Vous êtes sur la page 1sur 653

Yarn Technology and

Quality
By Zhang Shangyong ,Wang
Xungai

Chapter

1
Technology
Chapter 2
Chapter 3
Spinning
Chapter 4
Chapter 5

Fundamentals of Yarn
Yarn Evenness
Fibre Preparations for
Yarn Spinning Systems
Yarn Technology Extension

Chapter 1 Fundamentals of
Yarn Technology
1.1

Yarn count
1.2 Yarn twist
1.3 The designation of yarn structures

1.1 Yarn count


Introduction

Yarns come in different sizes. They can be


quite thick, or they can be very thin. Since
by their very nature textile yarns are soft
and squashy, the thickness of a yarn can
not be easily measured by yarn diameter.
But textile yarns are often sold on a weight
basis, so it is natural to express the size of
a yarn in terms of its weight or mass.

The two basic ways of doing this are by


indicating either how much a given length
of yarn weighs (the direct system)
or what the length of yarn will be in
a given weight (the indirect system)
.

These two broad yarn count systems are


expressed below

Weight of yarn
Direct yarn count
Given length
Length of yarn
Indirect yarn count
Given weight

Because a textile yarn is usually a very


slender assembly of tiny fibres, it is
conceivable that the weight of a yarn in a
given length will be very small while the
length of a yarn in a given weight will be
quite large.

Consequently, the yarn count figures would


get either incredibly small (direct system)
or large (indirect system) unless special
units are used. Over the years, many
different units have been used in different
sectors of the textile industry. This topic
describes these different units and the
conversions involved.

Objectives
At the end of this topic you should be able to:

understand the definitions for different


yarn count systems
know the conversion between yarn
counts
appreciate the effect of moisture on yarn
count results

Direct Count Systems


The direct systems are based on the weight
or mass per unit length of yarn. Some typical
direct systems are given below, together
with their definitions. Please note that while
the weight unit is gram, different lengths are
used in the definitions.

1.Tex (g/1000m)
This is the mass in gram of one kilometre, or
1,000 metres, of the product.
If one thousand meters of yarn weigh 20 grams or
one hundred meters of the yarn weigh 2 grams,
the yarn would be 20 tex. On the other hand, if
100 metres of yarn weigh 5 grams, then the count
of the yarn will be 50 tex.

2.Dtex (g/10,000m)
This is called deci-tex. It is the mass in gram of ten
kilometre, or 10,000 metres, of the product. It is a
smaller unit than tex (1 tex = 10 dtex), and is
usually used for fibres and filament yarns.
A 167 dtex polyester filament would weigh
167 grams for every 10,000 meters of the
filament.

3.Ktex (g/m)
This is called kilo-tex. It is the mass in gram of
one metre of the product. It is a much larger
unit than tex (1 ktex = 1,000 tex), and is
usually used for heavy products such as slivers.
If a sliver weighs 5 grams per metre, then the
count of this sliver would be 5 ktex.
The tex system (tex, ktex, dtex) is the preferred
standard system. By definition,
1 ktex = 1,000 tex = 10,000 dtex

4.Denier (g/9,000m)
Denier is also used extensively in the industry,
particularly for manufactured fibres and silk. It
is the mass in gram of nine kilometres, or 9,000
metres, of the product.
By definition,
1 dtex = 0.9 denier
If a 300 denier yarn is made up of 1.5 denier
individual filaments, there will be a total
number of filaments in the yarn.

Indirect Count Systems


Indirect count systems are not as straightforward
as the direct ones. In the early history of yarn
manufacture, different spinners, often
geographically and culturally isolated from one
another, devised their own ways of measuring
yarn thickness. Consequently, there are numerous
indirect count systems that have been, and
continue to be, used in the industry.

Some examples are given below, together


with the mass and length conversions
1.Commonly used
- Metric (Nm)
m/g
- English Cotton (Nec)
pound
- Worsted (Nw)

No. of 840 yard


hanks per

No. of 560 yard


hanks per pound

The metric count (Nm) is relatively


straightforward. It is the length in metre of
one gram of the product. For example, if one
gram of yarn measures 40 metres, then the
metric count of this yarn would be 40 Nm.
Similarly, if one pound of cotton yarn
measures 1,680 yards, or two hanks of 840
yards, the English cotton count of this yarn
will be 2 Nec. Please note that a hank of yarn
is an unsupported coil consisting of wraps of
yarn of a certain length.
The conversions between different units will
be discussed later.

2.Less commonly used


- linen , hemp
,ramie
- asbestos(

No. of 300 yard


hanks per pound
" " 50 "

"

"

"

- glass )

" " 100 "

"

"

"

- spun silk

" " 840 "

"

"

"

- raw silk
ounce

" " 1000 "

"

"

3.Occasionally used in the woollen industry


- Yorkshire skein
-West of England
- American cut
- American run
- Dewsbury
ounce
- Galashiels
24ounces

No. of 560 yard hanks


per pound
" " 320 "
" " "
" " 300 "
" " "
" " 100 "
" " ounce
" " 1 "
" "
" " 300 "

"

"

You may wonder how the strange length


units such as 840 yard hank and 560
yard hank came about. The first massproduction spinner the spinning-jenny
was able to spin yarns simultaneously
onto several bobbins and filled
the bobbins up at the same time. The
bobbins were changed after 840 yards of
cotton yarns were wound onto them. To
estimate the thickness of the yarns, the
spinner simply counted how many full
bobbins were needed to balance a weight
of one pound.

Conversion between Different Yarn Counts


It is often necessary to make conversions
between different yarn count systems. For
this purpose, the following mass (weight) and
length conversions are needed:
1 yard (yd)= 0.9144 m
1 pound (lb)= 0.4536 kg
1 ounce (oz)= 1/16 lb
1 dram 9dr)= 1/16
oz
1 grain (gr)= 1/7000 lb

Worked Examples
Question1:
What is the conversion factor between
worsted count (Nw) and tex ?
Solution:
According to definition, one worsted count
(Nw) = one 560 yard hank per pound, or
1 Nw

1 560 yard
pound

Since 1 yard (yd) = 0.9144 m and 1 pound


(lb) = 0.4536 kg, the above equation
1 560 0.9144 m 512.064 m 1.12892 m
becomes,
1 Nw

453.6 g

453.6 g

Therefore, for a yarn of Nw worsted count,


each gram of this yarn would measure
1.12892 times Nw meters. Since tex is the
mass in gram of a 1,000 meters of yarn, we
need the number of grams in 1000 m of the
yarn.
1000
885.8
No of grams per 1000 m tex

1.12892 Nw

Nw

The above equation885


can.8also be written as:
Nw

tex

So the conversion factor is 885.5.

Question2:
If a yarn is 20 tex, what is the worsted count of
this yarn?
Answer:
Using the conversion factor given above, the
885.8
worsted yarn count
44is
.3 Nw
20

Conversion between other count systems can


be worked out in a similar way. Table 1.1 lists
commonly used conversion factors. You may try
to work them out yourself.

Moisture and Yarn Count


Regardless of the yarn count system used, it is
necessary to measure the weight and length of a
yarn in order to determine its count. But most
fibres, particularly natural fibres, absorb moisture
from atmosphere. The weight of the yarn will be
different at different moisture level. The water
content in textiles can be expressed as either
moisture content or as regain
.

Their definitions are:


Mass of absorbed water in specimen (W )
Re gain ( R )
100
Mass of dry specimen ( D )
Moisture content ( M )

Mass of absorbed water in specimem (W )


100
Mass of original undried specimen (W D )

From these definitions, the conversion


between regain (R) and moisture content (M)
can be worked out according to the equation
below:
R
M
1 R

In commercial transactions, the mass


to invoice is worked out on the basis of
an agreed conventional regain level,
not on the actual regain of the yarns
(or other textiles) being traded. This is
very important. Because, in the
absence of an agreed conventional
regain level, smart sellers may take
advantage of the moisture absorption
property of their textiles and rip the
buyers off with large quantity of water
in their products.

The conventional regain levels, to be


used for calculation of the legal
commercial mass, have been
established by national or international
standards. These commercial regain
values are purely arbitrary values
arrived at for commercial purposes for
interested parties, and they often vary
from fibre to fibre and from country to
country.

In Australia, the conventional regain


rates for some fibres are given in Table
1.2Table 1.2: Conventional regain rate for selected fibres
Fibre
Wool and hair fibres
Combed (worsted)

Conventional regain (%)


18.25

Carded (woollens)
Cotton
Normal cotton

17

Mercerised cotton
Silk

10.5
11

8.5

Polyester
Staple fibre

1.5

Continuous filament

1.5

Conditioning the whole lot of yarns or other


textile materials to the conventional regain rates
given above is not practical, because of the time
required etc. In calculating the commercial mass
to invoice for a lot, the following procedures are
often followed:

(1) Extract a sample of mass (gw) from the lot


(whose total gross weight is GW)
(2) Determine the dry weight (dw) of the sample
by oven drying to completely evaporate the
moisture contained in it.
(3) Calculate the commercial mass to invoice
(cw), based on a conventional regain R%, by
means of the formula:
cw GW

dw 100 R %

gw
100

Example:
Suppose a lot of worsted yarn is to be shipped
to a buyer, and the gross weight of lot is 1000
kg. We now need to work out the commercial
mass to invoice for the lot of yarn.

Answer:
We first extract a small sample (say 500 grams)
from the lot. After oven drying of this small
sample, the dried mass becomes, say, 450
grams. For worsted yarn, the conventional
regain rate is 18.25% according to Table 1.2.
18.25
450 100
Therefore, the commercial
1000mass

to invoice
1064.25 ( kg )
500
100
should be:

Review questions
1.Suppose you have two cotton yarns. The
count of yarn A is 20 tex, and that of yarn B
is 20 Nec. Which yarn is a thicker one? You
need to justify your answer via proper count
conversions.
2.If a worsted yarn has a count of 40 Nw
(worsted count) at a regain level of 20%,
what would be the count of this yarn, in tex,
when it is oven-dried?

1.2 Yarn twist( )


Introduction
In the manufacture of staple fibre yarns(
), twist is inserted into the fine strand of
fibres to hold the fibres together and impart
the desired properties to the twisted yarns.
Without twist, the fine strand of fibres would
be very weak and of little practical use. A
change in the level of twist also changes many
yarn properties, such as strength and softness.
This topic describes the nature of yarn twist,
the effect of twist on yarn properties, as well
as twist measurement.

Objectives
At the end of this topic you should be able to:
understand the effect of twist on certain yarn
and fabric properti
appreciate the importance of surface twist
angle and of selecting the right twist factor
for different yarns
know how to calculate the twist contraction
know the basic rules that apply to twist
measurements

Nature of twist
Types

of twist
(1) Real twist
To insert a real twist into a length of yarn,
one end of the yarn should be rotated
relative to the other end, as indicated in
figure2.1(a)

(2) False twist(


When inserting false twist into a length of
yarn, both ends of the yarn are clamped,
usually by rollers, and twist is inserted with a
false twister between the clamping points, as
indicated in figure

If the yarn is not traversing along its axis,


the twist will be in opposite directions
above and below the false twister. If the
false twister is removed, the opposite
twists will cancel out one another, leaving
no real twist in the length of yarn. If the
yarn is traversing along its axis, then the
section of the yarn moving away from the
false twister would have no net twist, as
indicated in figure 2.1(b)

False twisting is a very important


phenomenon, which has considerable
practical implications in yarn
technology. False twisting is featured
in many key processes that we will
discuss later, including woollen ring
spinning, open-end rotor and friction
spinning, air jet spinning, and filament
yarn texturing.

Twist

direction
A twist can be either in Z direction or S direction
as indicated in figure 2.2, depending on the
orientation of the surface fibre in relation to
yarn axis.

S-Twist

Z-Twist

It is worth noting that twist direction affects


fabric properties. For example , This figure 2.3
shows two identical twill-weave fabrics with the
warp yarn of different twist direction.
S twist

Twill direction

Z twist

(A)
(B)
(Short arrows i ndicate direction of light
reflected from the warp and weft yarns)

Fabric A will be more lustrous than fabric B,


because light reflected by fibres in the warp
and weft is in the same direction. Fabric A will
be softer while fabric B firmer, because in
Fabric B, the surface fibres on the warp and
weft in the region of contact are aligned in the
same direction and they may get stuck inside
each other and reduce the mobility of the
intersection. Whereas for fabric A, the surface
fibres on the warp and weft in the region of
contact are crossed over, and they can move
about easily. The freedom of movement at the
yarn intersections is the key for fabric softness.

Self-locking

effect

Because of twist in a yarn, the fibres on yarn


surface take a roughly helical configuration
around the yarn. When the yarn is under
tension, these surface fibres are also under
tension. However, because of the helical
configuration, part of the tenon is diverted
radially, which creates a radial pressure. This is
illustrated in figure 2.4.

Figure 2.4 Helical fibre under tension (Lord 1981, p. 75)

The radial pressure tends to pack the fibres


together, increasing the normal force between
them, and so increasing their frictional resistance
to slipping past each other. The more tension is
applied to the yarn, the more it locks together,
hence 'self-locking'. An analogy is, when you
wind a string around your arm, as you pull the
string along the arm and away from each other,
the string bites deeper into the flesh.
Without twist, there wont be any self-locking
effect to prevent fibre slippage. Consequently the
yarn would have no strength.

Effect of twist level on yarn strength


The level of twist is usually expressed in number
of turns per metre (tpm). Number of turns per
inch or twist per inch (tpi) is also used in the
industry.

Fig. 2.5: Effect of twist level on the strength of staple (spun) yarn

More twist gives greater radial component to


any applied tension, so increases resistance of
fibres to slip and the strength of yarn increases
as a consequence. This is depicted by the
coherence curve in figure 2.5.
On the other hand, if a bundle of parallel
filaments is twisted, the twist will put the
individual filaments under torsional stress. This
stress weakens the filaments and the strength
of the filament would decrease as the level of
twist increases. This is depicted by the
obliquity curve in figure 2.5.

For staple fibre yarns, these two curves combine to


give the actual 'twist-strength curve' for a
staple fibre yarn as shown by the heavy line in
figure 2.5.
Figure 2.5 indicates that for staple fibre yarn,
increasing the twist level will increase yarn
It should be noted that for continuous filament
yarn, the obliquity curve applies. In other words,
twisting a continuous filament yarn only reduces
the yarn strength, regardless of the twist level
used. If a continuous multi-filament yarn is twisted,
the reason for the twist is to keep the individual
filaments together, not for strength.

Twist angle
This is the angle of fibres to yarn axis, and this angle
varies throughout yarn, from zero at centre to
maximum at yarn surface. The fibres on yarn
surface are the most important, as they bind the
others into the yarn (refer to self-locking effect
discussed earlier).
While it is not common practice to measure the yarn
twist angle, the surface twist angle made by the
surface fibres in relation to yarn axis is a very
important parameter. It determines the essential
yarn characteristics such as yarn softness, yarn bulk
etc, which in turn govern many essential fabric
properties. The following example illustrates the
point.

In figure 2.6, yarn 1 and yarn 2 have the


same twist level one turn each. But the
surface fibre on the thicker yarn is
obviously stretched more to accommodate
this twist. This would mean the thicker
yarn is more closely packed. As a
consequence, yarn 2 will not be as soft as
yarn 1. In other words, even though the
twist level is the same in these two yarns,
the yarn characteristics are quite different.

Therefore, we can not simply use twist level to


represent yarn character. However, the surface
twist angles of yarn 1 (1) and yarn 2 (2) are
different. They can better reflect the yarn
characteristics, regardless of the difference in
Yarn 1
Yarn 2
yarn thickness.
L

d1

d2

d1

d2

Figure 2.6 Two yarns of the same twist level, but different surface twist angles

Twist factor (Twist multiplier)


This is a very important factor that relates to the
angle of twist helix the surface fibres have in a
yarn. As we will see later, this factor is very
important for a spinner because of the following
reasons:
Like surface twist angle, it governs the yarn
characteristics
It is used to work out the twist to use in
spinning, in order to maintain the same surface
twist angle and similar yarn characteristics
when the yarn count is changed. The twist
worked out from twist factor is also needed for
setting up the spinning machine.

Relate twist factor to twist angle


Because it is much easier to measure twist level in
turns per metre than twist angle, we should relate
twist level to twist angle.From figure 2.6, we get,
tan =

d
L

Also from figure 2.6, the height (pitch) of one turn


of twist is L. Since the twist level is normally
specified as the number of turns per metre, the
twist level in one metre of the yarn would be:

so

1
twist =
L

tan = d x twist

(1)

We also know from experience that yarn


diameter is also very hard to measure,
because textile yarns by their very nature
are soft and squashy. On the other hand,
yarn count is normally used as we have
discussed in the first topic of this module.
But we can relate yarn diameter to yarn

count using the expression below:


cubic density = linear density
(Tex) /cross- sectional area (A)

Assuming a circular cross section for


the yarn, we get,
-3
-3
Tex x 10 (g/m) Tex x 10
(g/ m ) =
=
2
A( m )
d 2 /4
3

4 Tex x 10 -3

(2)
Solve ford:d =

Combining equations (1) and (2):

twist =

or

twist =

K
tex

tan

2 tex x 10-3 /

whereK = 0.5 tan

103 /

(t.p.m tex )

K is called the twist factor, and is


proportional to if remains constant.

Thus, K is a factor relating twist level to yarn


count. The derivation shows that if two yarns
have the same twist factor, they will have the
same surface twist angle, regardless of count.
Since surface twist angle is the main factor
determining yarn character, then twist factor
can be used to define the character of a yarn.

It is worth noting though there are minor


errors associated with the use of twist factor
for the following reasons:
The cubic density may be different for
different yarns. It is assumed in the above
calculation that this will not change for
yarns of the same surface twist angle.
Different fibres with different frictional and
other properties will create different yarn
character.

Nevertheless, the relationship we have


just derived between twist, twist factor
and yarn count is one of the most
important in the study of yarn technology.
This relationship is expressed in different
ways for different yarn count systems.
For the tex system
Twist (turns per metre ) =

Twist Factor ( K t )
tex

For the metric count (Nm) system


Twist (turns per metre ) = Twist Alpha ( m )

Nm

For English cotton (Nec) count system


Twist (turns per inch) = Twist Factor ( K e ) Nec

For worsted count (Nw) system


Twist (turns per inch) = Twist Factor ( K w ) Nw

Please note the unit for twist is also different


in the above expressions of twist factor. In
addition, twist factor is also known as twist
multiplier, twist alpha, or twist coefficient.

Choice of twist factors


Yarns intended for different end uses have
different characteristics. Since twist factor (like
surface twist angle) determines yarn
characteristics, the choice of twist factor is
often governed by the intended use of the
yarns. If maximum yarn strength is of the
utmost importance, one would obviously
choose the optimum yarn twist (see figure 2.5)
and the optimum twist factor for strength.

However, the end-use of yarn may be


such that other properties may be more
important. For example, a yarn to be used
for weft or for hosiery may be required to
be soft and bulky and therefore a low
twist factor is used. A yarn to be used for
the production of voile or crepe fabric will
necessitate the use of a high twist factor.

If one considers staple yarns for the


production of plied or cabled sewing
threads then soft twisted single yarns
are used and this results in the highest
strength in the final thread.
Another important feature to consider is
that the productivity for spinning yarns
of lower twist factor is higher. For these
reasons, the majority of yarns are spun
with a twist factor lower than the
optimum twist factor for maximum
strength.

Table

2.1 shows the twist factor most


commonly used for the various types of
yarns. Table 2.1: Twist factors most commonly used

Please note these are reference values


only, and the recommended values
vary from source to source.
Once a twist factor is chosen, the level
of twist required for the yarn can be
calculated for a given yarn count. This
twist level is then used to set up the
spinning machine for yarn production.

The distribution of twist in staple spun yarns


If someone twists your head, it is your
neck that suffers most. That is because
the neck is a thin place and offers little
resistance to being twisted. By analogy, if
a yarn of varying thickness is twisted, it is
usually the thin spot in the yarn that gets
twisted the most.

Invariably, yarns spun from staple


fibres (eg. wool, cotton) are not
perfectly uniform, and there are thick
and thin spots along the yarn length.
This variation in yarn thickness will
lead to variation in the twist level
along the yarn length, because twist
tends to accumulate in the thin place.

The fact that twist tends to accumulate in


the thin spot along the yarn has several
important implications:
(1) It exacerbates the variation in yarn
linear density. While variation in yarn linear
density is the fundamental cause of twist
variation, concentration of twist in the thin
places will make those places even thinner,
exacerbating the problem of yarn
unevenness.
(2)It improves the evenness of a fibre
assembly during drafting against twist

In the drafting stage of woollen ring


spinning, the woollen slubbing is drafted
while twist is inserted into the slubbing
(drafting against twist) to control fibres
during drafting. Because twist tends to
accumulate in the thin spots, the fibres in
thin regions in the slubbing are more difficult
to draft than those in the thick places, which
have less twist. As a result, the thick places
are drafted more than the thin places, thus
improving the evenness of the drafted
material. This is depicted in figure 2.7.

Thin place with


more twist

(a) Before drafting

(b) After drafting against twist

(3)It has implication for twist measurements


Because the twist level varies along the yarn
length, the twist measured at a short length of
yarn may not reflect the true average twist of
the yarn. Standard test procedures should be
followed to measure the yarn twist accurately.
The relationship between twist and yarn count
may be expressed by the following formula:
Twist

1
Tex

where p is usually greater than 1 but less than 2


for most yarns.

Twist contraction
When a bundle of parallel fibres is
twisted, the distance between the two
ends of a fibre will decrease, particularly
for fibres near the surface of the twisted
bundle. As a result, the overall length of
the twisted bundle is shorter than its
length before twist insertion. The
reduction in length due to twist insertion
is known as twist contraction.

The following formula is used to


calculate the amount of twist
contraction:
% contraction

Lo - L f
=
x 100 %
Lo

where Lo = original length before


twisting
Lf = final length after twisting

It should be noted that because of twist


contraction and the associated change in
length, the count of a yarn will change
slightly when twist in the yarn is
changed. Twist contraction increases yarn
count (tex), because the weight of the
yarn is distributed over a shorter length.
N
The followingNformula
can be used
=
f

where

1 - C

No = count (tex) before twisting


Nf = count (tex) after twisting
C = %contraction

Measurement of twist

Twist measurement is a routine test for


yarns. Because of the variation in twist
along yarn length as discussed earlier, care
should be taken in measuring the twist of
staple spun yarns. Some basic principles are
discussed here.

Sampling rules
The following rules should be observed when
measuring yarn twist:
a. Tests should not be limited to a short length
of the yarn package.
b. Beware of "operator bias" - tendency to
select either thicker or thinner regions.
c. Discard first few metres from package.
Being a free end, it could have lost twist .
d. Remove yarn from side of package, not over
end. Removing yarn over end will change
the twist level in the yarn.
e. Tension in Yarn during test
e.g. For single worsted yarns: 5 + 1
mN/tex

Principles

of measuring methods

The two common methods used in twist


measurement are straightened fibre method
and untwist/retwist method.
(1) Straightened fibre method(
This method involves counting of the
number of turns required to untwist the
yarns until the surface fibres appear to be
straight and parallel to yarn axis. This
method is mainly used for ply and
continuous filament yarns.

(2) Untwist / Retwist Method(

This is the common method used for


staple fibre yarns. It is based on twist
contraction (hence also known as twist
contraction method).

Review questions
1)"For a staple fibre yarn, the higher
the twist, the stronger the yarn". Is
this statement true? Why?
2)A yarn of 40 Nm (metric count) has
a twist factor of 3,000 . What is the
twist level, in turns per metre, of this
yarn?

3)If a yarn has an alpha metric (m) of


100, what is the twist factor in the tex
system (Kt) and the English cotton
count system (Ke)? You need to show
your working.
4)Assuming after twisting, the count of
a multi-filament yarn is increased from
150 dtex to 180 dtex, what is the
amount of twist contraction
experienced by this multi-filament
yarn?

1.3 The designation of yarn


structures

Introduction

This topic describes a method of indicating the


composition of yarns. The yarns can be spun
yarns or filament yarns. They may be single,
folded or cabled yarns. This topic is adapted
from Handbook of textiles standards for
students, published in 1988 by the Standards
Association of Australia.

Objectives
At the end of this topic you should be able
to:
describe a yarn according to its
designation
designate a yarn based on a detailed
description of the yarn

Systems and rules


Two systems can be used for yarn designation.
Single-to-fold notation (preferred)
This is the preferred system, where single
component of the yarn is described first,
followed by a description of how the
components are combined together to make up
the resultant yarn.
Fold-to-single notation
This notion is opposite to the single-to-fold
notation. The whole structure is described first,
followed by a description of its components.

General

rules
The following general rules should be
noted:
Use tex for staple spun yarns and dtex for
filament yarns
"fn" indicates n filaments in a single mono
(n=1) or multifilament yarn.
"t0" indicates components combined without
twist
"Rxyz tex" specifies the "resultant" count of
the yarn (xyz tex) in its final form.
specification after a semi-colon is optional
The following sections list examples of yarn
designations.

Single yarns( )

A single yarn, or singles yarn, may be a spun


yarn (or staple yarn), a mono-filament yarn, or
a multi-filament yarn. The ways of designating
these different single yarns are given below.
Spun yarns(
The details used in the designation of spun
yarns include:
Linear density (tex)
Direction of twist (S or Z)
Amount of twist (turns per metre)
For example, the designation
40 tex Z 660
describes a spun yarn that has a count of 40
tex, with a twist level of 600 turns per metre,
and the twist is applied in Z direction.

Mono-filament

yarns
The details used in the designation of
mono-filament yarns include:
Linear density (dtex)
Symbol f
Symbol t0 if not twisted; otherwise
twist direction and amount
For example, the designation 17 dtex
f1 t0 describes a mono-filament (f1)
yarn with a count of 17 dtex, with any
twist (t0) in the yarn.

Multifilament

yarns
The details used in the designation of multifilament yarns include:
Linear density
Symbol f
Number of filaments
Symbol t0 if not twisted; otherwise twist
direction and level
Resultant linear density
For example, the designation
140 dtex f40
t0 means a multi-filament yarn with a count of
140 dtex, consisting of 40 individual filaments
which are not twisted. Please note that the
linear density of each individual filament will
be .

Multiple

wound yarns
These are the yarns that have several
components wound up together, without inserting
any twist. This is also known as assembly
wound yarns
1.Multiple wound yarns with similar
components
The details used in the designation of such yarns
include:
Notation according to single yarn used
Multiplication sign, .
Number of single yarns laid together
Symbol t0
Example:
40 tex S155 2 t0

2.Multiple wound yarns with dissimilar


components
The details used in the designation of such
yarns include:
Notation according to single yarn used,
connected by the addition sign + and put
in brackets
Symbol t0
Example:(25 tex S420 + 60 tex Z80) t0 (Can you
describe this yarn?)

Folded or plied yarns


These are the yarns that have several
components twisted up together.
1.Folded yarns having similar components
The details used in the designation of such
yarns include:
Notation according to single yarn used
Multiplication sign, .
Number of single yarns twisted together
Direction of folding twist
Amount of folding twist
Resultant linear density

Example: 34 tex S600 2 Z400; R69.3 tex


(a singles yarn of 34 tex with a twist of 600
turns per metre in S direction is twisted
together with another yarn of the same
descriptions. The folding twist is 400 turns
per metre in Z direction, and the resultant
yarn count is 69.3 tex (slightly higher than
34 x 2 due to twist contraction.)

2.Folded yarn having dissimilar


components
The details used in the designation of such
yarns include:
Notation according to single yarn used,
connected by the addition sign + and put
in brackets
Direction of folding twist
Amount of folding twist
Resultant linear density
Example:(25 tex S420 + 60 tex Z80) S360;
R89.2
tex
(Can you describe this yarn?)

Cabled yarns
Cabled yarns have several components, which
can be either similar or dissimilar in
structures.
1.Cabled yarns having similar components
The details used in the designation of such
yarns include:
Notation according to folded yarn used
Multiplication sign, .
Number of folded yarns cabled together
Direction of cabling twist
Amount of cabling twist
Resultant linear density
Example:20 tex Z700 2 S400 3 Z200; R132 tex

2.Cabled yarns having dissimilar


components
The details used in the designation of such
yarns include:
Notation according to single and folded
yarns used, connected by the addition sign
+ and put in brackets.
Direction of cabling twist
Amount of cabling twist
Resultant linear density
Example:
(20 tex Z700 3 S400 + 34 tex
S600) Z200; R96 tex

Single-to- fold versus fold-to-single notations


So far we have used the preferred single-to-fold
notation for yarn designation. Examples of foldto-single notation are given below, together with
their equivalent single-to-fold notation.

Example one
133 dtex f40 S 1000; R 136 dtex (single-to-fold)
R 136 dtex f 40 S 1000; 133 dtex (fold-to-single)
This describes a multifilament yarn of 136 dtex
after twisting to 1000 t/m in the S direction.
Before twisting,
the count was 133 dtex, and the individual
filament
linear density is 133/40 = 3.3 dtex.

Example two

20 tex Z 700 x 2 S 400 x 3 Z 200; R


132 tex (single-to-fold)
R 132 tex Z 200 / 3 S 400 / 2 Z 700; 20
tex (fold-to-single)
This describes a cabled yarn built up
from a singles yarn of 20 tex with 700
t/m Z, plied with itself with 400 t/m S,
which is subsequently three-plied with
200 t/m Z twist.

Example three
(25 tex S 420 + 60 tex Z 80) S 360; R 89.2
tex (single-to-fold)
R 89.2 tex S 360 / (S 420 + Z 80); 25 tex +
60 tex (fold-to-single)
This describes a two ply yarn with dissimilar
components, plied together in the S
direction with 360 t/m.

Finally, the following points are worth


noting:

(1) The resultant count is of most use


to the user of the yarn
(2) In textile mills, abbreviated
notations are often used, i.e. R40/2 tex
(3) For indirect count systems, the
single yarn count is normally given, i.e.
2/20 or 20/2 (2-fold 20s); 3/2/60 (3fold, 2-fold 60s, cabled yarn).

Review questions
Describe

yarns with the following


designations:
(25 tex S420 + 60 tex Z80) S360; R89.2
tex
(25 tex S420 + 60 tex Z80) t0
(20 tex Z700 3 S400 + 34 tex S600)
Z200; R96 tex
20 tex Z700 2 S400 3 Z200; R132 tex

Chapt 2 Yarn
Evenness

2.1

Theoretical aspects of evenness

2.2Measurement

yarn evenness

and benchmarking of

Theoretical aspects of evenness


Introduction
The foundations for the study of yarn
evenness( ) were laid in a 1945 classical
paper by Martindale, entitled a new method
of measuring the irregularity of yarns with
some observations on the origin of
irregularities in worsted slivers( ) and
yarns" (Martindale 1945). For this reason, the
evenness theory has often been refereed to as
the Martindale theory.

Objectives
At the end of this topic you should be able to:
Understand

the statistical limit of evenness


(limiting irregularity)
Appreciate the effect of fibre fineness( )
on yarn evenness
Know the evenness-related calculations
Understand the effect of fibre processing on
sliver and yarn evenness

Perfectly even yarn( )


For a spun yarn( ) (or any other fibre
assembly) to be perfect even, we need two
conditions:
(1)The constituent fibres are uniform in
thickness
(2)The yarn has the same number of fibres in all
cross sections along its length
Figure 1.1 depicts such an ideal fibre assembly
with perfect evenness.

Fig. 1.1: A perfectly even fibre assembly with uniform fibres and butted fibre
ends

While the first condition may be achievable


with manufactured staple fibres, natural fibres
such as cotton and wool always exhibit
variations in thickness along fibre length.
To satisfy condition (2) would mean that the
fibre ends are butted together (Figure 1.1). In
other words, as one fibre terminates, another
must be introduced to take its place. This would
require control and manipulation of each fibre
in the fibre assembly by the processing
machinery. This is not possible with current
processing technology.

Because of the variable nature of fibres,


particularly natural fibres, and the difficulty with
individual fibre placement in the fibre assembly
using current technology, a perfectly even yarn
is unattainable in practice. Therefore, a real yarn
(or any other fibre assembly) would always have
some irregularity in linear density, because the
way fibres are arranged deviates from whats
required to make a perfectly even yarn. The
question then is how does the current fibre
processing machinery arrange fibres? Without
this knowledge, we can not possibly know what
would be the likely irregularity for such an
arrangement.

To answer this question, we need to look at the


whole fibre to yarn processes and examine
what each process does to the fibres. In the
Introduction to Fibre Science and Textile
Technology unit, we have discussed, separately,
the processes involved in manufacturing cotton
and worsted yarns. Let us now briefly recap the
key processes, from the perspective of fibre
arrangement, before we move onto the
theoretical aspects of evenness.

Fibre arrangements during fibre to yarn


processing
Fibres arrive at textile mills( ) in large
bales( ). It is a statement of fact that
fibres vary in properties, both within a bale and
between different bales. We can certainly not
persuade a sheep to produce identical wool or
make a cotton plant to grow identical cotton. In
addition, wool and cotton grown in different
regions exhibit considerable variations in
properties.

To produce a large quantity of uniform


yarns from variable fibres, blending
and mixing is essential. There are two
fundamental requirements of the
blended product:
The blend is homogenous, and
The blend is intimate

These two requirements have different


but complementary connotations. A
homogenous blend means that the blend
components are in the right proportion,
while an intimate blend means the blend
components lie side by side without
regions of concentration of just one
component. If a blend satisfy these
requirements, then fibres in the blend are
thoroughly mixed up, in the right
proportions throughout the bulk (the
whole lot or population).

