Blends

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 20

26 DYNAMIC COMPETITION BETWEEN FIBRE TYPES IN THE DYEING OF BLENDS

CHAPTER 3

Dynamic competition between fibre types in


the dyeing of blends

3.1 INTRODUCTION
Dyeing systems in which more than one fibre type is present have been almost
completely neglected by those researching the mechanisms of dyeing processes.
This is not surprising, because the complications that arise when attempting to
predict quantitatively the uptake of a single dye by one fibre type are
considerable. Theories of dyeing established on the basis of careful
measurements of one such dye at various depths under a variety of dyeing
conditions can seldom be transferred intact to other members of the same range
of dyes, especially if these are distinctly different in structure from the first dye
chosen. Further difficulties arise with the binary or ternary combinations of dyes
that are used routinely in practice to achieve the wide gamut of colours with
which design colourists and dyers must work. Two dyes undergoing absorption
by the fibre simultaneously rarely reproduce the dyeing rate curves that they give
alone. There is almost invariably an interaction between them, often (but not
always) resulting in a slower rate of uptake and a lower equilibrium exhaustion
for each of the dyes in combination, compared with the corresponding values
when applied individually.
An entirely different dimension is introduced when two fibres are present in
the same dyebath. A single dye may be distributed between them according to a
complex relationship that is determined by the differences in dyeability
characteristics between the two substrates, the dyeing conditions and the applied
depth of the dye. An unequal distribution that may be found in an early stage of
the process, arising from differing rates of uptake by the competing substrates,
may later undergo a levelling effect as dye is desorbed from the initially more
heavily dyed fibre and is taken up by the other fibre type for which the dye has
inherently higher affinity.
When two different substrates are present in the same dyebath, initial kinetic

26

Contents Home
INTRODUCTION 27

conditions and later the energetic characteristics of the system will dictate which
of the substrates a given dye will tend to favour. More dye will ultimately be
absorbed by the substrate to which it is more substantive. This favoured
substrate will require less energy for the sorption process because the dye–fibre
forces of interaction are stronger. When a combination of two or three dyes is
applied to two different substrates, however, it does not always follow that the
dyes present will all be absorbed preferentially by the same fibre component to
the same extent.
It is certainly possible in practice to choose a set of dyeing conditions in which
a trichromatic combination of dyes with similar dyeing properties will yield a
matching hue on both substrates at equal depth (a solid effect) or at markedly
different depths (a shadow effect). To arrive at such ‘ideal’ combinations of dyes
and dyeing conditions, however, implies considerable preliminary laboratory
work, especially when other important factors, such as fastness demands and
non-metameric matching, must be taken into account [1]. This task can be
facilitated by deriving characteristic thermodynamic parameters of the dyeing
system from the sorption behaviour of the individual dyes on the separate
homogeneous substrates. It is claimed that these parameters allow computation
of the sorption behaviour of the dyes in combination on more than one substrate
simultaneously [2].
The substantivity ratio (Df /Ds) for each dye on each substrate depends on
temperature, pH, liquor ratio, concentrations of dye and electrolyte, and
particularly in this case on the other dyes and substrates involved in the
competitive sorption process. According to the Gouy-Chapman theory, the
substantivity ratio is related to the electrolyte concentration or ionic strength and
the total amount of ionic charge imparted to the substrate by the absorption of
dye ions. This factor encapsulates the mutual restraining effects of the dyes on
one another. The three characteristic constants in the Gouy-Chapman equation
are as follows:
A0 is a function of the standard affinity and thus the substantivity ratio under
the relevant conditions.
A1 is proportional to the charge density and is a measure of the overall charge
on the substrate.
A2 is related to the specific surface of the substrate accessible for dye sorption.

The simplest dyeing system that may be considered as representative of the


dyeing of a blend is that in which one dye is distributed between two different
fibres that have broadly similar dyeing properties, so that the same dye will give
an economic colour yield on both. Even if solidity of colour between them is the

