there being, for instance, no permanent
change in the structure of the substrate during dyeing. The
variation with temperature of the amount of dye absorbed at
equilibrium (the equilibrium sorption value) allows thermodynamic quantities such as the standard affinity or free-energy
change, and heat and entropy of dyeing to be determined.
Application of these concepts to dyeing systems, however,
requires certain assumptions about the ideal behaviour of the dye
species in the solvent, the internal solution and substrate phases.
Assumptions have also to be made about the internal or available
volume in the substrate within which absorptionoccurs.
The free-energy change, lc , i n the transfer of one mole of dye
from its standard state in solution to its standard state in the
fibre is given by

ic

.lF

--TAT

. . . . (6)

where .lF is the standard enthalpy or heat-content change in the


process and 15 is the corresponding entropy change. The standard state of the dye may be arbitrarily defined for both phases.
Energies of bonding between dye and substrate for different dyes
may be compared only on this basis within the limits of the above
assumptions. High heats of bonding between dye and substrate
molecules indicate a large change in entropy or decrease in
randomness of dye molecules on absorption, but it has been
pointed out (33) that such correlations may be spurious when
LIZ'and .IF'are obtained from the same set of data, and confirmation is required by mathematical transformation.

lyer 'el a/. (34) have applied Eqn 6 to show that a relation
between A H ' and 4 3 does exist in the dyeing of cellulose with
Chlorazol Sky Blue FF (C.1. Direct Blue 1) and that 4Ecvalues
increase with increasing size of alkali-metal cations present
during absorption. Calculated values of activity of the dye in
solution were used together with a variable substrate-volume
parameter defined previously (35) as the product of the surface
area available for dye sorption and the thickness of the diffuse
double layer existing between the bulk dyebath phase and the
outer surface of cellulose. As A i r increases in this manner there
is a corresponding increase in the value of the entropy change
uhich is considered to be due to different packing arrangements
of dye molecules at the surface of the substrate. The interpretation of these results, however, in terms of a breakdown in water
structure in the vicinity of the cellulose in the presence of large
cations (as is done by the authors) must be treated cautiously,
since thermodynamic data are concerned only with differences
in initial and final states and can give no information on the
mechanism by which the final state is approached. It is interesting
in this connection to note that, in the dyeing of cellulose with the
leuco anion of a non-symmetrical vat dye-a process similar to
the application of direct dyes-stacking or association of the dye
anion occurs on the fibre, and this leads to an oxidised dyeing
in which associates are present before oxidation (36).
More attention has been paid recently to research on the
dyeing of wool and nylon with acid dyes. Interaction has generally
been considered as electrostatic bonding between dye anion
and positively charged sites such as protonated amine groups
existing in the substrate under acid conditions. This conclusion
has been deduced from studies of sorption isotherms where no
further increase in dye sorption beyond the value equivalent
to the number of charged sites has been obtained. The possibility
that van der Waals forces arising from dipole-induced dipole
interaction and dispersion forces also operate must not be excluded. Possible substrate changes, particularly when the attainment
of equilibrium is prolonged, leading to the exposure of sites at
which interaction with dye anions may occur, must also be considered
One approech to reduce the number of these parameters (37)
has been to study the absorption of dye anions that normally
have no substantivity for cellulose, by a cellulose substrate
modified by conversion of some of the hydroxyl groups to
8-aminoethyl groups. Any dye bonding occurring should therefcre be specific to the amino groups. Thermodynamic affinities
and heats of dyeing show that the dye anion has interacted with
the protonated amino group. Changes in accessibility compared
with unmodified cellulose are shown to be negligible, but the
modification reaction in which polyethyleneimines having
substantivity for the substrate (38) are formed may lead to a
substrate in which more than one type of adsorption site is
present. The assumption of lack of substantivity of these acid
dyes for cellulose and their precise mode of interaction with
amino groups have been called into question (39), but in a simpler
system (40), when the same dyes have been applied to aminopolypropylene under acid conditions, the correlation between
protonated amino groups and equilibrium sorption values has
been explained on the basis of a simple ion-exchange model.
Although Langmuir-type adsorption isotherms indicate a
probable stoichiometric relation, amounts of dye absorbed over
and above the limiting value (overdyeing) give rise to Freundlichtype isotherms. This phenomenon is attributed to the presence. of
associated dye species within the dyebath. Theoretical expressions
describing differences in the two types of isotherm for acid dyes
on nylon depend not only on the degree of association of dye in
the dyebath but also on the equilibrium established between
sorbed and mobile dye anions within the polymer phase (41).
This latter type of equilibrium behaviour has been considered

