Sie sind auf Seite 1von 13

12

COLOR FADING IN TEXTILE MATERIALS


Color fastness depends on dye durability, the substrate properties, dye bonding and dis-
tribution in the material, and physical factors such as light intensity, air composition,
humidity, and temperature. In relation to these factors a dye may increase the degrada-
tion rate of material (phototendering) and may also inhibit photochemical changes in
the material. The major findings on these above aspects of color fading are discussed be-
low.
1-26
Azo-dyes forma broad group of compounds differing in chemical structure. In spite
of differences in chemical structure, some degradation mechanisms are common.
Freeman
19
proposes the following mechanism of degradation for red disperse dyes:
Color fading in textile materials 203
which is complementary to the major photoreductive cleavage of the azo linkage:
In the first process, the dye loses the pendant hydroxyethyl groups, whereas in the sec-
ond process the hydrazo-compound is produced and it then cleaves to a substituted ani-
line.
It is also possible, but not yet proven, that azo-dyes are degraded photooxidatively
by singlet oxygen:
204 G. Wypych
Such a mechanism may affect not only the dye but also the substrate.
Anthraquinone dyes undergo oxidative processes leading to a mixture of degrada-
tion products:
Color fading in textile materials 205
Several reactions occur, including dehydroxyethylation, dealkylation, and oxidation.
Grosjeans
5
studies on the fading of alizarin pigments showed that ozone is princi-
pally responsible for these changes. Alizarin Crimson was studied in this work. It is a
calcium-aluminum lake of 1,2-dihydroxy-anthraquinone. The reaction of Alizarin with
ozone produces a complex mixture of components suggesting the following mechanism
of degradation:
The mechanism shows that ozone may cause a dramatic change in the composition of
dyes by reacting with their aromatic components.
Photooxidationof trimethylmethane dyes was also evaluated. This process has sev-
eral stages which are initiated by electron ejection from a photoexcited dye cation. The
radical then reacts with oxygen, forming a hydroperoxide. This reaction is typical of
many other dyes and can be written in the following form:
A similar effect is caused by bond scission or hydrogen abstraction:
or
then
206 G. Wypych
An excited dye molecule may participate in an electron transfer to oxygen which gener-
ates an anion:
Dyes, like other materials, are mostly affected by radical processes and photooxidative
reactions which ultimately lead to the destruction of chromophores. The reactions with
substituents also lead to color modification by affecting auxochrome groups.
Dyes which have protective properties usually increase the stability of materials in
relation to their concentration. The deeper the shade of the dyed material, the better the
protection. Dyes which increase the degradation rate of the matrix usually do so at any
concentration.
22
Such dyes should not be used in materials that must be light stable.
Frequently, TiO
2
is added as a delustrant. Dull fibers contain approximately 2%TiO
2
.
22
The anatase form of TiO
2
causes degradation but in spite of this, it is frequently added
because it is less abrasive to the processing equipment. Morphological studies
22
show
that the addition of TiO
2
causes the formation of cracks and cavities in materials ex-
posed to UV radiation. Due to these processes, dull and semi-dull fibers are substan-
tially less UV resistant.
The chemical structure of dye cannot alone be responsible for all the color changes
observed. Polymer morphology strongly affects azo-dye photofading, as seen from Fig.
12.1 and from the data in Table 12.1. Poly(ethylene terephthalate) films used in the ex-
periment differ in morphology, as explained in Table 12.1.
Sample 1 was an unoriented film, sample 2 a two-way stretched film, and sample 3
a longitudinally stretched film. Fig. 12.1 shows that the film having the most compact
packing or the highest birefringence (sample 2) experienced the lowest dye loss. This
demonstrates that light stability is governed primarily by the supermolecular order of
the polymer matrix. Several dyes were tested on the same set of substrates and in each
case sample 2 had the best dye retention, meaning that the polymer packing was an
overriding factor in lightfastness. Two factors are involved: polymer packing and the
crystalline structure of the polymer. Better polymer packing slows down the oxygen dif-
Color fading in textile materials 207
Table 12.1: Morphology of PET films [Data fromP. L. Beltrame, E. Dubini-Paglia, B. Marcandalli,
P. Sadocco, and A. Seves, J. Appl. Polym. Sci., 33(1987)2965.]
Sample Molecular orientation Trans conformation index Birefringence
1 0 0.86 0.005
2 4 10
-1
2.02 0.140
3 4 10
-2
2.88 0.009
fusion rate which, in viewof chemical mechanisms discussed above, should decrease the
fading rate. The crystalline structure of the polymer affects the aggregation of dyes.
Dyes more aggregated in amorphous polymers have higher lightfastness than
monodisperse dyes. It is therefore important to consider morphology in a broader con-
text, since all these factors contribute to lightfastness.