Contents Home
28 DYNAMIC COMPETITION BETWEEN FIBRE TYPES IN THE DYEING OF BLENDS

objective, there is no guarantee that the dyeing conditions to achieve this at the
target depth will ensure satisfactory penetration and fastness on both substrates.
It is often easier to adopt this approach when a shadow effect is desired, because
any inherent difference in dyeability between the substrates can be exploited in
adjusting dye uptake levels towards the target difference in depth. The serious
limitations of this simple dyeing system become clear when reserve and colour
contrast effects are considered. By definition it is usually extraordinarily difficult
to suppress the dyeability of the less dyeable component to zero in order to
reserve it as white or to dye it in another colour, which means introducing a dye
from another dyeing class.
Another deceptively simple approach to the dyeing of a binary blend of fibres
is to dye each in turn with an appropriate class of dyes in two completely
separate dyeing processes, using the optimum conditions of application in each
case just as if they were entirely separate substrates. At first sight, this appears an
ideal way in which to dye for solidity, reserve, shadow or contrast effects at will.
In practice, however, technical limitations do arise here too. The wet fastness
characteristics of the dyeing achieved in the first process must be such that no
significant desorption from it occurs during the second one. In other words, the
fastness to cross-dyeing must be excellent. Thus the order in which the two
processes are carried out is important and the dyeing that requires the higher
dyeing temperature is normally applied first.
The two-bath process is also less than ideal unless both classes of dyes are free
from cross-staining problems, i.e. a class selected to dye one of the fibre types
should not cause significant staining of the other fibre present. In a majority of
instances, however, cross-staining of one or both fibre types must be taken into
account and a clearing process introduced after one or both of the dyeing
processes. Thus the so-called ‘two-bath’ process may require several baths for
completion. It is easy to see why this technically straightforward possibility turns
out to be the least attractive from an economic viewpoint.
Between these fairly obvious extremes of the single-class and the two-bath
methods there are two other compromises that offer greater flexibility than the
former and economic savings over the latter. These are the simultaneous one-
bath and the one-bath two-stage methods. For clarification the four possibilities
are summarised in Table 3.1. These will be referred to frequently in later chapters
and it is convenient to use the four hyphenated abbreviations listed in the first
column to distinguish between them. In general, the progressive increase in cost
in moving down this list is compensated by a wider choice of suitable dyes and
greater freedom from the practical problems discussed in terms of typical
examples in the remainder of this chapter. Another general rule is that the greater

Contents Home
INTRODUCTION 29

Table 3.1 Summary of general dyeing methods for binary blends

Method Dyebaths Dye classes Stages

Single-class One One One


One-bath One Two simultaneously One
Two-stage One Two in sequence Two
Two-bath Two Two Two

the depth of shade required, the more likely is it necessary to move to a method
lower down this list.

3.2 THE DISTRIBUTION OF ACID DYES ON NYLON/WOOL BLENDS


Nylon/wool blends are often dyed with a single class of anionic dyes, which may
be levelling acid, milling acid or metal-complex types. Ensuring solidity of shade
is normally the main requirement and this calls for careful selection of
combinations of dyes with similar rates of dyeing and build-up characteristics.
Partitioning of the dyes between the two fibres can be influenced by many
factors, including dye structure, applied depth, dyebath pH, blend ratio and the
quality of the component fibres.
Wool and nylon contain both basic and acidic groups, amongst which by far
the most important are amino and carboxyl groups respectively. Just like the
parent amino acids from which all proteins are derived, both of these polymeric
amides show zwitterionic characteristics at pH values close to the isoelectric
point, i.e. the pH at which the fibre contains equal numbers of protonated basic
and ionised acidic groups. As the pH decreases below this point, the carboxylate
anions are progressively neutralised by the adsorption of protons and the fibre
acquires a net positive charge (Scheme 3.1).
+ +
H3N [fibre] COO– + H+ H3N [fibre] COOH

Scheme 3.1

Conversely, as the pH rises above the isoelectric point, the fibre becomes
negatively charged as a result of deprotonation of the amino groups by
adsorption of hydroxide ions or other simple anions (Scheme 3.2).

+
H3N [fibre] COO– + –OH H2N [fibre] COO– + H2O

Scheme 3.2

Contents Home
30 DYNAMIC COMPETITION BETWEEN FIBRE TYPES IN THE DYEING OF BLENDS

The rate of dyeing of anionic dyes on nylon is much more rapid than on wool,
particularly at 60–80°C and low applied depths, so that pale dyeings on nylon/
wool show a marked preferential dyeing of the nylon. Even at 1% applied depth,
for example, the uptake by nylon is much more rapid than by wool and the
difference in rate is greater for dyes with a lower degree of sulphonation. Figures
3.1 and 3.2 are rate-of-dyeing curves for disulphonated and tetrasulphonated
dyes (Figure 3.3) respectively on these two fibres, at 93°C and an initial pH of
4.2 in aqueous acetic acid [3]. As dyeing proceeds, the pH rises towards 5 as a
result of sorption of acetic acid, particularly by wool.