THE THEORY OF DYEING


by Marshall (42), particularly for Orange I1 applied to a wool
substrate at the isoelectric point. By using a model based on
DoMan equilibrium partition and applying the condition of
electrical neutrality in internal and external solutions, variations
in dye-sorption isotherms can be attributed to the influence of
dye Concentration on the equilibrium between absorbed dye and
mobile dye in the internal solution. In approximating estimated
values of this equilibrium constant over a range of concentrations, it was necessary to vary the equilibrium constant to give the
best fit to the experimentally determined isotherms. This variation
may be used to determine the activity coefficient and hence the
degree of association of the dyes within the internal aqueous
phase, and, when this was calculated, the values of the activity
coefficient were similar to those calculated for the external dyebath phase.
In the thermodynamics of dyeing hydrophobic fibres with
disperse dyes, isotherms that are linear up to the point of saturation with respect to the dyebath phase are usually obtained.
Since most disperse dyes have very limited solubility in water,
even at the dyeing temperature, experimental difficulties arise in
determining whether such solutions are monomolecularly dispersed or contain associated species. Recent research by
McDowell and Weingarten (18) has not revealed any conclusive
results on this point, but it was shown that at 120C non-linear
isotherms could be obtained with 1-amino-4-hydroxyanthraquinone on polyester film. The isotherms were linear up to the
saturation solubility of the dye in the aqueous phase, the gradual
approach to a maximum value for uptake within the film beyond
this point being attributed to changes in the substrate. The heats
and entropies of dyeing for a number of dyes, when determined
from absorption isotherms, were higher than the values determined from the temperature coefficient of the ratio of solubilities
of the dye in polyester to the solubilities in water. In thermodynar+ic terms this difference is explained (43) by the Fact that
the heat of dyeing determined from sorption isotherms is an
integral heat of dyeing where the total heat evolved is due not
only to the interaction between dye and substrate molecules but
also, in the later stages, to the interaction between entering dye
molecules and substrate that already contains dye molecules.
The heat of dyeing obtained from solubility ratios approaches the
value for a heat of interaction between dye molecules and a
substrate containing dye. In the limit the difference between the
two is equal to RT, which at 150C (the maximum dyeing temperature used) is 0.84 kcal/mole. The experimentally determined
difference is greater than this value and lies in the range 1.78-4.45
kcal/mole for the dyes considered. Two contributory causes are
possible. The first is that association of dye could occur in the
substrate and the second that contributions to the experimentally
determined heats of solution of the dyes in water may arise from
heats of solid-state transitions occurring in the dye-solid suspension. Whether disperse dyes are associated in the substrate is
still an open question (see below) and the possibility of solidstate transitions occurring in disperse dye suspensions needs to
be more fully investigated.

Chemical Reaction with Substrates and Fibres


Although use is made of conventional physico-chemical
concepts in elucidating the mechanisms of reaction of coloured
compounds with substrates, discussion of the subject has been
arbitrarily divorced from normal dyeing theory in the past
since reactive-dyeing mechanisms involve not only interaction by
physical ionic and dipole-induced dipole and dispersion forces
but also formation of covalent bonds with the substrate. With
reactive dyes, simultaneous hydrolysis of the dye by water in
both external and internal phases can occur. The resultant
changes in chemical structure of both dye and substrate during
dyeing therefore lead to more complex mechanisms of absorption
and fixation.

