You are on page 1of 31

Wool

Wool
Helmut Zahn, Deutsches Wollforschungsinstitut an der RWTH Aachen, Aachen, Federal Republic of Germany (Chaps. 1, 2.1, 2.2, 6 11, 13) Franz-Josef Wortmann, Deutsches Wollforschungsinstitut an der RWTH Aachen, Aachen, Federal Republic of Germany (Chaps. 2.3, 3.1) Gabriele Wortmann, Deutsches Wollforschungsinstitut an der RWTH Aachen, Aachen, Federal Republic of Germany (Chap. 8, 10) Karola Schafer, Deutsches Wollforschungsinstitut an der RWTH Aachen, Aachen, Federal Republic of Germany (Chaps. 3.2, 5.4, 5.5, 10, 13) Rainer Hoffmann, Hochschule Bremen, Bremen, Federal Republic of Germany (Chaps. 4, 5.1 5.4, 12) Robert Finch, Woolmark Europe, Inc., Ilkley, United Kingdom (Chaps. 11, 12)

1. 1.1. 1.2. 1.3. 1.4. 2. 2.1. 2.2. 2.3. 3. 3.1. 3.2. 4. 4.1. 4.2.

Introduction . . . . . . . . . . . . . . . . Denition . . . . . . . . . . . . . . . . . . Historical Aspects . . . . . . . . . . . . Biology . . . . . . . . . . . . . . . . . . . Chemical Composition . . . . . . . . . Structure . . . . . . . . . . . . . . . . . . Morphology . . . . . . . . . . . . . . . . - and -Keratin . . . . . . . . . . . . . Two-Phase Model . . . . . . . . . . . . Properties . . . . . . . . . . . . . . . . . Physical and Mechanical Properties Chemical Reactivity . . . . . . . . . . . Resources and Raw Materials . . . . Sheep Rearing and Breeding . . . . . Places of Origin and Types of Raw Wool . . . . . . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . .

2 2 2 3 3 6 6 7 9 10 10 13 15 15 16

5. 5.1. 5.2. 5.3. 5.4. 5.5. 6. 7. 8. 9. 10. 11. 12. 13. 14.

Production Processes . . . . . . . . . . Methods of Obtaining Wool . . . . . Raw Wool Scouring . . . . . . . . . . . Mechanical Processing . . . . . . . . . Chemical Processing: Pretreatment Dyeing and Finishing . . . . . . . . . . Environmental Protection . . . . . . . Quality Specications . . . . . . . . . . Testing and Analysis . . . . . . . . . . Storage and Transportation . . . . . Uses . . . . . . . . . . . . . . . . . . . . . Trademarks . . . . . . . . . . . . . . . . Economic Aspects . . . . . . . . . . . . Toxicology and Occupational Health References . . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . .

16 16 16 17 17 19 21 21 22 23 23 23 24 26 27

1. Introduction
Today, wool accounts for only ca. 3 wt % of world textile-ber production. However, if the value added by dyeing and printing is included, the share in monetary terms is considerably higher. The importance of wool is due to its unique physiological properties as a clothing material and the wide range of techniques for converting it into textiles, including felts and nishing processes. Some important methods and discoveries in modern biological sciences had their origins in wool research (partition and paper chromatography, the structure of - and -keratin, intermediate laments).

The production of wool and woolen textiles and the trade in these commodities have played an important role in political and economic history.

1.1. Denition
The natural ber wool, like other types of ne and coarse animal hair and also silks ( Silk), is an animal ber [54], and although the term wool can be used in conjunction with the names of various animals, e.g., angora wool, it is here understood to include only the hair of the various breeds of domesticated sheep (Ovis aries).

c 2005 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim 10.1002/14356007.a28 395

Wool sheep, bred in Spain, produced wool of very ne quality, and in the late 1700s Saxon merinos were exported via England and South Africa to Australia, where the climate proved to be very favorable for rearing this breed.

1.3. Biology
Wool bers and other animal hairs, including human hair, are formed in follicles sheath-like formations of the epidermis, at the base of which are papillae of mesodermal cells (Fig. 1). Follicles are dynamic organs in which various processes occur simultaneously: cell division, cell differentiation, upward movement of the keratinocytes, biosynthesis of wool proteins as gene products, self-assembly of the proteins to form supramolecular structures, and nally keratinization (cornication) by formation of crosslinked disulde and isodipeptide bridges. At the same time, cells of the inner and outer root sheath are catabolized [58]. The various stages of development of the growing wool ber in the follicle can be divided into six zones (Fig. 1). In the upper part of Zone 1, ne keratin intermediate laments (KIF) and trichohyalin granules are already forming. Between Zones 1 and 2, laments aggregate to form macrobrils, in which keratin-associated proteins (KAP) are deposited. In Zone 2, amorphous protein aggregates, ca. 30 nm in diameter, form in the ber cuticle cells. By the end of Zone 2 the cuticle cells are hardened, while the nal keratinization of the entire wool ber occurs in Zone 3.

Figure 1. Diagram of fully developed wool follicle (not to scale) [58] The six zones are: 1) Bulb; 2) Keratogenous zone; 3) Zone of nal hardening of the ber; 4) Zone of inner root sheath degradation; 5) Zone of sloughing of inner and outer root sheath cells; 6) Pilary canal. AG = aprocine gland; AP = arrector pili; B = bulb; CTS = connective tissue sheath; DP = dermal papilla; E = epidermis; GM = glassy membrane; IRS = inner root sheath; ORS = outer root sheath; SG = sebaceous gland

1.2. Historical Aspects [5557]


The wool of the sheep was the rst textile raw material used by humans for clothing purposes. The development of clothing began when people rst dressed themselves in the skins of mammals. Felts, which are more easily shaped, were produced later from hairs plucked or cut from the animal ( Felts). The development of spinning and weaving to produce fabrics was a still later development. The rst country to process and trade in wool was Babylon (Babylonia = land of wool). The oldest known wool fabrics, dating from the second half of the second millennium b.c., were found in Danish tree cofns. Sheep rearing reached a high point in the Middle Ages and the Renaissance. The merino

1.4. Chemical Composition [27, 31]


The protein ber wool consists of carbon, hydrogen, oxygen, nitrogen, and sulfur. The elemental analysis of wool (free of water) is as follows:
Carbon Hydrogen Oxygen Nitrogen Sulfur Ash 50.5 wt % 6.8 wt % 22.0 wt % 16.5 wt % 3.7 wt % 0.5 wt %

Except for the sulfur content, this composition is typical for all proteins. The high sulfur

Wool content is due to the high content of cystine, a double amino acid containing two sulfur atoms in a disulde bond:
HOOCCH(NH2 )CH2 SSCH2 CH(NH2 )COOH

The side chains, which in wool account for a considerable proportion (50 %) of the protein material, interact with each other, thereby stabilizing the peptide by forming links between the chains and rings within a chain:

The ash contains potassium, sodium, calcium, aluminum, iron, silicon, sulfate, carbonate, phosphorus pentoxide, and chloride. Water-free wool consists mainly (ca. 97 %) of wool proteins, the remainder being made up of ca. 2 % structural lipids (Fig. 2) [5962], ca. 1 % mineral salts, nucleic acids [63, 64], and carbohydrates. Industrially scoured wool has only small residual amounts of wool grease (ca. 0.5 %). The structural lipids, in contrast, are components of the cell membrane complex which binds together the cuticle and cortex cells. A characteristic of the structural lipids of the cuticle is the presence of a covalently bound branched-chain C21 fatty acid 18methyleicosanoic acid [6567]. Total hydrolysis of the peptide bonds in wool proteins yields 24 amino acids. The data given in Table 1 were obtained by combining data from acidic and from enzymatic hydrolyses of wool. The amino acids are classied in ve groups: acidic amino acids, basic amino acids, amino acids with hydroxyl groups, sulfur-containing amino acids, and amino acids with no reactive groups in the side chain. The total amount of amino acids with reactive side chain groups is 5395 mol/g, and of amino acids without reactive side chain groups 3450 mol/g. The sum of these two gures, after subtraction of the concentration of C-terminal amino acids (10 mol/g), gives a gure of 8835 mol/g for the concentration of peptide groups in the peptide chains of wool. Like all proteins, wool contains both cationic and anionic groups, and is therefore amphoteric. The cationic character is due to the protonated side chains of arginine, lysine, and histidine, and to the small number of free amino groups at the ends of the peptide chains. The amino groups of lysine, histidine, the amino end groups, and the thiol groups of cysteine are important sites for the covalent attachment of reactive dyes [68]. Anionic groups are present as dissociated side chains of aspartic and glutamic acids and as carboxyl end groups.

This structural formula is a schematic representation of ve such links between segments of two hypothetical peptide chains. From the top downwards it shows interactions between phenyl rings (for the role of aromatic rings as hydrogen bond acceptors see [69]), hydrogen bonds between an asparagine residue and a serine residue, a salt bridge between an arginine residue and a glutamic acid residue [70], a disulde bridge between two cysteine residues, and an isodipeptide bridge between glutamic acid and lysine [71, 72]. The disulde bridge plays an important part in stabilizing the wool ber, leading in particular to its relatively high wet strength, moderate swelling, and insolubility. A second covalent bridge is provided by the isodipeptide N -(glutamyl)lysine, which provides an additional stabilizing effect in the cornied cell envelopes of the cortex and cuticle. The so-called salt bridge is due to the electrostatic interaction between cationic and anionic side-chain groups. Hydrogen bonds are formed in proteins between

Wool

Figure 2. Structural lipids

as donors and

as acceptors, especially in the helical rod domains of the keratin laments. Hydrophobic effects stabilize the aggregation of nonpolar side chains, so reducing the area of the interface with water. The three-dimensional structure of the wool proteins is stabilized both by the hydrophobic effect and by a wide variety of electrostatic interactions between amino acid constituents [73]. Eight N-terminal amino acids have been analyzed by means of 2,4-dinitrouorobenzene: cystine, glycine, threonine, valine, alanine, serine, and glutamic and aspartic acids. Other Nterminal amino acids are present in N-acetylated

form. Thus, the acid hydrolysis of wool liberates 50 mol/g acetic acid [74]. S-Carboxymethylated soluble wool proteins can be obtained by exhaustive mercaptolysis and carboxymethylation of the resulting SH groups [7577]. The dissolved S-carboxymethylated wool proteins can be separated into the main groups listed in Table 2, and these fractions can be separated into denite proteins, some of which have been sequenced [78]. In two-dimensional polyacrylamide gel electrophoresis [79], the proteins are rst separated according to charge, and, in the second dimension, in the presence of sodium dodecylsulfate (SDS), according to molecular mass (Fig. 3). The heterogeneity of this class of proteins can thus be demonstrated by a relatively simple analytical process. The keratins are well separated from the sulfur-rich proteins be-

Wool
Table 1. Amino acid composition of ne merino wool [49]

cause of their high molecular mass, and the high glycine tyrosine proteins are nearly at the front of the SDS dimension because their molecular mass is very low. Some of the heterogeneity of the keratins is caused by phosphorylation [80]. The keratins in wool come from the microbrils, which together with some other protein laments are classied as intermediate laments [8184]. Intermediate laments form an important lamentary system in the cytoplasm of higher eukaryotic cells. They have a diameter of 8 10 nm, and are thus larger than actin la-

ments and smaller than the microtubuli [41] of the cell.