100

80
Exhaustion/%

60

40

Nylon
20
Wool

20 40 60 80 100
Dyeing time/min

Figure 3.1 Rate of dyeing of CI Acid Red 1 [3]

100

80
Exhaustion/%

60

40

20

10 20 40 60 80 100
Dyeing time/min

Figure 3.2 Rate of dyeing of CI Acid Red 41 [3] (for key see Figure 3.1)

Contents Home
THE DISTRIBUTION OF ACID DYES ON NYLON/WOOL BLENDS 31

COCH3 H O SO3Na
HN NaO3S N
H O N
N SO3Na
N NaO3S

NaO3S SO3Na
CI Acid Red 1 CI Acid Red 41

Figure 3.3 Disulphonated and tetrasulphonated acid dyes

The origin of the markedly different rates of dyeing lies in the differences in
hydrophobic character between the two dyes and between the two fibres. The
nylon polymer is much more hydrophobic than wool and so it attracts the more
hydrophobic of the two dyes preferentially. Thus the disulphonated dye shows
the most rapid rate of absorption on nylon, with a time of half dyeing (time for
50% of the equilibrium exhaustion) of only about 2 minutes. The
tetrasulphonated dye is relatively hydrophilic and is therefore absorbed more
slowly by either fibre, showing the slowest rate of dyeing on wool, with a time of
half dyeing of approximately 3 hours.
Although nylon absorbs acid dyes more readily than does wool, partition
between the two fibres is not constant at all depths since the saturation
concentration on nylon is much lower than that on wool. The saturation limit on
wool is not approached at the applied depth necessary to saturate the amine end
group content of the nylon. In pale depths both the initial uptake and the
ultimate exhaustion are higher on nylon. In full depths, on the other hand, the
initial strike still occurs on nylon but eventually the wool becomes more heavily
dyed because it has a much higher saturation uptake. At some intermediate depth
depending on dyeing conditions, there is a point at which both fibres are dyed to
the same depth even though the nylon reaches this equilibrium position more
quickly.
This critical depth is specific for the dye and is much higher for a
monosulphonated dye than for a disulphonated analogue, since a disulphonated
dye requires about twice as many amine end groups for a given tinctorial yield
(Figure 3.4). Blocking effects may occur if monosulphonates and disulphonates
are applied together, resulting in a heavier depth on the wool that is closer in hue
to the disulphonated dye. Above the critical depth the nylon becomes
progressively more difficult to dye with levelling acid dyes and the distribution
on the nylon/wool blend increasingly favours the wool.
The blocking of disulphonates by analogous monosulphonated dyes may be

Contents Home
32 DYNAMIC COMPETITION BETWEEN FIBRE TYPES IN THE DYEING OF BLENDS

NH
CO

H CH2
+
O
NH3 –O3S N CH2
(CH2)6 N CH2
NH CH2
CO CO
CI Acid Red 88 NH

CO
H O NH
+
NH3 –O3S N
(CH2)3
(CH2)6 N
CH2
NH
CH2
CO
CH2
CI Acid Red 13 SO3– NH3
+

Figure 3.4 Electrostatic and hydrophobic bonding of acid dyes on nylon

exploited by applying mixtures of selected mono- and disulphonated pairs of


similar hue, varying the proportions according to applied depth in order to
minimise the disparity in depth on the two fibre components. The preferred
monosulphonated dyes are mainly monoazo yellows and reds with anthra-
quinone blues. The disulphonates have lower saturation values on nylon and so
they can only be used for pale and medium depths. They are virtually all yellow
to red monoazo and disazo types, or violet to green anthraquinone derivatives.
Given the widely varying qualities of the two fibre types, as well as possible
variations in the blend proportions, it is not surprising that specific dyeing
conditions for solidity cannot be laid down, only guideline starting points for
initial experimentation [4]. Anionic agents capable of controlling the uptake of
anionic dyes by the nylon component below the critical depth are of two general
types:
(1) levelling or blocking agents that are preferentially absorbed by nylon and
act as a partial reserve for levelling acid dyes;
(2) retarding agents of higher relative molecular mass capable of controlling the
distribution of dyes of higher wet fastness on nylon/wool blends.

Contents Home
THE DISTRIBUTION OF ACID DYES ON NYLON/WOOL BLENDS 33

Typical examples of the moderately substantive agents used to control the uptake
of levelling acid dyes are shown in Figure 3.5. They compete with these dyes for
the basic groups and facilitate the approach to an equilibrium distribution that
does not alter even on prolonged boiling, unless further additions of agent are
made to shift the equilibrium more in favour of the wool. Such products improve
the coverage of any dye-affinity variation in the nylon but have insufficient
affinity to exert a blocking effect with premetallised or milling acid dyes. A novel
agent developed recently, however, permits either equalisation or a total reserve
to be obtained even with metal-complex or milling dyes [5].