19
Rattee has recently reviewed (44) the chemistry of these
reactions from a kinetic standpoint. In view of the heterogeneous
nature of the dyeing process, the initial model system used has
been one in which a homogeneous reaction phase-soluble
reactive dye, soluble alcohol acting as the substrate and wateris considered. Reaction of dye with alcohol is analogous with
fixation of dye on cellulose. The fixation and hydrolysis reactions
are bimolecular, although the total reaction normally behaves as
a pseudo-unimolecular reaction because water and alcohol are
present in excess. It can be shown that, under conditions where
the alcohol is very slightly ionised, the ratio of the rate constants
of fixation and hydrolysis, i.e. the reactivity ratio, Z , must be
constant at any temperature and be independent of pH conditions.
At higher pH values, the ionisation of the alcohol may become
significant and the reactivity ratio in this case is no longer constant but depends on pH. When the model is extended to include
cellulose, which contains ionisable primary and secondary
alcoholic groups, the situation becomes more complex, since
reactions may proceed at different rates in the fibre and in the
aqueous phase. The distribution of dye between the two phases
assumes a greater significance under these conditions and the rate
of diffusion into the fibre also plays a part. Hydrolysis occurs
both in the dyebath and in the internal water phase, but for the
purpose of this discussion the latter effect can be shown to be
minimal. Simultaneous reaction of the dye with the fibre and
diffusion of the dye within the fibre can be accommodated by
applying a simplified Danckwerts equation (44) i,n which the
efficiency of fixation E, defmed as the ratio of the rate of reaction,
dfldr, to the rate of hydrolysis, dhldt, is given by

where [ D ] Fis the constant surface dye concentration, [Ills is the


dye concentration in the dyebath, Z is the previously mentioned
reactivity ratio, [C-] is the concentration of ionised hydroxyl
groups in the cellulose and KH is the bimolecular hydrolysis
constant. The apparent argument therefore is that the prime
factor determining the efficiency of fixation is the ratio [D]F/[D],,
the substantivity ratio. The diffusion coefficient, D, and the
reactivity ratio have only minor effects, as do pH conditions.
The last-named play a secondary role since the hydroxyl-ion
concentration in the external phase ([OH -I,,) influences the value
of [C-1.
Strictly, this hydroxyl-ion concentration is related to the
hydroxyl-ion concentration within the internal aqueous phase
and the situation is in practice more difficult to interpret.
Changing the pH of the dyebath will cause a more or less
negative potential to develop at the substrate surface and,
where the surface potential becomes more negative, mutual
anionic repulsion between the surface and dye anion will cause
the value of [D]F to decrease. Assuming that the dyebath
concentration is constant, the substantivity ratio will also
decrease. This effect, however, can be mitigated by increasing
the pH, which causes an increase in the diffusion coefficient.
From a practical point of view, therefore, any process that leads
to a high degree of exhaustion of the dye will tend to give a high
substantivity ratio and consequently improve the overall efficiency
of the dyeing operation. It is interesting to observe (46) that the
influence of urea in improving the efficiency of reactive dyeing
cannot be due solely to an increase in dye solubility, since this
effect would reduce the substantivity ratio where the amount of
bonded dye [ D ] Fis unchanged. Improved efficiency is therefore
attributed to an increase in the diffusion coefficient on increasing
the concentration of dye in the bath.
Practical confirmation of the validity of the modified Danckwerts equation (7) cannot be obtained under conditions where
dye hydrolysis and fixation occur simultaneously. Sumner and
Taylor (47) have attempted to overcome this problem in a
serni-quantitative way by considering the dyeing behnviour of