2. Structure
2.1. Morphology [8589]
Wool bers consist of two cell types: cuticle and cortex cells. In coarse bers, there is also a central medulla (Fig. 4). The cuticle [87] consists of plate-shaped cells (scales) that overlap longitudinally and

Wool associated proteins (KAP) are intercalated, enveloping individual microbrils and their aggregates. An individual microbril consists of protobrils, which in turn consist of protolaments (Fig. 5) [89].
Table 2. Proteins of the wool ber [78] Type of protein Number of proteins Range of molecular mass, kDa 40 50 56

Keratins Acidic keratins (Type I) Basic keratins (Type II) Keratin associated proteins (KAP) High-sulfur proteins

4 4

80 100

Ultrahigh-sulfur proteins High glycine tyrosine proteins (Type I) High glycine tyrosine proteins (Type II)

10 5

11 16 19 23 16 69 69

Figure 3. Two-dimensional polyacrylamide gel electrophoretic separation of wool proteins [79]. The proteins were extracted from merino wool and subjected to electrophoresis at pH 8.9 in one direction and in the presence of SDS in the other. LS = low sulfur; HS = high sulfur; UHS = ultrahigh sulfur; HGT = high glycine tyrosine

The medulla in coarser wool bers consists of hollow cells with a skeleton of amorphous proteins and ne laments. The natural color of brown wool is caused by pigment granula of black to brown eumelanin and yellow to red pheomelanin.

peripherally, with the 1 m thick edges of the scales pointing in the direction of the tip of the ber. Each cuticle cell consists of four layers with different cystine and isodipeptide contents: epicuticle, a-layer, exocuticle, and endocuticle. Between overlapping cuticle cells is the cell membrane complex. The epicuticle consists of proteins and lipids, including covalently bound 18-methyleicosanoic acid, (Fig. 2). The cortex consists of spindle-shaped interdigitated cells [88]. These consist of ortho, para, and more rarely meso cells, all differing in their cystine content and staining behavior with silver salts. A cortex cell of wool contains around 5 8 macrobrils with a diameter of 300 nm at their widest point. Between the macrobrils are cytoplasmic and nuclear remnants of the keratinocytes. This intermacrobrillar material swells in water more than the macrobrils. A macrobril is a bundle of 500 800 keratin intermediate laments (microbrils). Keratin-

2.2. - and -Keratin


The microbrils are partly crystalline, and on X-ray structural analysis give reections on the meridian (0.51 nm) and equator (0.98 nm), as well as many long-period reections on the meridian and some on the equator. Astbury [90] designates the modication that results in this ber diagram (Fig. 6 A) as -keratin. Wool that has been stretched by ca. 50 % in water and then steamed gives a new ber diagram with a meridian reection at 0.34 nm and equatorial reections at 0.465 nm and 0.98 nm. This Xray diffraction pattern is due to the modication known as -keratin (Fig. 6 B). A new interpretation for the transition is given by CAO [96] and Kreplak et al. [97]. In -keratin [91], the peptide chains are twisted like a right-handed screw (-helix, Fig. 7). The stability of the -helix is due to intramolecular hydrogen bonds between the >C=O groups of the peptide bonds and adja-

Wool

Figure 4. Longitudinal and cross section of a ne merino wool ber [86]

Figure 6. Diagnostic X-ray reections [9] A) -Keratin pattern; B) -Keratin pattern

Figure 5. Schematic representation of the three distinct levels of brillar organization in keratin lament assembly and architecture [89]. I) The 2 nm protolament consists of two coiled-coil keratin polypeptides, presumably one acidic (white) and one basic (black) species; II) The 4.5 nm protobril may be a six-stranded subunit composed of three 2 nm protolaments (shown on left), or, alternatively, may be an eight-stranded subunit formed by the lateral aggregation of two four-stranded 3 nm protolament intermediates (shown on right); III) The 10 nm lament may comprise four (left) or three (right) protobrils, depending upon protobril substructure. The number of protobrils (four or three) per 10 nm lament is not unique but just represents one possible polymorphic form.

Wool is formed by cross-linking of the chains by hydrogen bonds between peptide groups of opposing chains. The sheet is folded at the CHR groups of the chains to such an extent that the meridian reection of the -keratin X-ray diffraction pattern at 0.34 nm corresponds with the geometry of the model. The phenomenon of the transformation has been applied in the new developed OPTIM technology (see Section 5.4).

cent >NH groups of amino acid units located on the next turn of the -helix and intrahelical salt linkages.

Figure 8. Pleated sheet structure of -keratin [92]

2.3. Two-Phase Model


The complex morphological and molecular structure of wool (Fig. 4) echoes the construction principle of all biological composite structures and many man-made materials, combining components with different properties in one material so as to maximize suitability for its purpose. Table 3 shows the stepwise differentiation of the morphological structure of the wool ber into the most important two-phase structures. With regard to elastic properties, the essential difference between the components is in their degree of order. If only the -helical central rodlike domains in the microbrils (intermediate laments) are regarded as microcrystalline, the crystalline phase accounts for 30 % of the ber [93]. The other components make up the noncrystalline phase, which includes the cuticle, the cell membrane complex, the intermacrobrillar matrix, the intermicrobrillar matrix, and 40 % of the microbrils, i.e., the nonhelical ends of the keratin molecules or of the dimers. On stretching the ber in the linearviscoelastic region ( < 0.8 %), the behavior of the crystalline phase (hx) is linearly elastic and

Figure 7. -Helix structure of -keratin [91]

The -helix contains 18 amino acid units in ve turns, i.e., 3.6 amino acid units per turn. To give the distance between successive turns of the helix that leads to the observed meridian reection (0.51 nm), the helical chain must itself be slightly coiled (superhelix, coiled coil). Two superhelices combine to form a left-handed, twostranded, rope-like assembly in which the superhelices are arranged such that the hydrophobic side chains on the outside of the helix interlink to form a stable buttonhole structure. These dimers are the actual physical structural elements of the intermediate laments (microbrils), and can be termed molecular twins. In the -keratin structure [92], the peptide chains are considerably stretched, as in silk broin ( Silk, Chap. 2.1.). The chains form socalled pleated sheet structures (Fig. 8). The sheet

Wool
Table 3. Morphology of wool as a combination of two-component structures (see also Fig. ) Composite system Wool ber Cortex Cortex cell Macrobril Type r/c f/m f/m f/m Component 1 cuticle cortex cell macrobrils microbrils Component 2 cortex cell membrane complex intermacrobrillar matrix intermicrobrillar matrix

r/c = ring/core structure; f/m = lament in matrix

that of the noncrystalline matrix (nc) is linearly viscoelastic. The modulus of elasticity of the helices is [94]
Ehx = 7.8 9.6 GPa

tional to its volume fraction. The elastomechanical behavior is satisfactorily represented by the simple Feughelman two-phase model [95]. The elastic and viscoelastic properties of the ber largely determine the crease resistance, dimensional stability, drape, and handle of a fabric.

This value is remarkably close to the modulus of elasticity of crystalline ice, which is also stabilized by hydrogen bonds (ice I, 0 C: E = 10 GPa). This fact emphasizes the ideal arrangement and cooperation of the short -helical segments in the microbrillar crystal. The timedependent modulus of elasticity of the dry (d) and wet (w) noncrystalline phase (nc) are calculated to be
d Enc =

3. Properties [2, 6, 17, 48]


(See also Fibers, 1. Survey; Fibers, 6. Testing and Analysis)

3.1. Physical and Mechanical Properties


[17] Wool bers with no medulla have a density of 1.31 g/cm3 at 25 C and 65 % relative humidity. The mean ber diameter varies between 16 and 40 m and the mean ber length between 2.5 and 25 cm, depending on type and origin.

6.1 6.6 GPa;


d w Enc /Enc 13

w Enc =

0.5 GPa

The ratio of the moduli of elasticity of the noncrystalline phase is in good agreement with the ratio of the corresponding torsion moduli (G d /G w = 10) [94]. In a lament matrix composite, the torsion behavior is mainly a property of the matrix. The small difference between the values for the viscoelastic modulus of the noncrystalline phase in the dry state and the modulus of elasticity of the helical laments is due to the fact that both components are stabilized mainly by strong hydrogen bonds which are broken under the inuence of water in the noncrystalline phase, while they continue to exist in the crystalline region. For small amounts of longitudinal strain in the linear-viscoelastic region, the relaxation behavior of a wool ber shows no ne structure ascribable to individual morphological components. This fact can be explained by the low contents of individual components and their chemical and structural similarity. For small deformations, all noncrystalline components behave as constituents of a homogeneous mixture, whereby the effect of each component is propor-

Figure 9. Moisture absorption of wool as a function of relative atmospheric humidity at 25 C [98]

Water Absorption. Wool is hygroscopic, and the amount of water taken up depends on the relative humidity of the air, temperature, and the history of the wool (Fig. 9) [98, 114, 115]. The adsorption and desorption curves exhibit hysteresis. At average relative humidities, the

10

Wool of the peptide chains, thereby increasing the swelling [99]. Many physical and mechanical properties of wool depend on its moisture content. Thus, for example, dry undamaged wool has an ultimate tensile strength of ca. 150 200 N/mm2 in its initial state and a wet ultimate tensile strength of only 70 80 % of this value. Whereas the tensile strength of wool decreases with increasing moisture content, the fracture strain increases. In the dry state, it is 35 55 %, and in the wet state 40 60 %. Stress Strain Behavior. Figure 11 shows the stress strain behavior of a wool ber in standard atmosphere of 65 % relative humidity at 20 C compared with that in water at 20 C. The stress strain curve can be divided into three regions which are affected to different extents by increasing humidity. After decrimping (not shown in Fig. 11), the tension in the ber increases very rapidly and almost linearly up to a strain of 1 2 %. This so-called initial region (A B) is often incorrectly referred to as the Hookean region. Above this, the elongation increases rapidly for small increases in stress. This section of the curve (B C) is known as the yield region. Point C lies between 25 and 30 % elongation. The third region of the stress strain curve is known as the post-yield region (C D), and is terminated by rupture of the ber (D). In this section of the curve, the increase in tension that accompanies a given increase in elongation is larger. The three characteristic regions of the stress strain curve are very distinct in water at 20 C, and the transitions at points B and C are sharp. The slopes in the initial, yield, and post-yield regions are in the approximate ratio 100 : 1 : 10. It has been shown that distinct domains of the keratin laments undergo the transformation in the three regions [116]. Noteworthy is that when the wool ber is stretched in water to point C at room temperature, i.e., by ca. 30 %, there is complete reversibility of mechanical properties if the time in the stretched state does not exceed ca. 1 h. If the ber is then relaxed and kept in water at room temperature overnight or at ca. 50 C for 1 h, not only does the ber return to its original length, but, on repeating the test, gives the same stress strain curve as in the rst cycle.

difference between the two curves is ca. 2 %. As with all textile bers, the moisture uptake of wool is accompanied by liberation of heat. Although wool is hygroscopic, its surface is hydrophobic towards liquid water and is therefore difcult to wet. This apparently contradictory behavior, which is an important factor in the physiological effects of wool as a clothing material, is due to the fact that the interior of the ber absorbs water vapor, whereas liquid water is repelled by the hydrophobic outer surface of the cuticle. Water absorption results in swelling, whereby wool shows considerable swelling anisotropy, i.e., longitudinal and radial swelling are different (Fig. 10) [10].

Figure 10. Effect of moisture absorption on swelling and moduli [10] a) Length swelling; b) Radial swelling; c) Torsional modulus; d) Extensional modulus

An increase in the amount of absorbed moisture from 0 % to 33 % leads to a longitudinal swelling of ca. 2 % and a radial swelling of 16 %. The extent of swelling depends on the pH and composition of the swelling medium. The degree of ber swelling is least at the isoionic point (pH 4.9), where there is no excess of ionic groups. Above and below this point, the number of stabilizing salt bridges decreases, and the wool carries excess negative or positive charges which cause electrostatic repulsion

Wool This high degree of reversibility is not exhibited by any other ber.
1 w1 w2 = + Tg Tg1 Tg2

11

The mass fraction of each component is represented by w, the indices 1 and 2 denoting dry wool and pure water respectively. By replacing w1 by (1-w2 ), the Fox equation becomes
1 1 1 = w2 Tg Tg2 Tg1 + 1 Tg1

If the reciprocal of the glass-transition temperature of the wool water system is plotted against the mass fraction of water, a straight line is obtained, conrming the validity of the Fox equation [101]. The glass-transition temperatures of the pure components can be calculated from the slope and the intersection of this line with the axis. The time-dependent recovery of wool bers after a xed torsional deformation as a function of temperature and water content has been measured. For dry wool, this gave T g1 = 447 K (174 C), and for water T g2 = 125 K (148 C) [100]. Figure 12 shows the glasstransition temperature of wool as a function of water content in the ber.
Figure 11. Tensile stress strain diagrams

In addition to this simple tensile test, tensile elasticity testing can provide further information about the behavior of the ber. Here the bers are stressed several times between constant elongation limits or constant tension limits. The test enables the permanent elongation and the elastic elongation to be determined. In this way, the elastic behavior of bers over the whole range of elongation up to the point of rupture can be determined. Glass-Transition Temperature. Knowledge of the glass transition temperature T g of wool [100] as a function of water content is of central importance for understanding its viscoelastic ber properties, especially under the varying conditions of moisture content and temperature that can exist during the manufacture and wearing of woolen fabrics, and also for an understanding of the aging and tempering processes of these fabrics. A simple equation due to Fox [101] represents the relationship between the glasstransition temperature of wool and its water content:

Figure 12. Glass-transition temperature of wool as a function of mass fraction of water in the ber [100]

The temperature dependence of the physical properties of wool water systems has been investigated by many authors using a variety of

12

Wool

methods. Wool in water has a T g of 5 C, signicantly below room temperature. Thus, water in wool eliminates aging and tempering effects because the thermal history of the wool is erased when T g is exceeded, in analogy with the behavior of other glassy polymers.