No less important than the relationship of the dye distribution to polymer morphol-
ogy, is the degree of dye bonding with the polymer. Fig. 12.2 shows the dye absorption
spectra of an azo-disperse dye in cellulose diacetate and polyamide films.
This comparison shows that the azo-dye fades faster in a polyamide film. The au-
thors of this experiment
17
suggest that the faster fading of the dye in the polyamide film
is related to its weaker affinity to the polymer. Asimilar comparison is made for alizarin
dye on cellulose and silica gel substrate (Figs. 12.3 and 12.4). This study
5
compared the
ozone resistance of azo-dyes on two substrates. Alizarin dye on cellulose changed very
little, whereas the same dye on silica gel became colorless.
The lightfastness of reactive dyes (Remazol) on cellulose was examined in compari-
son with their R
f
values on thin-layer chromatographic plates. It was established that
the lightfastness correlates with the R
f
value (Fig. 12.5). The higher the R
f
value, the
weaker the bonding with the substrate, and consequently the faster the fading. Probaly,
208 G. Wypych
Fig. 12.1. Dye loss vs. exposure time in Xenotest
450 of poly(ethylene terephthalate) films. [Adapted,
by per mi ssi on, f r om P. L. Bel t r ame, E.
Dubini-Paglia, B. Marcandalli, P. Sadocco, and A.
Seves, J. Appl. Polym. Sci., 33(1987)2965.]
Fig. 12.2. Optical density vs. wavelength for dye on
cellulose diacetate (CD) and polyamide (PA) films
exposed in lightfastness tester for 140 h. [Adapted,
by permission, from F. M. Tera, M. N. Michael, and
A. Hebeish, Polym. Degrad. Stab., 16(1986)163.]
the stronger bond facilitates energy transfer and improves lightfastness. The method,
although indirect, offers a means of evaluating bonding strength and can forma base for
modelling the lightfastness of various combinations of dyes and substrates.
Carpignano
15
proposedthat the lightfastness of dyes relative to their chemical structure
could be predicted by
13
C NMR measurement. In this study, the lightfastness of 38 azo
dyes correlated with their respective chemical shifts, giving a basis for comparison and
prediction. The performance of the dye alone does not warrant that a particular dye is
suitable for a specific polymer but rather should be investigated in order to choose the
one best suited to a substrate.
Fig. 12.6 explains the effect of dyes on a matrix. The application of disperse dyes to
polyesters improves their resistance to phototendering. Polyamide, even when dyed
witha photoprotective premetallizeddye, is less resistant to weathering thananundyed
specimen. The type of dye obviously plays an important role in protecting the matrix
(Fig. 12.7). The concentrationof dye is also very important (Fig. 12.8). The concentration
effect depends on the combination dye-substrate (Fig. 12.9).
Color fading in textile materials 209
Fig. 12.3. Reflectance spectra of ozone exposed
and control samples of Alizarin dye on cellulose.
[Adapted, by permission, from D. Grosjean,
P. M. Whitmore, C. P. De Moor, G. R. Cass, and
J. R. Druzik, Environ. Sci. Technol., 21(1987)635.]
Fig. 12.4. Reflectance spectra of ozone exposed
and control samples of Alizarin dye on silica gel.
[Adapted, by permission, from D. Grosjean,
P. M. Whitmore, C. P. De Moor, G. R. Cass, and
J. R. Druzik, Environ. Sci. Technol., 21(1987)635.]
The orange dye used in this study offers photoprotection and an increase in its con-
centration increases the retention of tensile strength in cotton. In contrast, blue dye in-
210 G. Wypych
Fig. 12.5. Lightfastness of reactive dyes vs. R
f
val-
ues. [Adapted, by permission, fromW. S. Ha and C.
J. Lee, Melliand Textilber., 67(1986)724.]
Fig. 12.6. Strength retention of dyed and undyed
polyamide-66 and polyester fabrics vs. Florida ex-
posure time. [Data from B. Milligan, Rev. Prog.
Color Relat. Top., 16(1986)1.]
Fig. 12.7. Photoprotective effect on wool fabric of
selected premetallized dyes after exposure to
50,000 Langleys. [Data from L. Benisek, C. K.
Edmondson, and J. W. A. Mathews, Text. Res. J.,
55(1985)256.]