20
reactive dyes when hydrolysis is at a minimum. This occurs under
slightly acid conditions for reactive dyes on cellulose. If, now, the
dyeing behaviour of dyes containing non-reactive residues but
similar in structure to reactive dyes is observed over a range of
pH values, it can be assumed that the changes in affinity and
rates of diffusion of these inert dyes will be paralleled by similar
but theoretical changes occurring with the reactive dyes. The
affinity and diffusion coefficients of the latter can therefore be
determined over a pH range by extrapolation. For two out of the
three dyes examined in this way, calculated values of D increased
and values of [ D ] F decreased with increasing alkalinity. In
contrast, the third dye exhibited a minimum in its D values and
amaximum for [ D l ~ apH
t I 1 *5.
Hydrolysis in the dyebath of dichlorotriazine reactive dyes in
which the chromogen is linked to the reactive residue through an
imino group may not be a simple second-order reaction but may
be complicated by the presence of a deprotonated imino form
(-i;j:)Bxisting as a result of acid-base equilibrium between the
latter and the imino form (-NH-) itself (48). The products of
hydrolysis of each form will be identical, although the rate
constants for hydrolysis for each form will differ. These constants
and the acid-base equilibrium constant cannot be determined
without a knowledge of the activity coefficients of each species.
Products of activitycoefficientsand rateconstants can, however, be
approximated by computer processes. When this approximation
is carried out, the derived activation energies of hydrolysis are
found to depend on temperature with a maximum curvature at
30C. This illustrates a common feature either in the dye structures or in the system as a whole. The observation of a minimum
in the mean activity coefficient in solutions of Orange 11 at this
temperature (49) determined by differential vapour-pressure
manometry is relevant in this context and may indicate structural
changes in the aqueous solvent in the vicinity of the dye anion
at this temperature.
More recently, the emphasis in reactive-dyeing theory has
shifted towards protein substrates and the investigation of substructures within the protein that can react with the dye. The
elucidation of reaction mechanisms is more difficult than with
cellulose, since reactions are possible with a greater number
of sites of different types, the possibility of fibre degradation
is higher and fibre morphology is more complex. Very little
quantitative information was available until Shore, in an admirable series of papers (50) and adopting the approach previously
taken for the reaction of dyes with cellulose, examined the rate
of reaction of a monochlorotriazine dye with a number of model
compounds related in structure to the amino-acid residues in proteins. If it is assumed that the reactivities or dissociation constants
of the groups in the protein are unaffected by their neighbours
then their reaction rates and activation energies will be comparable with those of model compounds in aqueous solutions.
By such comparison the order of relative reactivity in watersoluble proteins is cysteine thiol > N-terminal amino > histidine
> imidazolyl, etc., down to lysine amino and serine alcoholic
groups. This assumes that the availability of the groups in the
protein to the reactive dye is equally as great as their availability
as model compounds in a homogeneous solution, but this is not
very likely. By extending the study (51) to include water-insoluble
proteins of known composition the most important groups to
react were again shown to be the cysteine thiol, the primary amino
groups of lysine and N-terminal amino-acid residues. It is
important to note that the thiol groups are reactive over the
whole pH range, whereas primary amino groups exert an influence only under alkaline conditions. Conditions with respect to
pH in kinetic and thermodynamic studies will differ from those
adopted in the reactive dyeing of cellulose. It is not possible then
to adopt comparative techniques such as those of Sumner and
Taylor (47), since the dye reacts readily under neutral or slightly
acid conditions.

REVIEW OF PROGRESS IN COLORATION; JONES


By using a mixture of hydrolysed and reactive dye and applying
an equation resulting from a combination of Danckwerts'
equation and Hill's equation for diffusion into an infinite cylinder
with the condition of a saturated fibre surface, Shore has shown
(52) that under acid conditions the diffusion coefficient of both
dye species increases with pH. Above pH 4.0reaction of the dye
with wool predominates and below this pH reactive dye preferentially hydrolyses. Under neutral conditions the reactivity of the
dye for a wool substrate was less than the theoretical level,
suggesting that there may be specific chemical hindrance to the
reaction in the solid phase. A similar observation had been made
previously in comparing the rate constants for reaction of a dye
with a water-soluble alcohol and with cellulose (53).
Independent confirmation that hydrolysis of reactive dyes is
minimised in the pH range 4-6 had been given by Lewis and
Seltzer (54). In the pad-batch dyeing of wool at room temperature
with dichlorotriazines under weakly acid conditions, the degree
of fixation approached unity, particularly in the presence of
additives such as sodium metabisulphite. It is considered that the
high rate of reaction may be due to the formation of thiol groups
in the wool arising from the reduction of disulphide bonds. It
may also be added that in the absence of reducing agents, under
acid conditions, protonation of the triazine nitrogen atoms may
increase the activity of the dichlorotriazine dye and allow
increased reaction with un-ionised thiol groups. Simultaneously,
the hydrolysis reaction would be minimal since a very low concentration of hydroxyl ion would be present. The specific
reaction with thiol groups in model compounds and substrates
requires further investigation.
CHROME MORDANTING