In other words, within such an ideal


blend, all different fibres are arranged
in a completely random way, and all
the fibres have the same chance of
being found at any selected place in
the bulk. Achieving this task is a major
objective in fibre to yarn conversion.
But it is not an easy task, and has to
be carried out gradually.

In the initial blending of fibres from bales, small


tufts of fibres are picked up and combined to
make a homogenous blend first. For intimate
blend, the fibre tufts need to be opened out
into individual fibres, in the carding process
that follows.
A key objective of carding is fibre opening. Only
when fibres are opened out into individual
fibres can different fibres lie side by side to
achieve an intimate blend. After blending and
carding, fibres are more or less randomized(
). Preserving this randomness is a key
objective of the subsequent drawing process.

During gilling( ) of wool or drawing of


cotton, several slivers are doubled together
first and drafted to reduce its thickness.
Doubling is a random operation because no
deliberate attempt is made to compensate
for thick places by doubling them against
selected thin places. If the fibre ends in the
individual slivers are randomly distributed,
they will still be randomly distributed after
doubling. If drafting is done properly, this
randomness will persist into the drafted
sliver.

One problem with maintaining the


random fibre ends distribution is the
fibre length variation. If fibre length is
very variable or if there are many short
fibres in the slivers before drafting, the
short fibres tend to be drafted in tufts
rather than individually, and a drafting
wave appears in the drafted sliver. Since
a drafting wave is a practically periodic
variation in the number of fibres in the
cross sections along the sliver, it defeats
the randomness of fibre ends
distribution.

For this reason, some fibre control


devices, such as pressure bars( )
in drawing and faller bars( ) in
gilling, are often used to minimize the
drafting waves( ) and improve
the random distribution of fibre ends.
Similarly, in the roving process(
) and during the drafting stage of
spinning( ), fibres are also
controlled during drafting.

From this brief discussion of the fibre


to yarn conversion, we can see that
throughout the different processes
involved, random fibre distribution is
a key objective. If all processes
perform perfectly, we will end up with
a completely random distribution of
fibre ends in the resultant yarn.

We call this yarn an ideal yarn, and the


irregularity of this yarn the limiting
irregularity.. Limiting irregularity is
therefore the minimum irregularity that
we must expect from any real yarn or
other fibre assemblies. A thorough
understanding of the concept of
limiting irregularity is essential for the
understanding of yarn evenness in
particular, and yarn quality in general.

Limiting irregularity( ) (CVlim)


A common method of expressing the
irregularity( ) of a yarn is to use the
statistical term CV or coefficient of
variation( ). Obviously the higher the
CV value, the more irregular the yarn is. The
traditional way of obtaining the CV value is to
dissect a length of yarn into many short
sections of equal length, say 1 cm, and then
weigh each of the short sections. Assuming
we have dissected a yarn into n short
sections, and the weights of these sections
are: x1, x2, x3 ..... xn respectively.

From these readings we can easily calculate


the following statistics regarding the yarn:
The mean or average:
___

x1 x 2 x3 ....... x n xi

n
n

The standard deviation:

__

(1.1)

__

__

( x1 x) 2 ( x 2 x) 2 ..... ( x n x) 2
n 1

__

2
(
x

)
i x

n 1

(1.2)

The

coefficient of variation:

CV

s
__

100%

(1.3)

x
The

percentage mean deviation (known as


the U% value in textiles)

| x

__

x|

100 %

__

(1.3a)

The CV thus calculated will be the measured


CV, or effective CV. It is the actual CV of the
yarn concerned. The U% value is listed here for
completeness. Increasingly, it is the CV or CV%
value that gets used for this purpose. For a
fault-free yarn with random variations in
thickness or linear density, the following
relationship exists between the U value and
the CV value.

CV 1.25 U

(1.3b)

Modern instrument, such as the Uster


Evenness Tester, can measure the U
and CV values of a fibre assembly at a
high speed. More on evenness
measurement will be discussed later.

Coming back to the concept of limiting


irregularity, we have said before this is
the minimum irregularity that must be
expected from even an ideal yarn with
random fibre ends distribution. The
limiting irregularity is also expressed
as a CV value, denoted as CVlim here.
Early works in this area have derived
the following very important
expressions for the limiting irregularity
of various yarns with random fibre
ends distribution.

(1) Limiting irregularity of an ideal


yarn without fibre variability:
100
CVlim
(1.4)
n
where n is the average number of
fibres in yarn cross section.

(2)Limiting irregularity for an ideal


yarn with fibre variability:

CVlim

100 1 0.0001CV A

2
(1.5)

where CVA is the coefficient of


variation of fibre cross sectional area.

These expressions indicate that the


number of fibres in yarn cross section
is overwhelmingly the most important
factor that determines the irregularity
of a yarn. Irregularity increases with a
reduction in the number of fibres in
yarn cross section. The fibre variability
also has some effect on the irregularity
value. But different fibre types vary
considerably in terms of fibre
variability.

Cotton

fibre)

and synthetics( ) (staple

Synthetic staple fibres have very little fibre


variability, and cotton fibres have some small
fibre variability. For these fibres, the number of
fibres in yarn cross section can be worked as
below:
Yarn linear density (tex)
No of fibres in yarn cross sec tion
Fibre linear density (tex)
(1.6)

The fibre variability of synthetic staples may be


ignored and we can simply use equation (1.4) to
calculate the limiting irregularity of a assembly
of synthetic staple fibres.
CVlim ( synthetic staple)

100
n

(1.6a)

Because of the small fibre variability in cotton,


we can not simply use equation (1.4) to work
out the limiting iregularity. Instead, the following
equation is used to calculate the limiting
irregularity of cotton fibre assemblies.
CV lim

106
n

(1.6b)

Worked example:
A cotton yarn of 25s English cotton count (Ne)
g
consists of cotton with a micronaire value of 4.1(
inch
)
What is the limiting irregularity of this cotton
yarn?
Firstly we need to work out the number of fibres
in yarn cross section using formula (1.6). To do
that we need to use the same count unit, tex, for
both fibres and yarns.
From the first module, we already know the
conversion between
English cotton and tex count
590.5
tex
Ne
systems
(
)Therefore, the yarn count in
590.5
23.62 tex
25
tex is

The following shows how fibre fineness is


converted into tex:
4.1 g 4.110 6 g
10 6 g
10 6 g
10 6 g
g

1.61
1.61 2 1.61 5
0.161
0.161tex
inch
2.54 cm
cm
1000 m
10 m
10 1000 m

Using equation (1.6), the average number of


fibres in
yarn cross section is:
23.62 tex
n
146.7
0.161 tex

Applying equation (1.6b), we get the limiting


irregularity for this yarn:
CV lim

106
n

106

146.7

8.75 (%)

Wool

fibres
For wool fibre, it is the fibre diameter and
diameter CV that get measured, not the fibre
cross section area and its CV. In addition, the
average number of fibres in yarn cross section is
not as easy to get as the yarn count. The following
equation has been derived to calculate the
average number of fibres in the cross section of
worsted yarns consisting of 100% wool fibres,
assuming a fibre density of 1.31 g/cm3.
972 tex
n 2
2
(1.7)
D (1 0.0001CV D )
where:
tex = yarn count in tex
D = mean fibre diameter of wool (in micron)
CVD = coefficient of variation of fibre diameter

Since the bulk of the merino wool fibres has a


CVD of about 24.5%, the above equation is
often simplified 917
to: tex
n

D2

(1.7a)

It should be noted though this simplified


equation is based on the assumption that CVD
= 24.5%.
If we put equation (1.7) into equation (1.5) and
note that for wool, we have the
2 following
3.2 D 1 0.0005 CV D
expression
for
CVlim ( wool
) the limiting irregularity
(1.8) of wool
tex
assemblies:

This equation has the following important


implications:
For a given yarn count (tex), the finer
the fibre in the yarn, the less the yarn
irregularity. This is the main reason why
fine fibres are more expensive than
coarse fibres.

For fibres of a given fineness (D), the finer


the yarn, the more irregular it is. This
explains why for a given fibre fineness, there
is a limit on the finest yarn count. It is worth
noting that the concept of irregularity
applies to not just yarns, but fibre
assemblies in general. Therefore, for a given
fibre fineness, the irregularity of sliver will
be less than that of roving, and rovings
irregularity will be less than yarn irregularity.
This can also be explained by considering
the different number of fibres in those fibre
assemblies.

If you reduce the CV of fibre diameter, the


irregularity of the yarn decreases. Put
differently, if you reduce the fibre diameter CV
by 5, you may increase the fibre diameter by 1
micron without significantly affecting the yarn
irregularity. This is the so-called 5-to-1 rule of
thumb.

The equations for wool appear rather complex.


A simpler equation for wool is given below:
CVlim ( wool )

112
n

(1.8a)

However, this equation should be used with


caution, because it is based on assumption
that the CV of fibre diameter is 25%. If the
diameter CV deviates significantly from 25%,
the above formula will lead to error.

Fibre

blends
Blends of different fibres are common and their
popularity is increasing. How do we work out the
limiting irregularity of blend yarns then? This can
be tackled by considering the blend yarn as a
ply yarn consisting of two or more single yarns,
each having one fibre component. If fibres in the
blend yarn are randomly distributed, it is
reasonable to assume the fibres in each
component are also random. Therefore, we can
treat each single yarn the same as we have
treated the 100 cotton or 100% wool yarns.

Given the count of the blend yarn Tb, and the


blend ratio of fibre component Pi, the count of
each component Ti can be worked out
according to the formula below:
Tb Pi
Ti
100

(1.9)

Once we know the count of each component


yarn, the limiting irregularity of the blend yarn
of n fibre components is given as follows,
CVlim (blend )

(CV1 lim T 1) 2 (CV2 lim T 2) 2 ..... (CVn lim Tn) 2


Tb

(1.10)

Worked Example:
A wool/polyester blend yarn is manufactured
on the worsted processing system. The yarn
has a count of 30 tex and contains 45% wool
and 55% polyester. The fibre fineness for the
polyester staple is 2.5 dtex. The mean
diameter of the wool fibres is 22 micron, with
a CV of 25%. What is the limiting irregularity
of this blend yarn?

Solution :
Assuming the blend yarn is a ply of two single
yarns, or 100 wool and 100% polyester
respectively, we can work out the count of the
wool component (Tw) and the polyester
component (Tp) according to equation (1.9):

Tb Pw 30 45
Tw

13.5 (tex)
100
100
Tb Pp 30 55
Tp

16.5 (tex)
100
100

Using equation (1.6), the average


number of fibres in the polyester
component (np) can be worked out as:
np

16.5 tex 165 dtex

66
2.5 dtex 2.5 dtex

Since polyester staple has little variability in


fineness, we can then use equation (2.4) to
work out the limiting irregularity of the
polyester component:
100 100
CV p lim

66

12.3 (%)

For the wool component,


we can use
equation
2
2
3.2 D 1 0.0005 CV D 3.2 22 1 0.0005 25
(2.8),
CV w lim

21.95 (%)
tex

13.5

Finally, we can use equation (2.10) to work


out the limiting irregularity of the blend yarn,
CVlim (blend )

(CV w lim Tw) 2 (CV p lim Tp) 2


Tb

(21.95 13.5) 2 (12.3 16.5) 2


30

12 (%)

So, the blend yarn has a limiting irregularity of


12%.
Finally, since some textile mills still use the U
% value discussed earlier, the limiting U value
can be worked using
the simple equation
CV
lim
U

(1.11)
below:
lim
1.25

Index of Irregularity (I)


After the proceeding discussions on limiting
irregularity, we should be very clear in our minds
that the number of fibres in yarn cross section has a
decisive effect on yarn evenness. Because of this, a
coarse yarn would always be more even than a
thinner yarn made from similar fibres and under
similar processing conditions. Does this mean the
coarse yarn is intrinsically better, in evenness, than
the finer yarn then?

The answer is no. The limiting irregularity


provides a reference point. But this reference
point is not fixed, it changes with yarn count. A
fair comparison of yarn quality in terms of
evenness between similar yarns of different
counts is to see how close the actual CV of each
yarn is to its respective limit (the limiting
irregularity). This would be a measure of the
degree to which the mass variations of a yarn
deviates from the ideal yarn with random fibre
ends distribution.

The index of irregularity provides such a measure.


It is defined as the ratio between the actual
(measured, effective) irregularity and the limiting
irregularity for the yarn (or other fibre
assemblies)
I

CV

eff

CV lim

(2.12)

where I = Index of irregulaity


CVeff = Effective (actual, measured)
irregularity
CVlim = Limiting irregularity.

The index of irregularity is a dimensionless


parameter. In the ideal case, I = 1. Since the
actual CV of a yarn is almost always higher than
its limiting CV, the I value is usually greater than
1. The higher the I, the worse the yarn is in
evenness, regardless of the yarn count. Of
course, as for limiting irregularity, the index of
irregularity also applies to fibre assemblies other
than yarns.

Figure 1.2 shows changes in CV and I of the


fibre assemblies at different stages of the fibre
to yarn conversion. It is worth noting that the
trends for CV and I are quite different. The
index of irregularity (I) gradually decreases
with further processing. This indicates that the
fibre assembly is increasingly approaching an
Defective passage
idealCVeff
one. I
CVeff

I
1
Carded
sliver

Combed Drawn slivers


sliver
1st
& 2nd

Roving

Yarn

Fig. 1.2: Changes in CV and I values in a combed cotton yarn


production

In other words, with further processing, the


fibre ends distribution is getting more and more
random. As mentioned early, promoting
random fibre ends distribution is a key
objective of fibre to yarn processing. At the
yarn stage, the index of irregularity is
approaching one, suggesting that the yarn is
approaching an ideal yarn.

On the other hand, there is a general trend


for the effective (or actual) CV of the fibre
assemblies to increase during fibre to yarn
processing, with the CV of the resultant yarn
higher than the rovings and the slivers. This
is a reflection of the decreasing thickness of
the fibre assemblies, and reducing number of
fibres in the cross section of the fibre
assemblies. At the yarn stage, the number of
fibres in the cross section is the lowest,
hence the CV of the yarn is the highest.

This example again demonstrates the difference


between the CV value and the I value. The I value
provides a good indication of how close a fibre
assembly is to an idea one with random fibre ends
distribution. Because of this, the I value is often
used as a quality control parameter for assessing
the performance of drawing and spinning.

For instance, if the I value is obtained at


every processing stage, and an increase
in I value is found after the 2nd drawing
as indicated by the broken line in Figure
1.2, that immediately tells us that the
2nd drawing is a defective one and
should be fixed. If all processing stages
are under control, the I values should
progressively decrease from start to the
end of the processing as indicated by
the solid line for I values in Figure 1.2.

Unlike the CV% and U% values, the index of


irregularity (I) is independent of the count of
the fibre assembly. This makes it an ideal tool
for use in the control chart. For instance, if
the I value is obtained at the roving stage for
every processing lot and plotted on a control
chart, abnormalities may be easily identified
before the final spinning stage.

Reduction and addition of


irregularities
Figure 1.2 shows that the measured CV (CVeff)
of the cotton sliver gradually reduces from
carding to 2nd drawing, and then the CV
increases again after the roving and spinning
stages. Why is this the case then? What is
causing the increase and decrease in yarn
irregularity. To answer this question, we need to
learn the law of doubling and addition of
irregularity.

Law

of doubling( )

During drawing, many slivers are combined


(doubled) on the input side to feed the
drawframe. The law of doubling says that if you
combine (double) n slivers together, the overall
irregularity of the combined (doubled) sliver will
reduce according to the following law:
(1.13)
CV CV
______

where: CVI = CV of all n slivers at the input to


the drawframe
CV = Mean value of the CV values of all
the single slivers
___
CV CV .... CV
____

CV

n = number of doubled slivers.

Therefore, doubling always reduces the overall irregularity.


This is not difficult to comprehend if you consider the large
increase in the number of fibres in the cross section of the
doubled material. The doubled material is then subject to
drafting, which reduces its thickness. As long as the
drawframe is functioning properly, and the output sliver is
thicker than, or as think as, the average thickness of the input
slivers, the CV of the output sliver will be lower than the
average CV of the input slivers. This explains the decrease in
measured CV from carding to drawing in Figure 1.2.

Addition

of irregularity

At the roving and spinning stage, there is


no doubling. A sliver is drafted into a
thinner roving, and a roving is drafted to
yarn thickness during spinning. The net
result is a reduction in the number of
fibres in yarn cross section. In addition,
the process itself may introduce
additional irregularities to the drafted
material.

Mathematically, if a fibre assembly enters a


drafting process (roving, spinning) with an
irregularity of CVin, and emerges from that
process with an irregularity of CVout, then the
additional irregularity due to the process itself
(CVadd) can be worked out using the following
formula:
CV 2 out CV 2 in CV 2 add

or

CVadd CV 2 out CV 2 in

The added irregularity comes from two sources


reduction in the number of fibres in cross section
and imperfect drafting.
The following example will help understand the
concepts here.

Worked example
Eight slivers, with an average irregularity CV of
4%, were fed to a drawframe. The drawn sliver
has a CV value of 3%. What is the total
irregularity introduced during the drawing
process?
Solution :
Sliver 1problem can be graphically
The above
CVadd = ?
represented as:

CVout = 3%

Sliver 8
CVave = 4%

First of all, we need to know the CV of the input


material (CVin). According to the law of doubling
(equation 1.13), this can be easily calculated:
___

CVin CV
n

4
8

1.4 (%)

Now that we know CVin and CVout is already


given as 3%, we can calculate the CV introduced
during drawing (CVadd) using equation 1.14.
CVadd CV 2 out CV 2 in 3 2 1.4 2 2.65 (%)

As mentioned before, this added CV is due to


two factors reduction in the number of fibres in
cross section and imperfect drafting caused by
material and/or machine related reasons.

Review questions
1.In the calculation of limiting irregularity,
information on fibre length is not used.
This implies that fibre length has noting to
do with the theoretical yarn evenness. Yet
in practice, fibres with shorter length and
higher length variations usually make less
even yarns, other things being equal. How
do you explain this 'discrepancy'?

2. A yarn is composed of 40/60 wool/cotton blend


and has a linear density of 20 Nec (cotton
count). The cotton has a fineness of 3.8
micronaire () and the wool has an average
diameter of 19 m (1 m = 10-6 m) and a
diameter CV of 25%. What is the limiting
irregularity of this wool/cotton yarn? If the blend
ratio is changed to 20/80 wool/cotton, is the
yarn evenness likely to improve or deteriorate,
compared with the 40/60 wool/cotton blend
ratio? (you need to show your workings).
.

3.In his classical book "Studies of quality


in cotton", published by Macmillan and
Co., Limited in 1928, W. Lawrence Balls
described such a paradox - the weaker
the fibre, the stronger the yarn! Please
explain this paradox, using the
information provided in this topic

Measurement and
benchmarking of yarn
evenness( )

Introduction

Up till now we have used the term effective


CV, actual or measured CV of yarns. But how
do we measure the CV of a yarn or a fibre
assembly and what do we do with the
measured results?
This topic discusses evenness measurement
and making use of the measured results.

Objectives
At the end of this topic you should be able to:
Understand the principle of evenness testing
Appreciate the importance of spectrograms
as a diagnostic tool
Know the difference between Uster Statistics
and Yarnspec

Principle of evenness testing


As mentioned before, the traditional way of
evenness testing is to dissect the fibre assembly into
many short sections and weigh each section, and
then calculate the CV of the fibre assembly from the
weights of the individual sections. This is still a
reference method, by which the accuracy of other
methods is judged. Such cutting and weighing
method is a very tedious process as you can
imagine, considering that a sufficiently long length
of yarn should be measured to get a CV value
representative of the bulk material.

Zellweger Uster AG, a textile instrument


manufacturer based in Switzerland, has
produced generations of evenness testing
instrument for rapid measurement of the
evenness of various fibre assemblies. The latest
is the Uster Evenness Tester 4, although its
predecessor (Uster Evenness Tester 3) is still
widely used. A photo of the Uster tester 3 is
given in Figure 2.1.

Fig. 2.1: A photo of Uster evenness tester 3 (Zellweger Uster AG)

The Uster Evenness Tester(Uster )


measures mass variations along the length of a
fibre assembly. It is based on the capacitance
principle as depicted in Figure 2.2. The two
capacitors detect the mass variations or weight
per unit length variations of the fibre assembly
running between them. These variations are
transformed into a proportional electrical
signal. The signal processing unit will process
this signal, and work out the U% and CV%
value, as well as other useful information
concerning the mass variations. All the details
can be displayed or printed out.

Evenness test results


The Uster evenness tester provides a
considerable amount of information on the
evenness of a fibre assembly, including:
Single

overall results
Diagram
Spectrogram

Single

overall results

These include the U% and CV% values, the


index of irregularity (I), as well as the number
of imperfections (thin place, thick place, and
neps). All those parameters are expressed as
single numbers, which are easy to use,
particularly in a mill situation. These single
values provide an overall picture of yarn
evenness. However, if the results are bad, the
causes of the poor results can not be
identified from these single values

Diagram

A diagram is simply a trace of mass (linear


density) variation along a fibre assembly. For
instance, if you dissect a long length of yarn into
many very short sections and then weigh each
section, you will get many mass readings (xi) as
Mass
shown
in Figure
2.4.
x (individual mass readings)
i

Mean
mass

Fig. 2.4: A manually constructed diagram of mass variation

Length

From this diagram, many useful statistics (mean,


CV etc) can be obtained as shown in the Limiting
Irregularity section.

A diagram obtained from the Uster evenness


tester is given in Figure 2.5. The dip in the
middle of this diagram was actually caused by
a missing sliver, of about 120 m, in the input
material for a drawframe. This example
demonstrates the usefulness of diagrams in
identifying certain faults in the fibre assembly.
Basically, the diagrams can help identify
extreme thin and thick places, slow changes in
the mean mass value, step changes in the
mean value, periodic mass variations of long
wave length etc.
Fig. 2.5: A diagram showing an extreme thin
place in the sliver (Furter 1982, p.12)

Spectrograms( )

The single overall results are very useful


in that they provide evenness
information in concise single values.
These single values are easy to use for
comparison purpose in particular. For
instance, the CV% or index of
irregularity of one yarn is higher than
another similar yarn, we can say one
yarn is better than the other in terms of
yarn evenness.

But that is often not sufficient for


quality control purpose. Suppose we
now know from the single overall
results (eg. CV, I) that a yarn is not
good in evenness, and we want to find
out what has caused the irregularity in
the yarn. Once we know what has
caused the irregularity, we can then try
to rectify the problem. For this, we
need the spectrograms.

Before we discuss the spectrogram, it is necessary to


say a few words about the nature of mass variations
in a fibre assembly. We already know that random
fibre arrangements lead to mass variation, and this
variation can be precisely calculated as discussed in
the limiting irregularity section. If that is all the
variation we get, then we have nothing more to
worry about, because that is exactly what we aim for
in a yarn. Unfortunately we often get more than just
the random variations, for two common reasons:

(a) Variability in fibre length and the


presence of short fibres make fibre control
during drafting difficult, this leads to nonrandom variations in a fibre assembly. Such
non-random variation is called a drafting
wave. It is called a drafting wave because
the mass variation occurs in a more or less
periodic manner in the drafted material,
much like a wave of variations along the
length of the fibre assembly.

(b) There may be machine defects or


mechanical faults in the drafting
systems, which causes changes in
drafting speed and the actual draft
periodically, leading to rather strictly
periodic mass variations in the drafted
fibre assembly.

Here we need to reflect upon what has been


discussed on Roller Drafting in the Introduction
to Fibre Science and Textile Technology unit. For
roller drafting, as depicted in Figure 2.6, the
most important concept is the concept of
perfect roller drafting.
Front rollers

Back rollers

Slower

Faster

Ratch setting

Fig. 2.6: A simple roller drafting system

The concept of perfect roller drafting is:


every fibre in the drafting zone should
travel at the speed of back rollers until its
leading end reaches the front roller nip.
Then the fibre gets instantly accelerated to
the front roller speed.If this is what actually
happens in roller drafting, we will get a
drafted fibre assembly with random
variation of fibre ends only.

However, when there are many short fibres


in the drafting zone, these short fibres will
not move according to the requirements of
roller drafting. They float and swim
together in the drafting zone, the speed at
which they travel depends on the speed of
their neighbouring fibres. The end result is
some practically (i.e. not strictly) periodic
mass variation in the drafted material. Such
practically periodic mass variation caused by
floating short fibres is called a drafting wave,
and its wave length is approximately 2.5 to 3
times the average fibre length of the fibre
assembly.

With good fibre control, using pressure bars (on


cotton drawframe) and pinned faller basr (on
worsted gillbox), drafting waves can be
significantly reduced or eliminated.
A common machine defect or mechanical fault of
Fig. 2.7: Eccentric rollers cause periodic mass variations
drafting elements
is roller eccentricity, as
front
indicated Eccentric
in Figure
2.7.
bottom roller
Drafted material

Speed varies
with radius
R
Wave length
(Roller circumference)

Eccentric back
bottom roller

Drafted material

R
r
Wave length
(Roller circumference x draft)

Because of roller eccentricity, the surface


speed (v) of the eccentric roller varies as the
radius of rotation (r) varies (n, where n is the
roller rpm). If the front bottom roller is
eccentric, a larger radius of rotation (R) will
lead to higher roller surface speed, which
means increased drafting, resulting in over
draft or a thin section in the drafted material.
The opposite happens with the smaller radius
of rotation, and this cycle repeats for every
complete revolution of the eccentric roller. As a
result the wave length of the periodic variation
is exactly the same as the circumference of the
offending roller.

On the other hand, if the back bottom roller is


eccentric and front rollers are fine, then at the
larger radius of rotation R), the back roller
surface speed will be faster, leading to a
reduction in draft and hence a thicker section in
the drafted material. The opposite is the case at
the smaller radius of rotation (r). Not only that,
the periodic mass variation caused by the back
eccentric roller will be lengthened by a factor of
the draft used. In other words, the wave length
of the periodic mass variation caused by a back
eccentric roller will be equal to the roller
circumference multiplied by a factor of draft, as
indicated in Figure 2.7.

Periodic mass variations in a yarn


often result in unwanted patterning in
fabrics made from such yarns. They
also lead to increased ends down
during spinning and subsequent
processing. It is essential in yarn
manufacture to prevent the
occurrence of such mass variations in
slivers, rovings or yarns.

Furthermore, the presence of periodic or


practically periodic mass variations in a
fibre assembly does not necessarily
result in significant increases in the CV
% value or in the index of irregularity.
So the CV% value or index of
irregularity will not indicate the
presence of those mass variations. But
how do we know if a fibre assembly has
a drafting wave or periodic mass
variation then? This question leads us
back to discussion on spectrograms.

Hypothetically, if a yarn has mass


variations that resemble a sinusoidal
wave as shown in Figure 2.8(a), then a
mathematical (Fourier) transformation of
such a mass variation signal will reveal
the frequency (f) of such variation as a
sharp peak shown in Figure 2.8(b). For a
signal that is not as simple as just a
sinusoidal wave, it has been proven
mathematically that it can be constructed
by superimposing a series of sinusoidal
waves of varying frequencies.

Therefore, if the original mass variation


in the yarn is of a more complex shape
as shown in Figure 2.8(c), then the
same mathematical transformation will
reveal the frequency of each of its
sinusoidal components as shown Figure
2.8(d). The different amplitude reflects
the different share of the respective
component in the original signal.

Amplitude

Amplitude
(a)

(b)
Transformation

Time

(c)

Frequency
(d)

Transformation

Time

f1

f2 f3 Frequency

Fig. 2.8: Transformation of time domain signal to frequency domain

If the original mass variation is of a random


nature, then after transformation, there will be
many frequencies of similar amplitude. Further, if
there is a periodic mass variation in addition to
the random variation, then the frequency of that
periodic mass variation will show up as a sharp
peak after the transformation. Put differently, if a
mass variation signal is subjected to a
transformation and a sharp peak (chimney)
appears in the transformed signal, then we know
there is a periodic mass variation in the fibre
assembly. This is basically how spectrogram
works.

Since wave length is more useful than


frequency for textile purposes, the
spectrogram indicates the different wave
lengths (on a logarithmic scale) versus their
amplitude. Modern evenness testing
instruments, such as the Uster Evenness
Tester, provide diagrams as well as
spectrograms for the fibre assembly tested.
The diagram is a time domain mass
variation signal, while the spectrogram
represents the same mass variation in the
frequency domain.

Figure 2.9 shows the diagrams and


spectrograms of 3 different yarns
normal yarn with random variation
only, faulty yarn with additional
periodic mass variation, and faulty
yarn with additional drafting wave.
Fig.2.9: Diagrams (left) and spectrograms
(right) of 3 yarns (Uster Spectrograph,
Zelleger Uster, PE404)

With respect to interpreting a spectrogram, the


following simple rules can be used as a guide:
(a) A fault-free fibre assembly will give a typical
normal spectrogram (with neither chimnies
nor humps)
(b) A chimney on top of a normal spectrogram
indicates the presence of a periodic mass
variation in the fibre assembly. The wave length
of this periodic mass variation can be read off
the horizontal axis (noting the logarithmic
scale)

(c) A hump on top of a normal spectrogram


indicates the presence of a drafting wave in the
fibre assembly. The wave length of the drafting
wave is equal to 2.5 to 3 times the mean fibre
length.

Once we get the wave length of a periodic


mass variation from the spectrogram, and we
know this wave length is related to the
circumference of the offending roller, we can
then identify the roller and replace it with a
good one to solve the problem. For drafting
waves, the use of more uniform fibres and
proper fibre control during drafting will usually
solve the problem.
Spectrogram is therefore a very useful quality
control tool in a spinning mill.

Benchmarking yarn evenness


In management jargon, benchmarking is a total
quality management tool and denotes the
procedure of identifying and quantifying
topnotch or world-class performance
(benchmarks) in a particular business or product
category and comparing the data with the
performance of the own company or product.
Lets assume we have already produced some
yarns and we have tested the yarns for
evenness. Now we want to know how good our
yarns are. In other words, we want to benchmark
a product - our yarns.

There are several ways of benchmarking


yarn evenness, including:
Index

of irregularity
Uster Statistics
Yarnspec (for worsted yarn only)

Index

of irregularity
Table 2.1 shows a classification of worsted yarns
based on the index of irregularity of the yarn.
Table 2.1: Classification of worsted yarns based
on the index of irregularity

Since processing technology is improving and


so is yarn quality, the data in this table may not
reflect the quality of worsted yarns in the future.
Generally speaking, a good quality worsted yarn
should not have an index of irregularity greater
than 1.2 by todays standard.

Uster

statistics(Uster )

While the evenness index value is of use to the


yarn manufacturers for internal quality control
purpose, what matters to the users of yarn (i.e
the weavers and knitters) is the actual
irregularity in the yarn they are going to use.
For this reason, the Uster Statistics is of great
practical importance.

So what is the Uster Statistics then?


The following excerpts from the 1997
Uster Statistics Book (produced by
Zellweger Uster) answer this question
briefly:Almost half a century ago, in
1949, the first Uster Standards were
presented to the textile public in
numerical form. This started a new era
in the assessment of the technological
and commercial value of spun yarns.

Over the years, the Uster Standards


have developed into the Uster
Statistics, which have been regularly
updated until today and additional
quality parameters for sliver, roving,
and yarns have been introduced
progressively. Simultaneously, the
methods and procedures applied to
establish the Uster Statistics have
been gradually enhanced.

Today, the Uster Statistics represent the only


truly comprehensive survey of the quality of
textile materials produced in the major textile
hubs around the world and they constitute the
mainstay of global market intelligence related to
textile quality.
The Uster Statistics are first and foremost a
practical guide to good textile practices in the
field of yarn manufacturing.
The Uster Statistics just seem to have been
made for quality benchmarking on the corporate
level.

Uster Statistics 1997 provide data on the


following major types of yarn:
100% CO, carded, ring spinning 100%
carded cotton (ring spun)
100% CO, carded, rotor spinning 100%
carded cotton (rotor spun)
100% CO, combed, ring spinning 100%
combed cotton (ring spun)
100% CO, combed, rotor spinning 100%
combed cotton (rotor spun)
100% CO, carded, rotor spinning 100$
carded cotton (rotor spun)
100% WO, worsted spinning 100% wool
yarn (worsted ring spun)

Provisional data is also provided for the


following types of yarn:
100% PES, ring spinning 100% polyester (ring
spun)

100% CV, ring spinning - 100% Rayon (ring


spun)
100% CV, rotor spinning 100% Rayon
(rotor spun)
65/35, 67/33 PES/CO, combed, ring
spinning 65% polyester/35% cotton blend,
combed (ring spun) and 67% polyester/33%
cotton blend, combed (ring spun)

65/35, 67/33 PES/CV, ring spinning 65%


polyester/35% Rayon blend (ring spun) and
67% polyester/33% Rayon blend (ring spun)
50/50 PES/CO, rotor spinning 50%
polyester/50% cotton blend (rotor spun)
50/50 PES/CO, air-jet spinning 50%
polyester/50% cotton blend (air-jet spun)
65/35 PES/CO, air-jet spinning 65%
polyester/35% cotton blend (air-jet spun)
55/45 PES/WO, worsted spinning 55%
polyester/45% wool blend (worsted ring spun)

The key quality attributes listed for these yarns


are:
Yarn

count variation (between bobbins or


packages)
Mass variation (U% and CV%)
Imperfections (thick and thin places, neps)
Uster Hairiness Index
Tensile properties (strength and elongation)

Yarnspec

(for worsted yarns)


Yarnspec is a computer program developed by
Scientists at CSIRO Textile and Fibre Technology
in Geelong. Since the work was funded by
Australian wool growers, the program has been
specifically designed for the prediction of
properties of worsted yarns and the
performance of worsted spinning, based on the
properties of worsted tops and the spinning
conditions. The predicted results are what a
worsted spinner can expect in terms of spinning
performance and yarn quality if the operation
follows best commercial practice. In other
words, Yarnspec can be used to benchmark the
performance of worsted spinners.