3.2. Chemical Reactivity [5, 13, 27, 31]


Dry Heat. Dry wool free of chemicals is stable for hours at 150 C. Short heating periods, e.g., 185 C for 30 s, produce tolerable changes in the bers. Moist Heat. Depending on the temperature, pH of the liquor, and duration of heating, heating with water can lead to degradation of wool. Heating with water at 130 140 C leads to the disappearance of the -keratin X-ray diffraction pattern with chemical decomposition. Wool is most stable near the isoionic point (pH 4.9), but chemical degradation occurs even below 100 C. The cysteine in the wool proteins is decomposed rst, liberating hydrogen sulde which then attacks the cystine in the wool and accelerates the decomposition autocatalytically [102]. This is the cysteine cystine self-degradation and self-cross-linking mechanism (thiol disulde degradation reaction).

Third step: Self-cross-linking by reaction of cysteine and lysine residues with dehydroalanine residues. Formation of lanthionine and lysinoalanine cross-links. Beside cross-links based on sulfur, also further cross-links are formed in wool during heat exposure, e.g., linking via formation of dityrosine [117, 118]. Acid Degradation. Mineral acids can degrade wool proteins to an extent that depends on pH, temperature, duration of reaction, and the presence of salts and surfactants. Sensitive sites in the wool proteins include the side chains of asparagine and glutamine, from which ammonia is released, and the peptide bonds formed by serine, threonine, cysteine, aspartic and glutamic acids, and tryptophan. The liquors from acid-degraded wool contain ammonium salts, free amino acids, and peptides (wool gelatins) [103,104]. The hydrolysis of peptide bonds leads to weight losses and the liberation of end groups in the wool proteins. Carbonizing by heating wool impregnated with sulfuric acid produces an N O peptidyl shift in the serine and threonine units of the peptide chain. The reaction is reversible if the acid is removed after the process, but if the wool is stored in moist air in the O-peptidyl form, the peptide chain is slowly and irreversibly hydrolyzed:

B = amino groups of histidine or lysine residues and main-chain end groups. First step: self-degradation of cysteine residues by heat-induced -elimination. Formation of dehydroalanine residues and hydrogen sulde.

Second step: self-degradation of cystine residues by reduction of disulde bonds with hydrogen sulde. Formation of S-thiocysteine and cysteine residues.

Degradation by Alkali. Wool is sensitive to alkali. In alkaline solution, the alkali binds to

Wool the wool proteins with swelling and simultaneous degradation. Even at low hydroxide concentrations, the reversible binding is accompanied by ber damage. The alkali binds to the wool so rmly that it cannot be completely removed by washing with water alone. The alkali salts of the wool proteins must rst be decomposed by reaction with stronger acids. Wool containing alkali residues undergoes decomposition on heating in a stream of dry air, the most important degradation reactions being the decomposition of cysteine and cystine. Alkalis hydrolyze the primary carbonamide bonds of asparagine and glutamine, cleave peptide bonds in the main chains, and cause racemization of the amino acid residues:
Cysteine dehydroalanine + hydrogen sulde Cystine dehydroalanine + thiocysteine Dehydroalanine + cysteine lanthionine Dehydroalanine + lysine lysinoalanine Dehydroalanine + ammonia -aminoalanine Glutamine glutamic acid + ammonia Asparagine aspartic acid + ammonia Isoleucine alloisoleucine.

13

Reducing Agents. Reducing agents are used in bleaching, chlorine removal, stripping of dyes, discharge printing, production of permanent creases by ironing, and chemical setting. Reducing agents such as sodium mercaptoacetate and mercaptoethanol reduce disulde groups to thiol groups. A 0.5 M solution of sodium mercaptoacetate dissolves wool above pH 10 in a few hours by reducing the cystine bonds responsible for insolubility:
WoCH2 SSCH2 Wo + 2 HSCH2 COONa 2 WoCH2 SH + (SCH2 COONa)2

where Wo = peptide chains of the wool proteins. Reaction of the thiol groups with iodoacetic acid produces S-carboxymethylkeratins:
WoCH2 SH + ICH2 COONa WoCH2 SCH2 COONa + HI

Some of the wool proteins become soluble on degradation. The solubility of wool in alkali under standard conditions is used as an indicator of chemical changes in the ber. Wool that has been damaged by acids, reducing agents, or oxidizing agents becomes sensitized to alkali, and its solubility in a standard solution of alkali increases. Degradation reactions are also revealed by yellowing of the wool. The decomposition reactions depend on the type of alkali, the temperature and duration of the reaction, the pH, and added salts and auxiliaries. Cationic surfactants increase the decomposition of cysteine and cystine and formation of lanthionine catalyzed by alkali. Nonionic surfactants have no effect on alkali damage to wool [105].

Reversible reduction of cystine bridges without dissolving the ber can be achieved with thiols if the reaction conditions are sufciently mild. This is the basis of the cold permanent waving process. The disulde bonds in wool can also be reduced by aqueous solutions of sodium sulde. This reaction is of industrial importance in the production of lime wool, obtained by dehairing the skins of slaughtered sheep. The sultolysis of wool is an important chemical reaction, e.g., in permanent setting. In the reaction with sulte or hydrogensulte, 1 mol Bunte salt (S-sulfonate anion) and 1 mol cysteine are produced per mole of cystine:
WoCH2 SSCH2 Wo + SO2 3 WoCH2 SSO 3 WoCH2 S +

WoCH2 SSCH2 Wo + HSO 3 WoCH2 SSO 3

WoCH2 SH +

Maximum cleavage occurs in the pH range 3 6. On rinsing, the Bunte salt and cysteine groups recombine to give wool cystine. Oxidative sultolysis converts disulde into two S-sulfonate anions:
2 WoCH2 SSO + 2 HO 3 WoCH2 SSCH2 Wo + 2 SO2 + [O] + H2 O 3

In the reaction of wool with reducing agents, cleavage of the stabilizing disulde bonds leads

14

Wool products at all stages of manufacture. Such damage can be detected microscopically by the specic color reaction with lactophenol cotton blue and by splitting of the bers into multicellular brillation products, cortex cells, and macrobrils. These (microbial) degradation reactions are caused by bacteria or fungal proteases. In undamaged wool, mainly the endocuticle and cortical cell nucleus residues are degraded hydrolytically by proteases. Macrobrils remain intact. Damaged wool is attacked more strongly, especially at weathered ber ends. Enzymes are of industrial importance in wool processing, e.g., in the dissolution of skin residues from wool felts by selective degradation, and the removal of skin-scale proteins from washed wool. Enzymatic dehairing of the small hides from sheep and goats is an established process. The use of enzymes for pretreatment and nishing is described in Sections 5.3 and 5.4.

to an increase in solubility. A linear relation exists between alkali solubility and cystine content of oxidatively and reductively bleached wool. At the same time, the enzymatic degradability increases. The dry stength does not always decrease, but the wet strength does. Plasticity increases, which is important for the process of permanent setting by deformation followed by oxidation. Oxidizing agents attack the disulde bonds. The reaction proceeds via various cystine oxides (sulfoxides and sulfones) and leads nally to cleavage of the disulde bond with formation of cysteic acid. Cysteine, methionine, tryptophan, and tyrosine are also attacked. The extent of oxidation depends on the oxidizing agent and its concentration, pH, temperature, and reaction time. Of industrial importance are oxidation reactions with hydrogen peroxide (for bleaching), and with potassium manganate(VII), chlorine, peroxysulfuric acid, and peroxyacetic acid to give antifelting effects. In the chloramide reaction, peptide bonds are chlorinated:
CONH + Cl2 CONCl + HCl

4. Resources and Raw Materials


4.1. Sheep Rearing and Breeding [44]
Sheep growing is best suited to drier regions of the world. Steppe and savannah pastures are suitable, but tropical regions are not. Sheep have also been bred in countries with intensive agriculture, where the waste products from arable farming are used as feed. Climatic factors such as altitude, precipitation, and air temperature are important for the growth of sheep and their wool. Sheep acclimatize well, and their bodies can adapt to extreme conditions and moderate food supplies. Depending on the climatic and economic conditions, sheep-farming countries specialize in wool production (Australia and South Africa) or the production of wool and meat (New Zealand, Argentina, and Uruguay). The crossing of English mutton rams and merino yews produces the Corriedale sheep of New Zealand. In Australia, merino rams have been crossed with Lincoln yews. These crossbreds have become established in South America as well as New Zealand, and are good producers of meat and wool. With the introduction of frozen meat technology in 1882, the crossbreds became very

Photodegradation [106113]. Wool shows only limited fastness to light. On exposure to daylight wool is bleached after a short time. After prolonged irradiation or irradiation with UV light the wool ber yellows. Visible blue light (380 475 nm) is responsible for photobleaching, while maximum photoyellowing occurs with UV-B radiation (290 320 nm). Photodamage results in the formation of a variety of oxidation states of cystine residues, and mercaptans. Phenylalanine, tryptophan, tyrosine, histidine, and proline residues are other targets for photodamage reactions. Scission of the main protein chain occurs with formation of keto-acyl end groups. Progressive formation of nonspecic protein cross-links (e.g., dityrosine) has also been observed. Enzymatic Degradation. Under unfavorable conditions, wet wool is attacked by bacteria or fungi, and this can lead to damage to wool

Wool important. The meat is mainly exported to the United Kingdom.

15

5.2. Raw Wool Scouring (See also


Textile Auxiliaries, Chap. 4.3.) Wool obtained by sheep shearing contains varying amounts of impurities which must be removed in the scouring process. The pure wool content of raw wool is variable, the following being typical average values for Australian raw wool:
Pure wool Wool grease Wool suint Soil and vegetable matter Moisture 55 % 10 % 4% 6 20 % 12 %

4.2. Places of Origin and Types of Raw Wool


Places of Origin. Australian and New Zealand wool; South African wool (also known as Cape wool) including Karakul and indigenous wool; European wool, e.g., German, English, Scottish, Irish, and French; South American wool, e.g., Argentinian (e.g., Buenos Aires, Bahia Blanca, and Chubut wool), Chilean (e.g., Punta Arenas wool), Uruguayan (e.g., Montevideo wool) and other South American wools from Brazil, Peru, etc.; exotic wools, e.g., from North Africa, Iran, and other Asian countries (China, India, Mongolia, etc.). Types of Sheep. Merino wool from merino sheep or sheep of the merino type, mean ber diameter usually 25 m; crossbred wool from crossbred sheep, mean ber diameter usually > 25 m; comeback wool from sheep obtained by crossbreeding crossbreds with merino sheep; coarse wool from coarse wool sheep, mean ber diameter usually > 30 m.

5. Production Processes
5.1. Methods of Obtaining Wool [18]
By far the largest proportion of wool is obtained by shearing live sheep. In addition, biological methods are used to recover wool from the skins of slaughtered animals. The product is known as sweated, pulled, or Mazamet wool, after the French town of that name. In New Zealand, wool recovered from skins is traded under the name slipes. It is obtained by treatment with a mixture of lime and suldes. Reclaimed wool, obtained by disintegrating used woollen textiles, differs in having considerably shorter bers and lower quality and purity. In accordance with textile labeling regulations, it is labeled as pure wool.