Although not usually regarded as reactive dyeing in any sense,


the chrome mordanting of wool also entails direct chemical
reaction of the chromium with the fibre. Chromium is usually
applied as a hexavalent cation, in which form it is able to diffuse
into the wool. During application, the effect of heat accelerates
the reduction to trivalent chromium, probably by the action of
the substrate. The Cr(1II) can then react with the wool at a
comparatively low rate. Hartley (55) has considered the possible
reaction sites, reaction being excluded at amino, thiol or phenolic
groups (56), and concludes from pH changes during absorption
and from spectral data that the ionised carboxyl groups in the
wool take part in the formation of metal complexes.
This reaction is interpreted in terms of a probable first-order
nucleophilic reaction whereby the presence of at least one
strongly negative ligand, e.g. sulphate, in a water-saturated
6-co-ordinated chromium cation facilitates the loss of a water
ligand. This loss occurs as the rate-controlling step in the reaction
by the displacement of charge from the strongest negative ligand
to the central metal cation. The resultant penta-co-ordinated
transition-state entity is then in a sterically favourable orientation
to react with an ionised carboxyl group in the substrate. In
support of this reasoning, the more negative ligand remains
co-ordinated with the chromium after reaction. Although this
proposed mechanism still leaves room for 1 :I dye-chromium
complexes to be formed by ligand replacement, the existence of
this type of complex in wool has never been proved. The formation of 2:l dyeechromium complexes requires displacement of a
wool-carboxyl-ligand, and further research is necessary.
High-temperature Dyeing under Anhydrous Conditions
The process of padding a fibrous material with a dispersion
of a non-ionic dye, drying and then heating the treated fibre to a
high (190-220C) temperature for a short period to produce a
dyeing is by now familiar. The greatest use of such a process,
e.g. the Thermosol method, is in the coloration of polyestercellulose mixtures to dye the polyester component. Dyeing occurs

THE THEORY OF DYEING

21

under high-temperature anhydrous conditions. The mechanisms


by which dye molecules are transferred from the solid particle to
the substrate have been variously described as particle dissolution
in the substrate (57), a partial contact mechanism (58)or an
evaporation and dye-vapour absorption process (59). By similar
but independent experiments Datye (60)and Sumner et al. (61)
have shown that, on padding and drying, the dye particles and
solution are preferentially taken up by the cellulose component,
whereas in the fixation stage the dye is preferentially absorbed
by the polyester component. It follows, therefore, that the
vaporisation of the dye plays an important part in the transfer
mechanism. Contributions to the dyeing mechanism by transfer
of dye across dyed and undyed fibres through sites of contact or
by transfer through the medium of additives may or may not be
relatively important.
Both Sumner et al. and Datye have found that dispersing
agents or migration inhibitors have n o effect on the rate of
transfer. By determining the amount of dye absorbed when the
polyester substrate is placed at various fixed distances from the
source of dye, an approximately linear relation between the
amount absorbed and distance is found, from which the amount
absorbed at zero distance, i.e. when the dye source and polyester
are in contact, can be extrapolated. Since there is good agreement
between these extrapolated values and those experimentally
determined, for a number of disperse dyes over a range of
temperatures, Sumner et al. conclude that a single transfer
mechanism, viz. by vaporisation and absorption of dye vapour,
is adequate to explain the amount of dye absorbed. Confirmation
is given from linear plots of ln[D]/, where [D]/ is the amount
of dye absorbed in unit time, against reciprocal temperature,
It is assumed that the rate of absorption is controlled specifically
by the rate of diffusion d[D],/dt of the dye vapour in air. The
rate is directly related to the diffusion coefficient in air, D,and
to the vapour pressure of the dye p and inversely related to the
distance x between source and substrate, i.e.
In(d[D],/dr) In D Inp - In x
. . . . (8)
Sincc In D is proportional to reciprocal temperature and by
applying the Clausius-Clapeyron relation between p and T,
Ean 8 mav be exmessed as
In(d[D],/dr) In K C - In x --