R SO3Na
SO3Na
Sodium alkylbenzene sulphonate Sodium alkylnaphthalene sulphonate

R OSO3Na CH3(CH2)5 CH CH2CH CH(CH2)7COONa


OSO3Na

Sodium alkanol sulphate Disodium sulphoricinoleate

Figure 3.5 Levelling or blocking agents for nylon/wool dyeing (R = long-chain alkyl)

Retarding agents that control the initial uptake of these dyes by the nylon
component have higher affinity than the typical levelling agents defined in Figure
3.5 Many of these are also used as syntan aftertreating agents for dyed nylon.
These belong to a distinct group of condensates of formaldehyde with certain
sulphonated phenols, thiophenols or naphthylaminesulphonic acids [6]. The
rapid sorption of syntans by nylon is mainly attributed to electrostatic bonding
between negatively charged sulpho groups in the syntan and the protonated
amino groups in the fibre. Hydrogen bonding between uncharged polar groups
and hydrophobic interaction between nonpolar moieties in the syntan and the
nylon also contribute to the mechanism [4].
Maximum retarding effect is found when the syntan molecules are retained
close to the fibre surface, since any treatment leading to diffusion of the syntan
into the fibre interior tends to lower its effectiveness. When used with levelling
acid dyes, this type of agent does not prevent migration from wool to nylon on
prolonged boiling. With metal-complex and milling acid dyes, however, the
initial distribution is preserved on boiling because these dyes show only limited

Contents Home
34 DYNAMIC COMPETITION BETWEEN FIBRE TYPES IN THE DYEING OF BLENDS

migration. It is not normally necessary to employ levelling or retarding agents on


the rare occasions when chrome dyes are preferred, since these give quite good
solidity on nylon/wool blends at the heavy depths where they may offer an
advantage. Some of the syntans darken on exposure to light and this significantly
lowers the light fastness of the dyeings.
Considerable attention has been given in recent years to devising suitable
procedures for dyeing wool at lower temperatures and near the isoelectric point
(pH 4–5) in order to avoid the damage that inevitably occurs when dyeing at the
boil. The low-temperature dyeing of wool offers:
(1) optimum handle and durability;
(2) improved carding, spinning and weaving performance;
(3) brighter shades because of the lower degree of yellowing of the wool;
(4) shorter dyeing cycles, higher productivity and lower process costs;
(5) a better-quality end-product.

Problems associated with low-temperature dyeing include:


(1) slower diffusion into the interior of the wool fibres, resulting in inadequate
wet fastness;
(2) inadequate exhaustion and poor reproducibility.

The epicuticle of wool is the main barrier to penetration and this is mainly
responsible for the dyeing problems. These problems can be overcome by
damaging the fibre scales in a chlorination process before dyeing, but this lowers
wool quality too. The best results are achieved using monosulphonated 1:2
metal-complex and milling acid dyes.
Shade partition between wool and nylon in blend dyeing is dependent on
dyeing temperature. As already discussed, the tendency at the boil is for the
nylon to absorb dye more quickly than wool and to dye more deeply below the
critical depth. This is easily corrected by adding an anionic agent to retard uptake
by the nylon and achieve solidity of shade over a range of depths. At lower
temperatures (e.g. 80°C) the wool is generally dyed more heavily than the nylon
even without an anionic agent present. As dyeing proceeds there is some
migration from wool to nylon until the final partition is reached, but this is a
different equilibrium at the lower temperature than in conventional dyeing at the
boil. In general, partition favours the wool and solidity is difficult to attain. This
difficulty can be overcome, however, using a specific equalising agent in place of
the conventional anionic agent normally employed when dyeing at the boil [7].
When dyes of high substantivity are applied below the boil, surface dyeing of the
wool occurs to give dyeings of inadequate wet fastness.

Contents Home
THE DISTRIBUTION OF ACID DYES ON NYLON/WOOL BLENDS 35

Premetallised and milling acid dyes (relative molecular mass 700–1000)


demand more energy input for good fibre penetration. There is a close relation-
ship between dye substantivity, exhaustion and diffusion into the fibre and this is
mainly determined by the relative molecular mass and degree of sulphonation of
the dye. Addition of an equalising agent shifts the partition in favour of the nylon
at the lower temperatures, allowing solidity to be achieved at any temperature in
the 80–100°C range, depending on the amount of auxiliary required at a given
depth of shade. The optimum dyeing temperature is recipe-specific and a
temperature selector system has been developed for use with the specific
equalising agent. In pale depths (below ca. 0.6% total dye) the conventional
syntan-type retarder may be used at a relatively low concentration [8].

3.3 THE DISTRIBUTION OF ACID DYES ON NYLON/POL


YURETHANE
BLENDS
The polyurethane elastomeric fibres vary in their ability to absorb anionic dyes,
depending on their content of basic groups, but in general the equilibrium
distribution on a nylon/polyurethane blend tends to favour the nylon
component. The higher rate of dyeing of the elastomeric fibre at lower
temperatures, however, results in preferential dyeing of that component in the
early stage. There are two main methods of controlling this distribution, both of
them depending on the use of dyeing auxiliaries.
Anionic agents, such as sodium alkanesulphonates, sulphated castor oil,
disodium dinaphthylmethanedisulphonate or selected syntans, can be added to
promote dye uptake by the polyurethane at equilibrium by becoming absorbed
by the nylon and restraining subsequent migration of dye from the polyurethane
to the nylon. This method is more effective when dyeing with 1:2 metal-complex
or milling acid dyes of the monosulphonate type with good levelling properties.
The amount of agent required decreases progressively as the applied depth of the
dyeing increases.
Cationic agents (Figure 3.6) will form labile ionic complexes with typical
anionic dyes. This type of complex is absorbed more slowly and tends to favour
the polyurethane component more than does the parent dye. Acid dyes of the
disulphonate type with only moderate levelling properties can be readily
controlled on nylon/polyurethane materials by this method. The concentration of
cationic agent required increases with applied depth and the degree of
sulphonation of the dyes. A nonionic dispersing agent of the alkanol poly-
oxyethylene class is necessary to solubilise the dye–agent complex (Figure 3.7).