For single worsted yarns, the program requires


the following details as the input:
(a) Wool properties (from tops)
Fibre diameter (micron) and diameter CV
Hauteur length (mm), CV of Hauteur and
Fibre bundle tenacity (cN/tex)
(b)Processing details (spinning)
Spinning draft
Spindle speed (rpm)
Ring size (mm)
Traveller number

(c)Yarn details
Yarn count( )
Yarn twist( )
Dyed or undyed( )
The predicted outcome includes the following
details:
Yarn evenness( ) (I, CV%, U%)
Yarn Imperfections( ) (Thin
places/km, thick places/km, and neps/km) (
)
Yarn tenacity and breaking elongation% (
)
Spinning ends-down per 1,000 spindle
hours( )

For a worsted spinner, Yarnspec is a step ahead


of Uster Statistics for performance
benchmarking, because it takes into
consideration of the fibre properties used in
spinning the yarn. In addition, it provides
information on yarn strength as well as on the
critical spinning performance in terms of endsdown per 1,000 spindle hours. Yarnspec can
also be used to predict the properties of two
folded yarns.

Review questions
1.An ideal sliver of 70 mm mean fibre length is roller drafted
with a draft of 10 under the following three conditions:
(a)Perfect roller drafting
(b)Presence of a large number of uncontrolled short fibres
(c)An eccentric back drafting roller with a diameter of 3
cm.
Explain how drafting under each condition will affect the
evenness of the drafted sliver, and sketch and label the
spectrogram for each drafting condition.
2.A 50 tex worsted yarn of 100% wool is measured for its
evenness on the Uster evenness tester. If the CV of this
yarn is 15%, how good is this yarn in relation to world
production of similar yarns?

Fibre preparation for


spinning

Short Staple Processing


Introduction

Short staple fibres( ) refer to fibres less than


2 inches in length. Cotton is a typical example of
short staple fibre. The short staple system(
) is used to process cotton mainly,
cotton/polyester( ) blends are the next most
commonly processed fibres on the short staple
system. Other fibres, such as viscose( ), are
also processed occasionally using the system. Short
staple yarns make up the bulk of international yarn
market.
Since cotton is the dominant fibres used, the
emphasis of this topic will be on cotton processing.
The actual spinning of yarns is discussed in a
separate module.

Objectives
At the end of this topic you should be able to:
Know the flow chart of cotton processing
Understand the principles and objectives of
carding, drawing, and combing
Appreciate the differences in the process and
property of carded ring spun yarn(
), combed ring spun yarn( ),
carded rotor spun yarn( ), and
combed rotor spun yarn( ).

Cotton growing

Process
overview

Harvest
Cotton seed by product

Ginning

The process flow


chart for cotton
processing from
fibre to yarn is
shown in Figure
1.1.

Agricultural
Processes

COTTON LINT
Baling, HVI Classing

(Transport to textile mill)


Engineered Fibre Selection EFS
(Marshalling into mixes or laydowns)

Blow-room processes
(blend, open & clean)
Carding
Drawing
Lap forming
Combing
Drawing (x 2)

Rotor spinning

Roving
Ring spinning

Short staple yarns


Carded rotor spun yarn,
Combed rotor spun yarn,

Carded ring spun yarn


Combed ring spun yarn

Fig. 1.1: Fibre to yarn processing for cotton

Textile
Processes

The agricultural processes include cotton


growing, harvesting and ginning( ).
Cotton grown in different regions have
different properties. Modern cotton
harvesting uses machine pickers or
strippers. Since cotton fibres do not
mature at the same time and machine
picking is less discriminative than
traditional hand picking, large quantities
of impurities such as green bolls, leaf,
stick, and trash are also picked up during
cotton harvesting, together with the seed
cotton.

On a weight basis, seed cotton


contains approximately 35% fibre
(lint), 55% seed, and 10% trash.
Obviously the cotton seed and other
impurities need to be removed from
the fibres. This is largely done in a gin,
which removes all green balls and
cotton seeds, about 95% burrs, 92%
sticks, and about 85% fine trash.

The actual process (in a gin) that separates


cotton fibres and the seed they grow
attached to is called ginning. Other
machines are also used before and after
ginning to mechanically clean the fibres. The
ginned cotton is now known as cotton lint or
lint cotton. The fibres in the cotton lint vary
considerably in length because of fibre
breakage caused by the severe mechanical
actions during ginning and cleaning. The
cotton lint is then sampled and packed into
bales weighing 227 kg (or 500 pounds),
containing over 60 billion individual fibres.

Fibre samples are now tested on the High


Volume Instrument (HVI) for a range of
fibre properties, including strength,
elongation, length, uniformity, micronaire,
colour, and trash. The HVI system was
developed in the late 1960s and has been
increasingly accepted since the 1980s.
Before the introduction of the HVI system,
cotton in a bale was graded subjectively
by experienced cotton classers for
properties such as staple length, colour,
and trash content.

The results were then used to assign


cotton bales into lots or categories.
When the cotton was ready for
consumption, the bales were grouped
into mixes or laydowns. Bales from
different regions were mixed in
proportion to the number of bales in
each lot and fed into the opening line
machinery in the blow-room.

Today, objective measurement is widely


used. When the bales arrive at a textile mill
to start the textile processes, the test results
are used as a basis for fibre selection and
mixing according to the end product
requirements. The modern cotton mill will
"engineer" its yarn to meet specific end-use
requirements. The engineered fibre selection
(EFS) system, introduced by Cotton Inc.
(USA) in 1982, has been used increasingly
by cotton mills to facilitate this important
task.

It is most useful in bale management,


particularly for storing and retrieving
bales, for selecting bales with fibre
properties within specified ranges and
average values, for composing
consistent bale laydowns and for
predicting yarn strength and other yarn
properties based on tailer-made
regression analyses. The bulk of the
cotton bales consumed in America is
now managed at the mill level by the
engineered fibre selection system (EFS).

Adequate fibre blending and mixing is also vital


to ensure processing efficiency and yarn quality.
The cotton lint still contains some small trash
particles, which have to be removed by the
textile processes, such as carding and combing.
The textile processes also perform fibre opening,
fibre alignment, fibre mixing and attenuation to
get the fibres ready for spinning. As indicated in
Figure 1.1, depending on the particular
processing route followed, four major types of
cotton yarn may be produced carded rotor
spun, carded ring spun, combed rotor spun,
combed ring spun.

An overview of the key stages is given in Table


1.1.

It is important to keep in mind at this stage


that before fibres can be made into useful
yarns, they should be:
Free

from impurities
Well individualised and aligned
Well mixed
Of adequate length and strength

Knowing these requirements will help us


understand why the fibres need to go through
many textile processes before the actual spinning
process. For instance, in order to remove
impurities imbedded in fibres, we need to open
the fibres first to expose those impurities. Fibre
opening needs to be gradually carried out so as
not to stress and damage the fibres too much. In
fact, there are two opening stages:
Stage 1: Breaking apart (break large tufts of
fibres into small tufts)
Stage 2: Opening out (open small tufts into
individual fibres)

Individualising the fibres is very important. As


mentioned in the module on yarn evenness,
poorly separated fibres will travel in groups
during drafting, which will lead to reduced
evenness and increased imperfections in the final
yarn. For a yarn to have adequate strength, fibres
in the yarn should be well aligned in order to
share the applied load on the yarn. The different
degrees of fibre alignment in different yarns often
explain the differences in yarn properties.

Because of the variability that exists both within


and between fibres, fibres should be well mixed
before the actual spinning stage. There are two
basic requirements for a good fibre mix or blend:
Requirement 1: The blend (mix) is homogenous
Requirement 2: The blend (mix) is intimate

The first requirement entails that different


fibres are mixed in the right proportion, while
the second requirement can only be achieved
with different individual fibres lying side by
side.
Preserving the quality of fibres during
processing is also essential to ensure yarn
quality. Damage to fibre length and strength
will lead to reduced yarn strength.
With this overall picture in mind, we can now
discuss the individual textile processes applied
to fibres.

Objectives

The blowroom is the section of a cotton spinning


mill where the preparatory processes of opening,
blending and cleaning are carried out. The
blowroom machines blend, open and clean the
ginned cotton before feeding it to the cotton card.

The ginned cotton, still contaminated with some


impurities, arrives in the textile mill in compressed
bales, fibre properties often vary from bale to bale.
Blending is regarded as the most important process
in a cotton spinning mill. It reduces variation of
fibre characteristics, permits uniform processing
and improves yarn quality. In the blending process,
different cottons of known physical properties are
combined to give a mix with the required or predetermined average characteristics.

For example, the general formula for calculating the


theoretical fineness (micronaire, g/in.) is as follows
Wt
Wt
Fb

Wn
W
W1 W2

.....
F
F1 F2
Fn
where Fb is the fineness of a blend of n components;
Wt is the total weight of the blend; and W is the
weight of any one component and F is its fineness.
In terms of weight percentages, the above equation
becomes:
100

100
Fb

Pn
P
P1 P2

.....
F
F1 F2
Fn

where P is the percentage by weight of any one


component and F is its fineness.

The group of different bales to be mixed is


referred to as the "laydown", with laydowns
containing between 10 and 100 bales. In a
modern cotton spinning mill, automatic bale
pluckers are used to extract and blend fibres
from the laydowns. The bale plucker extracts,
at controlled rates, small tufts of cotton from
each bale (in the laydown) and delivers the
fibres through pneumatic ducting (hence the
name blow-room) to the subsequent opening
and cleaning machines.

In the opening and cleaning processes,


clumps of fibres from the bale plucker are
opened to smaller tufts to facilitate further
processing and to allow removal of
impurities, such as dust, sand, seed
particles, leaf and stem fragments and
motes (undeveloped seeds). A number of
machines are employed to perform the
opening and cleaning actions gradually, so
as not to cause too much fibre damage. In
other words, the opening action in blowroom is largely the breaking apart stage
mentioned in the previous section.

Opening-out or individualising the


fibres is not intended in the blowroom. But some fibre damage is
inevitable. In addition, the opening
processes also create neps, or highly
entangled fibres. At the end of the
blowroom processes, the opened and
cleaned cotton fibres are condensed in
a lap form to feed into the carding
process.

Blowroom

installations
In a typical blowroom installation, six
distinguishable zones can be identified as
indicated in figure 1.2. These are:
ZONE 1 - Bale opening
ZONE 2 - Coarse blending
ZONE 3 - Blending (Mixing)
ZONE 4 - Fine cleaning and dust removal
ZONE 5 - Intensive opening/cleaning (optional)
ZONE 6 - Card feed

If the cotton contains only few impurities, then


the zone of intensive cleaning is not necessary.
Figure 1.2 Operating zones in a typical blowroom (Klein 1987a, p.13)

The zone 1 opening machine is usually an


automatic bale plucker, an example of which is
given in figure 1.3. As mentioned in the previous
section, the bale plucker extracts, at controlled
rates, small tufts of cotton from different bales in
the laydown and delivers the fibres through
pneumatic ducting to the subsequent opening
and cleaning machines.

Figure 1.3 An automatic bale opening machine


(Fritz and Cant 1986, p.327, courtesy of Rieter
Machine Works)

The zone 2 coarse cleaning (and opening)


machine is usually a step cleaner (figure 1.4)
or a dual roller cleaner (figure 1.5). Both
cleaners feature widely spaced striker
elements working on fibres in relatively free
flight. So the action on fibres are gentle.

Figure 1.4 A step cleaner (Shrigley 1973, p91)

For the step cleaner, fibres enter through the


entry point (A). They are worked on by a series of
6 inclined beaters (B, D), under which is fitted a
grid-bar system (C). The grid bars support the
upward flow of fibres while allowing impurities to
fall through into the trash box (F). The action of
opposing spikes between adjacent beaters helps
opening fibres and dislodging impurities. The
baffle plates (E) prevent the beaters from
creating a circular air current by deflecting the air
and fibre to the next beater. The coarsely cleaned
fibres come out the step cleaner through exit (G).

Figure 1.5 A dual-roller cleaning (Shrigley 1973,


p93)

The dual-roller cleaner, shown in figure 1.5, has two


spiked rollers (beaters A and B) mounted
horizontally, with grid bars beneath each beater. A
condenser fan downstream the cleaner draws
fibres through the machine by suction. Fibres are
opened by the spikes gradually. As can be seen
from figure 1.5, the outlet is positioned higher than
the inlet, ensuring that only the relatively small
and light tufts of fibres can fly straight through. The
residence time of fibre materials within the
machine can be adjusted by the baffle plate D.
An example of zone 3 blending machine is shown
in figure 1.6. This machine comprises several
adjacent chambers into which the fibres are blown
from above. The chambers are filled successively,
but the fibre material is removed from all chambers
simultaneously. This gives good long-term mixing.
Figure 1.6: The Hergeth Hollingsworth multiple mixer (Klein 1987a, p.19)

Fine cleaning requires further opening of fibres to expose


the impurities. For this reason, finer opening elements are
used.
Dust removal is not confined to one particular zone. But
fine dust can only be effectively removed when the fibres
are relatively open. Dust removal equipment is usually
incorporated into the pneumatic ducting system, with the
dust separated from fibre by air suction through
perforated surfaces. The principle of a condenser type
dust extractor is shown in figure 1.7. The condenser also
helps even the flow of fibres, because the thin spots in the
fibre sheet laying on the perforated screen will allow more
air to flow, hence carrying more fibres to the thin spots.

Figure 1.7: Sketch of a condenser type dust extracter (Lord


1981, p.137)

Finally, the opened and cleaned fibres are fed


to a card through a chute feed device. An
example of a chute feed is shown in figure 1.8.
It is vital that a uniform sheet of fibres be fed
to the card. For this reason, chute feeds have
control systems to ensure fibres are uniformly
packed inside the chute.

Figure 1.8: An example of a chute feed (Lord


1981, p.138)

Cotton carding
Objectives

An old spinning mill adage states that "well


carded is half spun". Card is also referred to as
the heart of a spinning mill. This highlights the
importance of the carding process. The main
objectives of carding are:
Fibre opening/individualising
Fibre cleaning
Fibre mixing
Fibre aligning
Sliver forming

Understanding how a card achieves these


objectives requires knowledge of its operating
principle. Figure 1.9 shows a diagram of a
modern high performance flat-top card.

Fibres from the blow-room are supplied


pneumatically via pipe ducting (1) to the feed
chute (2) of the card. An evenly compressed fibre
batt of about 500 900 ktex (g/m) is formed in
the chute. A transport roller (3) forwards fibres
from this matt to the cards feed device (4),
consisting of a feed roller and a feed plate. The
licker-in or taker-in (5), covered with strong metal
teeth, snatches fibres from the feed device,
dislodges heavy impurities via the gaps of grid
segments (6), and carries the fibres to the main
cylinder (8). The suction ducts (7) carry away the
dislodged impurities. The sawtooth elements (or
clothing elements) on the main cylinder (8), which
has a higher surface speed than the taker-in, strip
the fibres off the taker-in (5) and carry them to the
main carding zone between the cylinder (8) and
the flats (10) (hence the name flat-top card).

The flats comprise some 80 120 carding bars


combined into a band moving on an endless
path. When in the carding zone, the teeth on
the carding bars and that on the main cylinder
act together to repeatedly tear apart fibres
into individual ones, to remove neps (highly
entangled fibres) and some impurities. As the
action carding bars emerge from the carding
zone, a cleaning unit (11) strips fibres, neps
and impurities from the bars, and the bars
then return for further action in the carding
zone. Obviously, the main carding zone is
where most of the cards objectives are
achieved. Extra carding bars (9, 12) are also
used to increase the level of fibre opening in
carding. The underside of the main cylinder is
enclosed by grids or cover plates (13).

Fibres coming out of the main carding zone are


individualised and aligned. They are carried by
the teeth on the main cylinder to meet the teeth
on the doffer (14), which has a slower surface
speed than the cylinder. The doffer snatches
some (not all) fibres from the cylinder surface
and combines the fibres into a web because of
its substantially lower surface speed relative to
the main cylinder. Fibres not snatched by the
doffer continue to travel with the main cylinder.
These fibres are called recycling fibres. The
recycling fibres will soon meet with fresh fibres
from the taker-in and together these fibres are
worked on in the main carding zone.

This is how fibre mixing is achieved in


carding. In fact, carding is the only
process where intimate mixing can be
achieved. A fibre may go around the
main cylinder many times and get
mixed with many fresh fibres before it
is finally removed by the doffer. The
stripping device (15) then removes the
fibre web from the doffer. The web is
brought together as a sliver and
compressed by a pair of calender rolls
(16). Finally, the coiler (18) deposits
the sliver into a sliver can (17).

The carding action of a flat-top card like this is


quite intensive, and may cause considerable
fibre damage and breakage, particularly for
long fibres. For this reason, flat-top card is not
used to card long fibres such as wool. Carding
also creates some neps, or highly entangled
fibres. While carding is supposed to align and
straighten individual fibres, most fibres in a
carded sliver have hooked ends. The reason for
this will be discussed later.

Card

clothing

The tern 'card clothing' refers to the large


number of pins or teeth covering the surfaces of
various rollers on the card. There are three
major types of card clothing - flexible fillet wire,
semi-rigid wire, and metallic sawtooth wire, as
carding angle
shown in figure 1.10.
Knee

Backing
material

(a) Flexible
fillet wire

(b) Semiregid wire


Figure 1.10 Different types of card clothing

(c) Metallic
sawtooth wire

The flexible fillet wire was developed to mimic the


flexibility of natural teazles used in the past. This
type of clothing is believed to minimise fibre
damage during carding. The fillet wire is designed
with a knee on each needle, so that when the
needle bends back as force is applied during
carding, the point of the needle does not project
too far outwards. Otherwise the points on two
adjacent surfaces will touch each other, causing
damage to the needles. In other word, the knee is
designed to prevent point-rise to avoid wire
damage. The needles are embedded in the
backing material. Dirt and short fibres tend to
accumulate under the knees of flexible fillet wire. If
not removed, carding efficiency will drop.
Removing the dirts and short fibres from the card

The semi-rigid wire is often used on the flats of


cotton cards. It is more rigid than the flexible
fillet wire and has no knee. During carding,
some short fibres and impurities will also
accumulate in the semi-rigid wire. Long fibres
have more contacts with the main cylinder
clothing and will tend to move with the main
cylinder. The short fibres and impurities
removed from the clothing of flats are called
flat strippings. Compared with flexible fillet
wire, the semi-rigid wire has the advantage of
not choking with fibres and correspondingly
eliminating less flat strippings.

The metallic wire, also known as sawtooth wire,


is a later development. It is less prone to damage
and allows higher production rate. Modern high
production cards are usually fitted with metallic
fillet wire. The carding angle as indicated in
figure 1.10 is the most important of the tooth.
The higher the carding angle, the more
aggressive the tooth is, and more fibres it will
hold during carding. For this reason, the taker-in
clothing usually has a very small carding angle,
or even a negative carding angle so that fibres
on the taker-in can be easily transferred to the
cylinder. Similarly, the doffer clothing usually has
a higher carding angle than the cylinder clothing
to allow fibre transfer from cylinder to doffer.

Basic

actions in carding
There are two basic actions in cotton carding:
carding (or working) action and stripping
action.
The tooth direction and relative surface speed
decide which action occurs between two
adjacent and clothed (or toothed) surfaces.

(a) Point-to-point carding


Each tooth has a point and a back, as indicated
in figure 1.11a.
If the tip of the tooth on one surface points to
the tip of the tooth on the other surface, a
point-to-point carding (or working) action
occurs (figure 1.11b).
point
back
back

point

(a) The Point and Back of a tooth

(b) Carding action (Point-to-point)

Fig. 1.11: Carding and stripping actions

(c) Stripping action (Point-to-back)

For instance, the teeth arrangements on flats


and main cylinder, and on main cylinder and
doffer (Figure 1.9) are typical point-to-point
arrangements. Therefore carding action
occurs between flats and main cylinder, and
between main cylinder and doffer.
It is through the carding action that fibre
opening occurs. Both surfaces contest for
fibres and as a result, fibres are separated.

The level of fibre opening in carding can be


represented by points-per-fibre. This is the ratio
of the total infeed fibres per unit time over the
number of working points available in the same
time. As the card production rate increases,
more fibres must pass through the card, this
would reduce the number of points-per-fibre,
hence the carding effect on the fibres. To
maintain the carding effect, extra working
points are added on modern cards.

(b) Point-to-back stripping


If the tip of the tooth on one surface points to
the back of the tooth on the other surface, a
stripping action occurs (figure 1.11). The point
strips fibres off the back.
For instance, a stripping action occurs between
the main cylinder and the taker-in. The teeth on
the main cylinder point to the back of teeth on
the taker-in, so the fibres on the taker-in are
stripped by the teeth on the main cylinder.
It is through the stripping action that fibres are
transferred from one surface to another during
carding.

Quality

of carded sliver
The important quality considerations for the
carded slivers are:
Fibre

length
Number of neps
Fibre alignment
Sliver evenness

(1) Fibre length


As mentioned earlier, the intensive carding
action causes considerable fibre damage and
breakage, leading to reduction in fibre length
and increase in short fibre content. Changes
in mean fibre length before and after carding
has also been used to estimate the level of
fibre breakage in carding, using the formula
below: Mean fibre length before carding main fibre length after carding
% Fibre breakage
ean fibre length after carding

(2) Number of neps


Neps are highly entangled fibres. Usually a nep
contains ten or more fibres. It is a very serious
problem in the textile industry.
The ease with which a fibre forms part of a nep is
related to its bending rigidity. Immature cotton and
fine fibres bend easily. They are prone to nep
formation during carding. Many neps often persist
into the final fabrics. Neps contain many immature
cotton fibres, which have less cellulosic materials
than mature fibres. During fabric dyeing, they do
not take up as much dye as the rest of the fabric,
causing a serious fabric fault known as white
specks.
Closer card settings between adjacent surfaces,
sharper teeth, and higher doffer speed (reducing
recycling fibres) can be used to reduce the
number of neps in carding.

(3) Fibre alignment


Ideally, fibres in a carded sliver should be
straight and parallel. This is not quite the case.
In a typical card sliver, the fibre configuration
may be:
20% fibres
50% fibres
fibre ends)
15% fibres
fibre ends)
15% fibres

- straight fibres
- having trailing hooks (hooks at trailing
-having leading hooks (hooks at leading
-having hooks at both fibre ends

You may wonder how fibres get hooked up


during carding.
Fibre hooks are created in carding by the doffer
mainly. The doffer picks up fibres on the
cylinder by allowing the fibre leading ends to
hook around its teeth. The teeth on the fastmoving cylinder surface then comb and 'brush
forward' the other ends of the fibres. Therefore,
the fibres on the doffer surface have a majority
of trailing hooks. This configuration persists to
the carded sliver.

The generation of trailing hooks by the doffer is


graphically illustrated below.
1

Main cylinder

Doffer
3
4

1
2
3
4
5

a fibre app roaching the doffer


fibre leading end picked up by doffer
fibre trailing end combed forward by cyli nder
trailing hook fibre formed on differ
the trailing hook persists to carded sliver

Card
sliver

Fig. 1.12: Formation of trailing hooks in carding

Further processing is therefore necessary to


help straighten up these fibre hooks.

(4) Sliver evenness


It would be very difficult, if not impossible, to obtain
an even yarn from irregular slivers.
Uniform feed to the card is essential for the
uniformity of card sliver. Modern chute feed regulates
the amount of fibres fed to the card to improve
uniformity. In addition, some cards are fitted with an
autoleveller or autolevelling system to ensure
uniformity. An autoleveller is a system fitted to
carding (and drawing) machines to automatically
reduce the variation of the linear density of the
output material. This is achieved by monitoring the
linear density of the input or output material and, if
necessary, changing the machine speed to
compensate for any deviation from a pre-set value.

Two main types of autolevelling systems have


been used on the carding machines: the openloop (or feed forward) and closed-loop (feed
backward) autolevelling systems. The principles
of these two types are indicated in Figure 1.13.
Material
input

Ref.
Signal

Material
output

Measuring
unit

Control
unit

Regulator
unit

An open-loop (feed-forward) control


system

Material
input

Measuring
unit

Regulator
unit

Material
output

Control
unit

A closed-loop control system

Fig. 1.13: Open-loop and closed-loop autolevelling systems

Ref.
Signal

For the open-loop aultolevelling system, the


linear density (or thickness) of the input
material is measured by a measuring unit. The
result is compared with a set value or the
reference signal. If there is any deviation, the
control unit will direct the regulator unit to
change the process speed to maintain a regular
output. In this case, the measured signal is fed
forward from input to output, and this
autolevelling system is also known as feed
forward autolevelling system. This system is
often used to correct short-term variations in
linear density.

For the closed-loop autolevelling system, the


linear density (or thickness) of the output
material is monitored by a measuring unit. It is
then compared with the set value or the
reference signal. If the measured value deviates
from the set value, the control unit will direct the
regulator unit to change the process speed so
that the deviation can be reduced. Since the
measured signal is fed backwards from output to
input, this system is also known as feedback
autolevelling system. This system is suitable for
correcting medium to long-term variations in
linear density.

Figure 1.13a shows an open-loop autolevelling


system on a cotton card . In this example, the
thickness of the feed stock is measured at the
feed roller. The thickness signal is then fed
into a programmable controller which
processes the signal and works out the correct
draft required to adjust the weight of the
material being processed, and commands a
servo-motor to change the relevant roller
speed to maintain a uniform output.

Figure 1.13a Open-loop autolevelling on a card


(Lennox-Kerr 1983, p. 46)

A closed-loop autolevelling on a cotton card is


shown in figure 1.13b. In this case, the
thickness of the output sliver is monitored by a
pneumatic measuring unit. The signal is fed
back to the feed roller so that it can either slow
down or speed up to change the draft.

Figure 1.13b Closed-loop autolevelling on a


card (Klein 1987a, p54. Courtesy of Zellweger
Uster).

Drawing
Converting bales of fibres to a thin strand of
fibres or yarns requires enormous fibre
attenuation. Put simply, attenuation (drafting)
is to make input material longer and thinner. In
this sense, carding can also be regarded as a
fibre attenuation process. Drawing continues
the fibre attenuation, it also performs several
other functions.

Objectives

The drawing process aims at achieving the


following objectives:
Attenuate the card slivers
Reduce the fibre hooks and improve fibre
alignment
Blend and mix fibres
Reduce the irregularity of card slivers by
doubling
Drawing usually implies the actions of doubling
and drafting. Doubling is the combing of
several slivers and drafting is attenuation.

By now we already know that fibres in card


slivers are by no means straight and parallel,
and there are many hooked fibres, particularly
trailing hooks, in the card slivers. Many of these
hooks should be straightened as fibres slide past
each other in the drawing process. Slivers from
different cards vary evenness and other
properties, and should be blended to reduce the
irregularity. Cotton and synthetics are often
blended in drawing in sliver form. Finally, when
card slivers are combined (doubled), attenuation
is necessary to reduce the thickness of the
drawn sliver. Drawing plays a crucial role in the
final quality of yarn, and a good understanding
of the fundamentals of drawing is essential.

Basics

of roller draftin
The basic elements of a roller drafting unit is
shown in Figure 1.14.

Figure 1.14 Basic elements of drafting

There are two sets of rollers, the front rollers


(top and bottom) and the back rollers (top and
bottom). The front rollers are also known as
the delivery rollers while the back rollers the
feed rollers. For drafting to occur, the front
roller surface speed VFR is faster than the back
roller surface speed VBR. The bottom rollers
are the driving rollers, which drive the top
rollers by frictional contacts. The top rollers
are loaded by spring or dead weight to apply
pressure to the fibre material running
between the two sets of top and bottom
rollers.

There are several important definitions with


respect to roller drafting:
Input count (tex)
(a)Material Draft=Output count (tex)
(b)Mechanical Draft

Output surface speed


=Input surface speed

(c)Drafting Zone = the region between the front


and back rollers, where drafting occurs.
A drawframe usually has at least two drafting
zones, and the total draft of the drawframe is not
the addition, but multiplication, of the drafts in
separate zones. For instance, if a drawframe has
3 drafting zones with drafts of D1, D2 and D3
respectively, then the total draft of the drawframe
should be .

(d)Ratch = distance between the nip points of


the front
and back rollers.
(e)Doublings = number of slivers fed to the
drafting system for one output sliver.
The material draft and mechanical draft are not
always equal. The material draft is the real
draft. The ratch is also known as the ratch
length or ratch setting. It is set according to the
length of the longest fibres in order to prevent
these fibres from being stretched to break.

Perfect

Roller Drafting
Assuming all fibres are uniform in length and
diameter, and straight and parellel to sliver axis.
The position of each fibre in the sliver will be
fully described by position of its fibre leading
end (FLE) and its length. Existing textile
processes can not arrange these fibres in such a
perfect manner that a sliver would have the
same number of fibres in its cross sections
along its length (we call this sliver a perfect
sliver). The best sliver that can be expected
under optimum processing conditions is one in
which the fibre leading ends (FLEs) are
randomly distributed. A sliver with random FLEs
distribution is called an ideal sliver.

For an ideal sliver to remain ideal after drafting,


the random FLEs distribution should be
maintained. This requires that the drafting
should increase the distance between fibre
leading ends by a factor exactly equal to the
draft used, which can only be achieved through
perfect roller drafting.
Achieving perfect roller drafting requires
individual control and manipulation of single
fibres during the drafting process, which is not
possible with existing technology. Nevertheless,
near-perfect roller drafting may be obtainable if
the pressure distribution in the drafting zone is
ideal.

Ideal

Pressure Distribution
The ideal pressure distribution indicated in
Figure 1.15 has the following features:

Figure 1.15 Ideal and real pressure distributions in simple roller


drafting

(a) Back roller pressure is enlarged, in order to


keep fibres move at back roller speed.

(b) Front roller pressure is more concentrated


or narrow, so that fibres under the influence of
the front roller pressure can change speed
near the nip of the front rollers.
(c) Pressure from back rollers decreases
gradually, so fibres are not held back while
they get accelerated at the front roller nip.

With an ideal pressure distribution in the


drafting zone, near-perfect roller drafting can
be achieved.
The pressure distribution provided by a simple
roller drafting system is indicated by the
broken line in Figure 1.15. Such a pressure
distribution does not facilitate a near perfect
roller drafting and often some fibre control
devices are needed in the drafting zone. Fibre
control during drafting will be discussed later.

Real

Drafting
In a real roller drafting situation, both the
pressure distribution in the drafting zone and
the slivers themselves are not as ideal as the
ones mentioned above. Fibres in a real sliver are
of different lengths and diameters. They are
often not straight and parellel to sliver axis.
Fibre leading ends may not follow a perfect
random distribution. There are also grouping of
fibres in the sliver due to fibre entanglements
and frictional contacts. Besides, there may be
variatiosn in roller speed and slippage between
fibres and rollers etc. All these factors
contribute to the fact that perfect roller drafting
is not achieved in real drafting.

Because the ratch of a drafting zone is set


according to the length of the longest fibres,
many fibres are not gripped by either the front
roller or the back roller nip for some part of
their journey through the drafting zone. Figure
1.16 shows a simple roller drafting unit with
three typical fibres of different lengths in the
drafting zone. The three fibres a, b and c will
behave differently during drafting.

Figure 1.16 Movements of fibres in the drafting zone

Fibre a: The front rollers exsert more pressure


on this fibre than the back rollers. It has been
accelerated to the front roller speed and is a
'fast-moving' fibre, travelling at the speed of
front rollers.
Fibre b: This fibre is under heavy control of the
back rollers. It is a 'slow-moving' fibre.

Fibre c: This fibre is not under direct control of


either roller nip. Fibres which are not gripped by
either nip are known as 'floating fibres'. The
movement of a floating fibre will depend on its
surrounding fibres. If it is surrounded by some slow
moving fibres and some fast moving fibres, it may
be accelerated before its leading end reaches the
front roller nip if the sum of the frictional forces
between it and the fast moving fibres is greater
than the sum of the frictional forces between it
and the slow moving fibres. A floating fibre like
fibre c may also accelerate to immediate velocities
depending on the ratio of the frictional forces.
Floating fibres are obviously short fibres. Another
feature of this floating fibre is that once it gets
accelerated, it moves faster than other longer
ones because there is less restraining force on its
short training end.

The following observations can be made from


this simplified description of fibre movements.
(1)Speed change zone
Not all fibres change their speeds, or get
accelerated at the front roller nip as required
by perfect roller drafting, because fibre
lengths are different and fibres at the edges of
the drafting zone are unpredictable (edge
fibres can not be controlled effectively). There
is a speed change zone near the front roller
nip. This speed change zone will be more
localised towards the front roller nip if fibre
length is more uniform.