As wool is bought and sold on the basis of its pure wool content, the method of determining this has been standardized (IWTO-19-76). Wool grease is highly soluble in organic solvents such as benzene, hexane, acetone, and perchloroethylene. Consequently, there have been many attempts to use these solvents in wool scouring. However, scouring with water or aqueous alcohol is also necessary to remove the wool suint, skin-ake protein, and soil impurities. Solvent-based processes are, therefore, of minor importance. Solvent scouring systems are the de Smet plant using hexane and isopropyl alcohol as solvents [146], and the Wooltech system using 1,1,2-trichloroethylene as solvent [147, 148]. However, the major wool scouring is performed today by emulsion processes. To evaluate the scouring result, the residual fat content, whiteness, ash content of the wool, and pH of the aqueous extract are determined (IWTO Methods, see Chap. 8). In scouring machines, vigorous motion gives better washing, but can lead to a higher level of felting and hence to ber shortening during later stages of the manufacturing process. This problem has been solved satisfactorily by the sievedrum scouring process. Several other processes have been developed to improve wool scouring. These processes meanwhile all use suction drum bowls in combination with rake systems. The New Zealand Minibowl process and the Australian Siro Scour process have been installed in some industrial plants, but are not yet established in Western Europe. In wool scouring, ca. 6 20 L of wastewater is produced per kilogram wool, depending on the process and raw wool, and even if most of the wool grease is removed, e.g., by occula-

16

Wool centrifuge spinning, pot spinning, etc., are being tested in pilot plants. So far, none of these processes is used commercially In the production of woolen yarn, the slubbing is directly produced by splitting up the carded web, and is then spun on a ring spinning machine. Both worsted and woolen yarns are converted to fabrics by the standard methods also used for other types of ber (weaving, knitting, woven and tufted carpet manufacture, etc.). The property of the wool ber of migrating in various directions within the ber bundle on mechanical treatment and in the presence of water and auxiliaries, which is due to the frictional difference, is utilized in the production of felt from loose wool (carded web) or woven materials ( Felts, 2. Felting Process, Chap. 2., Felts, 3. Pressed Felts, Chap. 3.). Fiber migration is also used to cause fulling of wool fabrics and carpet yarns.

tion with acid and polyelectrolytes, the remaining wastewater is still highly contaminated. The pollutants still present in the wastewater can be 80 90 % degraded in multistage chemical and biological plants. Evaporation and incineration plants are able to destroy > 99 % of the pollutants and recover energy, water, and detergents. This has provided a satisfactory solution to the problem of wastewater purication. Although it is no longer worth recovering potassium carbonate from the scouring water, recovery of the byproduct wool wax is economically viable ( Waxes, Chap. 2.3.3.). Wool-scouring plants are today mainly located in the wool-producing countries, but woolcombing plants (which are normally integrated with wool-scouring plants for cost and quality reasons) are found both in the wool-producing countries (advantage of proximity to the raw material) and in the processing countries (advantage of proximity to the market).

5.3. Mechanical Processing


Total annual wool production, expressed as pure scoured wool, is 1.4 106 t in 1998, of which ca. 600 000 t is converted into worsted yarn and 400 000 t into woolen yarn and carpet yarn [149]. To produce worsted yarn, the scoured wool is rst disentangled by carding, which removes most of the nonbrous impurities. The web from the carding machine is gathered to form the socalled sliver. Several slivers are combined on the rst drawing frame and stretched, thereby improving uniformity. This is followed by combing, restretching, and nal stretching. The back washing process once carried out before nal stretching is not now usually used. The combed slivers (tops) from the nal stretching process, which have linear densities of 14 24 g/m, are an important commerical form of wool. The top undergoes further combing passes to give uniformity and can be blended with other types of ber, e.g., synthetic bers. The roving is then produced by stretching on rubber rollers or on a yer, and this is then spun to produce worsted yarn. Ring spinning frames are generally used for this. Rotor spinning frames are now mainly used for short bers and blends of short bers. Several developments to increase the production of the ring spinning process, like

5.4. Chemical Processing: Pretreatment


Carbonization ( Textile Auxiliaries). The mechanical processes of carding scoured raw wool and of combing prior to spinning worsted yarn remove most of the vegetable matter. However, some of these impurities remain in the wool, so that the next stages of processing, e.g., the conversion of noil to wool yarn, cannot usually be carried out immediately. Noils and woolen fabrics, which sometimes also contain unacceptable amounts of vegetable matter, are therefore carbonized. This name is derived from the fact that the impurities, which consist mainly of cellulose, form dark to black carbonization products after impregnation with sulfuric acid and heating. These can then be removed from the ber bundle because of their brittleness. In recent research, the use of enzymes to remove vegetable matter is being investigated [150, 151]. If carbonized material is not dyed while it is still wet and contains acid, the fabric must be protected from the effects of moisture and light, and the storage time must be as short as possible. Also, the dyeing method used must be as mild as possible.

Wool Antifelting Treatments. Most of the antifelting treatments developed for wool modify the surface of the ber by chemical attack. This affects the scales such that the difference between the roughness of the ber surface in the tip-to-root and root-to-tip directions is reduced or even eliminated. At one time it was believed that antifelting treatments had to remove the surface scales entirely. It is now known that this is not so; a treated ber will often look no different from an untreated ber under the microscope [29]. Antifelting nishes work by partially removing, softening, or coating the scale layer, or by preventing movement of the bers relative to each other by spot welding. Treatment with adhesives can only be used on the nished fabric, whereas the other processes can also be used on tops. They include chlorination and treatment with manganate(VII) or peroxysulfuric acid. This must be carried such that only the scale layer is attacked and not the interior of the ber. Precise process control gives very good stability to washing accompanied by low losses in weight and strength. A process widely operated in Germany that can also be used on the top is a combination of relatively mild chlorination followed by treatment with a polyamide epichlorhydrin resin (e.g., Hercosett 125). This is today the most important process for the production of machinewashable articles of pure wool, and gives products complying with the specications of the IWS Super Wash mark. The chlorination stage of the process is carried out with hypochlorite, whereas the Kroy process uses gaseous chlorine. In 1994, the total quantity of wool sliver treated worldwide in accordance with the IWS Super Wash specications was ca. 25 000 t. The so-called Vantean process, an oxidative process developed in Japan, comprises the following steps: 1) Pretreatment with heavy metal ions (can be omitted) 2) Treatment with hypochlorite in a saturated, strongly acidic salt solution 3) Treatment with sulte 4) Stabilization with formaldehyde The process removes the cuticle almost completely, and gives the wool a luster resembling that of much more expensive ne animal hairs.

17

For ecological reasons, work is directed towards the development of a new antifelting process that does not use chlorine or chlorine compounds [152]. The most promising developments so far are treatment with peroxysulfuric acid and manganate(VII), but a commercial process of this type for tops treatment is not yet available. Research is being carried out into antifelting effects produced by plasma technology [119] or topochemical enzyme processes [120]. Recently, an environmentally friendly procedure for antifelting treatment of wool was developed, involving replacement of the conventional chlorination by a treatment with a low-temperature plasma, followed by application of a new developed polyurethane based polymer [153156]. The development in plasma technology is far advanced and tests with a pilot plant are running. A commercial process will be available in the next years. Bleaching of wool is not as important as the bleaching of natural cellulose bers. For many purposes, the natural cream color of wool is acceptable, e.g., for garments nished in this color or for those dyed in dark shades. For white articles and sometimes for pastel shades, the wool must be bleached, preferably with hydrogen peroxide (see Textile Auxiliaries). Naturally pigmented dark wool bers require special treatment. Although they are present in very small numbers in high-quality wools (a few naturally colored bers in 10 g wool), they have a detrimental effect on the appearance of garments dyed in pale shades. The colored bers can be selectively bleached by a process in which the wool is pretreated with an iron salt, carefully washed, and then bleached with hydrogen peroxide [121]. For carpet wools, bleaching during raw wool scouring, simply by adding hydrogen peroxide or reducing agents to the last bowl, is a process of some importance. This practice can lead to increased wool yellowing during dyeing but is commercially attractive because the whiter wool attracts a higher price. The highest level of whiteness can be obtained by a combination of bleaching with hydrogen peroxide and reductive bleaching with, e.g., Blankit products, which contain sodium

18

Wool can be dyed in most stages of its manufacture, i.e., as ock, slubbings, yarn, or piece. The dyeing of ock is of importance in the production of melange shades, of woolen yarn consisting of blends with other types of ber or of large batches of uniform shades. The dyeing of sliver is useful for the production of large batches of completely uniform color. Color discrepancies can usually be leveled by mixing slivers without additional dyeing. Dyeing of wool blends is described in [45, Chap. 9]. The dyeing and printing of wool are treated in detail in Textile Dyeing, Chap. 5.1., Textile Printing, Chap. 7.). The most important dye classes for wool dyeing are afterchrome dyes (ca. 25 30 %), acid dyes (ca. 25 30 %), 1 : 1 metal-complex dyes (ca. 7 %), 1 : 2 metalcomplex dyes (ca. 30 %), and reactive dyes (ca. 10 %) [160]. Chrome dyes continue to fulll a high proportion of wool dyeing capacity, particularly as a cost effective, high-wetfastness system for slubbings and loose wool. Chromium residues (especially Cr(VI)) in the dyed textiles can cause allergies and toxic skin reactions in sensitive individuals. Due to the allergic and toxicological potential of chromium the usage of afterchrome dyes for wool dyeing is strongly decreasing. Furthermore, residual chrome creates a problem in dyehouse wastewater. This has been minimized by the use of modied chrome dyeing techniques [122124]. Reactive dyes represent an alternative method of achieving the high wetfastness of chrome dyes whilst eliminating the problem of heavy metal contamination of dyehouse, scouring liquors, and dyed textiles [161]. The tenacity of wool is often reduced in stock-dyeing processes. As a result of this, many bers are broken in the card. Most of the reduction in ber strength is due to setting of bends or curvatures of the packed bers by the hot water treatment. Such set ber bends are nonuniformly strained by tensions and thus resist smaller loads than bers set in the straight state [129]. In addition to the minimum wool damage exerted due to their optimum application pH in the isoionic region, reactive dyes further enhance wool quality by chemical reaction with thiol groups, H2 S, and histidine side chains, and by acting as antisetting agents [125128]. Research studies have been performed to dye wool from nonaqueous media, i.e., from super-

dithionite and sometimes also an optical brightener. However, care has to be taken in this bleaching process as ber damage can occur. Bleached and, to an even greater extent, optically brightened wool undergoes more yellowing than unbleached wool on exposure to light. Solubility in alkali is a sensitive indicator of bleaching damage to wool (see page 13). However, it is possible that although the alkali solubility (30 %) may be twice its normal value (12 17 %), a decrease in breaking strength of only 2 5 % occurs. New Developments. In the last few years there has been a growing demand for elastic fabrics and easy care properties, not only for sports and leisurewear but for formal wear also, particularly for wool and wool rich trousers (Wool plus Lycra). The disadvantage of elastomeric bers in the fabric is that these bers are not dyeable to the fastness standards of the main bers in the cloth. Some natural stretch can be produced in wool bers by subtle yarn technology, however, a 20 % stretch requires the application of reducing agents. Sulte-based textile auxiliaries were developed that provide worsted wool fabrics with a natural stretch [157]. This permanent elasticity of yarns and fabrics is achieved through supercontraction of the wool ber. A new technology for reduction of the ber diameter of wool bers has been developed by Woolmark Company and CSIRO; both, process and wool bers have the trade name OPTIM [158, 159]. OPTIM ne (permanently stretched ber) reduces the diameter of wool by 3 4 m. Another variant, OPTIM max (temporarily stretched ber) is a ber which is able to shrink by 20 to 25 % so that a higher bulkiness and softness is achieved. The process combines a stretching and either permanently or temporarily set in stretched conguration ( transformation) so that a new wool ber with different chemical and physical characteristics is produced [135], due to the transformation OPTIM ne has a more silk-like character than untreated wool [159].

5.5. Dyeing and Finishing [37, 40, 45, 50]


Dyeing and Printing. Dyeing is the most important wet nishing process for wool. Wool

Wool critical carbon dioxide [162, 163]. Wool can be dyed from supercritical carbon dioxide with hydrophobic dyes, improved fastness properties are achieved by application of reactive disperse dyes [162, 163]. Setting. According to de Boos [130], setting is an essential feature of the nishing of wool fabrics. It is used to stabilize fabrics prior to scouring and dyeing, to provide the required thickness, handle and visual properties to fabrics, to provide some durability to the nal nish, and occasionally to minimize dimensional changes in subsequent tailoring and wear. For these reasons, setting is an integral part of good nishing and fabric quality. Cohesive set in wool bers is thought to involve rearrangement of the hydrogen bonds in the ber, and is imparted whenever wool is distorted at temperatures above the glass-transition temperature (Fig. 12) of the ber and cooled while distorted. This set conguration of the fabric is stable as long as the bers remain below their T g . Permanent set in wool requires the rupture of disulde bonds in addition to the hydrogen bonds in the ber. Most authors [131, 132] believe that thiol disulde exchange reactions at temperatures in excess of 60 C are required to impart permanent set. Others [131135] consider the transition temperature for the onset of thiol disulde exchange reaction as the temperature required for chemical cysteine degradation and H2 S formation (see page 13). The decisive role of cysteine thiol groups as initiators of the reactions that lead to permanent set is common to both theories. Thus, all industrially proven methods of minimizing dyeing damage due to permanent set are based on the same principle, i.e., blockage or oxidation of the cysteine thiol groups during dyeing [136, 137]. The effectiveness of reactive dyes or other reactive chemicals is due to the fact that they block not only the cysteine thiol groups, which decreases the rate of H2 S formation, but also amino groups of the wool proteins, preventing the catalysis of cysteine degradation reactions. Shrink-Resistant Finishes. Fabrics or garments can be treated with nonchlorination processes which utilize polymers alone ( Textile Auxiliaries, Chap. 7.6.5.).