Contents Home
36 DYNAMIC COMPETITION BETWEEN FIBRE TYPES IN THE DYEING OF BLENDS

CH3 CH3
+ H33C16 N+ CH3 H35C17 N + CH2
H33C16 N
Cl – CH3 Br – CH3 Cl –

Cetylpyridinium Cetyltrimethylammonium Stearyldimethylbenzylammonium


chloride bromide chloride

Figure 3.6 Cationic complexing agents for nylon/polyurethane dyeing

H O R (OCH2CH 2)x OH
CH3 O3S N
+
R N CH3 N
Alkanol polyoxyethylene
CH3
CH3
+
SO3– CH3 N R R (OCH2CH2)x OH
CI Acid Red 13
as bis-alkyltrimethylammonium complex CH3

Figure 3.7 Schematic representation of solubilised dye–agent complex (R = long-chain alkyl)

3.4 THE CROSS-STAINING OF WOOL BY DISPERSE DYES


Wool keratin is the most sensitive of textile fibres towards staining by dyes of all
types because it is a natural protein containing many different functional groups.
The staining of wool by disperse dyes is a serious problem in dyeing blends with
any of the ester fibres. Cellulose acetate suffers a loss in lustre if it is treated at the
boil, as is normally necessary to dye the wool component during the second stage
of a two-stage or two-bath sequence. Migration from the acetate to the wool
increases with dyeing time and temperature of the wool-dyeing stage. Migration
from cellulose triacetate to wool is slower under these conditions but treatment
at 105°C or at the boil with carrier, in order to attain full depth and penetration
on the triacetate, usually results in serious staining of the wool.
Polyester/wool presents the most difficult problem and studies have revealed
numerous factors that may affect the degree of staining observed in practice [9].
The stain on the wool is dull in hue and exhibits poor fastness to light and wet
treatments. Staining implies a loss of colour yield on the polyester and makes
shade matching more difficult because individual dyes differ in their propensity
to staining. Bulky, low-twist wool becomes stained more easily then fine, high-
twist yarns. Wool quality is another factor influencing the degree of staining.

Contents Home
THE CROSS-STAINING OF WOOL BY DISPERSE DYES 37

Dye structure (molecular size, number, type and distribution of polar groups)
has less influence on the initial level of staining than dyebath conditions and the
quality of the dye dispersion. The mechanism of wool staining involves:
(1) hydrogen bonding between proton donor groups (e.g. amino, amido,
hydroxy) in the disperse dye molecules and in wool keratin;
(2) interaction by means of secondary (dipolar and van der Waals) forces
between the dye molecules and wool;
(3) physical sorption of aggregated particles of disperse dyes on the scaly
surface of the wool fibre.

The preferred disperse dyes for these blends have low intrinsic saturation values
on wool, low tendency to aggregate at the dyeing temperature, rapid rates of
diffusion into polyester and high equilibrium exhaustion. Mechanical retention
in yarn crevices may play a part in the initial deposition, since particle size and
stability of the dye dispersion are important. Staining tends to decrease with the
concentration and anionic charge of the stabilising agents present in the dye
dispersion, but the magnitude of the effect is specific to the types of dye and
agent.
In a recent investigation of the kinetics of polyester/wool dyeing and the wool
staining problem, polyester and wool fabrics in the weight ratio 55:45 were used
to allow the disperse dye uptake by both components to be determined
independently [10]. Dyeings with CI Disperse Blue 185 (Figure 3.8) and CI Acid
Red 211 were carried out at the isoelectric point (pH 4.5) and wool damage was
assessed by measuring tensile properties and alkali solubility. Both disperse and
acid dyes are absorbed by wool far more quickly and easily than the disperse
dyes are taken up by the polyester at relatively low temperatures, because the
rates of diffusion in wool are so much more rapid. After only 30 minutes at
110°C, therefore, the amount of disperse dye that has stained the wool is more
than twice that absorbed by the polyester (Figure 3.9).