(2)Drafting wave due to floating fibres


Since the movement of a floating fibre depends
on its contacts with neighbouring fibres. Some
floating fibres may be pulled forward out of turn
by the neighbouring fibres that have already been
accelerated. Floating fibres accelerating out of
turn can cause adjacent fibres to also accelerate,
creating a thick place. The thick place is then
drawn forward by the front roller nip, leaving a
thin place behind. This process repeats to produce
alternatively thick and thin places in the drafted
fibre assembly. This is a practically periodic
irregularity, and is widely known as a 'drafting
wave', because it is caused by floating fibres
during drafting. The wavelength of a drafting
wave is about 2.5 times the mean fibre length.

(3)Drafting wave due to sliver elasticity


Due to crimp and poor orientation and entanglement of
fibres, slivers are elastic over small strains. They can
stretch under the drafting force. As the fast moving fibres
are withdrawn from the bulk of the sliver, their frictional
contacts with the sliver create a withdrawal force, which
then extends the sliver. This extension will cause the fibre
leading ends to reach the front roller nip and accelerate
ahead of their turn, and so produce a thick place. A thick
place means more fast moving fibres and higher withdrawal
forces and more extension, causing more fibres to
accelerate ahead of their turn until the sliver reaches the
limit of its extension. At this level of extension, the number
of fibres reaching the front roller nip settles back down to
the normal level, and so the number of fast moving fibres
(& withdrawal force) decreases. The reduction in withdrawal
force causes the sliver to retract, so fibres reach the front
roller nip later than expected, producing a thin place. Again,
this process repeats and drafting wave is produced. The
wave length of this drafting wave is also about 2.5 times
the mean fibre length.

(4)Periodic mass variations due to mechanical


faults
If there are mechanical faults in the drafting
system, such as eccentric drafting rollers,
periodic mass variation will result in the
drafted material. This phenomenon has been
discussed in the module on yarn evenness.
In summary, real drafting may deviate from
perfect roller drafting because of either
material related or machine related factors. In
practice, fibre control during drafting is
necessary to reduce this deviation.

Fibre

Control in Roller Drafting


The main aim of fibre control is to keep the
floating fibres at the speed of back rollers until
they reach the front roller nip (i.e. to prevent
fibres being accelerated out of turn), while still
allowing long fibres to be drafted.

Figure 1.17 Examples of fibre control for the Short Staple


Drawframes

Different yarn manufacture systems, and


different process in the same system, often
apply different control device in drafting. Two
examples of fibre control in short staple drawing
machines (drawframes) are shown in Figure
1.17. The control roller and pressure bar force
the fibre assembly (in the drafting zone) to take
a curved path, thus increasing the pressure on
fibres at the control roller or pressure bar. The
increased pressure helps to control fibre
movement during drafting.

Doubling

in Drawing
As mentioned in the beginning, drawing often
implies the actions of doubling and drafting. We
have already discussed drafting at length.
Doubling simply means combining several slivers
together as the input to a drawframe. According
to the law of doubling discussed in the module on
yarn evenness, if n slivers are doubled together,
the CV of the doubled material
1
will be reduced by
n
a factor of
or
_______
CV after doubling
______

CVbefore doubling
n

whereCV
is the average CV of the
individual slivers before doubling
before doubling

Usually 8 slivers are doubled up to give one


output sliver, as indicated in Figure 1.18. The
linear density of the output sliver is determined
by the amount of total draft applied in the draft
Draft zone
zone.
Measuring
unit

Break
draf t

Creel with feed rolers (variable speed)

Servomotor
(for speed
change)

Cans of input slivers

Signal
processing
unit

Autolevelling System

Figure 1.18 A drawframe with 8 doublings and an autolevelling unit

Main
draf t

Autolevelling

in Drawing
Fibre control and doubling are necessary in
drawing to improve the quality, particularly
evenness, of drawn slivers. As in carding,
autolevelling is often used in drawing to further
improve the evenness of drawn materials. The
principle of autolevelling has been discussed in
the carding section. An example of autolevelling
in drawing is shown in figure 1.18. This is an
open-loop or feed forward autolevelling system.
The input material is measured for linear density
or thickness by a measuring unit, the signal is
processed and compared with set value by the
signal processing unit. If deviation exists, then
the servomotor is instructed to change the speed
of the drafting rollers to adjust the draft in other
to reduce the irregularity of the output material.

Fibre

Straightening in Drawing
We already know that most fibres in card
slivers are hooked fibres, and one of the key
objectives of drawing is to straighten out these
hooked fibres.
Consider a trailing hook (T) and a leading hook
(L) in drawing as shown in Figure 1.19.
Back
rollers

Front
rollers

Back
rollers

Trailing hook i n the drafti ng zone


(straightens out easily)

Front
rollers

Leading hook in the drafti ng zone


(does not straighten out easily)

Figure 1.19 Fibre straightening during drafting

For the trailing hook, it will travel initially at the


speed of the back drafting rollers. Soon its leading
end, embedded in fast-moving fibres under the
influence of the front drafting rollers, will travel with
the fast-moving fibres at the front roller speed.
Since the hooked end of the fibre is still embedded
in a relatively thick body of slow-moving fibres
controlled by the back rollers, the difference in
speed between the leading end and trailing
(hooked) end will straighten out the hook. For the
fibre with leading hook (L), the hook can get caught
easily by the fast moving fibres and travel at the
front roller speed, while the unhooked trailing end
offers little resistance to its acceleration. As a
result, the leading hook (L) is likely to persist into
the output material. From this brief discussion, it is
clear that one passage through a drawframe only
effectively removes trailing hooks.

In a card sliver, the majority of fibre hooks are


trailing hooks. But as the card sliver is deposited
into a can and gets taken out to feed a drawframe,
it follows a first-in-last-out principle and a reversal
of hook direction occurs. This is known as natural
reversal of fibre direction. Because of this natural
reversal, most fibres (in the card slivers) entering
the first drawframe have leading hooks, which do
not get effectively straightened out as we have just
discussed. In addition, a short staple combing
machine (the comber) straightens out leading
hooks effectively (which is different from a worsted
comb for long staple fibres), and trailing hooks
must be presented to a ring spinning machine (the
drafting in ring spinning does not straighten out
leading hooks).

For these reasons, there must be an even


number of passages between the short staple
carding and combing machines, and an odd
number between the short staple carding and
ring spinning machines. You can see this
arrangement from Figure 1.1.
Figure 1.1 indicates that after two drawing
passages, the sliver can go directly to rotor
spinning to produce a carded rotor spun yarn.
However, if a high quality ring spun cotton yarn
is required, the sliver should go through a
combing stage, followed by further drawing,
roving and finally the ring spinning process.
Combing is discussed next.

Combing

Introduction

Combing is a key process that makes the


difference between an ordinary yarn and a quality
yarn. It enables the ultimate yarn to be smoother,
finer, stronger, and more uniform than otherwise
would be possible, at a cost of course.
The basic objectives of combing are:
(a) Removal of a pre-determined amount of short
fibres
(b) Removal of neps and impurities
(c) Straightening of the retained long fibres

The continuous assembly of long and


parallel fibres delivered by the combing
process is called a comb sliver. Just as
long and well-aligned fibre polymers
(molecules) make strong fibres, long and
straight fibres in the comb sliver will
make strong and smooth yarns.
The materials rejected in the combing
process is called noil. Noil contains short
fibres, neps and impurities. The amount
of noil produced may be expressed as
either percentage noil or tear ratio:

weight of noil
Percentage noil
100
weight of (noil comb sliver )

weight of comb sliver


Tear ratio
:1
weight of noil
The relationship between the two expressions
is given below:
100
Percentage noil
(tear 1)

or

100
1
Tear
percentage noil

For example, a 10% noil is equivalent to a tear


ratio of 9 : 1, and a 16% noil is equivalent to a
tear ratio of 5.25 : 1.

The amount of noil produced is of significant


importance. A higher noil means longer fibres in
the comb sliver, but less comb sliver will be
produced.
Yarns made from the combed sliver are called
combed yarns. Without combing, a carded yarn
would be produced. Combed yarns, consisting
of longer and more parallel fibres, are of better
quality and command a higher price than
carded yarns or yarns produced without the
combing stage. A brief comparison of combed
and carded cotton yarns is given in Table 1.2
below. `

The

comb

A combing machine is usually referred to as a


comber, or simply a comb. The short-staple
spinning mill uses only the rectilinear comb with
swinging nippers and stationary detaching rollers,
as originally conceived in 1845 by J. Heilman in
Alsace and further developed in 1902 by the
Englishman Nasmith and in 1948 by the Whitin
company.
The most common machine layouts used in
practice comprise single-sided machines with
eight heads. Double-sided machines with six-plussix head are also manufactured by the Platt Saco
Lowell company.

A typical comb head is sketched in Figure 1.20.


Top comb
Detaching
rolls

Input
sliver
lap

Nipper
jaws
Feed
rolls

Combed sliver

Cylinder comb
(Circ ular comb)
Figure 1.20 Sketch of a rectilinear cotton comb

The sliver lap, or the thick sheet of fibres formed


on a sliver lapper before combing, is fed to the
comb in an intermittent fashion. In each operating
cycle, the lap is advanced a short distance (4 to
6.5 mm) and then gripped (by the nippers) so that
a fringe of fibres is presented to the toothed
section of a cylinder comb. The fibre fringe is then
combed by the teeth on the cylinder comb and in
the process, short fibres, neps and impurities are
removed from the fringe. The short fibres, neps
and impurities are collectively called noils. When
the un-toothed portion of the comber roll comes
into contact with the fringe, the nippers open and
swing towards the detaching rollers to allow the
fringe to be drawn off by the detaching (or drawoff) rolls.

The closest distance between the nipping points


of the nippers and the detaching rollers is called
by several names gauge setting, detachment
setting, noil setting. This setting is the most
important setting on a comb. It has the largest
impact on the percentage noil and the mean fibre
length of the comb sliver. During the drawing off
process, the fibre fringe is pressed into the
needles of the top comb, so that the portion of
the fringe not combed by the cylinder comb can
now be combed by the top comb. The neps and
impurities will not be able to pass through the
closely pinned top comb and are removed by the
cylinder comb in the next combing cycle.

The detaching rolls bring the combed fringe to


the tail end of the previously combed material
to make a joint or piecing. The combed fibres,
from many combing heads, are then brought
together and consolidated into a sliver and
coiled into a sliver can.
The comb operates intermittently in cycles. Its
speed is described in terms of cycles per minute
or nips per minute. Modern cotton combs run in
excess of 350 nips/minute

Sequence

of operations
The sequence of operations of a cotton comb is
described below
(a) The feed rollers S move the sheet W 4 6.5
mm forward, while the nippers ZO/ZU are held
open (feed).
(b) The upper nipper plate ZO is lowered onto the
cushion plate ZU so that the fibres are clamped
between them (nipping).
(c) The combing segment (K), mounted on
rotating cylinder (Z), sweeps its needles or sawteeth through the fibre fringe (B) and carries away
anything not held by the nippers (rotary combing).
(d) The nippers open again and move towards the
detaching
rollers A (nippers forward).
Figure 1.21 Sequence of operations of a cotton comb (Klein, 1987b, p2-4 )

(e)Meanwhile, the detaching rollers A have


returned part of the previously drawn off stock
(web V) by means of a reverse rotation, so
that the web protrudes from the back of the
detaching device (web return).
(f)In the course of the forward movement of
the nippers, the projecting fibre fringe B is
placed upon the returned web V (piecing).
(g)The detaching rollers begin to rotate in the
forward direction again and draw the clamped
fibres out of the sheet W held fast by the feed
rollers (detaching).

(h)Before the start of the detaching operation,


the top comb F has thrust its single row of
needles into the fibre fringe. As the fibres are
pulled through the needles of the top comb
during detaching, the trailing part of the fringe
is combed, thus making up for the inability of
the cylinder comb to reach this part of the
fringe (combing by the top comb).
(i)As the nipper assembly is retracted, the
nippers open for the next feeding step. The top
comb is withdrawn. A new combing cycle
begins.

Fibre

selection in combing
The theory of combing deals with the key issue of
fibre selection in the combing process, i.e. what
goes into the noil and what goes into the comb
sliver. The percentage noil is largely a function of
the detachment setting and the feed distance per
combing cycle. It is worth pointing out here that
there are two types of feeding arrangements
concurrent feed and counter-feed. With
concurrent feed, the fibre sheet is fed forward
into the nippers while the nippers are swinging
towards the detaching rollers. With counter-feed,
the fibre sheet is fed forward during the return of
the nippers. The type of feeding also affects the
percentage noil in combing.

According to Charles Gegauffs noil theory, the


percentage noil (N%) are related to the
detachment setting (D), feed distance (F), and
the longest fibre length (L), according to the
formula below:
2
DF
2
N %
L

DF
2
N %

100

( For counter feed )

100

( For cocurrent feed )

It should be noted that these formulas are used


to reflect the relationship between percentage
noil and important comb settings, not to
calculate the actual percentage noils. The
implication of this relationship on the quality of
comb sliver is discussed in the following section.

Quality

issues in combing
The theory of combing or noil theory discussed
in the previous section provides a good starting
point on the quality issues in combing.
Research at Rieter has shown that the
percentage noil has a major impact on the
important quality attributes of the resultant
Improvement i n
yarns, yarn
asquality
depicted
in Figure 1.22.
(%)
100

Yarn imperfections
50
Yarn evenness
Yarn strength
5%

10%

15%

20%

Percentage
noil

Figure 1.22 The effect of percentage noil on yarn quality attributes

In a normal combing process for cotton, a


percentage noil between 10 to 20% is expected.
Combing with a noil percentage below 10% is
often referred to as upgrading combing, while
combing with a noil percentage about 20% is
known as super combing. Super combing is only
used when superfine combed yarns are to be
produced.
Now that we know the importance of percentage
noil, what are the major factors that affect
percentage noil in combing then? We will need to
refer to the noil theory to answer this question.

(1)The detachment setting (D)

As mentioned before, this setting is the closest


distance between the bite of the nippers and the nip
line of the detaching rollers. According to the noil
theory, the percentage noil increases as the
detachment setting increases. At larger detachment
setting, more fibres are removed into the noil and
the average fibre length in the comb sliver is longer.
Because of the increased fibre loss with increase in
detachment setting, the cost of production will be
higher. In practice, spinners need to find the
optimum detachment setting based on a balance of
quality and cost. The detachment setting on a
cotton comb normally lies in the range 15 to 25
mm.

(2)The feed

According to the noil theory, both the type of


feed and the feed distance affect the percentage
noil. For the same feed distance, counter-feed
results in higher noil percentage than concurrent
feed. In other words, there are more short fibres
in the comb sliver with concurrent feed than with
counter-feed. In upgrading combing where the
quality requirement is low and production rate
needs to be high, concurrent feed is often used.
On the other hand, if the quality requirement is
very rigorous, counter-feed should be used.
Some modern combs allow selection of the feed
type according to needs.

The feed distance affects the noil, the quality and


the production rate of combing. According to the
noil theory, noil increases with feed distance for
counter-feed, and decreases with feed distance
for concurrent feed. Increase in the feed distance
will increase production rate, because more
fibres go into the comb sliver at a larger feed
distance. But the cleanliness of the combed web,
i.e. its freedom from impurities and neps, will
deteriorate at a higher feed distance. Therefore,
a lower feed distance needs to be used for
higher quality requirements.

(3)Effect of fibre hooks

The cotton comb deals with relatively short fibres,


compared with worsted comb. The hooks on
short fibres are small. As a result, leading hooks
can be removed easily by the cylinder comb,
with little damage or breakage to the hooked
fibres. An exception is when that both limbs of a
leading fibre hook are held in the nippers while
the protruding loop is still long enough to be
engaged by the teeth on the cylinder comb. This
event is much rarer in short staple combing than
in long staple combing.

Trailing hooks can cause serious problems in short


staple combing. Some trailing hooks may persist
into the comb sliver or cause fibre breakage during
detaching, particularly if concurrent feed is used.
Consider a fibre with a trailing hook just lying in
the bite of the nippers. With concurrent feed, the
trailing hook will be pushed forward out of the
nippers as the nippers swing towards the
detaching rollers. The subsequent detaching action
may drag this fibre (and its trailing hook) into the
comb sliver, particularly if this fibre is near the
bottom of the fibre sheet where the top comb has
not yet penetrated. If this fibres trailing hook is
engaged by the top comb, fibre breakage is likely
to occur because of the high friction between the
hook and the top comb needle. With the counterfeed, no feeding occurs during the forward
movement of the nippers.

During detaching, the fibre fringe gets pressed


into the top comb in front of the nippers, and the
trailing hook in the fibre concerned is likely to be
combed by the top comb. In such case, the fibre
could be combed straight unharmed and
dragged into the comb sliver, or the fibre may be
broken because of the friction between the fibre
and the needles of the top comb. The fibre
breakage will generate two short fibres, one
proceeding to the comb sliver, and the other
blocked by the top comb but removed as noil in
the next combing cycle. This has two major
undesirable consequences short fibres in the
comb sliver and increased combing waste.

The trailing hook, or other forms of fibre


disorientation, may also carry its neighbouring
short fibres forward to be detached by the
detaching rollers, again increasing the number
of short fibres in the comb sliver. Large trailing
hooks will also reduce the effective length of
even a relatively long fibre. The comb is likely
to treat this fibre just like an ordinary short
fibre and remove it into the noil. In other words,
trailing hooks (or other forms of fibre
disorientation) can also increase the number of
long fibres in the noil, which of course is
undesirable.

From this discussion, it is clear that good fibre


alignment in the feed lap is essential for quality
combing. Trailing hooks are particularly
troublesome. This requires good lap
preparation before combing. It also requires
that an even (i.e. two) number of machines
between the card and the comb, as indicated in
Figure 1.1. This ensures leading hooks mainly
are fed into the cotton comb.

(4) Comb overlap effect in piecing


At the end of each combing cycle, a small tuft of
combed fibres is detached. This tuft is then
partially overlapped on the previously detached
tufts, like shingles on a roof or roofing tiles. Such
pieced structure, or the comb overlap effect, is an
inherent source of faults in the operation of
rectilinear combs. Because of this, the combed
sliver exhibits periodic mass variations along its
length, which can be revealed using the
spectrogram.
The combing process is normally followed by one
or two passages of drawing before spinning. The
drawing subsequent to combing will reduce the
comb overlap effect. If a combed ring spun yarn is
to be produced, a roving process is needed before
the ring spinning process.

Roving
Introduction

A roving is a fine strand (slubbing) intended to


be fed into the ring spinning machines (ring
frames) for making yarns. Rotor spinning
machine and other new spinning systems use
slivers as feed materials. But conventional ring
frames still use rovings as the feed material. A
roving is much thinner than a sliver, but thicker
than a yarn.
The main objective of the roving machine is to
further attenuate the drawn sliver (to make it
longer and thinner) and get it ready for
spinning.

The drawframe has already produced a sliver that


is clean, and consists of more or less parallel
fibres. Such a sliver satisfies the essential
requirements for yarn production. The question is
why is there a need for the roving process and why
cant we feed slivers to conventional ring frames?
There are two major reasons for this need. First, a
very high draft, in the order of 300 to 500, is
required to bring the thickness of a sliver into the
thickness of a yarn. Conventional ring frames can
not cope with such a high draft. Second, the
drawframe slivers are deposited in bulky sliver
cans, which are difficult to transport and present to
the ring frames as feed material. The much smaller
roving packages are better suited for the purpose.

Roving

frame

The commonly used roving machine for cotton


is a flyer frame (or speed frame) as shown in
Drafting Unit
Figure 1.23.Back
rolle rs
Front
rolle rs
Vd
Double
aprons

Flyer
Flyer
leg

Sliver can
(carded or
combed sli ver)

Dr
Bobbin
drive
Flyer
drive
Figure 1.23 Diagram of a roving frame

nb
nf

Roving
bobbin

Presser
arm

There are three basic steps in the operation


of the roving frame drafting, twisting, and
winding. These basic steps are exactly the
same as the basic steps required in spinning.
Consequently, an understanding of the
roving process will help us understand the
spinning process to be discussed in the next
module.

The input to this roving frame is a drawn sliver


(either carded or combed) from the last drawing
process. The sliver is drafted by a roller drafting
unit. Between the front and back rollers (the
drafting zone), the fibres pass between the double
aprons, which control the fibre movement during
drafting. The front nose of the double aprons is set
close to the front roller nip for good drafting
performance. You may recall the concept of
perfect roller drafting, which requires that fibres in
the drafting zone travel at the speed of back
rollers until the fibre leading ends reach the front
roller nip. The double aprons travel at about the
same speed as the back rollers, and they control
the fibres until they reach the front roller nip.

A small amount of twist (30 to 65 turns per


meter) is inserted into the drafted fibre strand
via the rotation of the flyer. The bobbin (on a
spindle) is driven at a speed different to that of
the flyer. The different in bobbin and flyer speeds
allows the slightly twisted fibre strand or roving
to be wound on the bobbin. If the rotations of the
bobbin and the flyer are synchronised, the roving
will not be wound up onto the bobbin. The flyer
arm through which the roving passes helps to
support the relatively weak roving due to its low
twist level.

In addition, a presser arm is attached to the


lower end of the hollow flyer leg (through which
the roving runs). This presser arm guides the
roving from the exit of the flyer leg to the
roving package. The roving is wrapped two or
three times around the presser arm. The friction
between the roving and the presser arm will
increase the roving tension at the winding on
point. This will give a compact roving package.
A compact package has more roving and is
more stable as well.

On the roving bobbin, each coil of roving material


is arranged very closely and almost parallel to one
another (parallel wind) so that as much material as
possible is taken up in the package. For this
purpose, the bobbin rail (not shown in the
diagram) with the package on it moves up and
down steadily. The build-up of roving package
leads to an increase in the wound length of roving
per coil. The speed of the bobbin rail movement is
reduced by a small amount after each completed
layer. With the increase in package diameter (Dr),
the bobbin rotation rate is also changed to
maintain a constant difference between the
surface speeds of the package and the flyer. This
speed difference is the winding on speed and
should be the same as the speed at which the fibre
strand is delivered by the front drafting rollers.

The working principle of the flyer roving frame can


be summarised as below:
roller drafting, delivers fibre strand at a
constant speed Vd
flyer rotate at nf (constant) to twist the strand

bobbin

nf
Twist level =
Vd

rotates at nb (different to nf) to wind on


the roving
- either bobbin lead or flyer lead, as long as
there is a difference in rotational speed
- winding on speed
Dr ( ndiameter
V W =roving
f - nb )
( Dr=
on bobbin)

Dr varies as the roving package builds up,

change nb to match VW with Vd

fibre strand is supported by a flyer arm (no


ballooning, best for thick weak strand)
- flyer speed is limited by the mechanical
design.

Quality

issues in roving
Since ring spun yarns are produced directly from
rovings, the quality of the roving is very
important. The roving process is essentially a
drafting process (not a drawing process because
there is no doubling). In fibre drafting, fibre
control is important. Good condition of the
double aprons, the right ratch setting (distance
between the front and back drafting rollers) are
important in ensuing good fibre control.

Rovings should be routinely sampled and tested


for evenness. Particular attention should be paid
to the spectrogram to see if any drafting wave or
periodic mass variation exists in the rovings.

The small amount of twist inserted in the roving


is necessary to ensure trouble-free transport of
the roving package, smooth unwinding of the
roving at the ring frame, and to prevent
accidental drafting of the roving during roving
winding. This twist should be as small as
practically possible for two reasons. First, if the
flyer rotation speed is fixed, a higher twist level
means lower delivery speed or lower production
rate. Second, high twist in the roving may cause
problem in drafting of the roving at the ring
frame, because fibres may not be able to slide
past one another freely. The machine
manufacturers will recommend the right level of
twist for different fibre materials used.

Review questions
1.The key processing stages for cotton include
opening and blending
carding
drawing
combing
Describe the objectives and principle of each of
these processes, using sketches if necessary and
use about 200 words for each process.
2.Four bales of cotton, of 500 pounds each, are to
be mixed together for the blow-room process. If
the cotton fineness in these bales is, 3.8, 4.1,
4.4, and 3.3 micronaire (mic., g/in.)
respectively, what would be the theoretical
fineness of the cotton in the mix? You need to
show your calculations.

3.On a draw-frame or drawing machine, why is it


often necessary to adjust the ratch setting
according to the fibres to be processed?
4.With reference to the two-zone drawframe in
figure 1.18 and assuming a break draft of 1.5
and a main draft of 4. If the average count of the
eight input slivers is 12 ktex, what is the count
of the single output sliver? If the delivery (front)
roller speed is 400 m/min, what would be the
approximate speed of the feed (back) rollers?
5.With reference to fibre hooks, explain why two
passages of drawing are necessary between
cotton carding and cotton combing. You should
consider the following points:
Hook generation in carding
Hook removal during drafting
Effect of hook direction on combing

Worsted Processing
Introduction
Long staple fibres are fibres longer than about
2 inches. Fibres such as merino wool, mohair
and alpaca fibres are typical long staple fibres.
Synthetic staples of similar length are long
staple synthetic fibres. Long staple fibres are
processed on the worsted processing system
mainly, to make worsted yarns.
This topic focuses on the principle and quality
of worsted processing of wool fibres.

Objectives
At the end of this topic you should be able to:
Understand the principles and functions of
worsted processing from raw wool to top
Appreciate the effect of raw wool quality on
the quality of tops
Know the applications of the TEAM formula in
top-making
Appreciate how fibre processing affects fibre
properties

Process overview
The worsted industry is more fragmented than
the cotton or short staple industry. The
processing from greasy wool to worsted yarn is
often carried out separately in different mills early stage processing (ESP) mill, top-making
mill and spinning mill. The early stage
processor cleans the greasy wool. The topmaker buys the clean wool from the early
stage processor and converts the wool into a
top (a sliver ready for worsted drawing and
spinning). The spinner sources the top from the
topmaker and processes it into worsted yarns.

Some mills engage in both early stage


processing and topmaking, others are
vertically integrated and do the whole
processing from greasy wool to yarn. Early
stage processing and top-making are also used
synonymously to refer to all the greasy wool to
top processing stages. Worsted processing
utilises relatively fine (< 27 micron) and long
(> 45 mm) virgin wool and other long staple
fibres. A typical sequence of worsted
processing of wool is given in Figure 2.1.

Raw Wool

Early stage processing

Opening & Blending


Scouring & Drying

SCOURED
WOOL

Worsted Carding
Intermediate Gillings
(usually 3)
Top-making
Combing
Finishing Gillings
(usually 2)
WORSTED
TOP
Spinni ng

Drawings
(2 to 5)
Spinning

Figure 2.1 A typical worsted processing sequence for wool

WORSTED
SINGLES YARN

Before wool processing can start, we need to first


of all source the raw wool. Sourcing the right raw
wool is vital and requires a good understanding
between the raw wool, wool top, worsted yarn
and fabric. Even though raw wool is the starting
point for wool processing, the decision to source
a certain type of raw wool is governed by the
intended end use of the fibre. Fabric
requirements govern yarn requirements; yarn
requirements govern top requirements, which in
turn govern the raw wool specifications. This
relationship is represented in Figure 2.2.

Fabric Requirements
Yarn Requirements
Top Requirements
Raw Wool Specifications
Figure 2.2 End use governs raw wool purchase

Traditionally, translating the end use


requirements to raw wool requirements calls for
considerable skills and experience. However,
tools have been developed in recent years to
facilitate this translation in the worsted industry.
The notable examples of such tools are TEAM
formulae and Yarnspec software developed by
the CSIRO Textile and Fibre Technology (formerly
known as CSIRO Division of Wool Technology).

The TEAM formulae and applications


Introduction

The bulk of Australian wool clip is now


scientifically sampled and the samples are
objectively measured before sale for a range of
properties. When objective measurements of
greasy wool started in Australia in the early 70s,
only Yield (i.e. the amount of clean fibre that can
be produced from the greasy wool), Vegetable
Matters (VM) and Mean Fibre Diameter (in micron)
were measured. While these measurements are
very important for raw wool sales, they can not be
used to adequately predict the processing
performance of the measured wool. In the late
70s, technology for additional measurements of
raw wool became available.

These include measurements of staple length,


staple strength and position of break using the
ATLAS instrument developed in CSIRO. The use
of these measurement results has become an
indispensable tool for modern wool processing
mills worldwide. There are three major
advantages associated with the use of
objectively measured and specified wool:
Maintaining control of the quality of wool
delivered to the mill,
Monitoring processing performance and
quality management in the mill, and,
Optimising wool blend selection and
minimising raw wool cost by taking
advantage of the wool price differentials.

To make use of these advantages, the wool


processing mills should have some knowledge
of the TEAM prediction formulae.
TEAM stands for Trials Evaluating Additional
Measurements. These trials were conducted
between 1981 and 1988 by the former
Australian Wool Corporation, the Australian
Wool Testing Authority Ltd (AWTA Ltd) and the
CSIRO Division of Wool Technology (now known
as CSIRO Textile and Fibre Technology). Over 20
mills in 12 different countries were involved in
the trails.

As a result of these trials, a series of prediction


formulae (known as the TEAM formulae) were
released, which can be used to predict the
processing performance of fully measured wool
in terms of the following:
- Hauteur
- CV of Hauteur
- Barbe
- Romaine

Hauteur and Barbe are two different measures


of the average fibre length. Their calculations
may be explained with the simple case of two
fibres indicated in figure 2.3. The average
length of these two fibres may be different,
depending on how we calculate it.
A1

L1

Fibre 1

A2

L2

Fibre 2

Figure 2.3 A simplified example for calculating the mean


fibre length

Numerical mean fibre length : L N

L1 L 2
2

Cross sec tion biased mean fibre length ( Hauteur ) : H

A1 L1 A2 L 2
A1 A2

Weight biased mean fibre length ( Barbe) :


( A1 L1) L1 ( A2 L 2) L 2
A1 L12 A2 L 2 2
B

( A1 L1) ( A2 L 2)
A1 L1 A2 L 2
where is fibre density.

For example, if L1 = 60 mm, L2 = 80 mm, A1


= 300 m2, A2 = 400 m2 , then the above
calculations will give: Ln = 70 mm, H = 71.4
mm, B = 72.8 mm.

The numerical mean length is the true average


length of the fibres. But it is difficult to obtain in
practice. Hauteur length has been the most
commonly used in the worsted industry. It can be
easily obtained from commercial instruments such
as the Almeter and WIRA fibre length meter.
Romaine is the term used in the worsted industry
to describe the amount of noil produced during
the combing process, expressed as a percentage
of the total (noil and combed sliver).
These prediction formulae can be applied to any
combing wool with full objective measurements,
whether that be individual lots or entire
consignments, of wholly measured Australian
wool. Computer softwares have also been
developed to assist with the calculations of the
predicted results.

The general formula for Hauteur


As discussed above, Hauteur is the (cross
section biased) mean length of fibres in the
wool top (a top is a sliver of parallel fibres,
obtained after the processes of scouring,
carding and combing of greasy wool). It is
normally measured by using an instrument
called Almeter. Hauteur value has a significant
effect on top price and the subsequent yarn
properties. Hauteur has traditionally been
estimated subjectively by wool buyers and
processors.

However, the TEAM project has demonstrated


that the theoretical Hauteur value may be
predicted from measurements on raw wool
according to the following general formula:
H = 0.52 L + 0.47 S + 0.95 D - 0.19 M* - 0.45 V 3.5
where H = Predicted Hauteur (mm)
L = Mean Staple Length (mm)
S = Mean Staple Strength (N/ktex)
D = Mean Fibre Diameter (micron)
M* = Adjusted Percentage of Middle Breaks (%)
3.5 = A Constant

NB 1: The value of M* is determined from the percentage


of staples which broke in the middle portion (PM) as
displayed on an AWTA Test Certificate for Staple Length
and Strength as follows:
for PM values between 0 to 45% then
for PM values between 46 - 100% then

M* = 45
M* = PM

For example, if a consignment has objective


measurements of Mean Staple Length (90mm), Mean
Staple Strength (40 N/ktex), Mean Fibre Diameter (21
microns), adjusted Percentage of Middle Breaks (50%),
and Mean VM Base (2.0%), then the Hauteur
predicted from the general formula is:
H = 0.52 * 90 + 0.47 * 40 + 0.95 * 21 - 0.19 * 50 - 0.45
* 2.0 - 3.5 = 71.7 (mm)

NB 2:
The TEAM formula for predicting
Hauteur is based on the processing results of
545 consignments combed at 20 different mills
worldwide. Raw wool test data for the
consignments fell into the following ranges:
Mean Staple Length
59 - 123 mm
Mean Staple Strength 23 - 60 N/ktex
Mean Fibre Diameter 17 - 31 microns
VM Base
0.1 - 10.0%
Application of the above formula to data which
falls outside these ranges should be treated
with caution.

The mill adjustment factor( )


All wool processing mills are different. They
may use different raw materials and have
different processing machinery installed.
While the general formula for Hauteur
gives a satisfactory general prediction for
mills, as adjustment to reflect performance
for individual mills are often needed,
particularly if a mill produces tops that are
consistently longer or shorter than the
length predicted by the general formula.
The TEAM prediction represents the world
average of processing performance.