19

Protection from Moths and Beetles [138 141]. See Insect Control, Chap. 14.4.. Permethrin-based insect-resist (IR) products account for ca. 90 % of the European market, based on the volume of treated wool. Permethrin-treated wool has poor fastness to moisture and light, so that treatment with these products is recommended mainly for carpets and for storage protection. Conventional wool-specic protection agents based on Sulcofuron are Mitin FF h.c. (80 % AS, Ciba Specialty Chemicals), and Mitin FF liquid (Ciba Specialty Chemicals). Wool treated with Mitin FF has very good fastness towards moisture and light, and is therefore suitable for articles from which high standards of colorfastness are demanded.

Mothproong agents are usually applied during the exhaustion dyeing process, i.e., are added to the dyebath and are taken up by the wool with the dye. Other possible methods include application during scouring. Flameproong. Wool is less ammable than cellulose and synthetic bers (see Textile Auxiliaries, Chap. 7.5.). Its high ignition temperature, high limiting oxygen index, low heat of combustion, and low ame temperature are particularly advantageous and are connected with the chemical and morphological structure of the wool ber, which has high nitrogen (16 %) and moisture (10 14 %) contents. Negatively charged titanium and zirconium complexes with -hydroxycarboxylic acids and uorides can be exhausted on the positively charged wool ber in acid conditions, the result being a signicant improvement in the natural ame resistance of wool that is fast to washing and dry-cleaning. The complexes with carboxylic acids are exhausted on the wool ber

20

Wool reusable. Wool wastes are biodegraded by bacteria, fungi, and insects (moths, carpet beetles) and can be used as long-acting nitrogen fertilizers [165]. Wool production and wool industry, however, are faced with wastewater limitations due the pesticide residues in the clip [166], the highly polluted wool scouring efuent (see Section 5.2) [167], AOX in the shrink-resist processes, chromium in the dyeing process [168, 169], and the application of insect resist (IR) agents [170]. Dyeing and nishing with minimum environmental impact is crucial to the manufacture of eco-textiles [171, 172]. Modern efuent treatment removes pollutants by ecologically tolerable processes. By strict compliance with dyeing instructions the chromium(VI) content in the efuent can be kept within tolerable limits. Legal and environmental aspects of the textile nishing industry are described in Textile Dyeing, Chap. 14.; Textile Auxiliaries, 1.4 Environmental Aspects, Chap. 1.4., Textile Auxiliaries, 4.4 Ecological Aspects of Fabric Pretreatment , Chap. 4.4. A detailed and authoritative account of the environmental aspects of wool production and wool processing is afforded by the published proceedings of the two quinquennial wool textile research conferences, held in 1990 and 1995 [42, 52].

at the boil, whereas the uoride complexes can be effectively exhausted at lower temperatures. The uoride complexes can also be applied by a pad batch rinse dry technique. Titanium complexes are more effective than the zirconium ones, probably because of better penetration of the ber by the smaller titanium complexes. However, titanium complexes cause yellowing of wool, which increases with light exposure. Zirconium complexes do not affect the shade of wool and are fast to light [142]. Lightfastness Treatment. The effect of light on wool is rst to bleach it (photobleaching) and then to cause yellowing (photoyellowing). For interior woolen textiles photobleaching is the main problem because window glass lters short-wave UV light which induces photoyellowing. However, visible light (especially blue light, i.e., light with a wavelength of 400 450 nm), which causes a marked photobleaching, passes window glass. A new approach was developed by WRONZ-Institut to prevent photobleaching by application of a sulfonated aromatic compound (Lanalbin APB) [164]. Photobleaching is especially marked when woolen textiles are also thermally stressed (e.g., in automobiles). The lightfastness of wool, including bleached and dyed wool, can be improved by the application of sulfonated UV absorbers, hydroxybenzophenones (Uvinuls) or hydroxybenzotriazoles (Cibafast W) from the dyebath [143145].

7. Quality Specications [48, 53, 173]


Testing methods are described in Fibers, 6. Testing and Analysis, 2. Tests of Form and Mass, Chap. 2., Fibers, 6. Testing and Analysis, 3. Tensile Properties, Chap. 3. Quality of wool or tops implies a statement of the type, origin, and spinnability of the raw wool. An important quality criterion is mean ber diameter (neness), which inuences the softness of hand and is of paramount importance in spinning. Price is determined mainly by neness. Wool neness is specied by the mean ber diameter in micrometers. In the past, there were various systems of quality specication, e.g., as used by spinners, merchants, and sheep farmers. The most important systems are the British (numbers ranging from 100s to ca. 36s), French

Hydrophobization and Oleophobization Treatments. If necessary, wool, like other bers, can be made water- and oil-repellent by treating it with silicones or uorinated hydrocarbons.

6. Environmental Protection
Like the other organic natural bers, cotton and silk, wool is an environmentally friendly, renewable textile raw material. Woolen materials are

Wool (numbers in the ranges 150 100, I VI, and C), and German [letters in the range AAAA (4 A) to A, B, etc., up to EE]. Table 4 compares the British system with mean ber diameter and neness. Recently British mills have agreed that the existing micron denitions should be extended for superne wools to 100s 200s (i.e., 17.0 to 13.5 m mean ber diameter), because of increased demand for ne fabrics for traditional tailoring.
Table 4. British quality numbers, mean ber diameter, and mean ber neness [44] British quality number Mean ber diameter, Mean ber m neness, dtex 90s 80s 74s 64s 60/64s 60s 58s 56s 50s 48s 46s 17 17 18 18 19 21 22 23 25 28 33 34 37 3 3 3.3 3.3 3.7 4.5 5.0 5.4 6.4 8.1 11 12 14

21

8. Testing and Analysis [46, 48]


Chemical methods of testing wool to assess damage at the various stages of manufacture are well established. The use of a combination of methods enables the causes of impairment of mechanical properties to be elucidated with a high degree of certainty. The methods include the determination of foreign substances (wool grease, spinning oils, ash, etc.), amino acids [178], and alkali or urea bisulte solubility. It is not possible to quote mandatory values of mechanical and chemical data for a given wool. It is only possible to collect typical values for undamaged wools [179]. These analytical data vary on wet nishing of the wool in various directions and to various extents. Only in the case of certain standard treatments, such as carbonization, peroxide bleaching, or chlorination, can denite tolerances be given which must not be exceeded after a correctly performed process in order to ensure a defect-free product. Reliable gures are obtained if samples are taken by standardized methods at the various stages of a treatment process and analyzed. Most chemical testing methods for wool have been standardized by the Technology and Standards Committee of the International Wool Textile Organization (IWTO) [180]. The most important criteria for the assessment of changes to wool caused by industrial processing, especially wet nishing, are mechanical properties such as ultimate tensile strength. The dry and wet strength and elongation of yarn bundles are determined (IWTO method 32-1982). The dry values for undamaged wool are in the range 10 12 cN/tex, but these can decrease on dyeing. Decreases by up to 15 % of the maximum ultimate tensile strength are acceptable for nishing processes. Efcient separation methods for amino acids and proteins by chromatography and electrophoresis (see Amino Acids, Chap. 6.) are applied in wool research for estimation of lanthionine, lysinoalanine, histidinoalanine, cysteic acid in hydrolysates of wool and S-carboxymethyl proteins in mercaptolytic digests. Onedimensional electrophoresis allows up to 50 distinct wool proteins to be separated (Fig. 13). Comparison of electrophoretic patterns of digests of dyed and nished wool textiles allow specic statements on the involvement

The British grade number gives the number of hanks of 560 yards (512 m) that can be spun from 1 lb (453.6 g) of wool (spinning limit). For ne yarns, i.e., those with on average ca. 35 bers in the cross section, wool-berdiameter distribution affects spinning performance, as well as yarn evenness and tensile properties. Effective neness, a parameter derived from yarn-evenness theory, is a better predictor of (ne) physical yarn properties than mean ber diameter alone [174]. There are two systems for obtaining an objective measurement of hand. In the Kawabata System, 17 fabric properties relevant to the hand are determined with four measuring devices. These enable statements to be made about the stiffness, elasticity, and surface condition of the fabric, and objective dimensionless hand indices can be calculated from these if required [175, 176]. In the FAST System (Fabric Assurance by Simple Testing), 14 fabric properties are determined using three compact instruments. When these are plotted on a multiaxial graph (the socalled ngerprint), conclusions can be drawn about the processing properties of the fabric during manufacture [177].

22

Wool tions in the bale. The wool pack material should not pose problems in dyeing and nishing.

of morphological components and chemical constituents, such as keratins versus keratinassociated proteins [181184].

10. Uses [32]


Merino and crossbred wools are mainly used in clothing, home textiles (carpets, domestic furnishings, and bedding), to some extent in automobile upholstery (most often blended with polyester) and in aircraft seatings [186188]. According to [189], the wool ber can now be found in applications as diverse as thermal insulation, weed mats, geotextile products, and as a means to soak up oil spills. Wool has been used for thermal insulation in houses in Europe for many decades and is being increasingly used for both homes and commercial buildings [190 194]. The natural property of wool to effectively absorb specic classes of widespread indoor air pollutants, namely aldehydes, SO2 and NOx , has led to new applications in the private and commercial sector [195197]. Wool is naturally re resistant, and the increased use is being driven by consumer demand for natural products in their homes. The moisture absorbing and desorbing properties of wool can also contribute to passive heating and cooling effects and, led, furthermore, to the use of wool for high-performance sportswear clothing. Of minor importance are special wool products, especially felts, e.g., for lter materials. Keratin hydrolysates can be isolated from wool by acid, basic, or enzymatic hydrolysis in combination with a reductive (or sometimes oxidative) treatment. These hydrolysates are used as ber protective agents during wool dyeing [198], for cosmetics, or for production of new textile materials [199]. l-Cystine can be isolated from hydrolysates of hair or wool (see Amino Acids, Chap. 3.2.4.).

Figure 13. Schematic electrophoretic protein-separation pattern of merino wool [183]

9. Storage and Transportation [185]


Wool must be stored and transported under normal atmospheric conditions. Bales of raw wool, scoured wool, tops, or any other form should be kept dry to avoid bacterial and fungal attack. The packaging material must allow some exchange of air to avoid damp and high-humidity condi-

11. Trademarks
The Woolmark Company was founded in 1937 as the International Wool Secretariat. Today it is known as The Woolmark Company and is the worlds leading wool textile marketing organization. Together with its three well known trademarks, it is synonymous with research and

Wool development, innovation, promotion, and quality assurance in pure new wool and wool blend products (cf. http://www.wool.com). The Woolmark symbol was designed by an Italian, Franceso Sarroglia to identify quality products made from pure new wool (Fig. 14). Textiles with this mark must consist of 100 % virgin wool. Admixtures of wool recovered by mild methods from the skins of slaughtered sheep are permitted. Equal status is given to ne animal hair from the alpaca, llama, vicuna, guanaco, camel, angora, rabbit, angora-mohair, cashmere and cashgora goats, and yak, provided that here also it is obtained by a mild process. The Woolmark was launched in 1964, and is now a registered trademark in over 140 countries. It is used by companies licensed by the Woolmark Company in more than 65 countries extending across a wide range of products clothes, carpets, home furnishings, washing machines and detergents. In 1970, the designation pure new wool (Reine Schurwolle) was included in the textile labeling law (TKG) of the Federal Republic of Germany, and became valid in the European Community in 1972. The Woolmark Blend symbol (originally called Woolblendmark) was introduced in 1971 to identify quality products made from wool rich blends (products comprising at least 50 % new wool). The Wool Blend symbol was introduced in mid1999 to extend the opportunities offered by The Woolmark Companys blends program; the Wool Blend symbol is used to identify products that contain at least 30 % (but less than 50 %) new wool (Fig. 14). The same high performance standards apply to all three trademarks. Before a manufacturer can use the Woolmark or any trademarks of the Woolmark Company he needs to apply for a license, and pay the relevant usage fee [200]. He also needs to be able to meet the quality standards that are required under the Woolmark program. Maintaining consistent high quality is a key feature of the Woolmark program, and it is important that all licensees have good in-plant quality assurance procedures in place. In addition to its three main brands, the Woolmark Company has introduced a series of programs, which highlight selected product attributes. These programs are backed by specications and clear product descriptions to ensure a clear and globally consistent message. Each pro-

23

gram is also identied by a specially designed sub-brand logo; the logos are featured on a range of promotion tickets, including Cool Wool, Merino Extrane, Pure Merino Wool, Machine Washable Wool, Total Easy Care Wool, Wool plus Lycra [201], Wool Cotton, Sportwool [159]. Further details are available from The Woolmark Company at http://www.woolmarkbrand.com.