O2N O NH2

CH3
HO O NHCH
CH2CH3

Figure 3.8 CI Disperse Blue 185

Contents Home
38 DYNAMIC COMPETITION BETWEEN FIBRE TYPES IN THE DYEING OF BLENDS

3
Dye uptake/%

1
Wool
Polyester

15 30 60 90 120
Dyeing time/min at 110 oC

Figure 3.9 Rate of uptake of disperse dye by polyester and wool [10]

As dyeing proceeds, desorption from the wool occurs at a rate determined by


the slow diffusion of the disperse dye into the polyester. The distribution
coefficient for a disperse dye between polyester and wool is controlled by the
difference in affinity values of the dye for the two substrates. After about an hour
at 110°C, an equilibrium partition is approached but the degree of staining of the
wool is not much less than after the first 15 minutes. As the disperse dye diffuses
slowly into the polyester, the wool stain slowly desorbs at a rate that keeps the
low concentration of disperse dye in the dyebath roughly constant.
After a dyeing time of 30 minutes, wool staining is considerable at all dyeing
temperatures up to about 115°C (Figure 3.10). However, above around 120°C
the diffusion of the disperse dye into the interior of the polyester fibre is
sufficiently rapid to give much lower staining of the wool. The pH dependence of
wool staining is not critical under the mildly acidic conditions preferred for
dyeing both components of this blend (Figure 3.11). At pH values below the
isoelectric point, wool staining is at a minimum. As the pH increases towards the
alkaline side, so does the degree of wool staining.
When tested in the presence of the premetallised CI Acid Red 211, the staining
of wool by the disperse dye was markedly greater [11]. This was attributed to
possible hindering of disperse dye desorption from the wool by the presence of
the metal-complex dye. Since anionic dyes of this kind interact strongly with
wool keratin they desorb again very slowly and incompletely. Sequestering agents
such as ethylenediaminetetra-acetic acid (EDTA) and citric acid are useful to

Contents Home
THE CROSS-STAINING OF WOOL BY DISPERSE DYES 39

3
Dye uptake/%

30 60 90 120
Dyeing temperature/oC for 30 min

Figure 3.10 Temperature dependence of disperse dye uptake [10] (for key see Figure 3.9)

4
Dye uptake/%

2
2 3 4 5 6 7 8
Dyebath pH for 60 min at 110oC

Figure 3.11 pH dependence of disperse dye uptake [10] (for key see Figure 3.9)

minimise wool staining, particularly the former. It is claimed that such agents
interact with certain disperse dyes (e.g. the 1-amino-4-s-butylamino system in CI
Disperse Blue 185) to hinder diffusion into wool (Figure 3.12). These agents have

Contents Home
40 DYNAMIC COMPETITION BETWEEN FIBRE TYPES IN THE DYEING OF BLENDS

low affinity for wool and formation of the dye–agent complex favours retention
of the disperse dye in the dyebath, allowing the polyester to absorb more dye as it
is released from the complex at higher temperatures and improving the ultimate
exhaustion.

+ –OOC
O2N O NH3 CH2
CH OH
–OOC CH2
CH3
HO O NH2CH
+
CH2CH3

Figure 3.12 Interaction between CI Disperse Blue 185 and citric acid

When dyeing wool-rich blends, the problems of wool staining and the
attainment of solidity at equilibrium are greatly aggravated. Uptake of disperse
dyes by the polyester in full depths can be 25% less on a 20:80 polyester/wool
blend than on a 50:50 blend [12]. If a 50:50 blend is dyed at a liquor ratio of
10:1, then the individual components are each actually at 20:1 with respect to
the dyebath. When the blend ratio is 20:80, however, the polyester is being dyed
at 50:1, making exhaustion far more difficult. To compound this, the liquor ratio
for the wool is only 12.5:1 and absorption of the disperse dyes by wool is
favoured preferentially. Less dye remains in the bath for dyeing the polyester
directly and this component becomes even more dependent on disperse dye
transfer from the wool [12].
Wool progressively loses its strength at elevated dyeing temperatures,
especially about 110°C, although at this temperature the damage is not severe for
dyeing times of 60 minutes or less. In the isoelectric region (pH 4–5) the keratin
structure is reinforced by electrostatic linkages between protonated amino
groups and carboxylate-containing amino acid residues. Hydrolysis of peptide
and disulphide bonds is also at a minimum under these conditions. The loss in
strength increases rapidly at pH values above 5 and temperatures above 110°C.
The sensitivity of wool to degradation makes the use of a carrier essential to dye
full depths on the polyester at a relatively low temperature. Carrier addition
lowers the degree of staining at a given temperature and also tends to lower the
temperature of maximum staining.
Carriers are able to minimise wool staining by accelerating the rate of uptake
of disperse dyes by polyester. It is believed that the sorption of carrier molecules
weakens the attractive forces between polyester segments, making them more