The constant (-3.5) in the TEAM formula can be


adjusted when a mill has monitored the processing
performance of 10 to 15 consignments. The mean
value of the difference between the Hauteur
predicted by the TEAM formula and the actual
Hauteur achieved can be calculated. This mean
value then becomes the "Mill Correction Factor"
and is added or subtracted from the constant in the
TEAM formula.
If changes are made to the mill processing
conditions, such as the installation of new
machinery, differences will be expected and
alterations to the Mill Correction Factor will be
necessary. It is also worth noting that the Mill
Adjustment Factors are commercially sensitive,
individual traders are unlikely to be advised of
them by the commissioning combers.

An example of calculating a Mill Correction


Factor is given in Table 2.1 below:

Processors can also calculate their own tailormade 'mill specific prediction formula' using a
regression equation based on the greasy wool
characteristics found to be the most important
for the mill. Any standard computer package
containing regression analysis can do this.
However, it must be stressed that new
formulae should only be developed on large
databases. When a small database is available,
the above approach for calculating the Mill
Correction Factor should suffice.

TEAM formulae for the predictions of CV of


Hauteur, Romaine & Barbe
Besides the formula for the prediction of Hauteur,
the TEAM projects also developed prediction
formulae for the Coefficient of Variation of
Hauteur, the Romaine and the Barbe.
The CV of Hauteur (CVH) is an important feature of
the top length distribution, since it may affect the
subsequent drafting and spinning performances.
Romaine is the term used to describe the quantity
of noil expressed as a percentage in relation to the
combined quantity of top and noil. It reflects the
combing efficiency. Barbe is a weight biased
measurement of fibre length in the top and is not
as widely used as Hauteur.

The three prediction formulae are:


CVH = 0.12 L - 0.41 S - 0.35 D + 0.20 M* + 49.3
Romaine = (-0.11) L - 0.14 S - 0.35 D + 0.94 V +
27.7
Barbe (B) = 0.73 L + 0.32 S + 0.96 D - 0.51 V 0.086 M* - 5.3
The symbols L, S, D, M* and V are as previously
defined for TEAM Hauteur Formula. The above
three formulae can also be adjusted to any
particular mill in the same way as the TEAM
Hauteur Formula, by adding or subtracting a mill
adjustment factor to the constants, when
appropriate.

It is also worth noting that the TEAM formulae


have been developed for general use and
therefore do not take into account variations in
processing performance between and within
specific mills. If the predictions are outside the
following ranges, care must be taken in
interpreting the prediction results:
Hauteur:
Less than 55 mm and greater
than 80 mm
CVHa:
Less than 40% and greater than 55%
Romaine* Less than 3% and greater than 12%

In summary, the TEAM project has demonstrated the


importance of additional measurements on raw wool.
Without these additional staple measurements, the
prediction of Hauteur and Romaine are not as
accurate and it is impossible to predict CVHa or other
top length distribution specifications important to the
spinner. Many topmakers and spinners worldwide are
willing to pay premiums to have their raw wool
measured for staple length, strength and position of
break. Wool users, whether mills, exporters, or
traders, should maintain a record of the greasy wool
measurement data and the difference between the
processing characteristics predicted by the TEAM
formulae and those actually achieved for each
processing consignment. By doing this, a 'mill
specific' database can be built and used to improve
the predictions for that mill.

Applications of TEAM formulae


The TEAM formulae have a number of
applications. The major ones include assisting
mill quality control and raw wool specification.
When a wool processing mill uses fully measured
Australian wool, the mill performance can be
monitored with the help of TEAM formulae. For
instance, the differences between actual and
predicted Hauteur of the tops produced in a mill
can be plotted on a time-series graph. By setting
boundary limits for the size of the deviations, a
control chart can be established to indicate
whether the mill's topmaking process is 'in
control'.

Another impartant application of TEAM is in


assisting with raw wool specification. As
mentioned in the overview, producing the right
worsted yarn requires the right tops, which in
turn require the right raw wool. To explain this
point, let's assume that we (a spinner) need to
spin 2/52 Nm weaving yarns for a fabric
manufacturer. The question is what is the right
raw material for efficient spinning of this yarn?

To answer this question, we need to get the right


tops first. The two critical parameters for a wool
top are the average fibre diameter and length.
Determining the average fibre diameter is
relatively easy, because the fibre should be fine
enough to ensure adequate average number of
fibres in yarn cross section. As discussed in the
module on yarn evenness, the minimum number of
fibres in the cross section of a worsted yarn should
be about 40. Below this limit, yarn quality drops
rapidly and spinning becomes inefficient due to
increased ends-down. Table 2.2 gives the
approximate diameter range used in majority of
commercial weaving and knitting yarns, and the
corresponding average number of fibres in yarn
cross section.

So for a 52/2 Nm weaving yarn, if we


wish to have 40 fibres on average in the
singles yarn cross section, then the
average diameter of the wool should be:
D2

917 19.2
,
40

D 21 ( micron )

Now that the average fibre diameter is


decided, the appropriate Hauteur (mean fibre
length) of the top is needed. For this purpose,
we may check previous record as to the likely
Hauteur value for a given fibre diameter
(micron) processed on our machinery. In the
absence of previous record, experience values
given in Table 2.3 can be used as a starting
point.

From this Table, a Hauteur value of at least


68 to 69 mm is necessary for the 21 micron
wool. If our machinery is in good condition,
the minimum Hauteur can be used.
Otherwise, higher Hauteur values should be
used to ensure efficient spinning. For the
current example, we can set the Hauteur
value at 70 mm.

By now we know that to spin a 2/52 Nm weaving


yarn, the top specifications should have an
average micron value of 21, and a Hauteur value
of 70 mm. The next step is to translate these
values into raw wool specifications. Again
determining the micron of the raw wool is the first
necessary step. In a typical top-making process, it
is normal that the average fibre diameter
increases by about 0.3 micron after top-making,
the reason for this is briefly discussed in the
section on combing. Again an individual mill's
past performance should be looked at in terms of
diameter increase (occasionally, fibre diameter
increase can be up to 1 micron). Keeping this in
mind, the average diameter of the raw wool
should be finer than that of the top.

We can use 20.7 micron for this example. Now


that the average micron for the raw wool is
determined, we need to know other raw wool
characteristics, such as staple length, staple
strength etc. This is where the TEAM formulae
play an important role. Using the TEAM
formulae, we can play with different
combinations of values for the parameters in
the TEAM prediction formula for Hauteur, such
as different staple length, staple strength, mid
breaks etc, to get the right Hauteur value (70
mm in this example).

One possible combination is:


Mean fibre diameter (D): 20.7 micron (already determined)
Mean staple length (L): 87 mm
Mean staple strength (S): 40 N/tex
Vegetable matter base (V): 1%
Percentage of mid breaks: 50%

Using the TEAM formula for Hauteur,


H = 0.52 L + 0.47 S + 0.95 D - 0.19 M* - 0.45 V
- 3.5
H = 0.52 x 87 + 0.47 x 40 + 0.95 x 20.7 - 0.19
x 50 - 0.45 x 1 - 3.5 = 70.3 (mm)

It should be pointed out here that the


Hauteur value predicted by the TEAM
equation is an average value. This means half
of the actual Hauteur values may fall above
the prediction, and the other half fall below
prediction. In other words, there is a 50%
probability for the actual Hauteur to be less
than that required. This is not an acceptable
situation. In practice, it is common to ensure
that the raw wool purchased will perform
better than predicted, so that there is only
about 5% chance of not meeting the
requirement.

It should also be pointed out though that in addition


to average fibre diameter and fibre length, fibre
diameter and length variations, short fibre content in
the top, are also important considerations.
Furthermore, the amount of noil produced during
combing is another important consideration. The
value of combing noil is only about 30% of the value
of a top. If we apply the TEAM formulae for Hauteur
CV and for Romaine to our example, we get:
CVH = 0.12 L - 0.41 S - 0.35 D + 0.20 M* + 49.3
= 0.12 x 87 - 0.41 x 40 - 0.35 x 20.7 + 0.20 x 50 +
49.3
= 46.1 (%)
Romaine = (-0.11) L - 0.14 S - 0.35 D + 0.94 V + 27.7
= (- 0.11) x 87 - 0.14 x 40 - 0.35 x 20.7 + 0.94 x 1 + 27.7
= 6.2%

The CV of Hauteur (CVH) may affect drafting


performance. Too high a CVH may lead to poor
evenness of yarns. For this reason, the experiences
have shown that CVH value should be less than
50%. The 46.1% CVH for our example is below this
figure. If we wish to have a lower CVH than 46.1%,
we can change some parameter in the CVH
formula, bearing in mind that any change will also
affect the predicted Hauteur and Romaine values.
As indicated earlier, computer packages are
available to assist with these calculations.
On the basis of this information, a specification of
the raw wool required for producing the 2/52 Nm
weaving yarn can be worked out. With a proper
specification, we will get the right raw materials.
Once we get the right raw materials, we can start
the processing. The first processing stage is
scouring, which is discussed next.

Scouring( )
A number of processes are carried out in a
scouring mill, including:
preparation of wool for scouring by opening
and blending,
scouring itself, and,
drying of scoured wool.

Raw wool contains a number of impurities. Some


of the impurities are removed in scouring, others
are removed in further processing. Table 2.4 lists
the impurities that are found in raw wool, and
the ways of removing these impurities. Figure 2.4
gives the average compositions of Australian
merino and crossbred fleeces.

Figure 2.4 Average composition of a Merino fleece (a) and Australian


crossbred (b) (Humphries 1996, p.90 [courtesy of International Wool
Secretariat])

The main objectives of scouring are:


To remove various impurities (grease, suint,
mineral) from wool in such a way that the scoured
wool is clean, full and open (with little felting of
wool).
To leave a small amount (about 0.5%) of residual
grease on the wool to facilitate further mechanical
processing.
Wool grease and suint are the key impurities removed
during scouring in the so called aqueous emulsion
scouring process using detergent and hot water.
Grease can also be removed by dissolving in organic
solvent using the solvent scouring process. But before
we talk about the actual scouring processes, we need
to understand the chemistry of these impurities.

The

chemistry of impurities( )
Suint
Suint consists of potassium salts of the various
lower fatty and amino acids, plus some inorganic
salts. Like the table salt used for cooking, suint is
soluble in water, particularly warm water (~ 30oC).
So removal of suint in aqueous scouring is not a
problem.

Grease/Wax
Grease is a mixture of higher fatty acids
(CnHmCOOH, or RCOOH) and alcohols. There is
about 2 - 15% free fatty acid in raw grease. At pH >
9, free fatty acid can be saponified (turned into
soap by decomposition with alkali). The Saponifying
(soap making) process is indicated below:

Na2CO3 + 2H2O ---> 2NaOH + H2CO3


NaOH + RCOOH ---> H2O + RCOONa
(soap!)
Grease has a relatively low melting point (38 to
43oC). In aqueous scouring, it is important to
raise the liquor temperature above this value to
facilitate grease removal. Since grease usually
forms a stable film around fibre surface. The
attraction between grease and wool needs to
be reduced to dislodge grease from fibre
surface. In aqueous emulsion scouring, this is
achieved with the help of scouring agents such
as detergent

Vegetable matter (VM)( )


Vegetable matter is cellulose material.
Scouring removes very little vegetable
matter from wool. A small amount of VM is
removed during the pre-scour opening, and
the rest is removed in the post-scour carding
and combing.

The

scouring agents( )
Detergent (Natural or Synthetic)
Detergents are surfactants or surface-active agents.
Surfactant molecules have a hydrophilic head and a
hydrophobic tail as indicated in figure 2.5 below.
Hydrophobic tail

Hydrophilic
head

Fig. 2.5: The molecule of a surfactant with a hydrophilic head and a


hydrophobic tail

In the presence of an interface (eg. air/water,


grease/water), the surfactant molecules are
adsorbed at the interface due to differing polarities
of head and tail. This then lowers the interfacial
surface tension. The importance of reducing
interfacial surface tension in scouring will be
discussed later.

The two main types of detergent are natural soap


and synthetic detergent.

Natural soap is made from action of alkali with a


fatty acid. It can also be created by the saponifying
process in scouring, as indicated in the section on
the chemistry of impurities. The water for wool
scouring should be soft water. Synthetic detergents
are not destroyed by lime salts. They can be used
under neutral pH (so less wool damage) but scouring
is more efficient in slight alkalinity. Synthetic
detergents have very high detergent efficiency,
even at low concentrations (when scouring fabrics in
the finishing stage, this may be a drawback.
Because a thorough degreasing often leads to a
worsening fabric handle. Most modern scouring mills
use synthetic detergents rather than natural soaps.

Water
The water used in aqueous scouring should be
of minimum hardness, particularly when natural
soap is used as the detergent. Hard water
containing calcium salts (lime) will reduce the
effectiveness of soap:
soap + calcium salts (lime) = lime soap
(insoluble!)
The insoluble lime soap can adhere firmly to
wool and cause difficulty in the subsequent
carding, combing and dyeing.

Aqueous wool scouring consumes a large quantity


of water. One kilo of greasy wool may require up
to 20 litres of water for scouring. If a scour has a
capacity of 1,200 kg/hour, then a staggering
24,000 litres of water will be consumed in every
hour of the scour's working life. Researchers and
industrialists have put considerable effort into
reducing water consumption in scouring and
treating scouring effluent.
Alkali( )
The addition of alkali adds to the efficiency of
scouring. It is usually in the form of sodium
carbonate (soda ash) rather than caustic alkali,
since caustic alkali attacks wool. As mentioned
before, alkali reacts with the fatty acid (in grease)
to produce soap in scouring.

The

principle of aqueous emulsion(


) scouring
As mentioned earlier, aqueous scouring
removes suint and grease mainly from wool,
and suint removal is relatively easy, because it
is soluble in water. Grease removal is more
complicated, and this section focuses on how
grease gets removed from wool in scouring.
The role played by detergent is highlighted in
this section. Generally speaking, there are
three stages involved in grease removal during
aqueous scouring.

Stage 1
This stage has the following functions:
- Wet out greasy wool
- Raise temperature above grease melting point
- Add alkali (demands of wool, saponification,
optimum scouring pH)
- Deliver detergent molecules to fibres by liquor
flow
and diffusion

Wetting is the first necessary step for aqueous


scouring. The more water molecules on fibre
surface, the better the fibre wetting. How easily a
greasy wool can get wetted depends upon the
fibre/water/air interfacial surface tensions.
Ironically, the wettability of water on wool textiles
is quite poor. When a drop of water is placed on a
surface of wool fibres, the water drop does not
spread out on the surface. The attraction between
molecules in the water drop is quite strong, which
tends to keep the water molecules together in a
ball shape. To increase the wettability of water, its
surface tension needs to be reduced so that the
water can spread out on the fibre surface to
achieve good wetting of the fibre.

Figure 2.6 depicts poor and good wetting


behaviour of water on a surface.

(poor wetting)

(good wetting)

Figure 2.6 Poor and good wetting behaviour of water on a surface

In relation to scouring greasy wool, we are


looking at a wool/liquor/air system as shown in
figure 2.7 below

Figure 2.7 Wetting the wool fibre with a drop of liquor - the wool/liquor/air system

In equlibrium, the three interfacial surface


tensions should be balanced.T We
have
wa - T Wl
.
T wa = T Wl + T al * cos , cos =
T al

This simple relationship can be used to explain how


detergent helps the wetting process. Detergents are
surfactants, which have the power of reducing
surface tension. When detergent is added to the
liquor, it greatly reduces the interfacial surface
tension between the liquor and its contact surfaces,
such as air and wool in this case. Therefore, with
the addition of detergent, both Twl and Tal will be
reduced. According to the above equations, should
decrease when Twl and Tal reduce. A lower means
the water drop is more spread out on the wool fibre,
hence better wetting. Therefore, from the change in
, we can tell how detergents help to wet the wool
fibres. In household washing, if you add detergent
to the water in the washing machine and then put
the clothing items in, the clothing items will get
immersed into the water more quickly than when
there is no detergent in the water.

Stage 2
This is the key stage, and has the following
functions:
Form surface film of detergent molecules

Gather grease into droplets


Sweep droplets from fibre by liquor flow
In this stage, we are looking at a
grease/wool/liquor system as shown in figure 2.8
below.

Figure 2.8 Removal of grease from wool - the grease/wool/liquor system

Again, in equlibrium, we have

T wg - T wl
.
T wl = T wg + T gl * cos(180 ), cos =
T gl

In the presence of detergent (surfactant) in the liquor,


both Twl and Tgl will be reduced. According to the
above equations, cos should increase (and should
decrease) when Twl and Tgl reduce. A gradually
reducing would mean the grease is rolling up as
droplets. When becomes zero, the grease ball
would come off the wool fibre easily with the help of
liquor flow. When the grease droplets are detached
from the wool, they are surrounded by surfactant
molecules. The hydrophobic tails of the surfactant
molecules will stick to the grease, while the
hydrophilic heads will stay in the liquor. In addition,
the like-charged hydrophobic heads on the surfaces
of the grease droplets will be mutually repulsive. This
keeps the grease droplets separate and suspended in
the liquor, without aggregating and re-depositing
back onto the fibre surface.

The grease removal process can also be

described without using the force balance


equations given above. When the scouring liquor
containing surfactant molecules comes in
contact with grease particles on fibre surface,
the water-hating tails of surfactant molecules
will compete for places in the grease, because
they dont like water molecules in the liquor. The
competition gets tougher and tougher as more
and more surfactant molecules try to stick their
tails into the oil. They only way of easing the
tension of competition is to create more surface
of grease, and the only way of doing this is by
breaking the grease apart and lifting the grease
away from fibre surface gradually.

Once removed from the fibre surface, the grease will


be surrounded by the surfactant molecules with the
tails inside the grease. In the mean time, the fibre
surface originally occupied by the grease will now
be occupied by the surfactant molecules, again
with their tails sticking to the fibres and heads
inside the liquor. The like-charged heads on the
fibre surface and on the grease surface repel each
other so they try to stay away from each other,
thus preventing the grease from being re-deposited
on the fibre surface. Similarly, the grease particles
broken apart by the surfactant will stay apart as
well. Therefore, after scouring, the scouring liquor
becomes an emulsion of suspended oil or grease
particles, which can be easily removed by rinsing.
For this reason, aqueous scouring is also known as
emulsion scouring or aqueous emulsion scouring. In
other words, surfactant helps to emulsify the oil or
grease to facilitate its removal. Similar principle
applies to house-hold washing.

The processes of grease or oil removal with


surfactant are illustrated below in Figure 2.9.
Water only

With surfactant
grease

Fibre

Figure 2.9 The process of grease removal from wool

Stage 3
This is the final stage which has the following
functions:
- Remove excess detergent and alkali from
wool
- Remove contaminants from liquor
Rinsing with fresh water achieves the first
function. Removing contaminants from liquor
requires complex effluent treatment system. In
fact, a well-known dilemma of wool scouring is
the environment friendliness-versus-cost
compromise.

Commercial

aqueous scouring systems


1.The basic configuration
In a typical process of aqueous emulsion
scouring, 5 or 6 scouring bowls are used. Figure
2.10 gives a simplified representation of a 6bowl aqueous wool scouring process.
Scouring agen ts

Greasy
wool in

Scoured
wool to
dryer
Bo wl 1
(De suint)

Bowl 2
(Scou r)

Bowl 3
(Scou r)

Bowl 4
(Scour)

Settling
tank
Centrifuge

Effluent Treatment

Grease Recovery

Figure 2.10 A 6-bowl aqueous scouring process

Bowl 5
(Rin se)

Bowl 6
(Rin se)

Fresh water in

The first de-suint bowl is used to remove watersoluble contaminants such as suint (or sheep
sweat) from the wool. The next three bowls
contain hot water, detergent and alkali for
grease removal, while the remaining two bowls
contain clean water for rinsing. Fibres are
propelled through each bowl and there is a pair
of squeeze rollers between the adjacent bowls.
Because of the scale structure on the wool
surface, excessive agitation of wool during
scouring will lead to felting of wool, which in turn
will lead to increased fibre damage during the
subsequent processes, carding process in
particular.

Fresh water is introduced from the last bowl for


rinsing, and flows backward to the scouring
bowls ('counter-current' flow). The water
temperature in the three scouring bowls is
usually set at about 55 to 60OC, with the
temperature in the rinsing bowls set at about 45
to 50 OC.
2.Scouring bowls
Three types of scouring bowls used in the
industry are given in figure 2.11.

(a) Conventional long bowl

(b) WRONZ Mini-bowl

(c) Fleissner suction drums

The conventional long bowl uses dunkers to push


the wool sheet into the scouring liquor. The ranks
then propel the sheet of wool through the bowl,
dislodging the mineral dirt from the fibres in the
process. The scratching of fibres against the
perforated screen (the false bottom) also helps dirt
removal. There is considerable agitation of wool
during scouring, which tends to increase the level of
wool felting. As mentioned earlier, wool felting is
very undesirable, because it leads to increased fibre
breakage in the subsequent processes such as
carding.
The WRONZ mini-bowl is the most widely used bowl
design. The agitation on the wool sheet in the bowl
is not as severe as in the conventional long-bowl. In
addition, the hopper shaped bowl facilitates fast
settling of the dirt in the bottom of the bowl.

The Fleissner suction drum type is the gentlest of the


three. Each bowl has several perforated rotating
drums half immersed in the scour bath. The scouring
liquor inside each drum is pumped into circulation
loop and then back to the bowl. When the wool sheet
enters the bowl, it is pushed into the liquor by a
rotating dunker (not shown in fig. 2.11c) first. The
sheet of wool then gets held against the suction drum
surface by in-flowing liquor and released as it
emerges at liquor surface on the other side. The sheet
of wool then flows to the next drum. Since there is
little mechanical agitation on the wool, wool felting is
minimised during scouring. This is the biggest
advantage of the suction drum system. But the
reduced agitation also means less dirt removal during
scouring. This is the well-known cleanliness-versusentanglement compromise in wool scouring. Reduced
entanglement often means reduced cleanliness as
well.

In addition to scouring bowls, the squeeze rollers


are also very important cleaning mechanisms.
Inadequate roller pressure will lead to scoured
wool with a high level of residual grease. Poor
opening of the wool before scouring will have a
similar effect. Fibre opening and blending before
scouring is discussed in the following section.

Fibre

opening( ) and blending( )


before scouring
Most consignments of greasy wool are made up
of a number of farm lots or inter-lots. Each lot
sold at auction should have a minimum number
of 4 bales of greasy wool. On average, each bale
weighs about 180 kg. The properties of wool in
the different lots are usually quite different.
Blending of the greasy wool in different lots is
therefore one of the most important functions in
the whole topmaking process. The blend should
be prepared to ensure a good mix across all the
different lots in a given consignment.

As an example, let's assume that we have a


consignment of 50 bales of greasy wool,
consisting of 4 lots of wool from different farms. It
would be very bad practice to feed the lots of
wool one after another, because fibre properties
differ from lot to lot. One way of mixing the lots
and feeding the bales of wool to a scour line is
sketched in figure 2.12. Other ways of mixing and
feeding the opening process has also been used
to minimise the irregularity in the final product.

Lot 1
(5 bales)

Lot 2
(20 bales)

Opening

Scouring

Lot 3
(15 bales)

Lot 4
(10 bales)

Figure 2.12 Blending of greasy wool bales before scouring

The typical opening systems used are the feed


hopper and drum opener, as indicated in figure
2.13.

(a) Feed hopper

(b) Double drum opener

Figure 2.13 Pre-scouring fibre opening systems


(Stewart 1985, p.5, p.7)
A usual combination of the opening systems is:
Feed Hopper 1 Double Drum Opener Feed
Hopper 2 Scouring

With the feed hopper 1, greasy wool from different


bales is placed on the feed apron in the hopper,
which forwards the wool to the spiked apron. The
spikes pick up tufts of wool and move them
upwards. Upon meeting the comb drum, large tufts
are separated into smaller ones and some small
tufts continue with the spiked apron until doffed off
by the doffer drum, others are returned by the comb
drum to the hopper. The retuned tufts will mix with
others already in the hopper, so some blending is
also achieved with the feed hopper opening system.
The double drum opener has two toothed drums
between which fibres are opened. There are usually
perforated screens underneath each drum, which
allow dirt and dust to fall away. So apart from fibre
opening, some cleaning is also achieved with the
double drum opener. Feed hopper 2 will deliver a
uniform, opened layer of greasy wool to the scour.

It is worth pointing out that all tri-pack wool

(three bales compressed into one for ease of


storage and transport) and some farm bales
which have had prolonged storage need to be
warmed up prior to scouring. Even for ordinary
bales, cold climates may warrant bale warming
before processing. Warming the bales will loosen
and open fibres in the bale, which leads to better
scouring and reduced fibre breakage. Bale
warming is achieved in a number of ways, the
most popular being steam or hot-air heating. A
recent development is to heat the bales with the
microwave technique.

Drying

of scoured wool
The wool leaving the last pair of squeeze rollers
has a moisture content of about 40%, or a regain
mass of water
regain

of about 66%
(
)It is typical to
mass of dry wool
dry the scoured wool to around 8 to 12% regain.

There are three main types of dryers available


for wool - suction drum dryer, conveyer dryer,
and Unidryer. They use either gas or steam for
heating. Figure 2.14 shows a side view of the
three dryers.
Figure 2.14 Side view of wool dryers (Teasdale
1996, p. 97)

1.Suction drum dryer


Upon entering the dryer, the sheet of wet wool is held on
rotating perforated drums by suction created by fans on the
end of each drum. An internal baffle in each drum blocks
half of its circumference so that wool is only held on half of
its surface by the suction. As the drums rotate, the sheet of
wool is passed from drum to drum so that the heated air
passes in alternate directions through the wool for even
drying. There is a 'counter-current' airflow inside the dryer.
In other words, the direction of air flow is opposite to the
direction of wool flow. Fresh (cool) air is fanned in from the
delivery end, which cools the dried wool at exit. The air is
recirculated to the drums via heating batteries. As the
heated air is drawn through the wool, it carries moisture
with it to dry the wool. The air is then heated again and
drawn through the layer of wool towards the wool inlet,
where the wet air is finally exhausted to the atmosphere.
Drum dryers usually have between 4 to 8 drums.

2.Conveyer dryer
This is the traditional hot air dryer where the
sheet of wool is carried through on a perforated
conveyer. Similar to the drum dryer, fresh (ccol)
air is introduced at the wool outlet to cool the
exiting wool. This air is then heated and blown
down through the layer of wool, carrying
moisture with in to dry the wool. The now moist
air is then heated again and blown through the
wool layer towards the wool inlet, where the
wet air is finally exhausted to the atmosphere.

Unidryer
This dryer was developed at the University of
New South wales. It carries wool between two
porous conveyer belts. The conveyer belts pass
next to a perforated screen. Again there is a
'counter-current' airflow inside the dryer, with
the fresh air coming into the dryer from wool
outlet and wet air exhausted at the wool inlet.
The direction of airflow in the two chambers is
opposite so that wool is dried evenly from both
sides. The unidryer is a powerful dryer and is
much smaller than the other two.

Worsted Carding
The same adage - "well carded is half spun", as
quoted in the cotton carding section also applies
to worsted carding. Carding is a vital process in
the fibre to yarn processing chain.
Objectives
The main objectives of worsted carding are:
-to disentangle and align the scoured wool,
-to remove the vegetable matters left in the
scoured wool,
-to intimately mix the fibres, and,
-to deliver the carded fibres in a continuous ropelike form called a card sliver.

Owing to the scale structure on wool surface, it


is unavoidable that some degree of fibre
entanglement occurs during aqueous wool
scouring. Carding is the only process that can
untangle and individualise the fibres. After
scouring and drying, the vegetable matters still
remain in the wool. The bulk of these foreign
matters are removed in carding. Carding also
achieves intimate mixing of wool fibres, which
is only possible with individualised fibres.
Considerable fibre breakage occurs during
carding, mainly because of the fibre
entanglement. To minimise fibre damage,
adequate fibre preparation between wool drying
and carding is essential.

Fibre

preparation before carding


After the wool is scoured and dried, the
moisture content and total fatty matter (TFM)
left in the wool are checked. This will allow
correct addition of oil/water to facilitate wool
carding. Usually the scoured/dried wool is gently
opened first by a simple opener as shown in
Beater
figure 2.15.
Doffer

Spiked
lattice
(wool from
dryer)

Oil/water
spray

Figure 2.15 A simple opener for scoured and dried wool

Wool to
storage bin
via pneumatic
duct

A mixture of processing oil and water is sprayed


on to the wool, preferably at the delivery end
just before the wool enters the pneumatic
transport ducts to the storage bins. The wool is
allowed to stand for a minimum of 12 hours in
the bins to allow oil and water to spread evenly
throughout the wool before carding. Insufficient
moisture in the wool will cause static problems
during carding, while too much oil will cause
wool lapping on the card rollers.

Roller-top

card
Unlike the flat-top card used for carding cotton
fibres (or other staples of similar length to
cotton), a roller-top card is used for carding wool
fibres. A simple roller-top card is shown in figure
2.16.

In the simple card depicted in figure 2.16, the


broken line represents the flow of fibres. The
incoming fibres are first picked up or 'licked in'
by the teeth of the lickerin or takerin.
stripper

Burr
beater

Feed
rollers

work er

Doffer
comb

cylinder
Licker-in

Doffer
Transfer
roller

Figure 2.16 A simple roller-top card

As the name implies, a roller-top card has


rollers, rather than flats, on top of the cylinder.
Roller-top cards are relatively gentle on fibres,
which is important in minimising damage to the
delicate fibres such as wool. The rollers are
clothed with metallic teeth pointing to certain
directions. As discussed in cotton carding, the
teeth direction, the relative speed between two
adjacent roller surfaces, and the rotating
direction of the roller surface govern the basic
actions (point-to-point carding or point-to-back
stripping) between the two adjacent rollers.

In the simple card depicted in figure 2.16, the


broken line represents the flow of fibres. The
incoming fibres are first picked up or 'licked in' by
the teeth of the lickerin or takerin. The sheet of
fibres travels with the lickerin and presents the
relatively exposed vegetable matters (VM) in the
fibres to the flicking action of the beater for
removal. Between the lickerin and the transfer
roller, a point-to-back stripping action occurs and
the sheet of fibres on the lickerin is stripped by the
teeth of the trasnfer roller. Another point-to-back
stripping action occurs between the cylinder and
the transfer roller, which allows the sheet of fibres
to be stripped off the transfer roller by the cylinder
teeth. The cylinder now carries the sheet of fibres
to the very important stripper/worker pair.

The relative surface speed (V) here is:


Vcylinder>Vstripper>Vworker. This, together with the
tooth direction and the surface rotating direction
as indicated in figure 2.16, gives the following
basic actions in the cylinder/worker/stripper unit:
-Cylinder/stripper: point-to-back stripping
action; cylinder strips the stripper.
- Cylinder/worker: point-to-point working action;
both surfaces will grab some fibres to untangle
them.
- Worker/stripper: point-to-back stripping action;
stripper strips the worker.

With this arrangement, the fibres carried by the


cylinder will 'by-pass' the stripper and proceed to the
working (or carding) action of the cylinder/worker
pair. This is where much of the fibre untangling and
alignment occur in the card. The fibres picked up by
the worker will soon meet the stripper and be
stripped by the stripper. The stripper then returns
these fibres to the cylinder for further action. The
fibres that are not picked by the worker will continue
their journey with on the cylinder surface, until they
reach the doffer. The doffer has a lower surface
speed than the cylinder, and a working or carding
action happens between them. This would mean that
the fibres entering the cylinder/doffer zone is further
opened, and a fraction of the fibres will be picked up
by the doffer teeth, and this sheet of fibre will then
be stripped off the doffer surface by the rapidly
oscillating doffer comb.

The fibres not picked up by the doffer will stay on


the cylinder surface as 'recycled fibre'. The
'recycled fibres' on the cylinder surface will soon
meet with the 'fresh fibres' that have just been
picked up by the cylinder from the transfer roller.
Together, the recycled and fresh fibres are then
presented to the cylinder/worker/stripper unit and
the process repeats. The 'looping' of fibres on the
cylinder/worker/stripper unit and the recycled
fibres meeting with fresh fibres on the cylinder are
also important for intimate fibre mixing. From this
brief description, we can see that the
cyliner/worker/stripper unit achieves the important
key objectives of fibre untangling, fibre alignment,
and fibre mixing. On a modern worsted card, many
such units are employed to ensure sufficient fibre
opening, aligning and mixing in carding.

Figure 2.17 shows a schematic diagram of a


modern worsted card.
Worker/stripper units
Worker
Burr
beater

Feed
rollers

Stripper

Licker-in

Main cylinder
(Sw if t)

Breast
cylinder

Doffer
comb
Doffer

Transfer
roller

Morel roller
(usually 2)

Figure 2.17 Diagram of worsted roller top card

A total of nine cylinder/worker/stripper units are


employed on the card to achieve the desired level
of fibre opening, aligning, and mixing. In addition,
the worsted card is often equipped with specially
designed burr removal rollers (eg. Morel roller)
and burr beaters to free the opened fibres from
burrs and other vegetable matters. The morel
roller is clothed with special teeth, whose gaps
are large enough to accommodate the fibres but
are too small for the vegetable matter, so that the
vegetable matter can be beaten off by the bur
beaters on top of the morel roller. The thin web of
fibres removed from the doffer by the doffer
comb is usually condensed into a sliver or rope
form, and deposited into a sliver can or coiler can
for further processing. Figure 2.17a shows a
sketch of a sliver can with coils of sliver inside.