Fig. 12.8. Strength retention vs. exposure time of
Reactive Orange II dyed cotton fabric. [Adapted, by
permission, from C. M. Ladisch, R. R. Brown, and
K. B. Showell, Text. Chem. Color., 15(1983)209.]
creases the degradationrate whenits concentrationincreases. The performance of a dye
depends on its type and color. It has long been accepted that colors such as orange, yel-
low, and red are the least stable and also most degrading to the matrix material. Al-
though, Fig. 12. 9 shows that dyes color is not the only prerequisite for its stability since
a blue color should be generally more stable than an orange. Several reasons are quoted
in the literature to explain this behavior, such as the participation of the dye in
photoreduction of the substrate during the dark cycle (some data puts this reasoning in
question), hydrogen abstraction due to n
*
transition followed by intersystem cross-
ing and triplet formation, and, most frequently, participation in singlet oxygen genera-
tion. Dye types may be classified according to their performance but, when classified by
dye process or general dye structure, the guidelines should be used cautiously since
some chemical structure differences may influence performance in a way different from
other dyes in the group. Generally, dyes which have a strong affinity (chemical, physi-
cal) to the material are likely to increase durability.
Fig. 12.10 illustrates one of the possible factors through which a dye enhances sys-
tem stability. The ability of a dye to quench luminescent impurities correlates with the
retention of polyamide-66 properties. The phosphorescence intensity of material in-
creases as polymer molecular weight and thus polymer viscosity increase.
Color fading in textile materials 211
Fig. 12.9. Total strength loss of cotton relative to re-
active dye type and its concentration. [Data fromC.
M. Ladisch, R. R. Brown, and K. B. Showell, Text.
Chem. Color., 15(1983)209.]
Fig. 12.10. Viscosity number of solution of
polyamide-66 film containing variable amount of
quenching dye (Acid Yellow135) after 200 h irradia-
tion vs. phosphorescence intensity at 400 nm.
[Adapted, by permission, from N. S. Allen, M.
Leward, and G. W. Follows, Eur. Polym. J.,
28(1992)23.]
Dyes not only screen the polymer from radiation but also improve abrasion resis-
tance of polyamide-66 fabric (Fig. 12.11). Note also that the type of dye influences the ef-
fect in a crucial way. The protective effect of dyeing and aftertreatment can be shown by
studying polymer relative viscosity (Fig. 12.12).
10
A reduction in relative viscosity indi-
cates a decrease in molecular weight caused by polymer degradation frombond scission.
Hydroquinone reacts with oxygen, forming quinone, which prevents oxidation of poly-
mer, as seen from Fig. 12.12.
Finally, it was demonstrated in several studies that there is a relationship between
dyefastness and the protection given to the substrate by the dye. Dyes having high fast-
ness offer better protection to the substrate. The protective effect is increased by the ad-
dition of UVabsorbers and stabilizers which work well in conjuction with dyes. Atypical
combination uses a benzophenone type of UVabsorber with a copper complex as a stabi-
lizer. Organic copper complexes used in small amounts (less than 50 ppm) decompose
peroxides. Since long protection is provided, the mechanism is probably self-regenerat-
ing.
22
This proves, once more, that lightfastness of the dye is achieved not only throughits
composition but also because of the matrix in which dye was incorporated. Interaction
produces the final effect. This point is illustrated by a study of metal containing
polydyes.
2
Dyes, which were a part of the backbone of the polymer, have shown to have a
212 G. Wypych
Fig. 12.11. Abrasion resistance of polyamide-66
fabrics relative to the length of exposure and dye
color. [Data from G. A. Horsfall, Text. Res. J.,
52(1982)197.]
Fig. 12.12. Relative viscosity of polyamide-6 dyed
with vat dyes. [Data fromC. D. Shah and D. K. Jain,
Text. Res. J., 54(1984)844.]
permanent, non-leaching nature. Solutions containing monomeric dyes had a relatively
short life.
In addition to the effect that the chemical composition of dye and polymer, dye dis-
tribution, and morphology of the polymer has on weathering, external sources, such as
irradiation wavelength and intensity, temperature, and humidity also play role. The
most important factor in natural weathering is the length of exposure. Fig. 12.13 shows
the effect of seasonal variations in the daylight duration on the strength retention by
polyamide dyed with photodegradant dye (Nylomine Blue C-G) and protective dye (Acid
Brown GL-W). Samples in Florida were exposed to the same amount of radiation energy
(21,000 Langleys). Summer exposure was twice as damaging as the same amount of en-
ergy inwinter exposure. All samples behaved ina similar manner. The difference insea-
sons should account for the differences in temperature and humidity, as well as for
differences intensity of radiation. Energy flow rate influences photochemical changes.
Energy, depending onthe intensity, may be dissipated or utilized for photochemical con-
version.
Fig. 12.14 displays the results of weathering studies done with artificial light. This
shows the predominant role played by radiation wavelength in the processes which
cause dye fading.