where K, C and N are constants and L is the latent heat of


sublimation of the dye at a total pressure of one atmosphere.
Plots of In[rate of transfer] or ln[D]+ at unit time against 1/T
should therefore be linear and of negative slope independent of x.
This has been found experimentally to be the case even when the
value of x is zero. On the other hand, the relation between [D]/
and x determined by Datye (60)was more curvilinear and extrapolation to zero distance could lead to results indicative of an
additional small contribution by a direct contact mechanism.
Dyeing by application of unsaturated dye vapour to both
polyester and nylon substrates at high temperature has been
adequately demonstrated (62).It is of interest to observe (5) that,
when model compounds such as p-nitroaniline and azobenzene
are applied to cellulose acetate by vapour-absorption methods
in the presence of unsaturated water vapour, the sorption
equilibrium values decrease as the concentration of water vapour
within the substrate increases. Since sorption occurs under
equilibrium conditions, the decrease cannot be attributed to a
reduced rate of dye transfer by the presence of water vapour, but
is possibly due to increased competition for sites between sorbate
and water or to changes in substrate structure influenced by the
presence of water. In conventional dyeing, the amount of water
present in the substrate cannot be controlled in this way. The
comparatively high saturation values and their rapid attainment
obtained in high-temperature fixation or solvent-dyeing processes
may therefore be explained by the absence of water in addition
to the high thermal energy of both dye and substrate molecules.

It has been noted recently (63)that the total amount of a dye


in a mixture absorbed by the substrate under heat-fixation
conditions can be less than that absorbed when the dye is applied
separately and that longer fixation times are required. In view of
these observations, it is possible that the vapour pressure of the
dye solid is reduced to that of a more stable polymorph, the
solid-solid transition taking place more readily in the presence
of a second vapour component. The question of the heat stability
and the possibility of structural transitions of disperse dye
particles under conditions of high-temperature fixation requires
further investigation, since no research has been published in this
field.
The Physical State of Dyes in Polymers
U p to this point we have considered the state of dyes in solution
and the kinetics and thermodynamics of dye transfer from the
dyebath to the substrate. Since this transfer process is dynamic
in nature, it is possible that further changes in the physical state
of the dye molecules in the polymer can take place outside the
environment of the transfer medium and more particularly when
the dyed material is heated, washed or exposed to light. It is
generally accepted that changes in colour occurring in the soaping
of vat and azoic dyeings can be attributed to the formation of
aggregates within the cellulose and some evidence has already
been given (36)that association of vat-dye anions takes place
within the substrate during dyeing.
At present, it is not known with any real certainty whether
disperse dyes in hydrophobic substrates exist as monomolecular
dispersions or whether they associate to fcrm large or small
aggregates. Giles et al. have drawn attention to this probleni and
consider that associates of non-ionic dye molecules are present in
dyed nylon, polyester and cellulose acetates. Thcy base their
conclusions on two main arguments. Firstly (64, 65) by measuring the rate of fading, as expressed by changes in optical density
of the dyed substrate with time when exposed to light, the order
of the fading reaction can be classified according to whether the
rate of fading changes approximately exponentially or linearly
with time. Some dyes may fade according to a combination of
these orders of reaction, whereas other (anomalous) dyes exhibit
an initial increase in the optical density of the dyed film. These
differences in reaction order or rate of fiiding are attributed to
the presence and growth of aggregates of dyc molecules within
the film on exposure to light. Secondly (66),for those dyes containing two absorpticn bands in the visible region of their spectra,
the ratio of the molar extinction coefficients of each band increases with increase in dye concentration in solution. Changes in the
ratio are related to changes in the degree of association as the
concentration of dye is varied. From similar changes in the spectra of dyed hydrophobic films arid dyed films exposed to light
for different periods, it is concluded by analogy that such films
contain both monomolecularly dispersed and aggregated dye
molecules and that the average aggregation number increases as
fading proceeds.
It has been pointed out (67)that, since the spectrophotometric
technique used to examine dyed films utilises monochromatic
light that is partially plane-polarised, changes in extinction ratios
may be due to dichroic effects within the film. It has also been
observed (26) that the dichroic orientation factor of dyed
filaments increases as the amount of dye present increases, since
the dye molecules first occupy the least oriented parts of the
substrate and the zones of higher orientation are occupied only
at higher concentrations of dye.
If the aggregation number of the dye increases during fading,
then there must be a contribution to rates of fading from a dyemigration mechanism. By studying the change in the dichroic
orientation factor of dyed polyester films with temperature, any
reversible dye migration taking place as a consequence of heat
treatment would be indicated by a reversible change in the