Important

settings in carding
(1)Card loading or production rate
The theoretical production rate of a worsted
card can be calculated using the formula below:

Theoretica l production rate (kg / h)

2.9 MFD (micron ) Card width (m) Swift speed (m / min)


1,000

This is the production rate at 100% efficiency,


and is related to the mean fibre diameter
(MFD), card width, and the swift surface speed.

If the card is in good condition and set properly,


it can produce quality products at this theoretical
production rate. Otherwise a lower rate is
necessary to reduce fibre breakage during
carding. The usual way of checking card
performance is to process the card slivers to
tops, and examine the length characteristics of
the tops on the Almeter and calculate the
coming yield and noil figures.

(2)Fresh fibre density (FFD)


Research at CSIRO has demonstrated that the
density of fresh fibres on the main cylinder (swift)
has a major effect on fibre breakage in the
carding process. The fresh fibre density (FFD) is
calculated using the formula below:
Fresh Fibre Densty ( g / m 2 )

Pr oduction rate ( kg / h) 1,000


60 Swift speed ( m / min) card width ( m)

By combining this equation with the formula for


production rate, we can derive the formula for
theoretical fresh fibre density (FFDt):
2.9 MFD ( micron )
Theoretica l fresh fibre density ( g / m )
60
2

If the actual fresh fibre density is significantly


higher than the theoretical fresh fibre density,
considerable fibre breakage may arise during
carding. On the other hand, if the actual fresh
fibre density is kept below its theoretical value,
increasing the card production rate will have
little effect on the quality of the carded sliver. It
is worth reiterating that the only way to confirm
card performance is to check the fibre length
characteristics on the Almeter and calculate
combing yield and noil figures at each different
setting.

(3)Roller settings
The clearance between adjacent roller surfaces
and the relative surface speeds are important
settings that affect carding quality. The clearances
gradually decreases from the feed to delivery end
of the card as the fibre materials become thinner.
The card manufacturer will advise on the best
settings for the particular type of fibre being
processed by its card. Incorrect settings may
reduce the mean fibre length, and increase the
number of neps in the carded sliver.
The reading material "The pressure on fibres in
carding" by Harrowfield, Eley and Robinson (1986),
reports relevant research carried out at CSIRO.

Like the cotton card, a properly set worsted card also


generates a majority of trailing fibre hooks in the carded
sliver, which are then straightened in the gilling (drawing)
processes following carding. Despite of the best effort,
fibre breakage and other less dramatic forms of fibre
damage are unavoidable in carding. Poor scouring leading
to increased fibre entanglement, excessive fresh fibre
density, and poor lubrication are the main causes of fibre
breakage in carding. In addition, highly entangled balls of
fibres (called neps) are also generated in carding. Card
surfaces with blunt teeth may lead to rolling of fibres
between adjacent surfaces and create neps in the
process. Grinding of card clothing to maintain sharpness
of the teeth will reduce this problem. Another possible
mechanism of nep formation in carding is the 'snap back'
effect. In carding, many fibres are stretched at high
extension rates. When one fibre breaks, the broken end
may 'snap back' and entangle with neighbouring fibres to
form neps. Most of the broken short fibres plus the neps
generated in carding are subsequently removed as waste
fibres (noils) in the combing process. This is also why the
combing yield and noil can reflect the performance of
carding.

Preparatory Gillings( )
Objectives
The main objectives of the gilling machine are to
further align the fibres in card sliver and to blend
the slivers from different cards.
A gilling machine is also known as a gillbox, or
simply a gill.
As discussed before, most fibres in the card
slivers have hooked ends, with the trailing hooks
at a majority. These hooks and other poorly
aligned fibres should be straightened out before
combing to increase the average fibre length
and reduce the percentage of noil during
combing.

Gilling

process
Gilling is basically a roller drafting process in
which fibre movements are controlled by pins
fixed on moving pinned bars (faller bars).
Figure 2.18 shows a schematic of a gillbox with
intersecting upper and lower faller bars
controlling fibre movement during drafting.
Such a gillbox is also called an intersecting
gillbox.

Figure 2.18 Schematic of an intersecting


gillbox (Grosberg and Iype 1999, p.14)

You may recall that perfect roller drafting requires


fibres in the drafting zone travel at the speed of
back rollers until their leading ends reach the front
roller nip, where they get accelerated to the front
roller speed instantly. In gilling, several slivers (eg.
from different cards) are combined as the input
material, which is drafted to produce a single
sliver at the output. During drafting, the faller bars
move at about the same speed as the back rollers
and the pins on the faller bars keep the fibres
moving at a similar speed. Once the leading end
of a fibre gets gripped by the nip of the front
rollers, that fibre gets pulled through the pins on
the faller bars and has its trailing end
straightened. So each passage of gilling
straightens fibre trailing ends mainly.

The distance between the front roller nip and the closest
faller bar is called the front ratch setting. This setting is
very important for gilling. If it is set too large, then many
fibres in the critical region near the front rollers are not
properly controlled by the pinned faller bars during
drafting. We already know that lack of fibre control during
drafting will lead to increased irregularity in the drafted
material. On the other hand, if the front ratch setting is
too small, pulling the fibres through the pins on the faller
bars may cause fibre breakage. In practice, the front ratch
setting is set at about half of the average fibre length.
Using this value as a starting point, the final setting
should be optimised based on sliver evenness results,
particularly the spectrograms obtained from the Uster
evenness tester. As discussed in the Module on Yarn
Evenness, the spectrograms allow us to identify the
presence of drafting faults such as drafting waves. If a
drafting wave is identified from the spectrogram, a closer
front ratch setting should be used to improve fibre control
and reduce the number of floating fibres during gilling.

Three

intermediate gillings
It is common practice in the worsted industry to
have three intermediate gillings between carding and
combing. You may ask why this is necessary. To
answer this question, we need to keep in mind the
following three points:

(1) Fibres in card slivers have a majority of trailing


hooks
(2) There is a natural reversal of fibre ends between
two
processing stages ('first-in-last-out')
(3) In feeding a worsted comb, fibres with leading
hooks tend to cause more fibre breakage in combing
(note the difference between worsted combing and
cotton combing!)

Figure 2.19 shows the configuration of fibre


hooks during gilling.

1st Gill

Card

Can

2nd Gill

Can

Comb

3rd Gill

Can

Can

Figure 2.19 Fibre configurations from card to comb

1st Gill

The card produces fibres with a majority of trailing hooks.


When the card sliver is deposited into a can and then taken
out to feed the 1st gill, there is a reversal of fibre ends, so
that the fibres entering the 1st gill have a majority of leading
hooks. Because gilling straightens trailing hooks only, the
fibres will emerge from the 1st gill still with the leading
hooks. Now the 1st gilled sliver is stored in a can and taken
out again for the 2nd gilling operation. The reversal of fibre
ends mean that fibre entering the 2 nd gill have a majority of
trailing hooks, most of which are straightened during this
2nd gilling. So after the 2nd gilling, most fibres are straight
except for a few which may still have some trailing hooks.
After a further can storage and removal from the can, the
sliver now enters the 3rd gill with a few fibres with leading
hooks, which can not be straightened and will persist to the
3rd gilled sliver. But when the 3rd gilling sliver is stored in a
can and taken out for combing, any remaining hooks in the
sliver would be trailing, which is fine as far as worsted
combing is concerned. Feeding a worsted comb with
leading hooks is likely to increase fibre breakage during
combing, as will be discussed in the following section.

From figure 2.19, you may think that the 3rd gill has
done nothing to the fibres. This is not quite true. In
gilling as in cotton drawing, there is a doubling
function as well. Many slivers are fed to a gill
together, and there is a doubling and blending
function by each gill, which improves the evenness
of the gilled sliver.
Five intermediate gillings between worsted carding
and combing have been tried before, but the
benefit is too marginal to justify the cost for two
extra gillings.
After the 3rd gilling, the slivers are ready for
combing. Combing is discussed next.

Worsted Combing( )
Objectives

Combing is a critical step in worsted


processing. Similar to cotton combing, worsted
combing achieves the objectives of:
(a) removing short fibres, neps, and impurities
(collectively known as noils)
(b) further mixing and aligning fibres
(c) forming a continuous rope-like comb sliver
Since longer and cleaner wool fibres make
better yarns, combing improves the yarn
quality and combed yarns command a high
price.

Combing process
A schematic diagram of a rectilinear worsted comb is
shown in figure 2.20. Up to 32 carded and gilled slivers
may be fed to the comb via a pair of feed rollers and
the feed gill assembly. Like the cotton comb, worsted
comb also runs in an intermittent fashion. In each cycle,
the following actions are performed:

(1) Feeding a short distance of slivers to the comb


(2) Initial combing of fibre leading ends by the cylinder
comb or comb cylinder
(3) Final combing of fibre trailing ends by the
intersector
comb or top comb
(4) Detaching the fully combed tuft and overlapping it
with previously combed ones
Figure 2.20 A side elevation of a worsted comb (Brearley 1964,
p50).

During the initial combing stage, the fibres are firmly


gripped by the nips or nipper jaws. Fibre leading ends
protruding from the nipper jaws are combed by the comb
cylinder (also known as the circular comb). Fibre penetration
of the pins is aided by a nipper brush (not shown in figure
2.20) attached to the upper nipper. In the initial combing
process, short fibres, neps and impurities are removed as
noils by the comb cylinder, which itself is then cleaned by
the noil brush. If some fibres have large leading hooks, there
is a possibility for the leading hooks to be engaged by the
teeth on the comb cylinder and fibre breakage will occur as
a result. This is why it is preferable that no leading hooks
exist in the feed stock. In contrast, the presence of leading
hooks in the feed stock to cotton combs is normal, since
there are only two passages of drawing between cotton
carding and combing. The difference is that for short staple
fibres such as cotton, the fibre length is much shorter than
long staples such as wool. Even if there are leading hooks,
the extent of the hooks will be considerably smaller and the
small hooks are unlikely to be engaged by the teeth of the
comb cylinder to cause fibre breakage.

In the final combing by the intersector comb (also


known as the top comb), the trailing ends not
combed by the comb cylinder are combed. The short
fibres, neps and impurities are held back by the
intersector comb, and will be removed in the next
combing cycle. The noils or combing noils (i.e. short
fibres, neps and impurities) represent combing
waste. The value of noils is about one third of that of
tops, so any increase in the percentage of noil is
going to cost the topmaker large sum of money. But
an increase in noil is usually accompanied with an
increase in the mean fibre length of the combed
sliver, because more short fibres are removed. This
improves the value of the tops. So choosing the right
percentage noil is a balancing act. A 6% noil is
typical in modern topmaking mills.

Each fully combed tuft of fibres is drawn off by


the drawing-off rollers and laid on top of the
previously combed ones, like shingles on the
roof. This overlapped web of fully combed fibres
is then consolidated by the calender rollers and
deposited in the sliver can. Because fibres in the
combed sliver simply overlap, the cohesion
between fibres is very small, therefore the
combed sliver is very weak. To improve the
strength of combed slivers, the slivers may be
crimped by a crimping mechanism before they
are deposited into the coiler can.

Figure 2.21 shows the worsted combing


in separate stages, noting that the upper
nipper has a brush attached to it to
assist with fibre penetration into the pins
of the circular comb.

Figure 2.21 A graphical representation of


worsted combing process (CSIRO Wool
Textile News, Feb. 1969).

Settings

on the comb( )
(1)Noil( ) setting

Noil setting is also known as detachment


setting or gauge setting. It is the closest
distance between the bite of the nippers and
the nip line of the detaching rollers. On a
worsted comb, this setting may vary from 26
mm to 40 mm. A large ratch setting will lead to
increased noil, and longer mean fibre length in
the top. This will be further discussed in the
following section on the geometrical model of
worsted combing.

(2)Feed( )
Normally about 12 to 32 slivers are fed to a
worsted comb, depending on sliver weight and
machine design. A practical rule of thumb is:
Rule of thumb for feed: Input ktex = 20 x Micron
For example, if the comb is processing 22
micron wool, then the total density of the feed
stock may be set at 20 x 22 = 440 ktex.
As mentioned earlier, the comb operates in an
intermittent fashion. On a worsted comb, the
feed length is usually adjustable between 4.5 to
9 mm per combing cycle.

(3)Production rate( )
The comb production rate can be calculated
using the formulas below:

A practical rule of thumb is:


Rule of thumb for production rate:
Prod.Rate = 1.4 x Micron
It is typical for a worsted comb to have a
production rate between 20 to 40 kg/hour.
Geometrical theory of worsted combing

Geometrical

theory of worsted combing


Belin and Walls (1963) developed a geometrical
model of fibre selection in worsted combing. This
model is represented in figure 2.22 below.
Nip of detaching
rollers
B

Nipper
position
E

Feed
mechanism

s ection combed
by circular com b

(a)
f

A
Noil Setting N

Top
comb

A
f

(b)
(a) Fibre b eard in the combi ng zone BD com bed by circular com b and about to be
presented to the detachingrollers. DE is the dead zone in front of the nipper,
where pi ns on the comb cylinder can n ot reach
(b) Fibre b eard advanced a distance f by the feed mec hanism F. Fibres with ends
inside detaching zone AB rem oved to combed sli ver, their tail ends combed by
the top comb C

Figure 2.22 Geometrical model of fibre selection in worsted combing

This model shows that for fibres not held by the nippers,
the combing action of the circular comb will remove them
as noils. These are the relatively short fibres, i.e. fibres
shorter than the noil setting (N). Longer but poorly aligned
fibres not gripped by the nippers will also be removed as
noil. After the initial combing by the circular comb, the
fibre beard is fed forward a short distance represented by
'f' in figure 2.22. Combed fibres with leading ends
reaching the detaching zone A'B (AB after feeding) will be
pulled through the top comb by the detaching rollers and
they will end up in the combed sliver. If a fibre has its
trailing end just gripped by the nippers and its leading end
just reaching the detaching zone A'B, this fibre will end up
in the combed sliver, even though its length is relatively
short (slightly longer than N - f). On the other hand, if a
relatively long fibre of length N is not gripped by the
nippers, it will end up as noil regardless of the fact that its
leading end is well inside the detaching zone A'B.
Therefore, it is inevitable that a few fibres in the noil are
longer than some fibres in the combed sliver. This will be
more so if the fibres are not well aligned before combing.
This also highlights the importance of pre-comb gillings
and the necessity to straighten fibres before combing.

For any fibre that is shorter than (N - f), there is no


way that this fibre will be able to get gripped by
the nipper and in the same time having its leading
end inside the detaching zone A'B, so this fibre will
always end up in the noil.
During detaching, the top comb is inserted just a
short distance in front of the 'dead zone' DE to
ensure that fibres are fully combed along their
entire lengths. The 'dead zone' exists because the
pins on the circular comb can not reach right up to
the nipping points of the nippers. In practice, this
'dead zone' length is reduced with the aid of a
nipper brush, which pushes the fibres into the pins
on the circular comb (figure 2.21). The impurities,
short fibres and neps blocked by the top comb
during detaching are removed as noil in the next
combing cycle.

Effect

of combing on fibre properties

In addition to the obvious changes in fibre length, combing


also changes fibre diameter slightly. With wool fibres, finer
fibres are usually shorter ones, and are also easier to break
during carding and combing. Since combing removes short
fibres mainly, the average diameter of fibres in the noil is
about 10% finer than the raw wool from which they have
been produced. With the removal of relatively shorter and
finer wools, the combed sliver has an average fibre
diameter that is about 1% coarser than the raw wool from
which the combed sliver is produced. In absolute terms, it
is usual for the average fibre diameter to increase by about
0.3 micron after combing. This increase should be
considered when selecting raw wool, as discussed in the
section on TEAM formulae and applications.
The reading material "ITMA'99 - long staple gilling and
combing" by Atkinson (1999) discusses the latest
developments in worsted gilling and combing, as exhibited
at the 1999 international textile machinery exhibition.

Worsted Top( ) Finishing


The worsted top finishing usually consists of
two gilling operations on the combed slivers.
Objectives

The objectives of top finishing are:


-Randomise comb overlap effect
-Mixing of fibres from different combs
-Producing tops of the right count, regain, and
package form
-Improving the evenness
An explanation of some of these objectives is
given in the following section.

The

process
As with gilling before combing, the top finishing
gilling processes combine drafting, doubling and
pin control.
As the combed slivers are removed from the
storage cans to feed the 1st finishing gillbox, they
follow the 'first-in-last-out' order (i.e. natural
reversal of fibre ends). Because of this natural
reversal, fibres in slivers fed to the 1st finishing
gilling are drafted in the reverse direction, i.e.
opposite to the direction of fibres as they came out
of the comb. This is known as 'reverse drafting'.
The reverse drafting helps to randomise the
overlapping fibre ends in the combed slivers,
leading to improved sliver strength due to the
increased inter-fibre cohesion. This is shown in
figure 2.23.

Combed
sliver with
overlapping
points
Overlapping points
Drafting
direction

Reverse drafting
randomises fibre
ends in the
combed sliver

Figure 2.23 Reverse drafting randomising comb overlapping effect

The first finishing gillbox feeds on slivers


from many combs. This doubling of slivers
(from different combs) in the gilling
process improves the evenness of the
gilled sliver, as discussed in the module on
yarn evenness. A special control
mechanism is often used in the finishing
gill to automatically level or regulate the
sliver evenness, such a mechanism is
called an autoleveller.

The principle of autolevelling has been discussed in the first


topic of this module. Figure 2.24 shows a gillbox with an
autolevelling system. The slivers are fed into the gillbox via a
condenser funnel and a pair of measuring rollers. This measuring
rollers are known as tongue and groove or shoe and groove
rollers. The bottom roller has a groove through which the slivers
run. The tongue or shoe roller is the top one, which sits in the
groove roller, pressing on the slivers. Variations in the amount of
fibre materials running through the tongue and groove rollers
will cause the tongue roller to move up or down. This movement
reflects the variation in material thickness and is automatically
recorded as a 'pattern line' on the memory unit. The memory
unit then feeds this information forward to the transmitter unit,
which controls a variable speed mechanism. This mechanism
can change the speed of the back rollers to vary the draft. For
example, if the tongue and groove measuring unit detects a
thick section running through, this information is recorded and
fed forward to the transmitter. When the thick section arrives in
the drafting zone, the variable speed mechanism will reduce the
back roller speed to increase the draft on this thick section.
Conversely, if the measuring unit detects a thin spot section
coming through, the draft on the thin section will be reduced by
increasing the speed of the back rollers.
Figure 2.24 Autolevelling gillbox (Brearley and Iredale 1980,
p.83)

After two finishing gillings, we get worsted


'tops'. The tops are usually packaged in balls or
in bumps. Balls are self-supporting cross wound
tops, while bumps are press-packed layers of
coiled tops.
The tops should be of the right moisture
content. The standard regain for tops is about
18%. It is normal to spray moisture on the
sliver at the delivery end of the first finishing
gill, to bring the regain up to about 19% first.
At the 2nd finishing gilling process, about 1%
of the moisture is lost through evaporation.

Quality

of wool tops
From a spinner's point of view, tops should be
produced to the spinner's specifications. These
specifications usually include requirements on
the following:
-Fibre diameter (micron) and CV of fibre
diameter
-Average fibre length (Hauteur) and CV of
Hauteur
-Count or size of the top
-Short fibre content

In addition, the tops should be free from


contaminations (coloured fibre, vegetable
matters, bale packing materials etc), and of
the correct shape and density.
The TEAM equations discussed before have
been routinely used by top-makers to select
the right wool for the tops, and for predicting
the important quality attributes of tops
according to raw wool specifications.

Worsted Drawing
Objectives

After the tops arrive at a worsted spinning


mill (or the spinning department of a vertical
worsted mill), they go through a series of
drawing processes. The main objectives of
worsted drawing are:
-Reduce the count of tops gradually before
spinning
-Mix tops of different properties
-Minimise irregularity in count, colour etc.

Drawing

process
The drawing sequence differs from mill to mill. A
typical sequence of producing a 21 tex worsted
yarn is given in Table 2.5

The roving frame used in the worsted


system is mainly of the rub finisher type,
although a flyer roving frame is also
used. A schematic of the rub finisher
roving frame is shown in Figure 2.25.

Fig. 2.25: Schematic of a rub finisher


roving frame (Grosberg and Iype 1999,
p.16)

The rub finisher roving frame uses roller drafting to


attenuate the input sliver. The roving straight after
drafting is weak, since no twist is inserted into the
roving. To increase the strength of the roving,
rubbing aprons are used to consolidate the roving
to increase fibre cohesion. Each roving package
normally has two rovings wound up side by side
without being twisted together (parallel winding).
The rovings are then used to feed a worsted ring
spinning machine.
Because the roving stage is the last process before
spinning, it is important that rovings are of good
quality. The Uster spectrograms are often used to
check whether there is any periodic mass variation
in the roving. If there is, it should be rectified as
soon as possible before further processing.

Once we have the rovings, spinning can


commence to make yarns. Spinning is
discussed in the next module.
While we have devoted this topic to worsted
processing, it is also important to know the
system of woollen processing. The reading
material Woollen processing system by
Osborne (1998) provides essential information
on the differences between worsted and
woollen processing.

Review questions
(1)If a 20 tex weaving yarn is to be produced with an
average number of 40 fibres in yarn cross section.
What would be the right average diameter of the
raw wool for the top-making process? You need to
show the calculations involved and consider
changes in fibre diameter during top-making.
(2)In your own words and use sketches if necessary,
explain how detergent helps to remove grease
from the surface of wool during the aqueous wool
scouring.
(3)An important unit on a card is the
cylinder/worker/stripper unit. Explain, with the
help of sketches, how this unit helps achieve the
objectives of fibre opening, aligning, and mixing
during carding.

(4)With reference to carding, gilling and


combing processes, explain why three
intermediate gillings are commonly used
in worsted top-making.
(5)With reference to the geometrical model
of fibre selection in worsted combing,
explain:
(a) Why is possible to find a few fibres in
the noil that are actually longer than some
fibres in the combed sliver?
(b) What happens to the noil% and the
mean fibre length of the combed sliver is
the feed 'f' is increased?

Staple Spinning and


Filament Texturing

Ring spinning
Introduction
Ring spinning has been and will continue to be an
important spinning system for making staple spun
yarns ( )from different fibres. Since its
invention in 1828, little has changed in terms of the
principle of ring spinning. Furthermore, the principle of
ring spinning for short staples such as cotton and for
long staples such as wool is exactly the same. So the
discussion in this topic applies to both short staple and
long staple ring spinning . This
topic discusses the three basic stages of ring spinning,
the physics of ring spinning, and the developments as
well as limitations of ring spinning. The detailed
differences in machine design for long staple and short
staple ring spinning are beyond the scope of this topic.

Objectives
At the end of this topic you should be able to:
Explain the basic principle of ring spinning
Know the features of ring spun yarns

Understand the theory of yarn balloon and


its implications
Appreciate the advantages and
disadvantages of ring spinning

The principle and process


A schematic diagram of a ring spinning process
is shown in Figure 1.1.

Figure 1.1
Diagram of a ring spinning system (Mathews &
Hardingham 1994, p.9)

It consists of a roller drafting unit


, a ring and traveller
assembly, and a bobbin
mounted on a spindle (driven by a tape). A
yarn guide (pigtail guide) is also used to
guide the yarn. To start ring spinning, a seed yarn
(on an empty bobbin) is threaded
through the traveller and the pigtail guide. It is then
brought to the nip of the front rollers where a thin
strand of fibres emerges. As the bobbin/spindle
rotates, the seed yarn is twisted and the twist flows
upwards to trap the thin strand of fibres emerging
from the front rollers. A continuous twisted strand of
fibres (i.e. the yarn) is thus formed. The newly
formed yarn is wound up onto the bobbin. To avoid
the newly formed yarn being wound onto just one
spot of the bobbin, the ring rail oscillates
upwards and downwards during spinning to build up

The three basic steps of ring spinning, i.e. drafting ,


twisting , and winding-on , are discussed below.

Drafting
The roving is drafted by a roller drafting unit on the
ring frame . Figure 1.2a shows the typical
drafting arrangement. They comprise three fluted
bottom rollers (a), against which are pressed
three top rollers (b) that carry the pivoted
weighting arm (c). The top rollers
are driven via frictional contacts by the bottom rollers, to
which the drive is applied. The three pairs of rollers form
two drafting zones . The break draft zone
formed between the back and middle pairs of
rollers has a small draft only, and there is little fibre control
in this zone.

Figure 1.2 The roller drafting arrangement (Klein1987, p.5)

The main draft zone is formed


between the middle and front pairs of rollers.
Fibre control is achieved by the revolving
double aprons (e) in this zone.

(a) The drafting arrangeme


(b) Cross-section through the drafting
arrangement

You may recall the concept of perfect roller


drafting. It requires fibres in the drafting zone
travel at the speed of back rollers until the
fibre leading ends reach the front
roller nip , where they get accelerated
instantly to front roller speed. Because of this
requirement, the apron speed is close to the
back roller speed (break draft is small), and the
aprons have a nose
which is very close to the front roller nip (see
figure 1.2b). The distance between the apron
nose and the front roller nip should be as small
as practically possible to ensure the best fibre
control during drafting.

Figure 1.2b also shows that the front top roller has
a slight overhang (a) relative to the front
bottom roller, while the middle top roller is set a
short distance (b) behind the middle bottom
roller. Such position is found to give smooth
running of the top rollers. In addition, the
overhang of the front top roller shortens the
spinning triangle (figure 1.3b), which tends to
reduce the rate of yarn breakage (ends-down) in
spinning. More on spinning triangle
is discussed in the following section on twisting.

Figure 1.3 Front top roller without overhang (a) and with overhang

Twisting

The essence of staple spinning is


about twist insertion. In ring spinning, twist
is inserted into the thin strand of fibres
emerging from the front roller nip to form the
yarn. During ring spinning, the spindle is
positively driven by a belt or tape at a constant
speed. The traveller is dragged around
the ring by the yarn being wound onto the
bobbin. The rotation of the traveller allows the
yarn between the traveller and the pigtail
guide to rotate at the same
speed. The persistence of vision will give us the
impression of a yarn balloon as the yarn
rotates at a high speed. It is the rotating balloon
that inserts the actual twist into the yarn.

As twist is generated in the yarn balloon, it travels


past the yarn guide towards the front roller nip. But
the twist can not quite reach the nip line of
the front rollers, because the fibres emerging from
the nip have to be diverted inwards to be twisted
around each other. So a small triangle of fibres,
without any twist, is formed between the front roller
nip and the fibre convergence point as shown in
figure 1.3. This triangle is called the spinning
triangle or twist triangle . It is also
known as the yarn formation zone. Because there is
no twist in this zone, it is a weak point and endsdown most often occurs in this region. For this
reason, a large triangle is not desirable. The height
of the spinning triangle is affected by the spinning
geometry and the twist level in the yarn. Overhang
of front top roller and high twist will reduce the
height, hence the level of ends-down in spinning.

Because of air drag on the yarn


balloon and friction between the traveller and
ring, the yarn balloon and the traveller rotate at
a slower speed than the spindle. As we will see
shortly, the balloon speed keeps changing as
spinning continues. Theoretically, we should
use the balloon speed to work out the twist
level in the yarn. But this is obviously difficult
because of the changing speed of the balloon.

In practice, the nominal twist level in the yarn


is calculated using the constant spindle
rotational speed rather than the balloon
speed. The discrepancy arising from this
approximation is quite small.
Spindle rpm (revs per min .)
Twist (turns per metre)
Yarn delivery speed (metres per min .)

(1.1)

The yarn delivery speed is surface speed of


the front rollers. It is also referred to as the
yarn production speed (rate).

Winding-on

As mentioned earlier, the yarn balloon rotates


at a slower speed than the spindle due to air
drag (resistance) on yarn and the friction
between traveller and ring. It is this difference
in rotational speeds of the balloon and the
spindle (bobbin) that allows the yarn to be
wound up onto the bobbin. Without this
difference in rotational speed or if the traveller
and spindle rotate in sync, there will be no
winding of yarn onto the bobbin. In addition,
the linear winding-on speed needs to match
the delivery (surface) speed of the front rollers,
otherwise the yarn will be too taut or too slack
during spinning.

Figure 1.4 shows a cross-sectional view of the


bobbin/ring/traveller assembly, with a yarn
being wound onto the bobbin via the traveller.

Figure 1.4 The cross-section of the bobbin/ring/traveller assembly

If the diameter of the bobbin at the yarn windon point is dwind-on, the linear winding-on
speed (Vwind-on) should equal the
circumference of yarn package ( dwind-on)
multiplied by the difference in traveller and
bobbin rotational speeds (nbobbin - ntraveller),
i.e.

Vwind on d wind on (nbobbin ntraveller )

This wind-on speed should match the speed at


which the fibre strand is delivered by the front
rollers. But as the yarn package builds up, its
circumference changes. With a constant
bobbin rotation speed (nbobbin), the traveller
needs to change its rotation speed (ntraveller)
so that the winding speed (Vwind-on) remains
constant and matches the front roller delivery
speed. The beauty of ring spinning is that the
traveller can self-adjust its rotational speed
during spinning. It's all done automatically by
the traveller itself.

Characteristics

of ring spun yarns


Ring spun yarns are characterised by their good
strength and smoothness , and their
relatively high hairiness .
Ring spun yarns feature near-helical fibre
configuration in the yarn structure. This is a
result of twisting a strand of well-aligned and
parallel fibres. In the whole processing stages
from carding right up to ring spinning,
fibre alignment has been an essential aim. By
having well-aligned fibres within the yarn
structure, the fibres are able to share the stress
in the yarn when a force is applied to stretch the
yarn. This is why ring spun yarns have excellent
strength. The good fibre alignment and the nearhelical configuration of surface fibres also impart
the yarn with a smooth and neat surface.

During ring spinning, the fibres on the two edges of


the spinning triangle must be strongly deflected to
get bound into the yarn at the convergence point
. The deflection is higher with a smaller
triangle (figure 1.3b). Not all fibres will be bound into
the yarn, particularly those fibres with a high rigidity
and low cohesion with
neighbouring fibres. As a consequence, some of
these fibres escape the twisting action and are lost as
fly , others may only get partially bound in with
their remaining parts projecting outside the yarn
surface as hair fibres. This is why ring spinning
produces relatively hairy yarns, especially when
spinning short and coarse fibres. The majority of
these hair fibres have their tail ends sticking out of
the yarn surface. This is because as the tail ends
emerge from the front roll nip, they are no longer
under any positive control and the centrifugal force
from the yarn rotation tends to throw these
uncontrolled tail ends out away from the yarn body,
making them protruding hair fibres.

Figure 1.5 shows a photo of the spinning


triangle and the hair fibres (trailing ends) on
the newly formed yarn.

Figure 1.5 Formation of trailing hairs in the spinning triangle (Wang et al 1999)

The hairiness is a desirable feature of staple


spun yarns. But too much of it can be a costly
nuisance. The latest compact spinning
technology, released at the 1999
international textile machinery exhibition in
Paris, eliminates the spinning triangle all
together by using a modified drafting
arrangement to compact the fibres before twist
is inserted. The compact spun yarns are very
smooth with few protruding fibre ends.

Twist

variation within yarn package


We mentioned in the previous section that as
the yarn package builds up, the traveller
adjusts its rotational speed automatically. This
would suggest that within a yarn package, the
twist level would be different. This is true. But
the difference is only marginal as the following
examples demonstrates.
Assume:
(1) Cop dimensions as in the
diagram below
(2) Front roll delivery speed is 15 m/min
(3) Spindle speed is 10,000 rpm

Since the linear wind-on speed = winding rpm


x circumference of the wind-on point = front
roll delivery speed, we have:
15
Winding rpm at A =
= 191 revs
-2
x 2.5 x 10
15
Winding rpm at B =
= 80 revs
-2
x 6 x 10

Ignoring the effect of up and down movement

of ring rail, we have:


Traveller speed at A = 10,000 - 191 = 9,809
rpm
Traveller speed at B = 10,000 - 80 = 9,920
rpm

9809
Therefore,
Twist at A =
= 654 twists/m,
15
9920
Twist at B
= 662 twists/m
15

The difference in twist is about 1% only. If the


effect of added potential twist due to unwinding
the yarn axially (at the next process) is taken
into account, then at the minimum diameter A,
more twists will be added and at the maximum
diameter B less twists will be added thus
bringing the twist levels more or less equal at
both points. Therefore, the effect of traveller
speed change (and cop build-up) on yarn twist
is very small.

Physics of ring spinning


The physics of ring spinning considers the
various forces acting on the yarn in the balloon,
as well as the forces acting on the traveller.
This consideration is necessary in order to gain
an insight into the nature of ring spinning, and
how different parameters/settings affect the
spinning performance.

Balloon

theory
Consider the section of yarn between the
pigtail and the traveller as shown in figure
1.6.

: angular velocity of balloon;


Tc: centrifugal force = m ds 2 y;
R: ring radius;

Figure 1.6 Forces acting on the balloon

m: yarn mass/unit length

H: balloon height
T: yarn tension

The centrifugal forces associated with the


rotation of the yarn makes this section of the
yarn balloon out to form a yarn balloon. Now
consider a small element, of length ds, in the
balloon. Ignoring the air drag and
other small forces, the only forces acting on
this small element would be the tensions T
T and T+dT at the two ends of the
element, plus the centrifugal force (m ds 2 y)
acting on the element.