Contents Home
THE CROSS-STAINING OF WOOL BY DISPERSE DYES 41

mobile at lower temperatures than when the carrier is absent, thus changing the
internal structure of the fibre and lowering the glass-transition temperature.
Hence carriers accelerate dye diffusion within polyester and so indirectly increase
the rate of transfer from wool to polyester. The chemical type of carrier exerts
only a marginal influence, the degree of staining tending to increase slightly in the
series: aryl esters < chlorobenzenes < aryl ethers < phenylphenols. Dyebath
conditions are more important. Unfortunately, carriers are harmful to the health
of operatives and to the environment.
The wool staining problem is an important criterion in the decision whether to
dye polyester/wool sequentially in the one-bath mode or by the two-bath method
with an intermediate clear to remove the disperse dye stain from the wool. In the
two-stage sequence, the wool is cleared after dyeing by scouring with a nonionic
detergent at pH 4–5 and 50–70°C. This causes no significant damage to the wool
but is only moderately effective and this process is only suitable for pale and
medium depths because of fastness limitations. Nevertheless, it offers shorter
processing times and higher productivity at lower cost. During the wool dyeing
stage, disperse dye can migrate from the surface of the dyed polyester and cause
back-staining of the wool [13]. There is no obvious relationship between the
degree of back-staining and the heat fastness class of the disperse dyes.
Full depths are usually dyed by a two-bath method. In the absence of the
anionic dyes the blend can be given an intermediate reduction clear after
applying the disperse dyes to the polyester. This is more effective than nonionic
scouring, especially for azo disperse dyes, but the wool suffers some loss in
strength and elasticity. When dyeing wool-rich blends, the two-bath sequence
gives better shade partition than the one-bath method, especially when the
polyester component is dyed first [12]. Two-bath dyeing allows a wider choice of
disperse dyes, since the wool staining is less important provided the intermediate
clearing is adequate and the wool is dyed at a lower temperature than the
polyester. Certain anthraquinone-based disperse dyes, however, are reduced but
not destroyed by the reducing conditions. The degradation products may still
discolour the wool, especially after reoxidising back to the quinone form during
the wool dyeing stage. In heavy depths the two-bath process gives improved
fastness to rubbing and perspiration [14]. These advantages are particularly
significant for the dyeing of deep navy and black shades on polyester/wool [15].

3.5 THE CROSS-STAINING OF WOOL BY BASIC DYES


Although ultimately less serious than staining by disperse dyes, the uptake of
basic dyes by wool in the initial stage of the one-bath dyeing process for wool/
acrylic blends with milling acid and basic dyes can be troublesome. Above the

Contents Home
42 DYNAMIC COMPETITION BETWEEN FIBRE TYPES IN THE DYEING OF BLENDS

second-order transition temperature, however, dyeing of the acrylic component


can take place. The thermodynamic affinity of a basic dye for the acrylic fibre is
much higher than for wool. By the time that the dyebath reaches the boiling
temperature, most of the basic dye initially taken up by the wool has been
transferred to the acrylic component. This transfer proceeds during treatment of
the fibre blend at the boil even after exhaustion of the basic dye from the dyebath
is virtually complete. The degree of initial staining of the wool by the basic dye
varies with the type of acrylic fibre present in the blend. This variation becomes
particularly important when the applied depth approaches saturation of the
dyeing sites in the acrylic fibre.
Cationic dyeing auxiliaries have been widely used in wool dyeing, ranging in
ionic character from the weakly basic alkylamine polyoxyethylene types to the
much more strongly basic fatty alkyl quaternary ammonium salts. These
products are used as levelling agents and it is not surprising that the cationic
retarders necessary when dyeing acrylic fibres with basic dyes behave in a similar
manner when added to the wool/acrylic dyeing system. The pronounced
retarding effect of a typical cationic retarder on the rate of dyeing of wool with a
milling acid dye of the disulphonated anthraquinone type (Figure 3.13) is
illustrated in Figure 3.14. The time of half dyeing is increased from about 22 to
33 minutes by addition of the agent [16].
Dyeing rate curves for this milling acid dye and a typical monoazo basic dye
(Figure 3.13) on a 50:50 wool/Orlon (DUP) blend in the presence (Figure 3.15)
and absence (Figure 3.16) of the cationic retarder demonstrate that the anionic

NaO3S CH2CH2CH2CH3

O HN

O HN

NaO3S CH2CH2CH2CH3
Cl
CH2CH3
CI Acid Green 27 CH3
O2N N N N
CH2CH2 N CH3
+
CH3 X–
CI Basic Red 18

Figure 3.13 Structures of typical milling acid and basic dyes

Contents Home
THE CROSS-STAINING OF WOOL BY BASIC DYES 43

100

80
Exhaustion/%

60

40

20 1% Dye, no retarder
1% Dye, 1% retarder

10 20 30 40 50 60
Dyeing time/min
60 70 80 90 100 100
Temperature/oC

Figure 3.14 Rates of dyeing of CI Acid Green 27 on wool [16]

dye is highly sensitive to the agent of opposite charge. Under these conditions the
agent increases the time of half dyeing from about 17 to 32 minutes for the acid
dye but only about 26 to 32 minutes for the basic dye.
These results indicate that in the initial stage of a wool/acrylic dyeing the
cationic retarder is either absorbed by the wool or forms with the acid dye a
labile complex that has lower affinity for wool than the parent dye. Either effect
will significantly decrease the rate of dyeing of the acid dye. It is also evident that