REVIEW OF PROGRESS IN COLORATION; JONES

22
dichroic orientation factor. Nakayama et al. (68) have shown
that, provided the amorphous polymer structure does not change
irreversibly, the orientation factor is in fact reversible. It is
concluded that, even if dye migration occurs at higher temperatures, the dye molecule reverts to the same type of absorption
site on cooling.
The view that disperse dyes are at least initially monomolecularly dispersed in hydrophobic fibres is favoured by Husy et al.
(69). In their more recent experiments they qualitatively show that
exposure to light of cellulose acetate dyed with an azo disperse
dye in which a trans+& rearrangement can take place is
accompanied by contractions in the dimensions of the substrate.
With dyes that d o not undergo this phototropic change, the
dimensions of the substrate remain unchanged. These results
indicate a close interaction between dye and substrate molecules
which is favoured more by a molecular dispersion than by an
associated state.

Conclusions
A reading of this review will show that no new theories of
dyeing have been postulated and that a comprehensive theory of
dyeing is still far from reality. In a recent survey, Valko (70)
suggests that, whereas thermodynamic studies of dyeing can
make useful contributions to the general theories ofintermolecular forces, diffusion processes and the influenceof parameters such
as concentration, temperature, electrolyte concentration and
polymer structure on these processes remain largely uninvestigated and would be of greater relevance to application methods.
As is shown in this review, however, when equilibrium studies are
carried out and assessed in conjunction with the growing amount
of information on the structure of water and aqueous solutions,
their relevance to dyeing theory should not be underestimated.
Dyeing theory has been previously retarded by lack of knowledge
about non-ideal behaviour, but it is now possible, e.g. by differential manometry or vapour-pressure osmometry, to determine the
mean activity coefficients of dyes in solution. This may be the
first step in determining activity coefficients of dyes in the
internal aqueous phase and inthesubstrate. Finally,muchresearch
is still needed on the heat stability of dye dispersions and solids
and on the thermodynamics and kinetics of heat-fixation
processes.
References
l McGregor and Peters, J.S.D.C.,84 (1968)267.
2 Milikevie, Text. Reserirch J., 39 (1969)677.

3 See, e.g., Hildebrand and Scott, The Solubility of Non-electrolytes


(New York: Reinhold Publishing Corpn, 3rd edn, 1950), p. 26.
4 Jones, Cibu Review, in press.
5 Thompson, Ph.D. Thesis (University of Leeds, 1969).
6 See, e.g., Franks and Ives, Quart. Rev., 20 (1966) 1;
Frank and Evans,J. Chem.Phys., 13 (1945)507.
7 Coates,J.S.D.C.,85(1969)355.
8 Green and Jones, Truns.Furuduy SOC.,63 (1967)1612.
9 Blandamer, Brivate, Fox, Symons and Verma, ibid., 63 (1967)1850.
I 0 Coates. Rigg and Smith, ihid., 64(1968)3255.
I 1 Niederer and Ulrich, Textilvcvedlurrg, 3 (1968)337.
I2 Malik and Saleem, J. Oil Col. Chem. Assocn, 52 (1969)551.
13 Wagner. TextilPmxis, 24 (1969)310,383.

14 Cockett, Rattee and Stevens, J.S.D.C., 85 (1969)461.


15 McDowell and Weingarten, MeIIiand Textilber., 50 (1969)814.

.-

16 Idem. ibid.. 50 11969)


59.
17 See, &g., Glasstone, Textbook of Physical Chemistry (London:
Macmillan & Co., 1948), p. 495.
18 McDowell and Weingarten. J.S.D.C.. 85 (1969)589.
19 Saito and Arai, Reit Govt Chem. Ind.Research Znst. Tokyo, 63
I