In equilibrium, the forces acting on each


element in the balloon should be balanced in
both X and Y directions. So we have
T cos (T + dT ) cos ( d ) 0
T sin Tc (T dT ) sin ( d ) 0
or

d (T cos ) = 0

2
d
(
T
sin

)
=
m
ds
y

If we assume the balloon is slim, i.e. , and yarn


tension (T) is constant and equals To (the
tension at the pigtail guide), we get the
following balloon equation.
R
m 2
y=
sin(
x)
sin(

m 2
TO

H)

TO

(1.2)

It should be stressed that this is the simplified


balloon equation. The detailed derivation need
not concern us here. What we are interested in
is the practical implications of this so called
balloon theory.

The shape of the balloon is sinusoidal, and its


amplitude (A) and wavelength () are:
R

A=

m 2

sin(

2
m
TO

(1.3)
H)

TO
m

(1.4)

A sine curve can contain one or


more points where the curve crosses the axis,
and these crossing points are called nodes
as shown in figure 1.7.

Figure 1.7 Formation of node in yarn balloons

For normal spinning, it is obvious from figure 1.7


(left diagram) that the balloon height
(H) should be less than half the wavelength
(/2). If the balloon height is more than half the
wavelength (i.e. H > /2), a node appears in the
balloon as indicated in figure 1.7 (right diagram).
What this means is that during spinning, the
yarn at the node would always want to be in the
space occupied by the yarn package, so that the
yarn at the node strikes the rotating package
during spinning, resulting in an end-break .
Therefore, this node should be avoided for
normal spinning.

We now know that it is impossible to operate a


ring spinning machine with a multiple-node
balloon (H > /2). However, means
that the yarn package height or size is limited.
Since, to spin yarn onto a large package, H
should be large, this requires a large to avoid a
node. A large is achievable (according to the
equation 1.4) through:
either a large yarn tension which may lead to
yarn breakage;
or a low spindle speed , which usually means
low production rate.

In practice, a BALLOON CONTROL RING


is normally used to restrict balloon expansion, so
that large yarn package may be used without
having to reduce the production or increase the
yarn tension.
Figure 1.8 shows a sketch of the ring spinning
process with a balloon control ring restricting the
balloon size. It should be pointed out though that
the balloon control (restriction) ring tends to hinder
doffing (removal of full package from spindle).
For this reason, balloon control rings are not
universally popular on modern ring frames fitted
with automatic doffing systems. Instead, these ring
frames use very small package sizes to reduce
balloon size and yarn tension.

Figure 1.8 Sketch of ring spinning with a balloon control ring

Forces acting on the traveller


Figure 1.9 shows the different forces acting on a
traveller during spinning.

RT

Tt = yarn tension at traveller; Tw


= yarn wind-on tension
N = normal force between ring and traveller; Fa = air drag on traveller
F = frictional drag between ring and traveller (= N ); mg = traveller
weight
Tc = centripetal forces on traveller = m R
(where m = traveller mass; R = ring radius; = spindle speed)
Figure 1.9 Forces acting on a traveller during spinning

All these forces must be balanced during


spinning. It is possible to derive equations for
this force balance to show the effect of variables
such as traveller weight, balloon size, spindle
speed, yarn count etc on
the spinning tension.
In
equilibrium:
tangential
forces = 0; radial forces = 0; vertical forces = 0
Ignoring Fa and mg, this gives the yarn wind-on
tension ,
2
Tw =

m R

sin
cos +
( cos - sin tan )
RT

(1.5)

Again, the detailed derivation need not concern us


here, and we should focus on the following
implications of this equation.
(1) Wind-on tension increases with the square of the
spindle speed (). Since wind-on tension is directly
related to spinning tension in the yarn above the
pigtail guide, increasing the spindle rotational speed
will drastically increase the yarn tension, which may
lead to increased ends-down. This limits the
maximum spindle speed in ring spinning. If spindle
speed is reduced to reduce yarn tension, the
production rates will drop.
(2)
Winding tension increases as package
diameter decreases. This puts a limit on the
minimum diameter of the empty bobbin.

This is not obvious from equation 1.5. But if we


look at figure 1.9a, we will see that as the yarn
package diameter decreases, the winding angle
or angle of lead () decreases. As the
angle of lead decreases, the tangential
component (T') of the winding tension
(Tw) reduces, since T' = Tw sin. But T' needs to
be sufficient to be able to move the traveller
around the ring during spinning. Therefore, as
the package diameter reduces, the winding
tension (Tw) will have to increase to maintain
the T' required to move the traveller.

Figure 1.9a Winding tension and winding angle as indicated in


figure 1.9

(3)Since larger balloon means higher yarn tension,


the winding tension increases as the balloon gets
longer. This limits the length of the bobbin.
Points (2) & (3) suggest that a yarn is most likely to
break when spinning at the bottom of an empty
bobbin when the winding angle () is the smallest.
This has been confirmed in practice. For this
reason, some machines operate at a lower speed at
the start (since spinning tension increases with the
square of spindle speed) to avoid the ends-down
during start-up.
This also explains the general rule used in practice
that the angle of lead () should not be less than
28O. This minimum angle of lead is equivalent to a
minimum ratio of (empty bobbin or tube
diameter)/(ring diameter) = 0.47.

(4)Winding tension increases as ring radius


(R) increases. This limits the size of the full
package that must fit inside the ring.
Points (2), (3) & (4) mean that there is a limit to
how much yarn can fit onto the yarn package
enclosed by the ring.

The package capacity is


approximately proportional to (ring diameter)2,
so a large ring diameter is desirable for
increased package capacity. But in practice, the
ring diameter is restricted by considerations of
the yarn tension, the minimum angle of lead
previously discussed, as well as other factors
such as power consumption , spindle
rpm, and traveller speed limitations (see
1
relationships
below).
Max. spindle
speed

Ring diameter

Max. linear traveller speed Ring diameter

This is why we do not see ring spinning machines


with very large rings and very small bobbins
(tubes). For coarser and stronger yarns, large ring
radius and yarn packages are used to allow for
more yarns on the package. For finer yarns, both
ring radius and the package size are smaller.
Most worsted yarns are spun using
rings between 45 and 75 mm in diameter at
spindle rotational speeds between 7,000 and
12,000 rpm. Short staple yarns are spun with
smaller rings, but at about twice the spindle
speeds used for worsted yarns.

(5)Winding tension increases with traveller


mass. Heavier yarns require a greater
centripetal force to keep them rotating.
Traveller mass is used as a variable to increase
the tension and generate the higher centripetal
force for heavier yarns. The traveller mass is
usually chosen according to the linear density
of the yarn being spun.
(6)Winding tension increases with an increase
in the frictional coefficient between the ring
and the traveller (RT).

While we talk about the winding tension here,


we should note that the spinning tension in the
yarn is directly related to winding tension, so
any factor that contributes to an increase in
winding tension will also increase the tension in
the yarn during ring spinning. In fact all the
implications for winding tension apply to tension
in the yarn or spinning tension.

Developments and limitations of ring


spinning

As indicated in the beginning of this module,


ring spinning is the most versatile spinning
process. It can spin yarns of a wide range of
counts (from very fine to coarse) from different
types of fibres (short as well as long staple
fibres). The quality of ring spun yarns has been a
benchmark against which the quality of yarns
produced on other spinning systems is judged.

The basic principle of ring spinning has not


changed much since its invention by Thorpe in
1828. But there have been numerous
developments of the ring spinning system,
particularly in the short staple sector. These
developments include:
Automatic doffing of full cops (bobbins)
Linkage to roving frame

Linkage to automatic winding machine

Developments of worsted ring frames are relatively


slow, because the market of worsted spinning is
much smaller than short staple spinning. Therefore
the incentive for manufacturers of worsted ring
frame is not very high.
Over the years, the ring spinning system has also
been modified to improve the properties of ring
spun yarns. Examples of such modifications
include:
Sirospun (see the reading Sirospun - A yarn
with character" by Waldauser)
Compact spinning (see the reading The
Suessen Elite Spinning system for long and short
staple fibres, courtesy of Suessen, Germany)

Ring spinning also has several major limitations.


These limitations include:
High power consumption
Small package size
Low production rate
Staple spinning is basically about twist
insertion. In ring spinning, twist insertion
requires the rotation of the whole yarn package
on the spindle. About 95% of the power used in
ring spinning is consumed by rotating the yarn
package to insert twist. This leads to the
relatively high power consumption for the ring
spinning systems.

The package size is limited in ring spinning due


to the need to reduce the balloon height and
yarn tension, as discussed in the section on the
physics of ring spinning.
Perhaps the most serious limitation of ring
spinning is its low production rate. We already
know the relationship between yarn twist level,
spindle speed, and yarn production or delivery
speed (equation 1.1). According to its end-use, a
ring spun yarn needs to have a certain level of
twist, which is determined before spinning is
started. From equation 1.1, we know that the
only way of speeding up the yarn delivery speed
or production rate is to increase the spindle
rotation speed.

We already know that any increase in spindle


speed will lead to significant increase in yarn
tension, hence the possibility of ends-down. In
addition, with the increase in spindle speed, the
traveller speed increases. This increases the
friction between the traveller and the ring.
Considerable heat is generated because of this
friction, which may result in traveller burning
during spinning. Because of these, spindle
speed can not be increased at will, and yarn
production is limited as a consequence.
Currently, the maximum spindle speed for short
staple ring spinning is about 25,000 rpm, and
that for long staple ring spinning is about
15,000 rpm.

The following example further illustrates this


point.
Suppose a spinning mill produces a
standard 49 tex yarn, with a twist factor of
3500 tpm , on 1,000 spindles operating at
15,000 revolutions per minute and 90%
efficiency. If the mill works 120 hours perweek,
calculate the weight of yarn produced per
week.
Machine production = Spindle production x No. of spindles x Machine efficiency
Note that:
So we need to know the production per spindle
first. This requires the yarn twist level.
Twist
factor of this unit that:
We know from
the first
module
Twist (tpm)
Tex

In this example, we know the twist factor


and yarn count . So the twist level in
the yarn is:
3500
Yarn twist

49

500 (tpm)

Using equation 1.1Spindle


we rpm
have:
15000
Yarn delivery speed (m / min)

twist

500

30 (m / min)

This is the production rate per spindle running at


100% efficiency. Since there are 1,000 spindles
running at 90% efficiency, the total production of
the ring frame will be:
Machine production = 30 m/min x 1,000 x 90% =
27,000 m/min.

Now the machine operates 120 hours a week, so


the weekly machine production will be:
Weekly machine production = 27,000 m/min x
(120 x 60 min) = 194,400,000 m = 194,400 km.
For a 49 tex yarn, each kilo meter of yarn weighs
49 grams. So a total of 194,400 km of yarn
would weigh about 9525600 grams or 9525.6 kg.
In other words, the average production per
spindle is about 9.5 kg per 120 hour week.
If finer yarns are produced, the weight will be
even less. This gives you an idea of the
production rate of ring frames.

Winding and folding after spinning


It is convenient here to discuss some of
the key processes immediately after ring
spinning - winding and folding.

Winding

We already know that ring spun yarns are wound onto small
bobbins or cops during spinning. Each bobbin contains only
a few grams of yarn. For transport, storage and further
processing, the small cops of yarn must be rewound onto
large yarn packages of the right density and structure. If the
yarns are to be dyed, then regular yarn packages of a low
density are necessary for even and good penetration of the
dye liquor. For weaving and knitting , fault free
yarns should be prepared on a large package of high
density. So the first process after ring spinning is yarn
winding (or rewinding). Today automatic winding machines
perform a number of important functions. These include
automatic change of the small yarn cops, automatic yarn
piecing , and yarn clearing . During yarn
clearing, yarn faults such as very thick and very
thin places are removed. Otherwise
these faults may cause problems in weaving or show up in
the final fabrics as defects.

The two basic winding mechanisms are: (a) a


package rotation mechanism
to form coils of yarn on the package;
and (b) a yarn traverse mechanism
to vary the position of wind. Two important
parameters about winding are the wind ratio
and wind angle (or angle of wind). The
wind ratio is defined as the number of yarn coils
wound on a package while the traverse
mechanism completes a full stroke in one
direction. In other words, the wind ratio is the
number of revolutions of yarn package per
traverse stroke. The wind angle is defined as the
angle contained between a coil of yarn on the
surface of a package and the diametrical plane of
the package as indicated in figure 1.10. Increasing
the wind angle will increase package stability but
reduce the package density.

Figure 1.10 Angle of wind and angle of crossing

Figure 1.11shows the principle of random cross winding


or drum winding . A
grooved drum is used as both package
rotation mechanism and yarn traverse mechanism. During
winding, the yarn package rotates via frictional contact
with the surface of the grooved drum, which is driven at a
constant speed. Because the grooved drum rotates at a
constant speed, the linear speed at which the yarn is
wound onto the package is also constant. As winding
continues, the package diameter grows, and the
rotational speed (rpm) of the package decreases to
maintain a constant yarn linear speed (linear speed =
package rpm x package circumference). As a result, the
wind ratio changes as the package builds up, but the
angle of wind remains constant (i.e. package density is
constant). When the wind ratio becomes an integer or half
integer, each succeeding wrap of yarn is laid exactly on
top of the preceding wrap, and a ribbon forms until the
wind ratio assumes a value that is sufficiently different
from the integer or half integer. This is the so called
ribboning effect in random cross winding.

Figure 1.12 shows a cone package


with ribbons. Ribbons are a
serious problem because they interfere with
smooth unwinding of the package, cause
localised abrasion of yarns in the ribbon, and
change the density of the package (ribbons are
much denser than the rest of the package). For
this reason, many random cross winders are
fitted with ribbon breaking devices to prevent
ribbon formation. These devices may oscillate
the yarn package or the grooved drum
sideways in a random manner, or introduce
Figure 1.11 Random cross winding (Lord 1981, p.557)
rational speed variations to the yarn package.
Figure 1.12 A cross wound cone package with ribbons

Another form of winding is precision cross


winding, in which the wind ratio is constant but
the angle of wind decreases as the package
builds up. The mechanisms of precision winding
are shown in figure 1.13. Instead of using a
groove drum to drive the package via frictional
contact, the package mandrel
is driven positively by a motor, and a
reciprocating yarn guide is used as the traversing
mechanism. Ribbon formation is avoided by
setting the package rpm so that the wind ratio is
not an integer or half integer. As the package
builds up, the angle of wind decreases so the
package density increases. Precision winding can
Figure 1.13 Precision cross winding (Lord 1981,
be used
to build a very dense package.
p.557)

Assembly winding is the winding of


two or more yarns on to a single package side by
side, without adding any twist. This is usually
done in preparation for the subsequent folding or
twisting operation.
Folding or twisting
Folding is a process of combining two or more
single yarns by twisting. It is also known as
twisting, plying and doubling. The resultant
yarn is a folded yarn , also known as plied
yarn or doubled yarn. Two-folding is a typical
process in the worsted industry.
You may wonder why it is necessary to twist
together two single yarns to make a two-fold
yarn. This is because a two-fold yarn has a
number of distinct advantages over its single
components, including:

(1)A balanced yarn can be produced


Single yarns are twist lively. In other words,
they always untwist when there is an
opportunity. The twist liveliness may lead to
snarling when the tension in
the yarn is insufficient. Twist liveliness may
also need to distortion of the resultant fabrics,
knitted fabrics in particular. However, if two
single yarns are combined and twisted
together in reverse direction, a balanced
folded yarn can be obtained.

(2)Improved abrasion resistance


As we have discussed already, ring spun
yarns are relatively hairy. The hair fibres often
cause problems in subsequent processes such
as weaving. Also in weaving, warp yarns are
subject to repeated abrasion and fibres may be
gradually rubbed away, leading to yarn
breakage. When two single yarns are twisted
together, their surface fibres are trapped
between the two single yarns. This improves
the abrasion resistance.

(3)Increased strength
The two fold yarn is stronger than its single
components.
(4)Reduced irregularity
The doubling reduces the irregularity
according to the law of doubling discussed
in the module on yarn evenne

Two-for-one twisting has been the


common method of producing a folded yarn.
Figure 1.14 shows a schematic of the two-for-one
twisting process. An assembly wound package
(i.e. two yarns assembled onto one
package without any twist) is usually used as the
stationary supply package. The supply yarn is
threaded through a guide mounted on a freely
rotating flyer and then passes through the hollow
rotating spindle. At the base of the spindle, the
yarn comes out forming a balloon, and then goes
onto the winding head via the yarn guide. Each
rotation of the spindle will insert one turn of twist
in the length of yarn within the spindle, plus
another turn of twist in the yarn balloon. As a
result, two turns of twist are inserted into the yarn
for each rotation of the spindle, hence the name
Figure 1.14 Two-for-one
twisting (Grosberg and Iype 1999, p.20)
two-for-one
twisting.

Review questions
1. Based on the discussion in this topic, sketch the appearance of a
typical ring spun yarn.
2. Rovings of 500 tex are used to feed a ring frame with 1000
spindles running at 20,000 rpm and 90% efficiency. A spinning
draft of 20 is used to produce the ring spun yarn. If 54
kilograms of yarn are produced each hour on the machine,
calculate the twist level in the yarn (t.p.m) and its twist factor
(t.p.m.). Include details of your calculation.
3. One of the limitations of ring spinning is the relatively small
amount of yarn on a full bobbin. Explain why we can not simply
increase the ring radius and use a small empty bobbin to allow
for a large quantity of yarn to be wound onto the bobbin before
the full bobbin is doffed.
4. Explain, with the help of sketch, the principle of Sirospun and 2for-1 twisting.

Rotor spinning
Introduction

Rotor spinning belongs to the family of openend (OE) spinning . Open-end


spinning systems are designed to overcome some
of the problems associated with ring spinning. As
discussed in the previous topic, twist insertion in
ring spinning requires the rotation of the whole
yarn package. In open-end spinning, only an end
of the yarn is rotated to insert twist, which
consumes much less energy than rotating a yarn
package. The most successful examples of the
open-end spinning concept are the rotor spinning
and friction spinning systems. This topic
discusses rotor spinning. Friction spinning will be
discussed in the next topic.

Objectives
At the end of this topic you should be able to:
Understand the basic concept of open-end
(OE) spinning
Know the principle of rotor spinning
Understand the differences between ring
spinning and rotor spinning

General concept of open-end


spinning

Open-end spinning is a relatively new


concept of spinning. The basic principle
of open-end spinning is illustrated in
figure 2.1 (Lord 1981, p.96).

Figure 2.1 The principle of open-end spinning (Lord


1981, p.96)

Like ring spinning, open-end spinning


involves the three basic steps of
drafting, twisting and winding-on.

Drafting

Very high draft is used to attenuate the


feed sliver (not roving) into individual fibres.
Such a high draft is usually by means of pinned
drafting (with toothed rollers) rather than by
roller drafting.
Because of the direct sliver feed, there is no
need to convert the sliver into roving first before
spinning, which is necessary in conventional
ring spinning.

Twisting

The individual fibres are collected at the yarn


open-end and twist is then inserted at the yarn
open-end.
Since only the yarn opne-end is rotated to
insert twist, open-end spinning is much more
energy efficient than ring spinning, which
requires the rotation of a massive yarn package
to insert twist. In addition, the twist insertion
rate in open-end spinning can be very fast. For
a given yarn twist level, this translates into fast
yarn delivery speed or high production rate.

Winding-on

In open-end spinning, twisting and winding are


separate operations so that yarns can be
wound onto a large yarn package.
In ring spinning, the package size is
restricted and the yarns from the many small
packages need to be joined up to make up a
large package.

In summary, open-end spinning has the following


major advantages compared to ring spinning:
elimination

of roving stage
high productivity and low energy consumption
large package size
Now that we know the basic principle of open-end
spinning and its advantages, we can proceed to
discuss the details of rotor spinning. As mentioned
in the introduction, rotor spinning is a successful
example of the open-end spinning concept.
We start with a brief account of the history of rotor
spinning.

Historical perspectives of rotor spinning


Compared with ring spinning, rotor spinning is a
relatively new spinning technology that has not yet
reached its maturity. A brief chronology of rotor
spinning developments is listed below:
1937

The first idea and basic rotor patented by


Berthelsen (Denmark).
1951 Meimberg (Germany) developed the invention
further and built the first spinning models.
1965 Rohlena and his group (Czechoslovakia) found
the correct combination of spinning elements and
showed the first commercially functional units in
Brunn.

1967

The Czech firm ELITEX exhibited its rotor


spinning machine (BD200) near the
international textile machinery exhibition (ITMA)
in Basel, Switzerland. The machine had a rotor
diameter of 75 mm, a rotating speed of 25,000
rpm, and a high twist multiplier (TM=6, or a
twist factor of about 5740 tpm.).
1970 First sale of BD200 in the West
1971 Invention of the twin disk rotor drive
allowing higher rotor speed (Suessen). With
smaller rotors (60 mm), the rotor speed
increased to 35,000 rpm.

1978 Introduction of 40-50 mm diameter rotors,


improved spinbox geometries, lower yarn twist
possible, first automatic yarn piecer and
package doffer fitted on the rotor spinner.
1989 Smaller rotors with speeds of 100,000 rpm.
1992 Quality yarns as fine as 13-15 tex
produced commercially at rotor speeds up to
120,000 rpm.
1999 Rotor speeds up to 150,000 rpm possible.
The development of rotor spinning technology
continues today. The ultimate aim is to produce
rotor spun yarns that match the quality of
comparable ring spun yarns, but at a fraction of
the cost of ring spinning.

Rotor spinning principle


Figure 2.2 shows the key elements inside the
spin-box of a rotor spinning system. Like
any other staple spinning system, the principle
of rotor spinning entails three basic steps of
drafting, twisting and winding-on.

Figure 2.2 Key elements inside the 'spin-box' of a rotor spinner

Drafting

When the feed sliver enters the spin-box, it is


presented to a toothed combing roll
(or comber roll) by a feed roller and feed
plate assembly. The teeth on the comber
roll comb the fibres in the sliver and because
the surface speed of the comber roll is much
higher than the feed roller, the density of fibres
on the surface of the comber roll is much less
than the density of the sliver. In other words,
drafting is performed by the comber roll on the
incoming sliver. The amount of draft exercised
by the comber roll is very high, and can be
calculated using the equation below:
Draft of comber roll

Surface speed of comber roll (m / min)


Sliver feed speed (m / min)

The housing of the comber roll has an


opening. Because trash and other impurities have
a higher density than fibres, they are ejected
through the opening by centrifugal forces, while
the flexible fibres can bend their way around the
comber roll until they are sucked into the rotor via
a transport tube. The transport tube is
tapered to allow acceleration of fibre flow through
it. This acceleration helps with fibre straightening.
This acceleration also means some fibre drafting is
performed by the transport tube. This draft is
usually quite small.
For a uniform flow of individual fibres into the rotor,
the feed sliver should have good uniformity in
linear density (e.g. a Uster CV% between 2.5 and
3.5), and the comber roll should be in good
condition.

Twisting

Once inside the gyrating rotor, the individual fibres


are thrown against the inner wall of the rotor
by the centrifugal force. The fibres slide
down the wall into the vee-shaped rotor groove as
shown in figure 2.3. As the rotor rotates, many
layers of fibre are collected around the rotor
groove to make up sufficient linear density for the
final yarn. This is the important doubling effect
in rotor spinning. The doubling or layering
of fibres tends to even out any short term
irregularities in the yarn, which makes the rotor
spun yarn surprisingly even. In contrast, there is
no doubling in ring spinning, the roving is
attenuated to yarn linear density during ring
spinning.

Once a sufficient number of fibres is collected in the rotor


groove, the fibres need to be taken out continuously
otherwise they will soon clog the rotor groove. To do this, a
seed yarn is first introduced into the rotor
through the navel (figure 2.3). Again, the centrifugal force
throws the seed yarn into the rotor groove. As the rotor is
rotating rapidly, the seed yarn rotates with it. This rotation
traps the loose fibres at the end of the seed yarn. At this
point, the seed yarn is pulled out, the fibres trapped to the
yarn end are peeled off the rotor groove by the outgoing
yarn. Since the peel-off point is rotating with the
rotor, twist is inserted into the out-going fibres.
Furthermore, the twist at the peeling point
extends a distance inside the groove to form the binding
zone or twist zone as shown in figure 2.4. Within this zone,
new yarn is formed. The stability of rotor spinning is
affected by the length of this twist zone, which is in turn a
function of the amount of twist at a given rotor speed and
the additional false twist induced by the navel
. Before we talk about the false twist, let us see how the
actual (real) twist in a rotor spun yarn is calculated.

Figure 2.3 A look inside the rotor

Figure 2.4 Formation of a twist zone inside the rotor groove (Deussen1993,
p.24)

We now know that the rotation of the peeling-off


point inside the rotor groove inserts twist into the
fibres to form a rotor spun yarn. The peeling-off
point rotates with the rotor at a very high rotational
speed (i.e. over 100,000 rpm). In addition, by
continuously withdrawing the newly formed yarn
from the rotor at the yarn delivery speed, the
peeling-off point also moves relative to and in the
same direction as the rotor, at the same speed as
the yarn withdrawal or delivery speed. In other
words, the real speed of the peeling-off point is
actually slightly faster than the rotor speed, by an
amount equal to the yarn delivery speed. But the
additional twist from this will be very small
compared with the twist from the rotor rotation.
The following example will demonstrate this point.

Suppose a rotor yarn is produced at 150 m/min by a


rotor at 120,000 rpm, and the diameter of inner
groove of the rotor is 30 mm. We wish to find out
the actual twist put into the yarn by the peeling-off
point inside the rotor groove.

Relative to the rotor, the peeling-off point travels at


the yarn delivery speed of 150 m/min. The
circumference of the rotor groove is 30 x = 94.2
mm (or 0.0942 m). Therefore, the rotational speed
of the peeling-off point (POP) relative to the rotor is:
rpm of POP (relative to rotor )

Linear speed of POP


150

1592 ( rpm)
Circumfere nce of rotor groove 0.0942

This rpm is about 1.3% of the rotor RPM, and the


twist put into the yarn from this additional source
will be about 1.3% of the twist due to rotor rotation.

This example shows that for practical purpose,


the twist in a rotor spun yarn can be calculated
from the rotor rpm and yarn delivery speed,
using the equation below.
Rotor rpm
Yarn twist (tpm)
Yarn delivery speed (m / min)

This will be the real theoretical twist in the


rotor spun yarn. What about the false twist we
mentioned a little earlier? This is briefly
discussed in the following paragraph.

The trumpet shaped navel is usually


made of wear-resistant ceramic material. The newly
formed rotor yarn is withdrawn through the stationary
navel to the winding mechanism. As indicated in figure
2.3, the yarn path is deflected 90O at the navel. The
tension in the yarn pushes the yarn against the navel
inner surface. Because of the friction between the
rotating yarn and the navel inner surface, the yarn rolls
on the navel surface. This rolling action produces a false
twist in the section of the yarn inside the rotor and this
false twist is in addition to the real twist from the rotor
rotation. Because of this additional twist, the length of
the binding zone or twist zone (figure 2.4) is increased,
which increases the stability of spinning. The yarn
(inside the rotor) is also stronger due to the additional
twist, which reduces the ends-down. Therefore, the
navel induced false twist plays an important role in rotor
spinning. It should be noted that the false twist does not
show up in the final yarn, which has real twist only. Also,
rolling of the yarn against the navel surface tends to
increase yarn hairiness.

Winding

The yarn withdrawn from the rotor is wound


onto a large yarn package, ready for use. Unlike
in ring spinning, twisting and winding functions
are divorced and this permits the use of large
package size. Figure 2.5 shows the key
elements of the winding mechanism. The yarn
package sits on a winding drum. The winding
drum is positively driven and friction contact
between the yarn package and the winding
drum drives the yarn package. A reciprocation
yarn guide ensures that the yarn is laid across
the package traverse.

Figure 2.5 Yarn winding mechanism


(Deussen 1993, p.12)
Figure 2.6 shows the whole process of rotor
spinning from sliver feed to yarn package.
Figure 2.6 Rotor spinning process (Deussen
1993, p. 6)

Characteristics

of rotor yarn
Ideally, fibres should be incorporated by twist into
the yarn in a helical configuration. In rotor
spinning, this is only possible if the fibres are laid
parallel inside the rotor groove away from the twist
zone and the peeling-off point (figure 2.4).
However, during the course of spinning, it is
unavoidable that some fibres actually land in the
twist zone or on the yarn catenary between the
peeling-off point and the navel. When this
happens, the fibres get wrapped tightly around the
already formed yarn and become the characteristic
wrapper fibres on the yarn surface. The
formation of a wrapper from a fibre landing in the
Figure 2.7
Formation
of wrapper fibres
rotor spinning
(Deussen 1993, p.24)
twist
zone
is depicted
inin figure
2.7.

While a longer twist zone makes spinning more


stable, it also increases the chance of wrapper
formation. The percentage of wrapper
formation can be approximated by the formula
below.
Mean fibre length (mm)
Percentage of wrappers
100
rotor diameter (mm)

From this formula, it is clear that long fibres and


small rotors will increase the chance of wrapper
formation. For this reason, rotor spinning has
been primarily used for short staple fibres such
as cotton and cotton blends. With long staple
fibres, a large rotor is necessary to reduce the
wrapper fibres and a slow rotor rpm has to be
used, this makes spinning uneconomical. This is
one of the reasons why rotor spinning has not
been successful in the long staple spinning
sector. In recent years, rotor technology has
been used to spin fine and short wool fibres.

Increasing the twist will also increase the


number of wrapper fibres. This is because an
increase in yarn twist will increase the length of
twist zone inside the rotor groove, causing the
newly arrived fibres to be held on the otherside
of the already twisted section and thereby
increasing the number of wrapper fibres.
Because the wrapper fibres simply wrap around
the yarn surface, they contribute little to yarn
strength. However, wrapper fibres tend to
increase the abrasion resistance and reduce
the hairiness of rotor spun yarns.

Sometimes, the middle section of a fibre gets caught by


the rotating yarn arm inside the rotor, as indicated in
figure 2.8. When this fibre gets wrapped around the
yarn, its two ends are twisted in opposite directions. So
it is possible to have both S-twist and Z-twist in the
wrapper fibres, even though rotor spun yarn is usually
spun with Z-twist only. Because of the presence of both
S-twist and Z-twist, it is difficult to completely untwist a
rotor spun yarn. The wrapper fibres, and difficulty to
completely untwist the yarn have been used to
differentiate between ring and rotor spun yarns. The
presence of S and Z twists in wrappers also explains the
fact that the measured twist of rotor spun yarns is lower
than the nominal machine twist calculated from the
rotor rpm and yarn delivery speed. The so-called the T
% value has been
usedtwist
to reflect
thetwist
degree of this
Measured
machine
T %
100%
difference,
as indicated
below.
Machine twist

The lower the T% value, the more orderly


and ring-yarn-like the fibre orientation is in the
rotor yarn. High T% values indicate a more
disorderly rotor yarn structure and the
presence of wrapper fibres. Rotor yarns of
100% cotton exhibit T% values between near
0 and -20%, while polyester/cotton blends
measure between -10 and -45% (Deussen
1993, p.27).
Figure 2.8 Formation of a wrapper with two ends twisted in opposite
direction (Nield 1975, p.34)

In ring spinning, fibre alignment is carried out right


up to the point of twist insertion, and fibres are
straight and parallel to each other when twist is
inserted. In rotor spinning, fibre alignment is
achieved to a less extent than in ring spinning. Poor
fibre alignment may occur as fibres impinge on the
sliding wale of the rotor, as they lie in the rotor
groove (unlike in ring spinning, fibres are not under
tension in the twist zone in the rotor groove). This,
together with the formation of wrapper fibres,
means that rotor spun yarns are usually weaker
than comparable ring spun yarns, and a higher
twist plus more fibres in yarn cross section is
necessary to increase the strength of rotor spun
yarns. A minimum of 90 fibres is necessary in rotor
spun yarns, compared with a minimum of 40 for
ring yarns. So in the fine yarn market, rotor
spinning can not compete with ring spinning.

Another important feature of rotor spun yarns


is their good evenness. As mentioned before,
fibres are deposited in the rotor groove in
layers. The number of doublings is quite large,
which tends to even out the short-term
irregularities and improve the evenness of the
resultant yarn. The number of doublings can be
calculated using the equation below:
No. of doublings (in rotor groove )

Yarn twist (tpm) rotor circumfere nce (mm)


1000

Table 2.1 summaries the characteristics of rotor


spun yarns in comparison with ring spun yarns.
Table 2.1 A comparison of rotor and ring yarns
Rotor Yarn Compared to Ring Yarn Reason
_________________________________________________________
Production rate
much higher Higher twist rate
Package size 20 times greater
No ring
Twist level
10-15% higher
Structure
Strength
15-20% weaker
Fibre orientation
Extensibility 10% higher Fibre orientation
Regularity
10-20% better
Doubling in rotor
Handle harsher
Wrapper fibres,
Higher twist
Abrasion
20-30% better
Wrapper fibres
Yarn count
coarser
Weaker
Fibre type
short staple mainly Wrapper fibres
Hairiness
Lower Higher twist,
Wrapper fibres
____________________________________________________________________

This comparison highlights differences made by


the presence of wrapper fibres. The fewer the
number of wrapper fibres in a rotor spun yarn,
the more closely the rotor spun yarn resembles
the structure of ring spun yarns. This has been
the aim that drives further developments in
rotor spinning technology.

Selection

or spinning components and


parameters
The key components for rotor spinning are the
combing roll, the rotor, and the navel. The
combing roll (or comber roll) opens the feed
sliver and individualises the fibres. The rotor is
the twist insertion element, while the navel
adds additional twist inside the rotor and
changes the surface texture of the resultant
yarn.