100 CI Acid Green 27


CI Basic Red 18
80
Exhaustion/%

60

40

20

10 20 25 30 35 40 45 50
Dyeing time/min
66 84 92 100 100 100 100 100
Temperature/oC

Figure 3.15 Rates of dyeing of wool/Orlon with 1% retarder [16]

Contents Home
44 DYNAMIC COMPETITION BETWEEN FIBRE TYPES IN THE DYEING OF BLENDS

100

80
Exhaustion/%

60

40

20

5 10 15 20 25 30 35
Dyeing time/min
58 66 75 84 92 100 100
Temperature/oC

Figure 3.16 Rates of dyeing of wool/Orlon with no retarder [16] (for key see Figure 3.15)

ca. 20% exhaustion of the basic dye occurs at temperatures below 80°C (i.e.
below the glass-transition temperature of the acrylic fibre) and this dye must be
absorbed by the wool. This effect is irrespective of whether the cationic retarder
is present or not, implying that complex formation between the acid dye and the
retarder is the most probable explanation of the mechanism [16].

3.6 THE TRANSFER OF DISPERSE DYES DURING THERMOFIXATION


OF POLYESTER/CELLULOSIC BLENDS
The pad–thermofix dyeing of polyester/cellulosic fabrics is one of the few systems
of dyeing of blends that has been subjected to theoretical study. During the early
stages of padding and drying, much of the dye applied becomes deposited on the
relatively more absorbent cellulosic component. Several possible mechanisms of
transfer of the disperse dyes from cellulose to polyester during the
thermofixation stage have been proposed, but it is now widely accepted that the
transfer proceeds through the vapour phase [17]. The extent to which this
transfer takes place depends on the time and temperature of thermofixation.
As heating of the polyester/cellulosic fabric is continued, the total amount of
disperse dye available for colouring the polyester decreases as a result of
volatilisation into the atmosphere inside the heating chamber and deposition on
the inner surfaces of the latter. Some disperse dyes of low relative molecular mass
(Mr) and relatively high volatility may suffer oxidative decomposition,

Contents Home
THE TRANSFER OF DISPERSE DYES DURING THERMOFIXATION 45

particularly if they contain sensitive substituents such as primary amino groups.


It follows that a critical combination of fixation time and temperature gives
optimum yield of a specific dye, when the supply of dye from the reservoir
provided by the cellulosic component is sufficiently depleted for these progressive
losses to begin to favour desorption rather than adsorption of vapour at the
polyester surface.
Low-energy disperse dyes (approx. Mr <300) suffer relatively serious losses
under the conditions required for optimum transfer and fixation on the polyester.
Maximum transfer for these dyes is found at 200–210°C. For most dyes of
intermediate energy (approx. Mr 300–400) a temperature of 210–220°C is
needed. Good fixation of high-energy dyes (approx. Mr >400) often requires a
temperature of 220–230°C, but optimum transfer is limited by the onset of
thermal degradation and yellowing of the cellulosic fibres, or even some
softening of the polyester.

3.7 REFERENCES
1. J Park and J Shore, Rev. Prog. Coloration, 12 (1982) 1.
2. O Annen, H Gerber and B Seuthe, J.S.D.C., 108 (1992) 215.
3. B C Burdett, C C Cook and J C Guthrie, J.S.D.C., 93 (1977) 55.
4. T M Baldwinson in Colorants and auxiliaries, Vol. 2 Ed. J Shore (Bradford: SDC, 1990) 568.
5. Anon, Dyer. 177 (Apr 1992) 31.
6. C C Cook, Rev. Prog. Coloration, 12 (1982) 73.
7. A F Doran, Dyer, 176 (Aug 1991) 49.
8. A F Doran, J.S.D.C., 109 (1993) 15.
9. R E Lacey, V S Salvin and W A Schoeneberg, Am. Dyestuff Rep., 50 (1951) 978.
10. J Wang and H Asnes, J.S.D.C., 107 (1991) 274.
11. J Wang and H Asnes, J.S.D.C., 107 (1991) 314.
12. A F Doran, unpublished work.
13. K Türschmann and K H Röstermundt, Z. Ges. Textilind., 71 (1969) 326.
14. K H Röstermundt, Deutscher Färber Kalender, 80 (1976) 247.
15. W T Sherrill, Text. Chem. Colorist, 10 (1978) 210.
16. D R Lemin, J.S.D.C., 91 (1975) 168.
17. C J Bent, T D Flynn and H H Sumner, J.S.D.C., 85 (1969) 606.

Contents Home

You might also like