12. Economic Aspects


Wool Production. Table 5 gives world wool production gures for the major producing countries during the period of the shearing seasons from 1995/96 to 1999/2000; Figure 15 shows percentage shares of world production. Table 5 shows actual weights of raw wool or lime wool. Wool production has grown only slowly over recent decades, as grazing land for sheep cannot be increased indenitely. A decline has even occurred since 1990/91 for reasons of climate, price (lucrative alternatives for farmers), and political economy (economic decline in Eastern Europe). Sheep rearing and hence wool production exist in almost all parts of the world, although at widely varying levels of intensity. In most countries, home-produced wool is processed and consumed internally, and therefore is not available on the international market. This also applies to important wool-producing countries such as China and the countries of the former Soviet Union. The suppliers of the raw material to the international wool textile industry are those countries that produce and export amounts of wool in excess of their own requirements. These principally comprise Australia, New Zealand, Argentina, Uruguay, and South Africa. Australia is the worlds largest producer and exporter of garment wools (merinos), exporting ca. 60 % of its wool production as raw wool, and thus being the most important supplier for the wool combers of the industrialized countries. New Zealand produces mainly scoured coarse wools that are exported for use in the production of home textiles, especially carpets. Argentina, Uruguay, and South Africa have also built their own wool scouring and combing plants, and export wool either in scoured loose form or as tops (scoured and combed wool). Table 6 lists

24

Wool

Figure 14. Woolmarks for pure new wool and new wool with other ber
Table 5. World wool production by the major producing countries (in 106 kg) (source IWTO). 1995/96 1996/97 1997/1998 1998/99 1999/2000

Raw wool weight Australia 697 725 700 684 671 New Zealand 269 275 266 252 256 Former Soviet Union 238 196 164 140 135 China 277 298 255 277 290 Argentina 81 78 68 70 63 Uruguay 79 85 78 63 58 South Africa 65 59 54 58 53 Others 826 804 811 812 814 World 2532 2520 2396 2356 2340 Merinoa 1077 1061 992 976 957 Crossbredb 626 620 582 557 551 Carpet woolsc 836 839 822 823 832 b c a Merino wool covers 24.5 m and ner. Crossbred wool covers 24.6 32.5 m. Carpet wool covers 32.6 m and coarser.

wool export gures, for the ve most important exporting countries for the seasons 1994/95 to 1998/99.

Figure 15. World wool production (1997/98)

Wool Prices. The price differences between the various wool qualities are considerable. The most important factors determining quality and therefore price are the neness and staple length of the ber. Not only the mean neness and length, but also the uniformity of distribution are of great importance. Other quality and price parameters are the content of vegetable matter and

pigmented bers, tensile strength, color, elasticity, crimpiness, and many others. Since the collapse in early 1991 of the minimum price system which imposed a lower limit on wool prices for 17 years in Australia, New Zealand, and South Africa, prices have been determined by the market forces of supply and demand. Figure 16 shows the development of the average auction price of Australian raw wool, the so-called market indicator (i.e., the weighted average of the qualities) in the sales seasons 1994/95 to 1999/2000. The market indicator during this period uctuated between 5.1 A$/kg (March 1999) and 8.6 A$/kg (March 1995). However, the price differences between two given qualities can vary greatly over time with supply and demand. This is illustrated in Figure 16, which also shows the price curve for 19 m wools. The difference in price between this and the market indicator varied between 0.80 A$/kg (March 1996) and 7.85 A$/kg (September 1994). The reasons for the relatively low prices for ne wools may be an increased supply of hunger-ne wools after a drought, and the lack of interest in such low-quality, albeit ne wools. A high price is evidence of a strong,

Wool
Table 6. Wool exports (raw wool, scoured wool, and sliver) by the ve major exporting countries (in 106 kg scoured wool) 1994/95 Australia New Zealand South Africa Argentina Uruguay Total 591 228 26 46 15 906 1995/96 534 196 29 32 14 805 1996/97 608 200 20 34 17 879 1997/98 556 192 24 25 8 805 1998/99 475 171 25 24 8 703

25

Figure 16. Price development of Australian raw wool a) Average auction price (market indicator); b) 19 m indicator

fashion-induced demand accompanied by a restricted supply.


Table 7. Consumption of shorn wool by the wool textile industry spinning mills in 1998 (in 106 kg scoured wool) 106 kg % Italy Germany UK Spain Other Western Europe Central and Eastern Europe Total Europe Africa (incl. South Africa) Middle East Total +Africa and Middle East North America Central and South America Total Americas India Other Indian Subcontinent Total Indian Subcontinent China and Hong Kong Japan South Korea Taiwan Total Far East Australia, New Zealand World 263.1 70.5 42.1 23.6 85.4 110.6 495.2 46.2 138.0 221.2 46.2 39.8 86.0 75.8 66.9 142.7 280.1 61.3 34.9 8.7 385.0 38.6 1372 42 18 11 6 22 8 36 3 10 16 3 3 6 53 47 10 73 16 9 2 28 3 100

Wool Consumption. Table 7 lists world wool consumption. Europe has the highest consumption of wool (36 % of the world total). This underscores the world importance of the European woolen industry. The four EU countries Italy, Germany, United Kingdom, and Spain account for 29 %, the rest of Western Europe 14 %, and Central and Eastern Europe 8 % of total world consumption for spinning mills. In second place is the Far East, with 28 % of the total world industrial demand for wool. For German wool imports, see Table 8. The comparatively minor importance of the wool textile industry in the United States, i.e., 3 % of the worlds consumption of virgin wool, is remarkable.

13. Toxicology and Occupational Health


The wool ber itself poses no health problems. On the contrary, woolen clothing protects against heat, cold, water, and sun as well as mechanical impact [203]. Wool products that have reached the end of their useful life may be recycled via the woolen

26

Wool

Table 8. German imports of wool and tops (in t) (Source: Comtext, Eurostat) Import Category Wool, incl. tops and noils (clean basis) Raw wool (clean basis) greasy (actual) scoured (actual) Animal hair Tops, etc. Noils and waste 1999 (full year) 63 302 33 800 33 969 12 434 1 732 23 944 5 588 Jan April 1999 19 874 10 705 8 873 4 671 819 9 136 1 938 Jan April 2000 18 826 10 924 9 417 4 520 1 059 8 378 1 748

process and can be used for production of wool materials of minor quality or technical products. Wool is biodegradable. However, pretreatments or presence of contaminants such as pesticides or textile nishing agents like mothproofing agents, ame retardants or dyes have impact on the ecotoxicology of wool materials. Prickle [204208]. No health risks are involved in wearing properly nished and dyed wool clothing, even if it comes in direct contact with the skin. Wool incompatibility is caused neither by toxic components of wool, nor is it a wool allergy; it is a purely mechanical irritation of the skin. Wool prickle is due to the coarser wool bers digging into human skin with enough force to excite pain receptors. Fabrics made from wool with no more than 5 % of bers larger than 30 m will not be perceived as prickly by most people. This mechanical skin irritation is not specic to wool, but can be caused by other natural and synthetic bers if sufcient coarse ber ends per unit area of skin are present. Azo Dyes [209211]. Properly dyed wool articles do not contain such azo dyes that could set free carcinogenic aromatic amines such as benzidine. Members of the Ecological and Toxicological Association of Dyes and Organic Pigment Manufacturers (ETAD) ceased the manufacture of benzidine-based dyes in the early 1970s due to the inability to ensure their safe handling in dyeing plants. However, because of the particular economic and technical merits of these dyes, they have continued to be manufactured in many countries. Provided the dyed goods have good fastness characteristics, there is not a concern about consumer risk.

Occupational Health. For the ecotoxicological potential of textile dyes, see Textile Dyeing, Chap. 14.7.. In the processing of wool, problems are caused by exposure of workers to noise and to respirable and alveolar dust [212, 213, 215]. There is a potential human health hazard to shearers and farm workers handling the treated sheep and wools. Users of lanolin products are at lower risk, especially if the commercially available reduced pesticide grades of lanolin are used where there is human skin contact. All the existing ectoparasiticide products are evaluated for potential hazard to shearers and farm workers for incorporation into the protocol of the Australian Occupational Health and Safety Commission. As measure of the safe exposure level the acceptable daily intake is compared with the potential exposure calculated from the amount of active ingredient present and the amount of wool grease covering the shearer [214]. However, the use of toxic agents for sheep dips, etc., is today discouraged and nonharmful alternatives are recommended. Also operational procedures are now strictly laid down to control and minimize any risks to farm workers, etc. Acknowledgement.The authors thank Hartwig Hocker (Director of the Deutsches Wollforschungsinstitut an der RWTH Aachen) for checking the manuscript and Klaus Wolf, Hans-Georg Hebecker, Peter Dufeld, and Kevin Russel for helpful suggestions.

14. References
1. H. Glafey, D. Kr ger, G. Ulrich: Technologie u der Wolle Chemische Technologie und mechanische Hilfsmittel f r die Veredlung von u Wolle, in Technologie der Textilfasern, vol. VIII, part 3 B, Springer Verlag, Berlin 1938.

Wool
2. E. R. Kaswell: Textile Fibers, Yarns, and Fabrics A Comparative Survey of their Behavior with Special Reference to Wool, Reinhold Publ. Co., New York 1953. 3. R. W. Moncrieff: Wool Shrinkage and its Prevention, National Trade Press, London 1953. 4. W. H. Ward, H. P. Lundgren, Adv. Protein Chem. IX (1954) 243 297. 5. P. Alexander, R. F. Hudson: Wool. Its Chemistry and Physics, Chapman & Hall, London 1954. 6. M. Harris: Handbook of Textile Fibers, Wiley-Interscience, New York 1954. 7. W. G. Crewther (ed.), Proc. Intern. Wool Text. Res. Conf., Australia 1955, vol. A F, Commonwealth Scientic and Industrial Research Organization, Melbourne 1956. 8. C. E. M. Jones (ed.), 2nd Quinquennial Wool Text. Res. Conf., Harrogate 1960; J. Text. Inst. 51 (1960) T 489 T 1552. 9. E. H. Mercer: Keratin und Keratinization, Pergamon, Oxford 1961. 10. W. J. Onions: Wool An Introduction to its Properties, Varieties, Uses and Production, E. Benn, London 1962. 11. P. Alexander, R. F. Hudson in Ch. Earland (ed.): Wool Its Chemistry and Physics, 2nd ed., Chapman & Hall, London 1963. 12. H. Doehner, H. Reumuth (eds.): Wollkunde, 2nd ed., P. Parey, Berlin Hamburg 1964. 13. W. G. Crewther, R. D. B. Fraser, F. G. Lennox, H. Lindley, Adv. Protein Chem. 20 (1965) 191 346. 14. R. Delerive, F. Maillard (eds.): Proc. 3rd Int. Wool Text. Res. Conf. Cirtel Paris 1965, Sections I IV, Institut Textile de France. 15. J. G. Cook: Natural Fibers, Handbook of Textile Fibers, vol. 1, Merrow Publ. Co., Watford, Herts 1968. 16. W. G. Crewther (ed.): Symp. Fibrous Proteins 1967 Butterworths Australia 1968. 17. E. G. Bendit, M. Feughelman in H. F. Mark, N. G. Gaylord, N. M. Bikales (eds.): Encylopedia of Polymer Science and Technology, vol. 8, Wiley-Interscience, New York 1968, pp. 1 44. 18. H. Behrens, H. Doehner, R. Scheelje, R. Wassmuth: Lehrbuch der Schafzucht, P. Parey, Hamburg Berlin 1969. 19. W. von Bergen: Wool Handbook, 3rd ed., vol. I, Wiley-Intersience, New York London 1969. 20. H. Spiebeg (ed.): British Wool Manual, 2nd ed., Columbine Press, Buxton 1969.