Figure 2.9 shows the tooth profiles of combing


rolls used for different fibre materials, and the
corresponding speed range.
Figure 2.10 shows the yarn count ranges in
relation to rotor diameters and speeds. Small
rotors of 30 mm diameter (K-230) and 31 mm
diameter (G-231) are used primarily for finer
yarns and higher rotor speed than large rotors.
In figure 2.11, the typical twist multipliers (TM)
used for knitting and weaving yarns of different
fibre compositions are given (note that twist
factor = 956.7 x TM).
Finally, an explanation of different navels and
their applications is given in figure 2.12.

Figure 2.9 Tooth profiles and speed ranges of


combing rolls (Deussen 1993, p.55)
Figure 2.10 Yarn count ranges for rotor diameters
and speeds (Deussen 1993, p.57)
Figure 2.11 Twist multipliers (TM) for weaving and
knitting rotor yarns (Deussen 1993, p.60)
Figure 2.12 Different navels and their applications
(Deussen 1993, p.62)

Review
1.Based

questions

on the discussion of the characteristics


of rotor spun yarns, draw a sketch depicting the
key features of a typical rotor spun yarn.
2.This question relates to a yarn being spun on
a rotor spinner under the following conditions:
input sliver 5.0 kTex
output yarn 50 tex
diameter of combing roll 70 mm
speed of combing roll 7,000 r.p.m.
rotor diameter 45 mm
rotor speed 100,000 r.p.
yarn delivery rate
200 m/min.

(a)Calculate the yarn twist level (t.p.m) and the twist factor of the rotor
yarn, ignoring the small additional twist due to rotation of the yarn "peelingoff-point".
(b)If the above yarn is made from fibres all of the same length of 25 mm,
what proportion of fibres will end up with some portion of their length in a
wrapper configuration?
(c)In order to reduce the number of wrapper fibres, would it be more
effective to use fibres which are 5 mm shorter, or use a rotor of 5 mm
greater in diameter? Show your reasoning.
(d)Sliver for the above yarn received 2 drawframe passages between
carding and spinning, with 6 doublings at each passage. Calculate the
number of doublings provided during spinning and the total doublings the
fibre assembly has received since carding.
(e)The fibre transport chute (between beater and rotor) has been designed
so that fibres enter the chute at the speed of the beater surface, and are
accelerated by air so that they leave the chute travelling 50% faster.
Calculate the average number of fibres lying across the chute inlet at any
one time and the corresponding value at chute outlet. (Take the linear
density of the fibres to be 0.25 tex.)
(f)What are the finest yarns (expressed in Tex and English cotton count)
that could be economically spun
from fibres of linear density 0.25 tex on
the rotor spinner. (Hint: consider the minimum nunber of fibres required for
a rotor spun yarn).

5 In about 1,000 words, compare and contrast the


technology of ring spinning and rotor spinning,
with reference to the economics and yarn quality
of each spinning system.

Friction spinning
Introduction
Friction spinning belongs to the family of open-end
spinning. Most patents related to friction spinning
were filed in the 1970s and 1980s, many of which
were from Dr Ernst Fehrer in Austria. Today friction
spinning is almost synonymous with the term
DREF(Dr Ernst Fehrer). It has been used to produce
yarns usually much coarser than ring and rotor spun
yarns at much higher production rate, and the yarns
have been largely used for domestic and industrial
applications.
This topic discusses the principle of friction spinning
in general, followed by a discussion of the DREF 2
and DREF 3 friction spinning systems.

Objectives
At the end of this topic you should be able to:
Explain

the principle of friction spinning


Understand the features of friction spun yarns
Know the differences between DREF 2 and
DREF 3 friction spinning systems

Principle of friction spinning


Friction spinning uses two friction surfaces to
roll up fibres into a yarn. A simplified sketch of
friction spinning is shown in figure 3.1. The
fibres flow freely to two rotating friction drums
(spinning drums, friction rollers, torque rollers).
The surfaces at the nip of the two drums move
in opposite direction to twist the fibres collected
in the nip. The yarn is formed from inside
outwards, by the superimposition of twisting of
individual fibres. The yarn is then withdrawn
from the nip to take-up package.

Figure 3.1 A sketch of friction spinning

In friction spinning, the yarn end in the nip of

the friction drums is tapered, just as the yarn


tail inside the rotor groove is tapered in rotor
spinning. Fibres are added continuously to the
tapered yarn end as the newly formed yarn is
withdrawn. This is illustrated in figure 3.2.

Figure 3.2 A tapered yarn end in the nip of the spinning drums

It can be envisaged that fibres deposited at


the thin end of the taper will end up in the
interior of the final yarn, while fibre deposited
at the thin end will stay on the surface.
The twisting rate in friction spinning is related
to the drum rpm, drum diameter and yarn
diameter as indicated below.
Drum diameter
Twist (tpm) Drum rpm
Twisting efficiency
Yarn diameter

Because of the very large ratio between the drum and


yarn diameters, the rotational speed of the drums
need not be high, provided adequate twist efficiency
is achieved. The twist efficiency is reduced due to the
slippage between the yarn in the nip and the drum
surfaces. It is possible to have a twist efficiency as low
as 40%. But even allowing for this, friction spinning is
still the most efficient way of inserting twist to fibres,
because twist is directly applied to yarn end.
Unlike ring spinning and rotor spinning, friction
spinning imposes very little tension to the yarn. So the
ends-down rate in friction spinning is very low and the
yarn can be withdrawn from the nip of the drums at a
very high speed, say 250 m/min. This makes friction
spinning more productive than ring and rotor spinning.

Similar to rotor spinning, friction spinning uses sliver feed


and tooth drafting. Fibres opened by a toothed roller are
directed towards the nip of the friction drums, at a very high
speed. The fibres should then impinge on the friction surface
that is entering the nip or the rotating mass of fibres in the
nip. Because the velocity of the entering fibres is much
higher than the surface velocity of the drum surface and the
rotating mass of fibres in the nip, fibres are decelerated as
they impinge on the drum surface or the rotating fibres in the
nip. This deceleration causes considerable fibre buckling just
before the fibres are incorporated into the yarn structure. As
a result, the fibre alignment in friction spun yarns is poor,
leading to poor strength of friction spun yarns. Having long
fibres does not help yarn strength much in friction spinning,
because the long fibres buckle more readily than short ones,
so their configurations within the yarn structure may not be
as good as shorter fibres. The poor yarn strength also means
that friction spinning can only produce relatively coarse
yarns.

With friction spinning, a core component can


be easily introduced in the nip to make a
composite yarn of a sheath/core composition.
Examples of this will be discussed in the
following section on DREF friction spinning
systems.
In this section, we have discussed the basic
principle of friction spinning, and the key
features of friction spun yarns. Next, we will
discuss the DREF 2 and DREF 3 friction
spinning systems developed by the Fehrer
company located in Linz, Austria.

DREF 2 friction spinning system


The DREF 2 friction spinning system was
introduced into the world market in 1977. It is
designed for coarse yarn counts in the 100 tex
to 4,000 tex range. The DREF 2 system is
primarily used for the recycling of all types of
textile waste fibres and mixtures with 10 -120
mm fibre lengths, and the spinning of technical
and other yarns for specialised applications,
such as blankets, cleaning rags and mops,
yarns for secondary carpet backings etc.

A diagram of the DREF 2 friction spinning


system is shown in figure 3.3.

Figure 3.3 DREF 2 friction spinning system (Fehrer AG)

As mentioned in the previous section, toothed


drafting is used in friction spinning. With the
DREF 2 system, the feed slivers are opened and
drafted by the teeth of a carding drum. The
individualised fibres are then stripped from the
carding drum by centrifugal force, supported by
an air flow. Gravity and air flow then carry the
fibres into the nip of two perforated spinning
drums. Assisted by air suction through the
spinning drums, the fibres in the nip are twisted
by friction on the two drum surfaces to form the
yarn. The yarn is then withdrawn by the take-up
rollers at delivery speeds of up to 250 m/min,
and wound onto a large yarn package.

A filament core can be easily introduced into


the nip of the spinning drums via the core
feeding (figure 3.3), to make a composite yarn
of a sheath/core structure. During spinning, the
filament core gets false twisted by the spinning
drums, while the staple fibres are deposited on
the false twisted filament to make a sheath. The
staple fibres are twisted as usual. But as the
filament core emerges from the nip of the
spinning drums, the false twist in it is removed
automatically, and the sheath fibres receive a
reserve twist in the process. The resultant
composite yarn has the characteristics of a
twistless filament core surrounded by a sheath
of helical wound fibres of varying helix angles.

The core/sheath effect can also be achieved


without the filament component. As indicated in
the previous section, the yarn in the nip of the
spinning drums has a tapered end, and fibres
deposited in the thin end of the taper are likely
to end up in the core position of the resultant
yarn. For example, fibres in the left-most card
sliver in figure 3.3 are likely to stay as core fibres
in the yarn, surrounded by sheath fibres from the
remaining two card slivers. This preferential fibre
arrangement facilitates an economic use of a
variety of raw materials. A 'core' sliver of waste
fibres may be used with other 'sheath' slivers of
high quality virgin fibres to make a quality yarn
with reduced raw material cost.

It is worth mentioning that even if a


filament core is not to be used as part of
the final yarn, a filament is often used to
help start the spinning process. Once
started, the filament is then cut to allow
the process to continue without it. This
also applies to the DREF 3 friction
spinning system that is discussed next.

DREF 3 friction spinning system


After the introduction of DREF 2 into the world
market in 1977, Dr Ernst Fehrer began work on
the development of the DREF 3 friction
spinning system, which was first presented to
the public at the 1979 international textile
machinery exhibition (ITMA'79) in Hanover. In
1981, DREF 3 entered the global textile
machinery market. DREF 3 is designed for the
manufacture of multi-component yarns in the
medium count range (25 - 667 tex). The yarns
have been used in a wide range of industrial
applications, including fire-resistant protective
clothing, aircraft and contract carpeting,
conveyor and transport belts, composites for
Figure
3.4 shows
a diagram of the DREF 3 friction
spinning system.
aviation
and
automotive
industries
etc.
Figure 3.4 DREF 3 friction spinning system (Fehrer AG)

There are a number differences between DREF


2 and DREF 3. First of all, toothed drafted is
achieved by two toothed drums in DREF 3
rather than just one carding drum in DREF 2.
Second, another roller drafting unit (Drafting
unit I in figure 3.4) is now added in DREF 3.
This drafting unit will deliver parallel fibres that
will form a core of parallel fibres in the final
yarn, surrounded by the sheath fibres from
drafting unit II (figure 3.4). There is also the
option for the introduction of a filament core as
in DREF 2. Therefore, composite yarns of three
different components can be engineered on
the DREF 3 system.

Figure 3.5 shows a side view of the DREF 3


system.

Figure 3.5 A side view showing the principle of DREF 3

For good doubling effect , fibres


slivers are used as feed slivers. The overall
density of the feed slivers is very high, which
means they have to be fed to the toothed
drafting rollers (carding drums) at a very slow
speed. To minimise fibre damage, the distance
between the clamping line of the last pair of
feed rollers and the line of the narrowest
clearance between the two toothed drafting
rollers is set at about one fibre length. In
addition, this distance is adjustable to cater
for raw materials of different fibre length.

As with DREF 2, fibre buckling occurs


as the individual fibres impinge upon the surface
of the spinning drum and the mass of fibres
already in the nip of the two spinning drums.
This leads to poor fibre orientation in the yarn,
which reduces yarn strength. But in DREF 3, the
core fibres are of parallel configuration. So DREF
3 yarns should have higher tenacity than DREF 2
yarns under similar conditions.
The percentage of core/sheath components can
be easily adjusted with the DREF friction
spinning systems.

Review

questions

1.Describe in your own words the principle of


friction spinning.
2.Friction spun yarns do not have wrapper
fibres on the yarn surface. Why are friction
spun yarns weaker than comparable rotor
spun yarns?
3.Compare and contrast the differences and
similarities between DREF 2 and DREF 3
friction spinning systems.

Air jet spinning


Introduction
In the early 1960s, the DuPont company (USA)
patented a method for producing what was called a
fasciated yarn , which is composed of a core
of more or less parallel fibres, wrapped around by a
small proportion of surface fibres. A false twist air
jet was used as the twister, hence the name air jet
spinning or air vortex spinning. Because the speed
of air vortex can be extremely high, the twisting
rate is very high with air jet spinning, which then
leads to high yarn delivery speed or high
production rate. Today, air jet spinning is almost
the most productive staple spinning technologies.

DuPont did not pursue its patented method to


commercialisation. Instead, further research
and development on air jet spinning were
carried by other companies. Since the 1980s,
the name most widely associated with air jet
spinning has been Murata, Japan.
This topic discusses DuPont's basic idea of air
jet spinning to make fasciated yarns. This is
followed by a discussion of Murata Jet Spinning
(MJS) and the associated development

Objectives
At the end of this topic you should be able to:
Understand

the basic principle of the Murata


Jet Spinning system
Appreciate the characteristics of air jet spun
yarns
Know the developments in air jet spinning

Principle of air jet spinning


As mentioned in the introduction, air jet
spinning makes a fasciated yarn with a parallel
core wrapped around by some surface fibres.
The original DuPont process uses one air jet
only. Figure 4.1 shows the principle of the
DuPont method, together with a comparison of
fasciated yarn versus conventional ring spun
yarn.
Figure 4.1 Schematic diagram of the fasciated yarn spinning system (Hunter 1978, p.16).

The feed sliver is drafted by a roller drafting unit (now


shown in full). The drafted fibres are presented as a flat
bundle to the aspirator and then pass the air jet twister
(torque jet). Because the fibre strand is nipped
between the delivery rollers (on the right) and the front
drafting rollers (on the left), only false twist is inserted
into the fibre strand by the air jet twister. At the air jet
twister, the main bundle of fibres are false twisted, but
some fibres at the edges of the fibre ribbon will escape
the twisting effect to some extent.

As soon as the fibres emerge from the air jet,


the main bundle of fibres will untwist to
cancel out the false twist in the bundle.
Because of the increased fibre contact with
the main bundle, the edge fibres will also
'untwist' with the main bundle, and the
amount of untwisting is greater than the
initial false twist these edge fibres received
from the air jet. As a result, the net result is
that the edge fibres will be given a real twist
in the opposite direction to that of the original
false twist. This difference in twist direction is
also depicted in figure 4.1.

An important feature of this process the high


rate of twist, leading to much higher rate of
yarn delivery than the ring and rotor spinning
systems. Roving stage is also eliminated
because the jet spinning systems can spin
directly from slivers. Jet spun yarns are usually
weaker than comparable ring spun yarns.
The jet spinning process relies heavily on the
clever manipulation of the edge fibres which
refuse to receive the full initial false twist. This
aspect is further developed by Murata in its
own jet spinning systems. The Murata Jet
Spinning concept is discussed next.

Murata jet spinning (MJS)


The Murata jet spinning (MJS) system uses two
air jets rather than one. The direction of air
vortex of the two jets is opposite. The first jet is
employed specifically to manipulate the edge
fibres while the second jet (main jet) is used as
the false twist jet. Figure 4.2 shows a schematic
of the Murata jet spinning system.
Figure 4.2 Schematic of Murata jet spinning (MJS) process
(Murata).

A drawframe sliver is fed from can (1) to a


roller drafting unit (2). A high draft in the
range 100-200 is used to attenuate the sliver
to a thin strand of fibres. The fibre strand
then proceeds to the two air jets (3 and 4).
The direction of air vortex in these two jets is
opposite to each other. The intensity of air
vortex in these jet is also quite different, with
the second jet (4) having an air vortex of
much higher intensity. The angular velocity
of the air vortex inside the 2nd jet (4) is more
than 1 million rpm. Owing to the intensity of
its air vortex, the 2nd jet (4) is the actual false
twist jet and will affect the main bundle of
fibres from the drafting unit.

The first jet (3), on the other hand, will affect the small
number of edge fibres. Because the air vortex inside the
1st jet (3) rotates in the opposite direction to that of the
2nd jet, the edge fibres are twisted by the 1 st jet in the
opposite direction to the main fibre bundle. As soon as
the main fibre bundle and the edge fibres emerge from
the 2nd jet (4), the main bundle untwists to cancel out
the false twist it received from the 2 nd jet. In the same
process, the edge fibres also 'untwist' with the main
bundle. Because the direction of 'untwisting' is the same
as the direction of the twist these edge fibres received
from the 1st jet, the edge fibres actually receive a boost
of real twist, allowing them to wrap tightly around the
now parallel main fibre bundle. The distribution of twist
in the fibre strand is depicted in figure 4.3.
Figure 4.3 Distribution of twist in the whole fibre strand (Klein 1993, p.14)

Since the installation of the 1st Murata jet


spinner in USA, this technology has achieved
considerable market penetration. Today, it is
used to spin yarns in the count range Ne 10 to
Ne 80 (7.5 to 59 tex), at a yarn delivery speed
of about 200 m/min. Fibres used include 100%
cotton, synthetics, and their blends (fibre
length up to 2 inches).
Compared with ring spun yarns, the Murata air
jet spun yarns are usually weaker, stiffer and
harder, but they have a lower tendency to
pilling and snarling.

Developments in air jet spinning


Two major developments in this field are from
Murata - the Murata Twin Spinner (MTS) and
the Roller Jet Spinner (RJS).

Murata Twin Spinner (MTS)


The Murata twin spinner produces an
assembly wound air jet spun yarns directly on
the spinning frame .

Figure 4.4 shows the process of Murata twin


spinner. The principle of yarn formation is
exactly the same the Murata jet spinning
discussed in the previous section. Rather than
making one air jet spun yarn, two single air jet
spun yarns are produced from two feed slivers.
Upon emerging from the air jets, these two
single yarns are brought together, cleared if
necessary by a yarn clearing device, and
assembly wound onto a large package. The
assembly wound package is usually twisted
with a two-for-one twister to form a quality 24.4or
Murata
Twin Spinner
(Murata)
foldedFigure
yarn
doubled
yarn.

Roller Jet Spinner (RJS)


This is a relatively new development. Again the
basic principle is similar to Murata jet spinning.
But only one jet is used, the 2nd false twist jet
is replaced by a pair of twisting rollers, called
the balloon rollers. The rotation of the balloon
rollers not only inserts false twist to the main
fibre bundle, they also drive forward the yarn at
a predetermined delivery speed. A schematic
diagram of the balloon rollers is shown in figure
4.5.

Figure 4.5 The balloon rollers used in RJS

The roller jet spinner (RJS) can achieve a yarn


delivery speed of bout 400 m/min, making it
the most productive staple spinning system in
commercial production. A notable feature of
the roller jet spun yarns is the very low yarn
hairiness level. This is because of the rolling of
surface fibres on the yarn by the surfaces of
the balloon rollers.
The reading material "Murata: RJS Spinning for
Coarse Counts" by Murata Machinery Ltd briefs
explains the principle of roller jet spinning, and
compares the relevant yarn and fabric
performances.

Review

questions
1.Rotor spun yarns have wrapper fibres and
Murata jet spun yarns also have wrapping
fibres binding the yarn together. What is the
main difference between the structures of
yarns produced by these two spinning
systems?
2.The following diagrams show the process
flow-chart for making two-folded (two-plied)
ring and rotor spun yarns from cotton fibres.
Please sketch the processing flow-charts for
making two-folded cotton yarns from Murata
Jet Spinner (MJS) and Murata Twin Spinner
(MTS).

3 There are a number of spinning systems


that are not discussed in this module.
These spinning systems, while not as
commonly used as the ones discussed in
this module, are nevertheless worth
knowing. Please consult the relevant
library resources at Deakin and class
handouts, explain (with the help of
sketches) the working principle of each of
the following spinning systems:
(1)Wrap spinning
(2)Self-twist spinning
(3)Mule spinning
(4)Woollen ring spinning

Filament Yarn Texturing


Introduction

Manufactured fibres have been used


increasingly, often at the expenses of a declining
share of natural fibres. Manufactured fibres can
be produced in various forms - tape, monofilament, multi-filament, and tow. Figure 5.1
shows an overview of the production and
subsequent processing of manufactured fibres

Figure 5.1: An overview of manufactured fibre production and processing

Different dies or spinnerets are used to extrude the various forms of


manufactured fibres. For example, to make a tape yarn, the polymer
material is extruded as a thin sheet first, which is then slit into flat and
narrow tape yarns for applications such as carpet backing, sacks and
packing bags. Manufactured fibres can also be extruded as a think bundle of
continuous filaments known as a tow, which is subsequently cut or stretch
broken into staple fibres for processing on their own or in blends with other
fibres on conventional short staple or long staple processing systems. But
most commonly, manufactured fibres are extruded as continuously multifilament yarns. The majority of multi-filament yarns that we use today are
polyester and nylon. In making these yarns, the polymer chip is melt by heat
and extruded through a series of tiny holes in a spinneret. The filaments are
then brought together and drawn to align the molecular chains in each
filament to improve the strength of the resultant multi-filament yarn.
Depending on the level of drawing, the resultant multi-filament can be a
partially oriented yarn (POY) or a fully oriented yarn (FOY). Up to this stage,
the yarn is often referred to as producer filament yarn, smooth filament
yarn, and flat filament yarn. Such a yarn is generally unsuitable for apparel
applications. Owing to the smooth surface of individual filaments, the
filaments are closely packed together, with little bulk and extension. Fabrics
made from these yarns are slippery, have poor breathability and handle. To
overcome these drawbacks, the filaments are either converted into staple
fibres first for processing on conventional wool or cotton processing
machinery, or are textured through a filament yarn texturing process.

For information on converting continuous filaments to


staple fibres, please raed the reading "Tow-to-sliver
conversion and bulked acrylic yarn production" by Oxtoby
(1987, p.232). Please note that "tow-to-sliver" conversion
is also known as "tow-to-top" conversion. Furthermore, a
similar technique to bulked acrylic yarn production has
also been used to produce bulky yarns from wool fibres.
This topic focuses on the texturing of continuous filament
yarns.
Textured yarns in comparison with the original smooth
filament yarn show increased bulkiness, porosity, softness,
and some of them also possess high elasticity. Garments
made from textured yarns are comfortable to wear. They
are noted for good draping qualities. They have good air
permeability and thermal insulation properties. They are
more absorbent and allow less static build-up than the
original smooth filament yarn. Textured yarns may be in
different forms, depending on the texturing process used.

The basic principle of texturing is to introduce


arcs, crimps, or loops into the smooth
structure of continuous filament yarns.
Different processes have been developed for
this purpose. The two commonly used
processes are false-twist texturing and air-jet
texturing.

Objectives
At the end of this topic you should be able to:
Understand

the process of 'tow-to-top'


conversion and the production of bulked acrylic
yarns
Understand the objectives of filament yarn
texturing
Know the principle and process false twist
texturing
Know the principle and process of air jet
texturing

False-twist Texturing
Principle

and process

Filament yarns like polyester and nylon and thermoplastic. They


soften and can be easily deformed with the application of heat; and
upon cooling, the filaments remain in their deformed state (i.e. the
filaments become heat-set). Normally if filaments are twisted,
torsional stress develops in the filaments and the filaments will
want to untwist to release the stress. However, if such filaments are
twisted and heat-set (i.e. heated above their glass transition
temperature and then cooled below their glass transition
temperature), their torsional stress will be relaxed and the
filaments will remain twisted even though the external force is
removed. If these filaments are then untwisted, stress will develop
again in the individual filaments. If these stressed filaments are
allowed to relax, they will seek the minimum-energy-state (or
least stressed state!) by forming adjacent helices and snarls. These
helices and snarls prevent the individual filaments from staying as
closely packed as before and the filaments will occupy a much
greater volume than before. In other words, the filaments become
bulky and textured. This process is illustrated in Figure 5.2

Fig. 5.2 Principle of filament texturing by twisting, heat-setting and


untwisting

In the early days of filament yarn texturing, the


three steps of twisting, heat-setting, and untwisting
are carried out in separate stages as indicated in
figure 5.3. This traditional process has now become
obsolete. But the principle has been used in modern
continuous false twist (FT) texturing processes.

Figure 5.3: Traditional filament yarn texturing with separate twisting,


heat-setting, and untwisting stages

In principle, continuous false-twist (FT) texturing is


very similar to a dynamic false-twisting process, as
indicated in figure 5.4. The main difference is that
in FT texturing, heating setting is involved. This is
also why FT texturing can only work on
thermoplastic filaments. Non-thermoplastic
filaments, such as glass filaments, can not be falsetwist textured.

Fig. 5.4 Difference between stationary false twisting, dynamic false


twisting, and false-twist texturing

A typical false-twist texturing process is shown in Figure 5.5.

Without the second heating process, the


textured yarn is called a stretch yarn, which is
bulky and stretchy. The stretch yarn is
suitable for hosiery and sportswear (tracksuiting, stretch pants, and swimwear etc). If
less stretch is required, the yarn can pass
through a 2nd heater under controlled
tension. This results in a modified stretch (or
set) yarn, which retains certain bulk, but with
much reduced extensibility, suitable for
outwear applications.

Figure 5.6 shows the possible processing routes from filament


extrusion (spinning) to textured yarn.

1 - spinning,

2 - drawing,

3-

texturing
Route A: 3-step process (Spin undrawn yarn conventionally + Draw
+ Texture)
Route B: Spin-Draw + Texture
Route C: Spin + Sequential Draw Texture
Route D: Spin + Simultaneous Draw Texture
Route E: Spin-Draw-Texture (under development)
Figure 5.6 Possible processing routes from extrusion (spinning) to
textured yarn (Hes and Ursiny 1994, p.24)

False

twist devices

A key element of a false-twist texturing


system is the actual false-twister. The false
twister should satisfy the following
requirements:
Grip yarn well to rotate it
Allow yarn to travel through
Insert twist at high speed, and
Easy to thread up

(A) Pin-spindle false twister


This is a hollow spindle with a horizontal pin
made of ceramic or sapphire. The filament is
threaded across the pin. The hollow spindle is
driven via frictional contact with large rolls at
speeds up to 1,000,000 rpm, and each
rotation of the spindle (and the pin) will insert
one turn of twist into the filament. Figure 5.7
shows a pin false-twist spindle.
Figure 5.7: Pin spindle false twister (McIntyre
and Daniels 1995, p. 247)

(B) Stacked disk type false twister


This is the most widely used false twister in
filament yarn texturing. It consists of three sets
of stacked disks mounted on three shafts (figure
5.8), through which the filament yarn runs. The
rotation of the disks not only inserts twist into
the filament, but also drives the filament
through the disks. By setting the surface speed
ratio between the disk and the yarn (the D/Y
ratio) correctly, equal yarn tension can be
achieved at both sides of the false twister. For
easy threading of the filament, one disk shaft is
movable while the other two are fixed.
Figure 5.8 Stacked disc type false twister (McIntyre and
Daniels 1995, p. 144)

(c) Crossed-belts false twister


As shown in figure 5.9, this false twister has
two belts crossed at a specific angle. The
filament yarn is twisted and driven between
two belt surfaces by the rotation of the belts.
It is said to give a soft yarn texturing with
little yarn damage. This type of false twister is
difficult to work with fine denier filaments (eg.
<78 dtex).
Fig. 5.9: A crossed-belts type false twister

(D)Ring false twister (disk-sandwich twister)


This false twister consists of two off-setting
rings or disks rotating in opposite direction,
with the filament yarn running between the
rings. One of the rings is rigid while the other
is flexible. At the yarn input end, a presser
presses on the flexible ring so that the yarn
can be twisted as the rings rotate. Similar to
the crossed-belts, the rotation of the rings
also provides a 'driving force' that drives the
yarn through the rings. Figure 5.10 shows
two diagrams of a ring false twister.
Figure 5.10: A ring false twister (Demir and Behery
1997, p.90)

(E) SZ simultaneous texturing twister


This is a new development, released by
Muratec (Japan) at the 1995 international
textile machinery exhibition (ITMA'95) held in
Milan.
It works on two filaments simultaneously,
inserting S twist in one filament and Z twist in
the other. The two filaments are then
combined and wound onto the same package.
The resultant textured yarn is claimed to have
Figure high
5.11 shows
the SZ twister and
the actual
texturing process using
this type of
bulkiness,
and
is torque-free
(balanced).
twister.
Figure 5.11 A SZ simultaneous twister (a) and the texturing process (b) (Courtesy
of Muratec, Japan)

Air-jet Texturing
Principle

and process

Air-jet texturing is a versatile process. It works


with both thermo-plastic (eg. Polyester and
nylon) and non-thermoplastic (e.g. rayon, glass
filament) filaments. In air-jet texturing, yarn
morphology is modified without disturbing the
internal structure of individual filaments. This is
achieved by creating loops and air pockets in
the yarn by opening up the yarn structure,
buckling the filaments, and locking up the
structure again.

The principle of loop formation in air jet


texturing can be described as:
Overfeed the filament yarn into an air nozzle
Open the feed yarn (or parent yarn) in a
turbulent air stream
The air stream displaces the filaments, and
convert the excess length into loops
interlace filaments to stabilise the loop
structure

Figure 5.12 depicts the principle of air-jet


texturing.

Fig. 5.12: Sketch of an air jet texturing process

Wetting of the filaments before the air nozzle is


used for the following reasons:
to reduce between-filament friction
to reduce friction between filaments and
nozzle wall
to improve separation of filaments
to get better texturing effect with smaller and
more even and frequent loops
A yarn with good textured effect is shown in
figure 5.13 below.

Figure 5.13: An air jet textured yarn with good texturing effect

Figure 5.14 shows example photos of dry and wet textured


yarns, while figure
5.15 shows a series of high-speed still photograph of yarn
being textured under wet conditions.
Fig. 5.14: Photos of dry and wet textured yarns (Demir &
Behery, 1997, p.276)
Figure 5.15: High-speed still photograph of yarn being
textured under wet conditions (Demir & Behery, 1997, p.249)

As can be seen from figure 5.13, air jet


textured yarn closely resembles a spun yarn,
with the protruding loops mimicking surface
hairs of a spun yarn. For this reason, air jet
textured yarns have found applications in a
wide range of products, such as jackets, shirts,
blouses, suits, outwear, furnishing fabrics etc.

Air nozzles
Many different air jet texturing nozzles have
been developed and the development is
continuing.
The reading material "Air-jet texturing: Effect
of jet type and some process parameters on
properties of air-jet textured yarns" by Kothari
and Timble (1991, p.29) gives a good account
on the history of air-jet development, as well
as on the test of air-jet textured yarns.

Other

possibilities with air-jet


texturing
Apart from being a versatile texturing process,
air-jet texturing also offers considerable scope
for engineering quite different yarns.
First of all, the linear density of air jet textured
yarns can be easily changed by changing the
level of overfeed into the air nozzle. A higher
overfeed will lead to a heavier textured yarn.

Secondly, co-texturing is possible by feeding


two or more filaments yarns together. By
having different overfeeds for the different
yarns, a core-and-effect yarn can be produced.
The yarn with a lower overfeed will stay in the
centre as the core while that with a higher
overfeed will stay predominantly on the
surface. Figure 5.16 shows the process of
producing a core-effect yarn using three feed
yarns.
Figure 5.16: Core-effect textured yarn production (Demir & Behery, 1997, p.214)

The loops of air-jet textured yarn can also be


broken after textured with a loop breaker as
shown in figure 5.17. In this process, the air
textured yarn wraps around several rolls in
succession so that protruding loops of the
incoming yarn are rubbed by the outgoing
yarn and thereby broken up. The resultant
yarn is called a Texspun yarn ,
because the free fibre ends of this textured
yarn give the yarn a very spun-like
appearance.

Figure 5.17: A Texspun process

Intermingling/Interlacing

In staple spun yarns, twist is used to hold the


fibres together in the yarn. In multi-filament
continuous yarns, there is very little cohesion
between individual filaments in the yarn, and
filaments separate easily. Even after texturing,
the yarns still lack inter-filament cohesion.
Consequently, the tendency for individual
filaments to separate has caused problems in
subsequent winding and weaving processes.
While twist can be used to impart interfilament cohesion, it is not a very efficient and
is costly. The favoured approach in the
synthetic fibre industry is the intermingling or
interlacing process.

So what is the intermingling process then?


Intermingling is a process of imparting interfilament cohesion by entwining the filaments
instead of or in addition to inserting twist. The
entwining is usually achieved by passing the
yarn under light tension through the turbulent
zone of an intermingling or interlacing jet
(nozzle).
A simple intermingling nozzle is shown in figure
5.18. It consists of a yarn channel, and an air
inlet in the centre of the channel. The
compressed air impinges on the traversing yarn
vertically and entwining the yarn at regular
intervals.

Figure 5.18: Simplified representation of the intermingling process

Intermingling has become a very efficient and


low-cost way of imparting cohesion to multifilaments. It has been used in many fields
where inter-filament cohesion is required.
Figure 5.11(b) shows the use of an interlacing
nozzle in false-twist texturing. The use of
intermingling in other processes is depicted in
figure 5.19.
Figure 5.19: Applications of the intermingling
process (Demir and Behery 1997, p.310)

REVIEW QUESTIONS
1.Compare and contrast false twisting
texturing with air-jet texturing. You should
make reference to differences in filament
input, the texturing process, and the
resultant yarn. You can use sketches to help
explain the points.
2.Briefly describe the objective, principle, and
process of filament intermingling.

Vous aimerez peut-être aussi