Color fading in textile materials 213
Fig. 12.13. Strength retention vs. month in which
exposure of full dull polyamide-6 to 21,000 Lang-
leys begins. [Data fromG. A. Horsfall, Text. Res. J.,
52(1982)197.]
Fig. 12.14. Fading time to 10% fade vs. light wave-
length for dye on polyamide film. [Adapted, by per-
mission, from F. M. Tera, M. N. Michael, and
A. Hebeish, Polym. Degrad. Stab., 16(1986)163.]
The temperature at whichdegradationoccurs has predictable effect ondegradation
rate. Although the authors of the relevant papers did not control all essential variables
of the degradationprocess, it is obvious that, as inany other chemical reaction, tempera-
ture increase also increases the rate of photochemical processes. This concerns both re-
actions related to chromophore degradation and the degradation rate of the matrix in
which the dye is tested. Fromthe studies of polyamide fibers,
22
it was found that a tem-
perature increase (during processing and exposure) above 60
o
Ccauses formation of deg-
radation products which increases absorption in the UV range. These products cause
yellowing but also act as photosensitizers. If the same material is exposed to heat in the
absence of oxygen no acceleration of degradation occurs. These observations led re-
searchers to increase the temperature in testing of lightfastness of automotive fabrics.
The humidity effect is complex. Water may participate in photochemical reactions, gen-
erating very reactive and mobile radicals. The presence of water accelerates
hydroperoxide formation and causes hydrolytic processes. Additionally, water swells
the polymer matrix and acts as a plasticizer, which increases the mobility of
macromolecules and small molecules, thus increasing the probability of effective en-
counters between excited molecules and other neighboring molecules. Also, the trans-
port of energy and radicals is improved. The swollen structure of the material enhances
diffusion and stimulates oxidative processes. These factors work in combination, which
complicates the investigation.
REFERENCES
1. P. L. Beltrame, E. Dubini-Paglia, B. Marcandalli, P. Sadocco, and A. Seves,
J. Appl. Polym. Sci., 33(1987)2965.
2. C. E. Carraher, V. R. Foster, Q. J. Linville, and D. F. Stevison, Polym. Mater.
Sci. Eng., 56(1987)401.
3. G. MBon, C. Roucoux, A.Lablache-Combier, and C. Loucheux, J. Appl.
Polym. Sci., 29(1984)651.
4. R. L. Feller, R. M. Johnston-Feller, and C. Bailie, J. Coat. Technol.,
59(1987)93.
5. D. Grosjean, P. M. Whitmore, C. P. De Moor, G. R. Cass, and J. R. Druzik,
Environ. Sci. Technol., 21(1987)635.
6. P. C. Crews, Text. Chem. Color, 19(1987)21.
7. G. A. Horsfall, Text. Res. J., 52(1982)197.
8. B. Milligan, Rev. Prog. Color Relat. Top., 16(1986)1.
9. W. S. Ha and C. J. Lee, Melliand Textilber., 67(1986)724.
10. C. D. Shah and D. K. Jain, Text. Res. J., 54(1984)844.
11. M. A. Herlant, Am. Dyest Rep., 75(1986)38.
12. S. P. Watts and A. Sundaram, Text. Dyer Printer, 20(1987)13.
13. A. Wehlow, Text. Prax. Int., 43(1988)277.
14. C. M. Ladisch, R. R. Brown, and K. B. Showell, Text. Chem. Color,
15(1983)209.
214 G. Wypych
15. R. Carpignano, P. Savarino, E. Barni, G. Viscardi, S. Clementi, and
G. Giulietti, Anal. Chim. Acta, 191(1986)445.
16. K. Ogino, H. Uchiyama, and M. Abe, Colloid Polym. Sci., 265(1987)52.
17. F. M. Tera, M. N. Michael, and A. Hebeish, Polym. Degrad. Stab.,
16(1986)163.
18. F. M. Tera, M. N. Michael, and A. Hebeish, Polym. Degrad. Stab.,
17(1987)13.
19. H. S. Freeman and W.N. Hsu, Text. Res. J., 57(1987)223.
20. N. S. Allen, Rev. Prog. Color Relat. Top., 17(1987)61.
21. P. M. Latzke, Melliand Textilberichte, 7(1990)502.
22. G. Reinert, F. Thommen, and J. Isharani, Text. Chem. Color, 230(1991)31.
23. G. E. Krichevsky, G. T. Khachaturova, and O. M. Anissimova, Int. J. Polym. Mater.,
13(1990)63.
24. N. S. Allen, M. Ledward, and G. W. Follows, Eur. Polym. J., 28(1992)23.
25. W. B. Achwal, Colourage, 1(1991)29.
26. P. P. Klemchuk, Polym. Photochem., 3(1983)1.
Color fading in textile materials 215

Das könnte Ihnen auch gefallen