27

21. K. R. Makinson in H. F. Mark, N. G. Gaylord, N. M. Bikales (eds.): Encyclopedia of Polymer Science and Technology, vol. 15, Wiley-Interscience, New York 1971, pp. 41 79. 22. L. Rebenfeld, H. H. Ward (eds.): Proc. 4th Int. Wool Text. Res. Conf. Berkely 1970, Parts I, II, Applied Polymer Symposia, no. 18, Interscience Publishers, New York 1971. 23. H. Rath: Lehrbuch der Textilchemie einschlielich der textilchemischen Technologie, 3rd ed., Springer Verlag, Berlin 1972. 24. R. D. B. Fraser, T. P. MacRae, G. E. Rogers: Keratins Their Composition, Structure and Biosynthesis, Charles C. Thomas, Springeld, Ill. 1972. 25. J. H. Bradbury, Adv. Protein Chem. 27 (1973) 111 211. 26. K. Ziegler (ed.): Proc. 5th Int. Wool Text. Res. Conf. Aachen 1975, vols I IV, Deutsches Wollforschungsinstitut an der RWTH Aachen 1976. 27. R. S. Asquith (ed.): Chemistry of Natural Protein Fibers, Plenum Press, New York London 1977. 28. H. E. Schiecke: Wolle als textiler Rohstoff, Fachverlag Schiele & Sch n, Berlin 1979. o 29. K. R. Makinson: Shrinkproong of Wool, in Fiber Science Series, vol. 8, Marcel Dekker, New York Basel 1979. 30. D. P. Veldsman (ed.): Proc. 6th Quinquennial Int. Wool Text. Res. Conf. Pretoria 1980, vols I IV. 31. J. A. Mclaren, B. Milligan: Wool Science The Chemical Reactivity of the Wool Fibres, Science Press, Marrickville, Australia 1981. 32. J. Leeder: Wool Natures Wonder Fibre, Australasian Textiles Publishers, Belmont/Victoria/Australia 1984. 33. E. Wang, D. Fischman, R. K. H. Liem, T.-T. Sun (eds.): Intermediate Filaments, Annals of the New York Academy of Sciences, vol. 455 New York 1985. 34. M. Sakamoto (ed.): Proc. 7th Int. Wool Text. Res. Conf. Tokyo 1985, vols I IV, Soc. Fiber Science and Technol. Japan. 35. I. Bearpak, F. W. Marriott, J. Park: A Practical Introduction to the Dyeing and Finishing of Wool Fabrics, Society of Dyers and Colourists, Bradford 1986. 36. M. Feughelman in H. F. Mark, N. M. Bikales, C. G. Overberger, G. Menges, J. I. Kroschwitz (eds.): Encyclopedia of Polymer Science and

28

Wool
Engineering, vol. 8, Wiley-Interscience, New York 1987, pp. 566 600. G. Ebner, D. Schelz: Textilf rberei und a Farbstoffe, Springer Verlag, Berlin 1989. G. E. Rogers, P. J. Reis, K. A. Ward, R. C. Marshall (eds.): The Biology of Wool and Hair, Chapman & Hall, London New York 1989. M. Ried: Chemie im Kleiderschrank, Rowohlt Verlag, Reinbeck 1989. M. Peter, H. K. Rouette: Grundlagen der Textilveredlung, Handbuch der Technologie, Verfahren und Maschinen, 13th ed., Deutscher Fachverlag, Frankfurt 1989. R. D. Goldman, P. M. Steinert (eds.): Cellular and Molecular Biology of Intermediate Filaments, Plenum Press, New York London 1990. G. H. Crawshaw (ed. ): Proc. 8th Int. Wool Text. Res. Conf. Christchurch New Zealand 1990, vols. I IV, Wool Res. Org. New Zealand. R. C. Marshall, D. F. G. Orwin, J. M. Gillespie, Electron Microsc. Rev. 4 (1991) 47 83. H. Zahn, B. Wulfhorst, H. K lter, u Chemiefasern/Textilind. 41 (1991) 521 552. D. M. Lewis (ed.): Wool Dyeing, Society of Dyers and Colourists, Bradford 1992. H. Zahn, J. Knott in EEC (ed.): Chemical Methods for Characterization of Wool at Different Stages of Processing, COMETT-EUROTEX, Guimaraes, Portugal 1992. F.-J. Wortmann: Thermo- and hydroplastische Eigenschaften von Wollfasern, Forschungsberichte des Landes Nordrhein-Westfalen, no. 3245, Westdeutscher Verlag 1992. W. Bobeth (ed.): Textile Faserstoffe Beschaffenheit und Eigenschaften, Springer Verlag, Berlin 1993. R. Kellner, F. Lottspeich, H. E. Meyer (eds.): Microcharacterization of Proteins, VCH Verlagsgesellschaft, Weinheim 1994. M. Bona: An Introduction to Wool Fabric Finishing, Verlag Textilia, Biella, Italy 1994. P. H. Greaves, B. P. Saville: Microscopy of Textile Fibres, Microscopy Handbooks 32, Royal Microscopical Society, BIOS Scientic Publ., Oxford 1995. M. Bona, G. Grupallo: Proc. 9th Int. Wool Text. Res. Conf. Biella Italy 1995, vols. I V, Citta degli Studi Biella and Int. Wool Sec. J. P. Laker, F.-J. Wortmann (eds.): Metrology and Identication of Speciality Animal Fibres, European Fine Fibre Network, occasional publication no. 4, Maculay Land Use Research Institute, Aberdeen, 1996. ISO 6938-1984: Textiles-Natural Fibres-Generic Names and Denition, International Organization of Standardization, 1984. P.-A. Koch, Text. Rundsch. 19 (1964) 189 203. W. F. Leggett: The Story of Wool, Chem. Publ. Co., Brooklyn N.Y. 1947. N. Hyde, National Geographic 173 (1988) 552 591. R. E. Chapman in J. Bereiter-Hahn, A. G. Matoltsy, K. S. Richards (eds.): Biology of the Integument 2, Springer Verlag, Berlin 1986, p. 298. D. E. Rivett, Wool Sci. Rev. 67 (1991) 1 25. C. A. Anderson, J. D. Leeder, Text. Res. J. 55 (1985) 416 421. H. E. Crabtree, P. Nicholls, E. V. Truter, Proc. Anal. Div. Chem. Soc. 16 (1979) 235. A. Schwan, J. Herrling, H. Zahn, Colloid Polym. Sci. 264 (1986) 171 175. J. Kalbe et al., Biol. Chem. Hoppe Seyler 369 (1988) 413 416. P. F. Hamlyn, G. Nelson, B. J. McCarthy, J. Text. Inst. 83 (1992) 97 103. D. J. Evans, J. D. Leeder, J. A. Rippon, D. E. Rivett in [34] vol. I, pp. 135 142. P. W. Wertz, D. T. Downing, Lipids 23 (1988) 878 881. A. K rner, H. H cker, D. E. Rivett, Fresenius o o Z. Anal. Chem. 344 (1992) 501 509. U. Altenhofen, H. Zahn, ACS Symp. Ser. 49 (1977) 162 175. M. F. Perutz, Faraday Discuss. Chem. Soc. 93 (1992) 1 11; Philos. Trans. R. Soc. London A 345 (1993) 105 112. J. B. Speakman, M. Hirst, Nature (London) 128 (1931) 1073 1074. R. S. Asquith et al., Biochim. Biophys. Acta 207 (1970) 342. B. Milligan, L. A. Holt, J. B. Caldwell in [22], part I, pp. 113 125. S. K. Burley, G. A. Petsko, Adv. Prot. Chem. 39 (1988) 125 189. H. Beyer, U. Schenk, J. Chromatogr. 39 (1969) 491 495. D. R. Goddard, L. Michaelis, J. Biol. Chem. 106 (1934) 605 614; 112 (1935) 361 371. W. G. Crewther, R. D. B. Fraser, F. G. Lennox, H. Lindley, Adv. Protein Chem. 20 (1965) 191 346. J. M. Gillespie: pp. 95 128 in [41]. L. M. Dowling, L. G. Sparrow, TIBS 16 (1991) 115 117.

37. 38.

54.

55. 56. 57. 58.

39. 40.

41.

42.

59. 60. 61. 62. 63. 64. 65. 66. 67. 68. 69.

43. 44. 45. 46.

47.

48.

70. 71. 72. 73. 74. 75. 76.

49.

50. 51.

52.

53.

77. 78.

Wool
79. R. C. Marshall, Text. Res. J. 51 (1981) 106 108. G. E. Rogers, E. S. Kuczek, P. J. Mackinnon, R. B. Presland, M. J. Fietz in [38] pp. 69 85. 80. B. R. Herbert, A. L. P. Chapman, D. A. Rankin, Electrophoresis 17 (1996) 239 243. 81. H. Thomas, A. Conrads, K. H. Phan, M. van de L cht, H. Zahn, Int. J. Biol. Macromol. 8 o (1986) 258 264. 82. M. van de L cht, Melliand Textilber. 68 o (1987) 780 786. 83. A. Conrads, H. Thomas, K.-H. Phan, H. Zahn, H. H cker, Naturwissenschaften 75 (1988) o 100 101. 84. W. G. Crewther, L. M. Dowling, P. M. Steinert, D. A. D. Parry, Int. J. Biol. Macromol. 5 (1983) 267 274. 85. J. A. Swift in [27] pp. 81 146. 86. M. G. Dobb et al., J. Text. Inst. 52 (1961) T 153 170. 87. K.-H. Phan, H. Thomas, E. Heine in [52], vol. II, pp. 19 30. 88. G. E. Rogers, Ann. N.Y. Acad. Sci. 83 (1959) 378 399. G. E. Rogers, J. Ultrastruct. Res. 2 (1959 b) 309 330. 89. R. Eichner, P. Rew, A. Engel, U. Aebi, Ann. N.Y. Acad. Sci. 455 (1985) 381 402. 90. W. T. Astbury, A. Street, Phil. Trans. R. London Soc. A 230 (1931) 75. W. T. Astbury, H. J. Woods, Phil. Trans. R. London Soc. A 232 (1933) 333 394. 91. L. Pauling, R. B. Corey, H. R. Branson, Proc. Nat. Acad. Sci. USA 37 (1951) 205 211. 92. L. Pauling, R. B. Corey, Proc. Nat. Acad. Sci. USA 39 (1953) 253. 93. F.-J. Wortmann, H. Deutz, J. Appl. Polym. Sci. 48 (1993) 137 150. 94. F.-J. Wortmann, H. Zahn in: Domke Festschrift, Lehrstuhl f r konstruktive u Gestaltung der TH Aachen, Aachen 1982, pp. 157 161. 95. M. Feughelman, Text. Res. J. 29 (1959) 223 228. 96. J. Cao, J. Mol. Struct. 553 (2000) 101 107. 97. L. Kreplak, J. Doucet, F. Briki, Biopolymers 58 (2001) 5269 533. 98. J. B. Speakman, C. A. Cooper, J. Text. Inst. 27 (1936) T 183 T 196. 99. E. El d, E. Silva, Z. Phys. Chem. Abt. A 137 o (1928) 141. 100. F.-J. Wortmann, B. J. Rigby, D. G. Phillips, Text. Res. J. 54 (1984) 6 8. 101. T. G. Fox, Bull. Am. Phys. Soc. 1 (1956) 123. 102. H. Zahn in [52], vol. I, pp. 1 16.