11968)
103.
.~~
--,

20 Apperley, J.S.D.C.,85(1969)562.
21 Weisz and Zollinger, Trans. Furaduy SOC.,63 (1967)1801 ;64(1968)

1693.
Standing, Warwicker and Willis, J. TextileZnst., 38 (1947)T335.
Murfet, M.Sc. Thesis (University of Lee&, 1969).
McGregor, Peters (R.H.) and Ramachandran,J.S.D.C., 84 (1968)9.
Brown and Peters (A. T.),Amer. DyestuffRep.,57 (1968)49.
Blacker and Patterson, J.S.D.C.,85 (1969)598.
Suzawa and Saito, Bull. Chem. SOC.Japan, 41 (1968)539.
Suawa, Saito and Shinohara, ibid., 40 (1967) 1596.
Bell, Text. ResearchJ., 38 (1968)984.
Mayer, Amer. Dyestuff Rep., 56 (1967) 869;
Mayer and Sulflow,MeIIiand Textilber., 49 (1968)813.
31 Petropoulos and Roussin, J. Chem. Phys., 47 (1967) 1491, 1496;
48 (1968)4619 ;50 (1969)3951.
32 BirdandPatel, J.S.D.C.,84(1968)560.
33 McGregor, ibid., 83 (1967)477.
34 Iyer, Srinivasa and Baddi, Text. ResearchJ., 38 (1968)693.
35 Iyer and Baddi, Proc. Symp. Contributions Chem. Synth. Dyes and
Mechanism of Dyeing, Univ.ofBombay,(1968)36.
36 Bender and Foster, Trans. Farahy SOC.,64 (1968)2549.
37 Porter and Evans, Text. ResearchJ., 37 (1967)643.
38 Roberts and Rowland, ibid., 39 (1969)686.
39 Rattee, ibid., 38 (1968)205.
40 Ferrini and Zollinger, Helv. chim. Acta, 50 (1967)897.
41 MiliCeviC, Textilveredlung,3 (1968)607.
42 Marshall, J.S.D.C.,85(1969)251.
43 de Boer, The Dynamical Character of Adsorption (Oxford:
Clarendon Press, 2nd edn, 1968), p. 48.
44 Rattee, J.S.D.C., 85 (1969)23.
45 Danckwerts, Trans.Farmby SOC.,46 (1950)300.
46 Kissa, Text. ResearchJ.,39 (1969)734.
47 SumnerandTaylor, J.S.D.C.,83(1967)445.
48 Rattee and Murthy, ibid., 85 (1969)368.
49 Aston, private communication.
50 Shore,J.S.D.C.,84(1968)408,413.
51 Idem, ibid., 84 (1968)545.
52 Idem, ibid., 85 (1969)14.
53 Hildebrand and Beckmann, Melliand Textilber., 45 (1964) 1138.
54 LewisandSeltzer,J.S.D.C.,84(1968)501.
55 Hartley, ibid., 85 (1969)66.
56 Meckel, Textilveredlung,2 (1967) 715;
Hartley, Australian J. Chem.,21(1968) 2723.
57 Meunier, Iannarone and Wygand, Amer. Dyestuff Rep., 52 (1963)
1014.
58 Price, ibid.. 54 (1965)13.
59 Datye, Pilkar and Rajendran, Proc. Symp. Contributions Chem.
Synth. Dyes and Mechanism of Dyeing, Univ. of Bombay, (1968)
57.
60 Datve. Textilveredlunp.4 11969)562.
61 Beni, Flynn and S d & , J.S.D.C.,85 (1969)606.
62 Jones, Teintex, 34 (1969)594.
63 Kern, Kissling and Herlant, Textilveredlung,3 (1968)595.
64 Giles, Shah and Yabe, Text. ResearchJ.,38 (1968)467.
65 Giles. Johari and Shah. ibid..38 (1968) 1048.
66 Gilesand Shah, Trans.Far&y Soc., 65 (1969)2508.
67 Patterson, private communication.
68 Nakayama, Okajima and Kobayashi, J. Appl. Polymer Sci., 13
(1969)659.
69 Husy, Merian and Schetty, Text. Research J.,39 (1969)94,
70 Valko, ibid., 39 (1969)759.
22
23
24
25
26
27
28
29
30

You might also like