29

103. H. Baumann, L. M chel, Text. Praxis Int. 29 o (1974) 507 510. 104. H. Baumann, Schriftenreihe des Deutschen Wollforschungsinstitutes 85 (1981), 216 246. 105. H. Meichelbeck, H. Knittel, Fette, Seifen, Anstrichm. 73 (1971) 25 29. 106. D. Goddinger, K. Sch fer, H. H cker, Wool a o Technol. Sheep Breed 42 (1994) 83 89. 107. D. Goddinger, K. Sch fer, H. H cker in [52], a o vol. II, pp. 92 108. 108. G. J. Smith, J. Photochem. Photobiol. Bull. 22 (1994) 145 149. 109. H. Zahn, J. Knott, Nova Textil 13 (1989) 42 46. 110. F. G. France, I. L. Weatherall in [52], vol. II, pp. 100 108. 111. G. J. Smith, K. R. Markham, W. H. Melhuish, Photochem. Photobiol. 60 (1994) 196 198. 112. K. Sch fer, J. Soc. Dyers Colour. 107 (1991) a 206 211. 113. C. M. Pande, J. Soc. Cosmet. Chem. 45 (1994) 257 268. 114. I. C. Watt, J. Macromol. Sci.-Rev. Macromol. Chem. C18 (1980) 169 245. 115. F.-J. Wortmann, P. Augustin, C. Popescu, J. Appl. Polym. Sci. 79 (2001) 1054 1061. 116. F.-J. Wortmann, H. Zahn, Text. Res. J. 64 (1994) 737 743. 117. K. Sch fer, I. M llejans, H. H cker, Melliand a u o Textilber. 78 (1997) no. 7/8, 506 508, E 108 E 110. 118. K. Sch fer, I. M llejans, J. F hles, H. H cker, a u o o Melliand Textilber. 78 (1997) no. 10, 727 732, E 164 E 167. 119. H. H cker in [52], vol. III, 319 324. o 120. E. Heine, H. H cker, Rev. Progr. Color. Relat. o Top. 25 (1995) 57 63. 121. A. Bereck, Rev. Progr. Color. Relat. Top. 24 (1994) 17 25. 122. A. C. Welham, J. Soc. Dyers Colour. 102 (1986) 126 131. 123. W. Mosimann, K. Schaffner, Textilveredlung 14 (1979) 12 15. 124. P. A. Dufeld, R. R. D. Holt, Textilveredlung 24 (1989) 40 45. 125. H. Baumann, J. Soc. Dyers Colour. 90 (1974) 125 129. 126. N. Werkes, Melliand Textilber. 70 (1989) 52 63. 127. D. M. Lewis, Melliand Textilber. 67 (1986) 717 723; Proc. 8th Int. Wool Text. Res. Conf., vol. IV, Christchurch, New Zealand 1990, pp. 1 49. 128. K. J. Dodd, C. M. Carr, M. Baird, J. Soc. Dyers Colour. 110 (1994) 190 194.

30

Wool
154. H. Thomas, M. Hedler, T. Merten, V. Monser, H. H cker, Proc. Int. Wool Text. Res. Conf. o 10th, Aachen, (2000) PL-3. 155. B. Jansen, F. K mmeler, H. Thomas, H.-B. u M ller, Proc. Int. Wool Text. Res. Conf. 10th, u Aachen, (2000) PL-6. 156. H. H cker, Proc. Int. Wool Text. Res. Conf. o 10th, Aachen, (2000) PL-7. 157. Anonym., International Dyer 185 (2000) no. 4, 21. 158. Anonym., Wool Record 157 (1998) (Issue 3644) 21. 159. B. O. Bateup, Proc. Int. Wool Text. Res. Conf. 10th, Aachen, (2000) KNL-5. 160. K. Hannemann, H. Flensberg, DWI Reports 117 (1996) 287 293. 161. K. Hannemann, Wool Record 159 (2000) no. 6, 48 51. 162. M. Giemann, K. Sch fer, H. H cker, a o Textilveredlung 34 (1999) no. 11/12, 12 19. 163. A. Giehl, K. Sch fer, H. H cker, J.-P. a o Mullender, J. Knott, Proc. Int. Wool Text. Res. Conf. 10th, Aachen, (2000) DY-2. 164. G. Carnaby, Proc. Niches in the World of Textiles, 77th World Conf. of the Textile Inst., Tampere, SF, May 22 24, 1996, vol. 1 (1997). 165. [32], Chap. 7. 166. G. Mack, Wool Record (1995) Dec., 5. 167. R. Hoffmann, G. Timmer, K. Becker in [52], vol. I, pp. 66 77. 168. P. A. Dufeld, R. R. D. Holt, J. R. Smith, Melliand Textilber. 72 (1991) 938 942. 169. H. Thomas, K.-H. Phan, H. H cker Report 17, o IWTO Harrogate Meeting, 1995. 170. J. D. Verity, Wool Record (1995) Mar., 25. 171. G. Horstmann, J. Soc. Dyers Colour. 111 (1995) 182 184. 172. H. G. Hebecker: Textil und Umwelt, Textilverband Schweiz, Z rich 1995, u pp. 24 25. 173. H. J. Henning, G. Mazingue, Text. Prax. Int. 35 (1980) 51 54, 179 185, 298 304. 174. G. J. J. B. de Groot, J. Text. Inst. 86 (1995) 33 44. 175. S. Kawabata, M. Niwa, J. Text. Inst. 80 (1989) 19 50. 176. S. Kawabata, K. Ito, M. Niwa, J. Text. Inst. 83 (1992) 361 374. 177. C. F. Allan, A. G. de Boos, N. G. Ly, Melliand Textilber. 81 (1990) 614 618. 178. F.-J. Wortmann, G. Wortmann, H. Zahn, Text. Res. J. 65 (1995) 669 675. 179. H. Zahn, Melliand Textilber. 50 (1969) 807 813.

129. B. Gullbrandson, Text. Res. J. 28 (1958) 965 968. 130. A. G. de Boos: The Conicting Requirements of Wool Setting for the Finisher and Tailor, Report No. G 61, CSIRO Div. Text. Ind., Belmont/Victoria 1987. 131. A. Robson, M. J. Williams, J. M. Woodhouse, J. Text. Inst. 60 (1969) 140 151. 132. H. Meichelbeck: Die Rolle der SH-Gruppen bei der Quervernetzung des Keratins durch Thio therbildung, Dissertation TH Aachen a 1963. 133. H. Zahn in [52], vol. I, pp. 1 16. 134. H. K. Rouette, DWI Reports 117 (1996) 93 112. 135. J. B. Speakman, J. Soc. Dyers Colour. 52 (1936) 335 346. 136. P. G. Cookson, K. W. Fincher, P. R. Brady, J. Soc. Dyers Colour. 107 (1991) 135 138. 137. S. M. Smith et al. in [52], vol. III, pp. 113 121. 138. I. Hammers-Page in W. Loy (ed): Taschenbuch f r die Textilindustrie 1994, Fachverlag u Schiele und Sch n, Berlin, pp. 426 436. o 139. T. Shaw, M. A. White in M. Lewin, S. B. Sello (eds.): Chemical Processing of Fibers and Fabrics, Handbook of Fibre Science and Technology, vol. 2, part 2, Marcel Dekker, New York 1984. 140. R. J. Mayeld, Text. Prog. 11 (1982) 1 10. 141. K. Wolf, I. Hammers, U. Altenhofen, Melliand Textilber. 66 (1985) 596 600. 142. L. Benisek, J. Text. Inst. 65 (1974) 102 108. 143. W. Mosimann et al. in [42], vol. IV, pp. 239 249. 144. G. Reinert, Textilveredlung 26 (1991) 86 92. 145. G. Reinert, E. Schmidt, R. Hilker, Melliand Textilber. 75 (1994) 606 614. 146. G. V. Barker, F. Davin, Proc. Int. Wool Text. Res. Conf. 8th, Christchurch, vol. 3 (1990) 91 106. 147. P. Hopkins, DWI Reports 119 (1997) 128 137. 148. P. Hopkins, Textile Magazine 26 (1997) no. 1, 7 8. 149. Wool Fact File 1998/99, published by Woolmark (Europe) Ltd., Ilkley, UK. 150. E. Heine, A. Ruers, H. H cker, DWI Reports o 122 (1999) 466 470. 151. E. Heine, A. Ruers, H. H cker, DWI Reports o 123 (2000) 475 479. 152. B. M. M ller, Rev. Prog. Colouration 22 u (1992) 14 21. 153. H. Thomas, F. K mmeler, H.-B. M ller, E. u u Prinz, U. Rott, J. Salge, M. Wolf, T.-J. Zahn, DWI Reports 122 (1999) 365 370.

Wool
180. International Wool Secretariat (ed.): IWTO Specications, International Wool Secretariat, Raw Wool Services Department, Valley Drive, Ilkley, Yorkshire, LS29 8PB, England. 181. R. C. Marshall, H. Zahn, G. Blankenburg, Text. Res. J. 54 (1984) 126 128. 182. H. Zahn, I. Souren, U. Altenhofen, Melliand Textilber. 65 (1984) 467 471. 183. C. Wilrich, G. Wortmann, F.-J. Wortmann, H. H cker in [52], vol. II, pp. 289 295. o 184. P. Seiler, C. Wilrich, G. Wortmann, F.-J. Wortmann, DWI Reports 117 (1996) 637 648. 185. W. Gocht, DWI Reports 117 (1996) 399 400. 186. J. D. Leeder in [32], pp. 24 31. 187. L. Benisek, P. R. Harnett, M. J. Palin, Melliand Textilber. 68 (1987) 878 888. 188. M. Ried in [39], pp. 197 198. 189. A. Pierlot: Wool in J. C. Salomone (ed.): Polymeric Materials Encyclopedia, CRC Press, Boca Raton, Fl., in press. 190. G. Fisher, Tech. Text. Int. 1993, Feb., 8 12. 191. H. Zahn, G. Blankenburg, Deutsche Schafzucht 85 (1993) 31 33. 192. G. Blankenburg, I. Hammers-Page, DWI Reports 114 (1995) 61 72. 193. E. Claen, G. Wortmann, R. Augustin, N. Effertz, F.-J. Wortmann, Proc. Int. Wool Text. Res. Conf. 10th, Aachen, (2000) OT-P5. 194. P. Offermann, DWI Reports 121 (1998) 247 257. 195. H. L. Fraenkel-Conrat, H. S. Olcott, A. Mohammed, Arch. Biochem. 24 (1949) 278. 196. S. M. Causer, R. C. McMillan, Australian Textiles 14 (1994) 48. 197. G. Wortmann, R. Sweredjuk, G. Zwiener, F. Doppelmayer, F.-J. Wortmann, DWI Reports 124 (2001) 378 386.

31

198. K. Sch fer, H. H cker, Melliand Textilber. 77 a o (1996) 402 406. 199. C. Jacques, J. P. Bruggeman, G. Vilarem, Proc. Int. Wool Text. Res. Conf. 10th, Aachen, (2000) OT-P7. 200. Bekleidung und W sche 18 (1989) 14 18. a 201. A. Kloeden, Wool Record (1995) Nov., 6, 51. 202. H. Menkens, Wool Record (1995) Dec., 45. 203. G. Reinert, R. Hilker, E. Schmidt, F. Fuso, DWI Reports 117 (1996) 259 273; Text. Res. J. 66 (1996) 61 70. 204. G. R. S. Naylor, C. J. Veitch, R. J. Mayeld, Text. Res. J. 62 (1992) 487 493. 205. G. Wortmann, A. K rner, DWI Reports 115 o (1995) 1 135. 206. G. Chalmers, Wool Record (1993), July, 41. 207. F. Klaschka, Melliand Textilber. 75 (1994) 193 202. 208. T. L. Diepgen, B. Salzer, A. Tepe, O. P. Hornstein, Melliand Textilber. 76 (1995) 1116 1121. 209. E. A. Clarke, D. Steinle, Rev. Prog. Color. Relat. Top. 25 (1995) 1 5. 210. J. Gr tze, F. Hoffmann, U. Sewekow, A. u Westerkamp, Melliand Textilber. 76 (1995) 875 883. 211. R. A. Moll, Melliand Textilber. 76 (1995) 993 995. 212. W. Beckmann, ITB Veredlung 1 (1990) 50 59. 213. R. G. Love, M. Muirhead, H. P. R. Collins, C. A. Soutar, Br. J. Ind. Med. 48 (1991) 221 228. 214. I. M. Russell in [42], vol. III, pp. 165 174. 215. V. Schmitz, K. Sch fer, H. H cker, Melliand a o Textilber. 80 (1999) no. 7/8, 596 600.

Xanthan Gum

Polysaccharides

You might also like